10.08.2013 Views

ION KINETICS AT THE HELIOSPHERIC TERMINATION SHOCK PIN ...

ION KINETICS AT THE HELIOSPHERIC TERMINATION SHOCK PIN ...

ION KINETICS AT THE HELIOSPHERIC TERMINATION SHOCK PIN ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

BOSTON UNIVERSITY<br />

GRADU<strong>AT</strong>E SCHOOL OF ARTS AND SCIENCES<br />

Dissertation<br />

<strong>ION</strong> <strong>KINETICS</strong> <strong>AT</strong> <strong>THE</strong> <strong>HELIOSPHERIC</strong> TERMIN<strong>AT</strong><strong>ION</strong> <strong>SHOCK</strong><br />

by<br />

<strong>PIN</strong> WU<br />

B.S., University of Science and Technology of China, 2001<br />

M.A., Boston University, 2005<br />

Submitted in partial fulfillment of the<br />

requirements for the degree of<br />

Doctor of Philosophy<br />

2010


© Copyright by<br />

<strong>PIN</strong> WU<br />

2010


First Reader<br />

Second Reader<br />

Third Reader<br />

Approved by<br />

Nathan A. Schwadron, Ph.D.<br />

Associate Professor of Astronomy<br />

Boston University<br />

W. Jeffrey Hughes, Ph.D.<br />

Professor of Astronomy<br />

Boston University<br />

Dan Winske, Ph.D.<br />

Laboratory Fellow<br />

Los Alamos National Laboratory


Men encourage her to accomplishments yet she remains eternally feminine.<br />

—Chien-Shiung Wu


For my grandparents


Acknowledgements<br />

I am deeply grateful to my advisor Professor Nathan Schwadron, who trains me; to my<br />

mentors Drs. S. Peter Gary and Dan Winske, who coach me; to my advisory committee<br />

Professors W. Jeffrey Hughes, Harlan Spence and Alan Marscher, who guide me; to<br />

everyone in the Astronomy department and everyone in the ISR-1 division of the Los<br />

Alamos National Laboratory, who accompany me; to my friends at home and abroad,<br />

who believe in me; and mostly, to my family, who love me.<br />

My thanks extend to the National Aeronautics and Space Administration (NASA)<br />

Interstellar Boundary Explorer (IBEX) mission, which funds my dissertation at Boston<br />

University; and to the NASA Solar and Heliospheric Physics SR&T Program, which<br />

funds the Los Alamos portion of this dissertation, performed under the auspices of the<br />

U.S. Department of Energy (DOE).<br />

vi


<strong>ION</strong> <strong>KINETICS</strong> <strong>AT</strong> <strong>THE</strong> <strong>HELIOSPHERIC</strong> TERMIN<strong>AT</strong><strong>ION</strong> <strong>SHOCK</strong><br />

(Order No. )<br />

<strong>PIN</strong> WU<br />

Boston University Graduate School of Arts and Sciences, 2010<br />

Major Professor: Nathan A. Schwadron, Associate Professor of Astronomy<br />

ABSTRACT<br />

A number of interesting phenomena were revealed after the two Voyager spacecraft<br />

crossed the termination shock in 2004 and 2007, respectively. This dissertation presents<br />

theoretical and computational studies to address ion heating, energy dissipation and ion<br />

speed distributions at the quasi-perpendicular termination shock. In our model, the<br />

termination shock is distinguished by the presence of a significant fraction of pickup ions,<br />

interstellar atoms which have become ionized and subsequently accelerated up to solar<br />

wind flow speeds. Using the Los Alamos hybrid simulation code, we demonstrate that<br />

the heating of pickup ions is dependent on the phase of gyration about the local magnetic<br />

field when they encounter the termination shock. The temperature of solar wind ions is<br />

raised by a larger factor than that of the pickup ions because some solar wind ions are<br />

specularly reflected. An analytic model for energy partition is developed based on the<br />

Rankine–Hugoniot relations and a polytropic energy equation (the Multicomponent<br />

Rankine–Hugoniot model). The polytropic index γ is varied to improve agreement<br />

between the model and the simulations. When the pickup ion relative density is 0%, the<br />

polytropic index is 5/3. As the pickup ion relative density increases toward 40%, the<br />

vii


polytropic index increases toward 2.2, suggesting a fundamental change in the character<br />

of the shock. We infer that the pickup ion relative density is about 25%, and that pickup<br />

ions gain the larger share (~90%) of the dissipated energy near the nose of the<br />

termination shock, consistent with Voyager 2 observations. Further, we explore the<br />

consequences of four different assumptions regarding upstream pickup ion velocities,<br />

each corresponding to a different thermalization level. The downstream ion speed<br />

distribution is found to be almost identical for each case, with two Maxwellian<br />

components providing a good fit. In addition, the downstream heated ion spectrum scales<br />

with the solar wind speed, with smaller spectral indexes corresponding to faster solar<br />

wind speeds, consistent with the Interstellar Boundary EXplorer (IBEX) inference. The<br />

IBEX team plans to use the constructed spectra to help interpret their measurements.<br />

Finally, we validate our simulations with numerical tests and observations.<br />

viii


Contents<br />

1 Introduction............................................................................................................... 1<br />

1.1 Motivation ........................................................................................................ 1<br />

1.2 Upstream Conditions........................................................................................ 8<br />

1.2.1 Solar Wind .......................................................................................... 8<br />

1.2.2 Pickup ions.......................................................................................... 9<br />

1.3 Observations of the Termination Shock......................................................... 14<br />

1.4 Kinetic Physics of the Termination Shock ..................................................... 20<br />

1.5 Research Background and Goals of this Dissertation .................................... 26<br />

2 Hybrid Simulations for an Idealized Termination Shock ................................... 32<br />

2.1 The Hybrid Simulations ................................................................................. 32<br />

2.1.1 Computer Simulation Methods......................................................... 32<br />

2.1.2 Previous Simulations of the Termination Shock .............................. 35<br />

2.1.3 Methodology of Hybrid Simulations ................................................ 38<br />

2.1.4 The Los Alamos Hybrid Simulation Model...................................... 41<br />

2.2 Simulation Results Overview......................................................................... 47<br />

2.3 Kinetic Structure of the Termination Shock................................................... 51<br />

2.4 Qualitative Examination of Heating via Phase-Space Plots........................... 58<br />

2.5 Heating and Energy Partition ......................................................................... 65<br />

2.6 Pickup Ion Energization ................................................................................. 70<br />

2.6.1 Simulated Pickup Ion Kinetics.......................................................... 71<br />

ix


2.6.2 Gyrophase-Dependent Acceleration Model for Pickup Ions............ 76<br />

2.7 Discussion ...................................................................................................... 82<br />

3 Analytic Model for an Idealized Termination Shock .......................................... 85<br />

3.1 Baseline Analytic Formula............................................................................. 85<br />

3.1.1 Heating of a Transmitted Ion ............................................................ 85<br />

3.1.2 Analysis of Upstream Mach Numbers.............................................. 86<br />

3.2 Multicomponent Rankine–Hugoniot Model .................................................. 87<br />

3.2.1 Modified Compression Ratio............................................................ 87<br />

3.2.2 Component Pressures........................................................................ 88<br />

3.2.3 Reflection Efficiency ........................................................................ 91<br />

3.2.4 Energy Partition ................................................................................ 96<br />

3.2.5 Downstream Mach numbers ............................................................. 99<br />

3.3 Downstream Multi-Ion Speed Distributions ................................................ 101<br />

3.3.1 Dimensional Analysis ..................................................................... 102<br />

3.3.2 Downstream Two-Maxwellian Speed Distribution ........................ 103<br />

3.3.3 Refined Downstream Two-Maxwellian Speed Distribution........... 108<br />

3.4 Discussion .................................................................................................... 110<br />

4 Hybrid Simulations for a More Realistic Termination Shock.......................... 115<br />

4.1 Variation of Pickup Ion Velocity Distributions ........................................... 115<br />

4.1.1 Analytic Derivation of Pickup ion Beta and Speed Range............. 116<br />

4.1.2 Simulated Downstream Speed Distribution.................................... 123<br />

4.2 Variation of Alfvén Mach Number .............................................................. 132<br />

x


4.3 Variation of Shock Normals......................................................................... 134<br />

4.4 Discussion .................................................................................................... 136<br />

5 Accuracy and Convergence.................................................................................. 139<br />

5.1 Cell Size ....................................................................................................... 139<br />

5.1.1 Cases without Pickup Ions.............................................................. 140<br />

5.1.2 Cases with Pickup Ions ................................................................... 143<br />

5.2 Super-Particles per Cell................................................................................ 146<br />

5.3 Discussion .................................................................................................... 149<br />

6 Theory/Model Connection to Observations........................................................ 154<br />

6.1 Comparison with Voyager 2......................................................................... 154<br />

6.1.1 Structure of the Termination Shock................................................ 154<br />

6.1.2 Core Ions......................................................................................... 156<br />

6.2 Implication for IBEX.................................................................................... 158<br />

7 Summary................................................................................................................ 162<br />

7.1 General Overview......................................................................................... 162<br />

7.2 Review of Scientific Context ....................................................................... 163<br />

7.3 Future Work ................................................................................................. 167<br />

List of Journal Abbreviations...................................................................................... 172<br />

References...................................................................................................................... 173<br />

Curriculum Vitae.......................................................................................................... 188<br />

xi


List of Tables<br />

Table 1-1. Voyager 2 Observations in the Upstream Region. .......................................... 15<br />

Table 1-2. Voyager 2 Observations in the Downstream Region. ..................................... 16<br />

Table 1-3. Voyager 2 Observations of the Termination Shock. ....................................... 16<br />

Table 1-4. Voyager 2 Observations of the Foreshock Region.......................................... 16<br />

Table 1-5 Voyager 1 (V1) Observations of the Termination Shock................................. 19<br />

Table 2-1 Results Calculated from the Hybrid Simulations (MA = 8, β sw = 0.05). ......... 67<br />

Table 4-1 Results Calculated from the Hybrid Simulations (MA=8, β sw =0.05, β PUI =8.53)<br />

...................................................................................................................... 125<br />

Table 4-2 Properties of the two downstream Maxwellians from the four simulations<br />

(MA=8, β sw =0.05, β PUI =8.53)..................................................................... 128<br />

Table 4-3 The goodness (R) of the two-Maxwellian and the refined two-Maxwellian fits<br />

for the four simulations (MA=8, β sw =0.05, β PUI =8.53) .............................. 131<br />

Table 4-4 The downstream temperatures of the simulations ( sw<br />

β =0.05, φ =20%<br />

Vasyliunas-Siscoe distributed upstream pickup ions).................................. 133<br />

xii


List of Figures<br />

Figure 1-1 An artist's rendition of the termination shock and its relation to our<br />

heliosphere and the interstellar medium. Adapted from Walt Feimer, National<br />

Aeronautics and Space Administration (NASA).............................................. 2<br />

Figure 1-2 The heliosphere inflated by the solar wind in the region surrounding the Sun<br />

in the interstellar medium. The heliosphere is formed by the supersonic solar<br />

wind as it expands radially from the Sun. At the termination shock, the solar<br />

wind abruptly slows and heats as it turns toward the tail of the heliosphere.<br />

The heliopause is boundary that separates the solar wind flow and the<br />

interstellar flow, where the pressure of the solar ions balances the pressure of<br />

the interstellar ions. In this model, the interstellar wind is also supersonic and<br />

a bow shock forms upstream of the nose of the heliosphere. Adapted by Stone<br />

[2001] from Zank [1999].................................................................................. 4<br />

Figure 1-3 The Parker Spiral. Beyond 20 AU, the interplanetary magnetic field is nearly<br />

toroidal as the solar rotation wraps up the field. Hence, this magnetic field is<br />

nearly perpendicular to the solar wind flow and the termination shock normal.<br />

Figure adapted from Steve Suess, NASA. ..................................................... 11<br />

Figure 1-4 A conceptual illustration of the pickup process, and the subsequent velocity<br />

distribution in the solar wind frame (upper panel) and the spacecraft frame<br />

(bottom panel). The pickup process ends when the pickup ion’s guiding<br />

center moves with the solar wind flow velocity............................................. 12<br />

Figure 1-5 Magnetic field profile from one of the Voyager 2 observed termination shock<br />

crossing. Adapted from: Bulaga et al. [2008]................................................. 15<br />

Figure 1-6 The magnetic field profile of a perpendicular supercritical shock from a hybrid<br />

simulation. The foot, ramp, and overshoot are indicated. From Wu et al.<br />

[1984]. ............................................................................................................ 24<br />

Figure 2-1 A one-dimensional representation of the simulation cells, the mesh points<br />

(center of each cell), and the simulation domain. Cell 1 and Cell NX2 are<br />

ghost cells that are necessary for the specification of boundary conditions.<br />

The simulation domain ranges from Cell 2 to Cell NX1................................ 40<br />

Figure 2-2 The upper panel is a representative setup of the one-dimensional simulation<br />

for the termination shock. The bottom panel shows the time evolution of the<br />

magnetic field B z profile from a 0% pickup ion simulation. Since the<br />

simulation is run in the downstream rest frame, the shock propagates to the<br />

xiii


left. In the bottom panel, x is normalized by c / ω pi where ω pi is the ion<br />

plasma frequency and time t is normalized by i Ω where Ω i the ion cyclotron<br />

frequency based on the upstream magnetic field. From Wu et al. [2009]...... 43<br />

Figure 2-3 Density profile (upper panel) and magnetic field profile (bottom panel) from a<br />

0% pickup ion simulation. The ion density n i is normalized by the upstream<br />

ion density u n ; the magnetic field B z is normalized by the upstream magnetic<br />

field u B ; and x is normalized by / c ω pi . From Wu et al. [2009]................... 49<br />

Figure 2-4 The shock strength (bottom panel) and the downstream flow speed (upper<br />

panel) as a function of pickup ion relative density......................................... 50<br />

Figure 2-5 By and Ez from the 0% (black) and the 20% (blue) pickup ion simulations. By<br />

is normalized to the upstream magnetic field Bu and Ez is normalized by<br />

vABu/c. The simulation dimension x is normalized by ion inertial length<br />

c/ω . ............................................................................................................. 51<br />

pi<br />

Figure 2-6 Profiles of Bz (top panel), Ex (middle panel) and Ey (bottom panel) from the<br />

0% (black) and 20% (blue) pickup ions simulations. The magnetic field is<br />

normalized by the upstream magnetic field Bu; the electric fields are<br />

normalized by vABu/c, and the spatial length x is normalized to the ion inertial<br />

length c/ωpi. .................................................................................................... 53<br />

Figure 2-7 Pickup ion (red), solar wind ion (green) and all physical ions (black) density<br />

and velocity profiles. Panel a, b, c, and d are from the 0% pickup ion<br />

simulation (where pickup ions are test particles). Panel a’, b’, c’ and d’ are<br />

from the 20% pickup ion simulation. In Panel a and a’, all densities are<br />

normalized to the upstream density (nu). In Panel b and b’, each density is<br />

normalized to its upstream value (red: nPUI / n u, PUI , green: nsw / n u, sw,<br />

black:<br />

n/ n u ). The spatial length x is normalized to the ion inertial length c/ωpi. .... 55<br />

Figure 2-8 The phase-space plots of solar wind ions in the 0% pickup ion simulation. The<br />

panels on the top are the x v – x , y v – x and v z – x plots for a) transmitted solar<br />

wind ions and b) reflected solar wind ions. The panels on the bottom are v x<br />

velocity distributions and x v – v y phase-space plots for c) upstream solar wind<br />

ions, d) downstream solar wind ions, e) downstream transmitted solar wind<br />

ions, and f) downstream reflected solar wind ions. The shock is marked by a<br />

dashed line in both Panel a) and b). The solar wind phase-space plot evolves<br />

from the upstream Panel c) into the downstream Panel d) passing the shock.<br />

Empirically, we can separate the downstream population d) into the<br />

xiv


transmitted ions in Panel e) and the reflected ions in Panel f). The velocities<br />

are all normalized by the upstream Alfvén speed v A and x is normalized by<br />

c / ω pi . From Wu et al. [2009]........................................................................ 60<br />

Figure 2-9 The phase-space plots of the 0% pickup ion simulation where pickup ions are<br />

treated as test particles. Solar wind ions are plotted in the panels (a,c,d) on the<br />

left and pickup ions are plotted in the panels (b,e,f) on the right. Upper panels<br />

are magnetic field profile B z –x, x v –x phase-space plots, and v y –x phasespace<br />

plots for a) solar wind ions and b) pickup ions. Lower panels are v x<br />

velocity distributions and x v – v y phase-space plots for c) upstream solar wind<br />

ions, d) downstream solar wind ions, e) upstream pickup ions, and f)<br />

downstream pickup ions. The velocities are all normalized by the upstream<br />

Alfvén speed v A and x is normalized by c / ω pi . From Wu et al. [2009]. ..... 62<br />

Figure 2-10 The phase-space plots of the 20% pickup ions simulation. Solar wind ions<br />

are plotted in the panels (a,c,d) on the left and pickup ions are plotted in the<br />

panels (b,e,f) on the right. Upper panels are magnetic field profile z B –x, v x –x<br />

phase-space plots, and v y –x phase-space plots for a) solar wind ions and b)<br />

pickup ions. Lower panels are v x velocity distributions and x v – v y phasespace<br />

plots for c) upstream solar wind ions, d) downstream solar wind ions, e)<br />

upstream pickup ions, and f) downstream pickup ions. The velocities are all<br />

normalized by the upstream Alfvén speed v A and x is normalized by c / ω pi .<br />

From Wu et al. [2009]. ................................................................................... 64<br />

Figure 2-11 The simulated temperature jumps (blue: solar wind; red: pickup ion) as<br />

compared with adiabatic temperature jumps. The polytropic indexes are<br />

calculated from Equation (2.10). In the 0% pickup ion simulation, pickup ions<br />

are treated as test particles. They are subject to the electricmagnetic fields but<br />

make no contributions to the source terms..................................................... 67<br />

Figure 2-12 Velocity phase-space plots (vx–vy) for pickup ions (left panels) and solar<br />

wind ions (right panels) within 20 c/ωi of the shock front from the 0% pickup<br />

ion simulation. All velocities are normalized by vA....................................... 72<br />

Figure 2-13 Phase space evolution of the pickup ions (first row of the vx–vy panels) and<br />

of the solar wind ions (second row of the vx–vy panels) as they cross the 20%<br />

pickup ion shock. The kinetic shock structure from Section 2.3 is zoomed in<br />

for comparison. Each panel of the evolution series corresponds to the color<br />

marked spatial region of 20 c/ωpi. The center of each adjacent panel is 5 c/ωpi<br />

apart................................................................................................................ 74<br />

xv


Figure 2-14 Pickup ion trajectories from the 20% pickup ions simulation in velocity<br />

space perpendicular to the background magnetic field. Both the x v and v y are<br />

normalized to the upstream Alfvén speed v A . Each panel shows a<br />

representative ion trajectory from time zero (indicated by an asterisk) to the<br />

end of the simulation. The arrows indicate the sense of temporal progress of<br />

each trajectory. From time zero, the color of each trajectory changes from<br />

black through blue, green, yellow, orange and, near the end of the simulation,<br />

to red. From Wu et al. [2009]......................................................................... 77<br />

Figure 2-15 Schematic interpretation of pickup ion energy gain (downstream rest frame).<br />

Left side of the figure corresponds to a representative trajectory of a<br />

“reversed” ion. Right side of the figure corresponds to a representative<br />

trajectory of a “crossing” ion. Both curves correspond to gyromotions with<br />

increasing Lamour radii. The electric field Ex and the electric field Ey in the<br />

shock region (foot, overshoot) are also shown. The downstream oscillation<br />

these fields are not shown. Instread, thick dash lines mark the average<br />

downstream fields........................................................................................... 79<br />

Figure 2-16 Energy distribution histogram of the downstream pickup ions. Blue line: the<br />

distribution for “reversed” ions. Black line: the distribution for “crossing”<br />

ions. The pickup ion energy is normalized to the upstream solar wind<br />

dynamic energy. ............................................................................................. 82<br />

Figure 3-1 Compression ratio r S , pressure jump Pd / P u and downstream pickup ions<br />

PUI<br />

thermal pressure ratio Pd / P d as a function of the pickup ion relative<br />

density φ . Dashed lines are theoretical predictions when γ is set to be 5/3,<br />

the solid lines are theoretical predictions when γ is 2.2. The diamond<br />

symbols mark values from simulations. From Wu et al. [ 2009]. .................. 94<br />

Figure 3-2 Solar wind reflection efficiency ε ref (%) as a function of the pickup ion<br />

relative density φ . The dashed lines correspond to γ =5/3 and the solid lines<br />

correspond to γ =2.2. Adapted from Wu et al. [2009]. ................................. 96<br />

Figure 3-3 Termination shock energy partition η (%) as a function of the pickup ion<br />

relative density φ . The dashed lines correspond to γ =5/3 and the solid lines<br />

correspond to γ =2.2. Red diamond symbols mark the percentage of heating<br />

the pickup ions gain from our simulations. Adapted from Wu et al. [2009].. 98<br />

Figure 3-4 Upstream and downstream Mach numbers as functions of pickup ion relative<br />

density. The three panels on the right are subsets of the same quantities of the<br />

three panels on the left: Alfvén Mach number M A (top panel), sonic Mach<br />

xvi


number cs M (middle panel), Magnetosonic Mach number M MS (bottom<br />

panel). In the right panels, we have a better view of how the solid red lines<br />

differ from the dashed lines. The black lines are the upstream Mach numbers<br />

as a function of the pickup ion relative density φ ; the red lines are the<br />

analytically calculated downstream Mach numbers at γ =2.2 and the red<br />

dashed lines are the analytically calculated downstream Mach numbers at<br />

γ =5/3. From Wu et al. [2009]. .................................................................... 100<br />

Figure 3-5 Downstream solar wind (top panel) and pickup ion (bottom panel) thermal<br />

speeds as functions of the pickup ion relative density and the polytropic index<br />

γ for a MA=8, β sw =0.05 perpendicular termination shock. The vertical line<br />

marks the 20% shell distributed pickup ion case, in which the γ =1.95<br />

solutions predict that the downstream solar wind thermal speed is 1.93 vA and<br />

the downstream pickup ion thermal speed is 12.20 vA, consistent with the<br />

simulation fitted thermal speeds (Chapter 4). .............................................. 106<br />

Figure 3-6 Downstream Maxwellian-1 (top panel) and Maxwellian-2 (bottom panel)<br />

thermals speeds as functions of the shell distributed pickup ion relative<br />

density and the polytropic index γ for a MA=8, β sw =0.05 perpendicular<br />

termination shock. For the 20% pickup ion case, the γ =1.95 solutions predict<br />

that the downstream Maxwellian-1 thermal speed is 1.39 vA and the<br />

downstream Maxwellian-2 thermal speed is 12.26 vA, consistent with the<br />

simulation fitted thermal speeds (Chapter 4). .............................................. 109<br />

Figure 4-1 The distribution function fPUI ( v⊥ ) versus v⊥ for the pickup ion shell (black),<br />

the pickup ion sphere (green), the Vasyliunas–Siscoe distributed pickup ions<br />

(orange), and the Maxwellian pickup ions (red) at the same β PUI =8.53. The<br />

panel on the left is ploted in linear-linear scale and the panel on the right is the<br />

same plot in linear-log scale. The pickup ion’s perpendicular velocity vperp is<br />

normalized to the upstream velocity uu . From Wu et al. [in preparation]... 121<br />

Figure 4-2 The magnetic field profile (Bz) of the four simulations. Black: the simulation<br />

with a pickup ion velocity shell. Green: the simulation with a pickup ion<br />

velocity sphere. Orange: the simulation with the Vasyliunas–Siscoe<br />

distributed pickup ions. Red: the simulation with the Maxwellian distributed<br />

pickup ions. From Wu et al. [in preparation]. .............................................. 125<br />

Figure 4-3 The upstream and the downstream speed distributions in their respective<br />

plasma frames. Black: the simulation with a pickup ion shell. Green: the<br />

simulation with a pickup ion sphere. Orange: the simulation with the<br />

Vasyliunas–Siscoe distributed pickup ions. Red: the simulation with the<br />

xvii


Maxwellian distributed pickup ions. Upper left panel: upstream speed<br />

distribution in log-log scale. Upper right panel: upstream speed distribution in<br />

log-linear scale. Lower left panel: downstream speed distribution in log-log<br />

scale. Lower right panel: downstream speed distribution in log-linear scale.<br />

From Wu et al. [in preparation].................................................................... 127<br />

Figure 4-4 The fitting (orange line) of the downstream distributions (in their respective<br />

plasma frames) with the two maxwellian components as derived from Section<br />

3.3.3. The left panel is plotted in log-log scale and the right panel is plotted in<br />

the log-linear scale. The dotted lines are downstream speed distributions from<br />

the simulations. Black dots: the simulation with a pickup ion ring shell. Green:<br />

the simulation with a pickup ion sphere. Orange: the simulation with the<br />

Vasyliunas–Siscoe distributed pickup ions. Red: the simulation with the<br />

Maxwellian distributed pickup ions. Right panel: log-linear scale. Left panel:<br />

log-log scale. The perpendicular speed vperp is in the unit of the upstream<br />

Alfvén speed vA. From Wu et al. [in preparation]........................................ 130<br />

Figure 4-5 The fitting of the downstream distributions (in their respective plasma frames)<br />

with the two maxwellian components as derived from Section 3.3.2. The left<br />

panel is plotted in log-log scale and the right panel is plotted in the log-linear<br />

scale. The dotted lines are downstream speed distributions from the<br />

simulations. Black dots: the simulation with a pickup ion ring shell. Green:<br />

the simulation with a pickup ion sphere. Orange: the simulation with the<br />

Vasyliunas–Siscoe distributed pickup ions. Red: the simulation with the<br />

Maxwellian distributed pickup ions. Right panel: log-linear scale. Left panel:<br />

log-log scale. The perpendicular speed vperp is in the unit of the upstream<br />

Alfvén speed vA. From Wu et al. [in preparation]........................................ 131<br />

Figure 4-6 The downstream ion speed distributions (in their respective plasma frames)<br />

from the simulations: MA=4 (blue), MA=6 (green), MA=8 (black), MA=12<br />

(orange) and MA=16 (red). The left panel is in the log-log scale and the right<br />

panel is in the log-linear scale. From Wu et al. [in preparation].................. 133<br />

Figure 4-7 The magnetic field profile (Bz, Bx) of the three simulations with different<br />

shock nomal. Black: the simulation with θBN = 90°. Orange: the simulation<br />

with θBN = 80°. Blue: the simulation with θBN = 70°. ................................... 135<br />

Figure 4-8 The downstream ion speed distributions (in their respective plasma frames)<br />

from the three simulations: θBN = 90° (black), θBN = 80° (orange) and θBN =<br />

70° (blue). The left panel is in the log-log scale and the right panel is in the<br />

log-linear scale. ............................................................................................ 136<br />

xviii


Figure 5-1 Magnetic field profiles of the three simulations without pickup ions: a) H=<br />

c/ω pi (top panel), b) H=0.5 c/ω pi (middle panel), and c) H=0.25 c/ω pi<br />

(bottom panel). ............................................................................................. 142<br />

Figure 5-2 The compression ratios as a function of time calculated from the average<br />

density of a 50 c/ω pi wide region that is 10 c/ω pi away from the overshoot in<br />

the downstream. The three simulations are presented: a) H= c/ω pi (blue), b)<br />

H=0.5 c/ω pi (black), and c) H=0.25 c/ω pi ( red curve).............................. 143<br />

Figure 5-3 Magnetic field profiles of the three simulations at the same resistivity but<br />

different cell sizes: a) H= c/ω pi , b) H=0.5 c/ω pi , and c) H=0.25 c/ω pi .<br />

Panel c) shows an unphysical fast moving shock caused by the small cell size.<br />

This effect can be eliminated with a larger resistivity as shown in Panel c’),<br />

where the cell size is as small as Panel c) but the resistivity is 5 times bigger<br />

the previous 3 panels. ................................................................................... 144<br />

Figure 5-4 The compression ratios as a function of time calculated from the average<br />

density of a 50 c/ω pi wide region that is 10 c/ω pi away from the overshoot in<br />

the downstream. The four simulations are presented: a) H= c/ω pi (blue), b)<br />

H=0.5 c/ω pi (black), c) H=0.25 c/ω pi (dashed red curve) and c’) H=0.25<br />

−4<br />

c/ω with higher resistivity (red curve) η = 10 (4 π / ω ) ....................... 146<br />

pi<br />

Figure 5-5 Magnetic field profiles of three simulations with different numbers of particles<br />

per cell. Top panel: simulation with 25 particles per cell; middle panel:<br />

simulation with 100 particles per cell; bottom panel: simulation with 400<br />

particles per cell............................................................................................ 148<br />

Figure 5-6 Entropy (black) of the system at the end of the simulation run: 20% pickup ion<br />

shell, MA=8, β sw =0.05, β PUI = 8.53. The temperature profile (blue) is scaled<br />

and overplotted. ............................................................................................ 152<br />

Figure 6-1 Comparison of our simulated magnetic field profile (orange) with the Voyager<br />

observed TS-3 magnetic field profile (black). Voyager 2 observing time is<br />

label on the x axis on the bottom and the simulation length is marked on the x<br />

axis on the top. Voyager 2 Data are provided by Leonard F. Burlaga......... 155<br />

Figure 6-2 The fitting of the simulated downstream distribution (MA=8, β sw =0.05, and<br />

β PUI =8.53) with one maxwellian component. The left panel is plotted in loglog<br />

scale and the right panel is plotted in the log-linear scale. The circles are<br />

xix<br />

pi


downstream speed distributions from the simulations. The lines are the fitted<br />

function. The perpendicular speed vperp is in the unit of the upstream Alfvén<br />

speed vA........................................................................................................ 158<br />

Figure 6-3 The modeled x direction (r direction in the RTN coordinate system) energy<br />

spectra of the sunward directing ions in the shock frame (left). Partial spectra<br />

of Figure 6-3 that fall into the IBEX-Hi detector limit (right). We fit the<br />

spectra with power law functions (solid lines). Blue: MA=4, black: MA=8, and<br />

red: MA=16. .................................................................................................. 159<br />

Figure 7-1 The three upstream components in the solar wind frame. The solid curve<br />

represents the Maxwellian solar wind; the dotted curve represents the<br />

Vasyliunas–Siscoe pickup ions; the dashed curve represents the powerlaw tail.<br />

...................................................................................................................... 169<br />

xx


List of Abbreviations<br />

ACR Anomalous Cosmic Ray<br />

AU Astronomical Unit<br />

BU Boston University<br />

ENA Energetic Neutral Atoms<br />

GPDA Gyrophase-dependent Acceleration<br />

IBEX Interstellar Boundary Explorer<br />

MHD Magnetohydrodynamics<br />

NASA National Aeronautics and Space Administration<br />

PIC Particle-in-cell<br />

RTN Radial-Tangential-Normal<br />

SDA Shock Drift Acceleration<br />

xxi


List of Symbols<br />

B magnetic field<br />

Bj the jth component of magnetic field<br />

c speed of light<br />

d as subscript or superscript, denotes downstream<br />

DT simulation time step<br />

E electric field<br />

Ej the jth component of electric field<br />

e elementary charge<br />

19<br />

1.6022 10 −<br />

× C<br />

e as subscript, denotes electron species<br />

f distribution function<br />

fi distribution function of the jth particle/species/component<br />

j as superscript and subscript, denotes ion species/component<br />

kB Boltzmann constant<br />

MA Alfvén Mach number<br />

Mcs sonic Mach number<br />

MMS Magnetosonic Mach number<br />

mj mass of a jth species particle<br />

n number density<br />

23<br />

1.38 10 −<br />

× J・K -1<br />

nj number density of the jth species/component<br />

P pressure<br />

xxii


Pj pressure of the jth component<br />

rL Larmor radius (gyroradius)<br />

rS compression ratio (shock strength)<br />

T temperature<br />

Tj temperature of the jth species/component<br />

u flow velocity<br />

u as subscript or superscript, denotes upstream<br />

ud downstream flow velocity<br />

uu upstream flow velocity<br />

v individual ion velocity<br />

vA Alfvén speed<br />

vcs sound speed<br />

vth thermal speed<br />

β plasma beta<br />

β j plasma beta for the jth component<br />

γ polytropic index<br />

γ j<br />

polytropic index for the jth component<br />

ε ref solar wind ion reflection efficiency<br />

ε 0<br />

permittivity of free space<br />

xxiii<br />

12<br />

8.8542 10 −<br />

× F・m -1<br />

η when expressed without subscript, denotes resistivity<br />

η j<br />

the percentage of dissipated energy that heats the jth component


θ BN shock normal angle<br />

κ power law index<br />

λ D Debye length<br />

μ 0 permeability of free space<br />

τ temperature jump<br />

xxiv<br />

4π10 −7<br />

× H・m -1<br />

τ j<br />

temperature jump of the jth component<br />

ϕ pickup ion relative density<br />

Ω cyclotron frequency<br />

Ωj cyclotron frequency of the jth species/component<br />

χ d downstream pickup ion relative pressure<br />

ω p wave frequency<br />

ω pj plasma frequency of the jth species


1 Introduction<br />

1.1 Motivation<br />

Our solar system lies within the heliosphere, an asymmetric bubble inflated into the<br />

interstellar medium by the outflowing solar wind. The boundary of the heliosphere is the<br />

heliopause (located at ~120 astronomical units 1 ), where the solar wind pressure balances<br />

the interstellar wind pressure. The heliopause is therefore a separator between the solar<br />

wind and the instellar medium. Inside the heliopause, there is another heliospheric<br />

boundary: the heliospheric termination shock (located at ~90 AU). This shock, like all<br />

collisionless heliospheric shocks, is relatively thin, about 10 -5 heliosphere radius (10 7 km)<br />

in thickness. At the shock, the interstellar medium starts to have an effect on the solar<br />

wind. The solar wind is slowed from a supersonic flow in the upstream region to a<br />

subsonic flow in the downstream region and then deflected back to the tail.<br />

The relationship between the termination shock, the heliosphere and the interstellar<br />

medium is illustrated in Figure 1-1. In the figure, upstream is sunward of the termination<br />

shock; the heliosheath lies downstream, beyond the termination shock. The nose of the<br />

termination shock is in the direction of the solar system's motion through the local<br />

interstellar medium. If this relative motion is supersonic, a bow shock will form in the<br />

interstellar medium. In this case, the heliosheath is divided into two regions: the inner<br />

heliosheath (the region between the termination shock and the heliopause) and the outer<br />

heliosheath (the region between the bow shock and the heliopause). The inner heliosheath<br />

1 1 astronomical unit (AU) =1.49×10 13 cm, approximately the Sun–Earth distance.


2<br />

is the sheath of primarily solar wind ions and the outer heliosheath is the sheath of<br />

primarily interstellar material.<br />

Outer<br />

Heliosheath<br />

nose<br />

Inner<br />

Upstream Downstream<br />

Figure 1-1 An artist's rendition of the termination shock and its relation to our<br />

heliosphere and the interstellar medium. Adapted from Walt Feimer, National<br />

Aeronautics and Space Administration (NASA).


3<br />

A more physical image of the termination shock and the heliospheric environment<br />

was produced by Zank [1999] through a Magnetohydrodynamic (MHD) simulation and<br />

was later adapted by Stone [2001], as shown in Figure 1-2. This model assumes the<br />

existence of a bow shock in front of the heliosphere. The interstellar wind is relatively<br />

cool upstream of the bow shock (dark blue), just as the solar wind is relatively warm<br />

upstream of the termination shock (yellow). Both plasmas are heated at their respective<br />

shocks, although the heating of the interstellar wind (light blue) is less than that of the<br />

solar wind (red) because it is expected that the bow shock is considerably weaker than the<br />

termination shock. The heated interstellar wind and the heated solar wind encounter at the<br />

heliopause (green region), where the two flows mix and where the temperature falls in-<br />

between the temperatures of both flows.<br />

A closer examination of Figure 1-2 shows that the termination shock forms a comet-<br />

like oval instead of a perfect sphere. The nose of the termination shock is compressed as<br />

the solar system moves into the local interstellar medium. On the opposite site, the tail of<br />

termination shock extends into the local interstellar medium. This tail feature is more<br />

pronounced for the inner heliosheath, as the downstream solar wind flows deflect and are<br />

dragged tailward (the flow directions are indicated with black arrows). The solar wind<br />

plasma inside the termination shock is generally cooler than the solar wind plasma in the<br />

inner heliosheath, because the termination shock slows the flow and heats the plasma.<br />

The Voyager spacecraft are currently in the inner heliosheath. Their trajectories are<br />

marked by the blue arrow (Voyager 1) and red arrow (Voyager 2). Both Voyagers<br />

encounter the shock in the upwind direction near the nose of the termination shock. The


4<br />

Voyager measurements of the termination shock will be summarized in Section 1.3 of<br />

this chapter.<br />

Figure 1-2 The heliosphere inflated by the solar wind in the region surrounding the Sun<br />

in the interstellar medium. The heliosphere is formed by the supersonic solar wind as it<br />

expands radially from the Sun. At the termination shock, the solar wind abruptly slows<br />

and heats as it turns toward the tail of the heliosphere. The heliopause is boundary that<br />

separates the solar wind flow and the interstellar flow, where the pressure of the solar<br />

ions balances the pressure of the interstellar ions. In this model, the interstellar wind is<br />

also supersonic and a bow shock forms upstream of the nose of the heliosphere. Adapted<br />

by Stone [2001] from Zank [1999].


5<br />

Unlike the Earth’s bow shock which has been extensively measured in the past 50<br />

years with satellite observations, the termination shock is not well explored or understood<br />

due to its distant location and large size. Voyager 1 and Voyager 2, launched in 1977, are<br />

the only spacecraft to reach the termination shock in 2004 and 2007, respectively.<br />

However, the Voyagers are not able to measure the suprathermal ions (~keV). Voyager<br />

1’s plasma instrument failed early before the spacecraft crossed the termination shock.<br />

Voyager 2 has a working plasma instrument, but the resolution of this instrument<br />

prohibits the measurement of the suprathermal ions, because the instrument is not able to<br />

distinguish signals from background noise in this energy range. The first indirect<br />

measurement of the suprathermal ions was made by the Interstellar Boundary EXplorer<br />

(IBEX) [McComas et al., 2009], through the remote measurement of energetic neutral<br />

atoms (ENAs).<br />

The IBEX spacecraft was launched on October 19, 2008 as an Earth orbiting satellite.<br />

The spacecraft’s low energy sensor IBEX-Lo measures neutral atoms ranging from 2 eV<br />

to 2 keV. The spacecraft’s high energy sensor IBEX-Hi measures energetic neutral atoms<br />

(ENAs) ranging from 300 eV to 6 keV. When IBEX’s orbit is outside the Earth’s<br />

magnetosphere, the spacecraft measures ENAs originated behind the termination shock in<br />

the heliosheath. The important fact about these ENAs is that they are originally<br />

termination shock heated ions. The ions undergo charge exchange with interstellar neutral<br />

atoms in the inner heliosheath. As a result of the charge exchange, the termination shock<br />

heated ions are converted into ENAs. As neutrals, the ENAs are unconfined by magnetic<br />

fields. The ENAs which have a sunward directed velocity will propagate into the


6<br />

heliosphere toward IBEX’s location. Therefore, IBEX’s measurement is a line-of-sight<br />

integration of the ENAs. To interpret the ENA measurements from IBEX, one needs to<br />

understand the termination shock heated ions. Our theoretical study of the ion kinetics at<br />

the termination shock is thus motivated by the IBEX mission to take important first steps<br />

toward understanding these ions.<br />

Collisionless shocks heat plasmas, both electrons and ions. Unlike the Earth’s bow<br />

shock, which is dominated by the thermal solar wind ion, the termination shock has an<br />

additional ion component, the suprathermal pickup ions. The presence of these ions<br />

implies that the physics of the termination shock is significantly different from that of the<br />

Earth’s bow shock. The important physical questions then are: How is the shock affected<br />

by this suprathermal ion component? What is the partition of dissipated energy between<br />

the solar wind component and the pickup ion component? How is each ion component<br />

energized? The related processes that provide answers to these questions are not only<br />

important to the termination shock but also generic to plasma physics. For example, other<br />

astrospheres 2 may also have more than one ion component at astrospheric termination<br />

shocks. Although the ions of these astrospheres operate at different energies, some of the<br />

scaling concepts and physical insights we can gain from studying the heliospheric<br />

termination shock are extendable to astrospheric termination shocks, particularly for the<br />

astrospheres that are similar to the heliosphere.<br />

2 Like the heliosphere, an astrosphere is a bubble blown by hot plasmas (stellar wind) originating from a<br />

single massive Sun-like star into the interstellar medium. Analagous to the heliospheric termination shock,<br />

an astrospheric termination shock is where the stellar wind becomes subsonic.


7<br />

We emphasize here that in this dissertation we do not consider the role of energetic<br />

particles (anomalous cosmic rays with energy ranging from ~100 keV to the MeV level<br />

and galactic cosmic rays with energy ranging from the MeV level to the TeV level),<br />

which are widely believed to be accelerated by diffusive shock acceleration at various<br />

astrophysical shocks (e.g., supernova remnants) [Kirk and Dendy, 2001]. Rather, to<br />

explain Voyager observations and to complement the recent IBEX observations at 2 eV<br />

to 6 keV [McComas et al., 2009, Funsten et al., 2009, Schwadron et al., 2009, Fuselier et<br />

al., 2009], we focus on the modeling aspects of the thermal and suprathermal ions (~95%<br />

protons) at the termination shock. These ions and the associated magnetic fields, when<br />

extended to three spatial dimensions, will provide the groundwork upon which to build<br />

energetic particle models. Similar work has been done from different perspectives. Lucek<br />

and Bell [2000] perform kinetic cosmic ray simulations coupled to a single fluid<br />

magnetohydrodynamic code. Giacalone and Ellison [2000] investigate the behavior of<br />

energetic particles using hybrid simulations that treat pickup ions as test particles.<br />

Although we have not considered or included energetic particles, our treatment of the<br />

suprathermal pickup ions throughout this work is self-consistent (meaning that pickup<br />

ions are not treated as test particles but as particles fully participating in the dynamics).<br />

With energetic particles in mind for the future, we focus here on understanding the<br />

fundamental scaling of the termination shock and the relative importance of the ion<br />

components.


1.2 Upstream Conditions<br />

8<br />

Upstream of the termination shock, the primary charged particles are electrons, the solar<br />

wind ions and the pickup ions. In this work, we assume all ions are protons. This<br />

assumption is valid for most quiet time solar wind conditions, when helium 3 and other<br />

heavy ions constitute only a very small fraction of the upstream solar wind population,<br />

e.g., the Helium to proton number density ratio is He/H


9<br />

articles predict that the termination shock location falls between 80 AU and 100 AU. This<br />

prediction was confirmed by the Voyagers’ crossings.<br />

1.2.2 Pickup ions<br />

In contrast to the terrestrial bow shock, the termination shock contains a significant<br />

component of very hot pickup ions. Pickup ions are created from interstellar neutral<br />

atoms that drift in through the heliosphere, unaffected by the solar wind's electromagnetic<br />

fields. These atoms become ionized through charge-exchange (electron transfer) with<br />

solar wind ions or by photo ionization. The ionization process is very fast. The newborn<br />

proton maintains its initial velocity, since the interstellar medium velocity uISM~25 km/s<br />

(very small compared with the solar wind bulk velocity usw~400km/s). After ionization,<br />

this new proton is picked up by the solar wind flow through interactions with the<br />

interplanetary magnetic field, as described in the pickup process detailed below.<br />

Consider a newly ionized pickup ion. This new proton starts to gyrate around the<br />

magnetic field. In the simplest case when the solar wind flow is perpendicular to the<br />

magnetic field (occurs at a large distance from the Sun, as implied by the Parker spiral as<br />

shown in Figure 1-3), the proton gyrates around the interplanetary magnetic field with an<br />

initial velocity ~usw as seen in the solar wind frame (top panel of Figure 1-4). In the<br />

spacecraft frame (bottom panel of Figure 1-4), the proton starts with a very small velocity<br />

(~25 km/s) as compared with the solar wind flow. Therefore, the proton “sees” a<br />

motional electric field caused by the moving solar wind plasma


10<br />

<br />

Esw = − usw × B.<br />

(1.1)<br />

A new proton gains energy from the motional electric field until its guiding center<br />

reaches a velocity that is equivalent to the solar wind flow velocity so that it becomes part<br />

of the bulk flow and no longer “sees” the motional electric field. In the spacecraft frame,<br />

this process appears as if those new protons are picked up by the solar wind flow. Thus,<br />

the protons received the name “pickup ions”. As a result, the pickup ions form a velocity<br />

ring by the end of the pickup process (in a gyroperiod), as shown in Figure 1-4. It should<br />

be noted that in reality, the magnetic field is not strictly perpendicular to the solar wind<br />

flow. The pickup ion’s gyrovelocity in the solar wind frame should be modified by a<br />

factor “sinθ”, where θ is the angle between the magnetic field and the solar wind flow.<br />

This factor can be neglected at large heliospheric distances because of the Parker spiral.<br />

Due to the initial deviation of the ring distribution from a Maxwellian, hydromagnetic<br />

waves in the solar wind are rapidly amplified by various anisotropic phase-space<br />

instabilities. These waves mediate the coupling of pickup ions parallel velocity to the<br />

solar wind flow [Burgess, 1997]. In addition, pre-existing solar wind<br />

turbulence/fluctuations enhance this coupling too [e.g., Burgess, 1997; Intriligator et al.,<br />

1996]. Upstream of the termination shock, it is widely believed that the pickup ions at<br />

least form a shell distribution due to the strong pitch angle scattering [Burgess, 1997]. A<br />

ring distribution is unlikely unless the mass loading is extremely strong within a few<br />

pickup ion gyro radii upstream of the termination shock.


Shock normal<br />

11<br />

Figure 1-3 The Parker Spiral. Beyond 20 AU, the interplanetary magnetic field is nearly<br />

toroidal as the solar rotation wraps up the field. Hence, this magnetic field is nearly<br />

perpendicular to the solar wind flow and the termination shock normal. Figure adapted<br />

from Steve Suess, NASA.


u <br />

B <br />

Esw <br />

usw<br />

Solar wind frame<br />

sw<br />

vy<br />

vy<br />

usw<br />

Spacecraft frame<br />

12<br />

pickup ring<br />

solar wind<br />

Figure 1-4 A conceptual illustration of the pickup process, and the subsequent velocity<br />

distribution in the solar wind frame (upper panel) and the spacecraft frame (bottom<br />

panel). The pickup process ends when the pickup ion’s guiding center moves with the<br />

solar wind flow velocity.<br />

usw<br />

solar wind<br />

usw<br />

vx<br />

pickup ring<br />

2usw<br />

vx


13<br />

While the solar wind ions have a relatively low temperature (a few eV upstream of<br />

the termination shock), pickup ions are created with a high effective temperature (~1<br />

keV). Hence, pickup ions are crucial to this study, because they dramatically increase the<br />

effective temperature of the upstream flow. It is also important to note that the increase of<br />

upstream flow speed (fast solar wind) significantly increases pickup ion energies. The<br />

effect of this on the shock will be discussed in Chapter 4.<br />

Although the pickup process upstream of the shock creates a ring-shaped ion velocity<br />

distribution initially, this anisotropic distribution is unstable to the growth of waves.<br />

These waves pitch-angle scatter the pickup ions into a shell distribution. From there the<br />

pickup ions are further scattered into broader distributions: first a velocity sphere, then a<br />

velocity distribution that described by Vasyliunas and Siscoe [1976] (which we refer to as<br />

the Vasyliunas–Siscoe distribution), and finally a Maxwellian distribution. The exact<br />

pickup ion velocity distribution has not been measured upstream of the shock, due to the<br />

Voyager spacecraft’s inability to resolve pickup ions. For the simulations described in<br />

later chapters, we assume that the upstream pickup ion distributions are a velocity shell, a<br />

velocity sphere, a Vasyliunas–Siscoe distribution and a Maxwellian distribution,<br />

respectively. Each velocity assumption corresponds to a successive level of<br />

thermalization experienced by the pickup ions as they flow away from the Sun, as well as<br />

a successive level of the pickup mass loading process along the path. By exploring the<br />

outcomes of these different assumptions, we can determine the importance of this<br />

unknown parameter and understand the kinetics.


14<br />

1.3 Observations of the Termination Shock<br />

To date, the only in-situ measurements of the termination shock have been made by the<br />

Voyager spacecraft. Voyager 1 crossed the termination shock in December 2004 at a<br />

heliocentric distance of 94 AU and a heliospheric latitude of 34.1 ° [Stone et al., 2005]. In<br />

August 2007 Voyager 2 crossed the termination shock at 84 AU and -27.5 ° in<br />

heliospheric latitude [Decker et al., 2008]. At the times of their respective crossings, both<br />

Voyager 1 and Voyager 2 carried operating magnetometers and instruments to measure<br />

energetic charged particles, but only Voyager 2 carried a functional plasma instrument<br />

(energy range 10-5950 eV). However, because of the plasma instrument’s low energy<br />

resolution (29%), Voyager 2 was not able to measure the relatively low flux of<br />

suprathermal ions (~keV ions). Therefore, there are no direct observations of the partition<br />

of dissipated energy at the termination shock.<br />

From the Voyager measurements, Richardson et al. [2008] conclude that the<br />

termination shock is a quasi-perpendicular, supercritical shock. Burlaga et al. [2008]<br />

further show that the shock’s magnetic profile has the typical supercritical shock feature 4 :<br />

a foot, a ramp, an overshoot, and an undershoot, as illustrated in Figure 1-5.<br />

Richardson [2008] provides a detailed analysis of the Voyager 2 plasma data at the<br />

shock and into the heliosheath. Both the Richardson et al. [2008] and the Richardson<br />

[2008] papers demonstrate that there were substantial fluctuations in the plasma<br />

properties, as shown in the following tables that summarize Voyager 2 observation from<br />

4<br />

The physical importance of these structures as to supercritical shocks will be explained in Section 1.4<br />

when we describe the kinetic physics.


15<br />

Richardson et al. [2008], Richardson [2008], Li et al. [2008], and Burlaga et al. [2008].<br />

In the tables, TS-2 is the second crossing when the termination shock is moving<br />

outwards, and TS-3 is the third crossing when the termination shock is moving inward.<br />

Figure 1-5 Magnetic field profile from one of the Voyager 2 observed termination shock<br />

crossing. Adapted from: Bulaga et al. [2008]<br />

Table 1-1. Voyager 2 Observations in the Upstream Region.<br />

Crossings nu (cc) uu (km/s) MMS,sw-core ≈ MA β Bu (nT)<br />

u<br />

TS-2 0.0013 (321.0, 11.3, 1.1)<br />

TS-3 0.0013<br />

4.9±0.1 0.053 (-0.012, -0.063, -0.022)<br />

(321.0, 11.3, 1.1) 8.8±1.2 0.053 (-0.012, -0.063, -0.022)


16<br />

Table 1-2. Voyager 2 Observations in the Downstream Region.<br />

Crossings nd (cc) ud (km/s) MMS, sw-core, d ≈MA, d β Bd (nT)<br />

d<br />

TS-2 0.0033 (184.7, 11.9, 5.8) 1.1±0.1<br />

TS-3 0.0022 (167.6, 29.7, 8.9) 2.8±0.4<br />

Table 1-3. Voyager 2 Observations of the Termination Shock.<br />

1.143 (-0.003, -0.132, -0.022) )<br />

0.508 (-0.006, -0.122, -0.013)<br />

Crossings θBN rS=nd/nu uS (km/s) wS (km) wramp (km) η τ<br />

sw<br />

TS-2 82.8°±3.9° 2.38±0.14 94.0±3.4 300,000<br />

TS-3 74.3°±11.2° 1.58±0.71 -67.9±17.3 100,000<br />

Table 1-4. Voyager 2 Observations of the Foreshock Region.<br />

- ~15% ~10 – 20<br />

6000 ~15% ~10 – 20<br />

rforeshock (AU) uu,foreshock(km/s) |ΔEsw,ram|/ Esw,ram<br />

75 400-300 (in three steps) 40%<br />

In the tables, n denotes the plasma density; u denotes flow velocity, subscripts “u”<br />

and “d” denote upstream and downstream; MMS is the magnetosonic Mach number<br />

defined as MMS=uu/(vA 2 +vcs 2 ) 1/2 , where vA is the Alfvén speed and vcs is the sound speed<br />

(caution: in the observations, vcs can only be inferred from the core solar wind, which in<br />

reality is expected to be much smaller than its true value.); MA is the Alfvén Mach<br />

number defined as MA=uu/vA; β is the plasma beta of the observed core solar wind ions;<br />

B is the magnetic field; θBN is the shock normal (defined as the angle between the<br />

magnetic field and the shock normal); rS is the shock strength (compression ratio); uS is<br />

the shock velocity (where a negative value means that the shock is moving inward); τ is<br />

the temperature jump of the measure core solar wind ions; wS is the width of the shock;


17<br />

wramp is the width of the ramp (notably for TS-3, w ramp ~1 c/ ω pi, u [Burlaga et al., 2008]);<br />

η represents energy partition and η sw is the amount of ram energy that is converted into<br />

solar wind heating; rforeshock (AU) is the location from the Sun of the foreshock<br />

measurements; uu,foreshock is the bulk speed of the upstream solar wind in the foreshock<br />

region which drops from 400 km/s to 300 km/s in three steps; |ΔEsw,ram|/ Esw,ram is the<br />

amount of solar wind ram energy that has been lost in the foreshock region. The vectors<br />

in the table are expressed in the Radial-Tangential-Normal (RTN) coordinate system (R,<br />

points from the Sun to the spacecraft; T, the Sun’s rotation vector crossed into R; N,<br />

completes the right-handed triad). For Li et al. [2008], we cite the mean values from their<br />

Table 1 and 2 in order to be consistent with Richardson et al. [2008], Richardson [2008],<br />

and Burlaga et al. [2008]. The resolutions of the Voyager 2 measurements are:<br />

• Magnetic field: 48 s [Burlaga et al., 2008],<br />

• Plasma: 192 s [Richardson, 2008],<br />

• Energy: 29% [Richardson, 2008].<br />

There are some inconsistencies between those published values, due to fluctuations<br />

and the limited resolution of the instruments onboard Voyager 2. For example, the<br />

downstream flow speeds of TS-2 and TS-3 inferred by Li et al. [2008] seem at odds with<br />

each other. As TS-2 is both more perpendicular and stronger than TS-3, the downstream<br />

flow speed of TS-2 should be smaller than that of TS-3. However, this is not the case in<br />

the tables. Also, the inferred temperatures<br />

2<br />

Td = βdBd /(2 μ0kBnd)<br />

are 174,000 K<br />

(downstream) and 5000 K (upstream) for TS-2 and 100,000 K (downstream) and 5000 K


18<br />

(upstream) for TS-3. However, Richardson [2008] measured 11,000 K over a wider range<br />

upstream and 181,000 K over a wider range downstream. For the above reasons, we<br />

adopt an estimate that the temperature jump of the core solar wind is on the order of ~10<br />

– 20 as listed in the last column of Table 1-3.<br />

In a plasma, the total sound speed is dominated by hot ions. The magnetosonic Mach<br />

number is the coupled Mach number of the sonic Mach number and the Alfvén Mach<br />

number. Excluding the hot ions, the Voyager-measured magnetosonic Mach number<br />

actually approximates the entire flow’s Alfvén Mach number (MMS,sw-core≈ MA), as<br />

marked in Table 1-1 and Table 1-2. The reason is that the Voyager measured core solar<br />

wind has a very small plasma beta and the sound speed calculated from these low energy<br />

core solar wind ions is only a small fraction of the total sound speed of the flow.<br />

The Alfvén Mach number M A = u miμ0 n / B,<br />

for a strictly perpendicular shock, is<br />

MA,d=MA,u/rS 1.5 (see Section 3.2.5 for derivation). Taking TS-3, if rS=1.58 (as in Table 3),<br />

then MA,d=4.35 (does not agree with Table 2); if MA,d=2.8 (as in Table 2), then rS=2.14<br />

(upper limit of the compression ratio in Table 3). If TS-3 has a MA twice that of TS-2’s as<br />

Table 2 shows, assuming that in the upstream TS-2 and TS-3 has the same conditions<br />

(which is not true for MA in Table 1), then nd,TS-2/nd,TS-3≈ Bd,TS-2/Bd,TS-3≈2 (1/1.5) =1.58.<br />

Comparing with the observations (Table 2), the density ratio (0.0033/0.0022=1.5) is in<br />

agreement but the magnetic field ratio (0.132/0.122=1.1) seems insufficient. Another<br />

observational inconsistency is that the upstream MA of TS-3 is about twice the upstream


19<br />

MA of TS-2, while these two upstream regions have similar magnetic fields, densities,<br />

and upstream velocities.<br />

The<br />

B<br />

B<br />

nd ,<br />

nu ,<br />

(downstream normal component of magnetic field over upstream normal<br />

component of magnetic field) is 2.10 for TS-2 and 1.94 for TS-3, as can be calculated<br />

from Table 1-1 and Table 1-2. This shows that the shocks should be very similar, despite<br />

the fluctuations and different versions of calculations, inferences and interpretations from<br />

different publications.<br />

Voyager 1 observations at the termination shock are listed in Table 1-5.<br />

Measurements from Voyager 1 are much more limited than those from Voyager 2, due to<br />

the instrument limitations. To start with, we are concerned with the “average” properties<br />

of the shock Voyager 2 observed: compression ratio ~2, plasma density np ≅ 0.0013 cm -3 ,<br />

magnetic field B ≅ 0.063 nT (upstream), upstream flow speed uu=300 km/s and upstream<br />

temperature 5000 K, which corresponds to Alfvén speed vA= 38 km/s (calculated from<br />

v<br />

A<br />

2<br />

Bu<br />

= , ignoring pickup ion density contribution to the upstream density nu,),<br />

μ nm<br />

0<br />

u i<br />

Alfvén Mach number of MA ≅ 8, βsw ≅ 0.05 and solar wind proton temperature Tp ≅ 0.43<br />

eV (upstream).<br />

Table 1-5 Voyager 1 (V1) Observations of the Termination Shock.<br />

encounters r (AU) rS uu (km/s) ud (km/s) θBn wS (km) τ Td (k)<br />

V1 94 0.4<br />

2.6 +<br />

−0.2<br />

200 100 - -<br />

-<br />

-


20<br />

There are three important puzzles posed by the Voyager 2 measurements. First,<br />

Richardson [2008] found that solar wind ion heating accounts for only ~15% of the flow<br />

dissipation. Richardson [2008] postulated that pickup ions account for most of the<br />

dissipation. Second, Richardson et al. [2008] and Li et al. [2008] concluded that<br />

downstream of the termination shock, the heliosheath flow speed remains super-<br />

magnetosonic. Third, the peak of anomalous cosmic rays (ACR, ~100keV – MeV)<br />

predicted by diffusive shock acceleration [e.g., Pesses et al., 1981; Jokipii 1986] is not<br />

observed [Stone et al., 2005] and the ACR enhancement is anisotropic in two opposite<br />

directions for Voyager 1 and Voyager 2 observations, respectively.<br />

We emphasize that the density jump and magnetic field jump are ~2, indicating that<br />

the shock crossed by Voyager 2 is relatively weak by comparison with the strong shock<br />

limit of 4. Although suprathermal ions are too tenuous to be measured by the plasma<br />

instrument, Richardson [2008] infers that pickup ions make up roughly 20% of the sheath<br />

plasma and have energies of about 6 keV in the heliosheath.<br />

1.4 Kinetic Physics of the Termination Shock<br />

Because this dissertation focuses on kinetics, we group shocks, based on the nature of the<br />

collision process, into gas kinetic shocks and collisionless plasma shocks.<br />

In general, a shock forms when the flow encounters an obstacle with a speed that is<br />

faster than the fastest wave in the flow medium. For a gas, this speed is the sound speed;


21<br />

for a plasma, this speed may be the magnetosonic speed (fast mode wave speed), the<br />

Alfvén speed, or the slow mode wave speed.<br />

Pure gas kinetic shocks, involving no electromagnetic forces, are dissipated by<br />

collisions between gaseous particles. Because those particles are neutral molecules or<br />

atoms, the Rankine–Hugoniot relations (the conservation rules of mass, momentum and<br />

energy) can be formulated within hydrodynamics [Gombosi, 1994].<br />

Collisionless plasma shocks such as the Earth’s bow shock, however, are not<br />

dissipated by binary collisions. Plasmas of interplanetary space are considered<br />

collisionless, with a mean free path around 1 AU or more (much larger than their average<br />

Coulomb interaction range). In fact, a half century ago, it was hotly debated whether<br />

collisionless shocks even existed or not. The debate was not resolved until the<br />

observation of the Earth’s bow shock [Sonett and Abrams, 1963].<br />

For shocks in collisionless plasmas, energy dissipation is provided by mechanisms<br />

other than collisions. Here, we only address the quasi-perpendicular shocks, which are<br />

relatively thin shocks with a shock normal nearly perpendicular to the magnetic field.<br />

Consider such a quasi-perpendicular shock with a single, relatively cold, upstream ion<br />

component. At the shock, the dynamic energy of the plasma is dissipated mainly by<br />

cross-field current driven instabilities [Winske, 1984] if the shock is subcritical (i.e., the<br />

Alfvén Mach number is below a critical number). If this shock is a supercritical shock, it<br />

is instead dissipated mainly by ion reflection [e.g., Paschmann et al., 1982], as is the case<br />

in the Earth’s bow shock. The first study of such supercritical shocks was made by


22<br />

Marshall [1955] in which he found that a flow of fast Mach number 2.76 into a cold<br />

MHD fluid required more dissipation than the maximum possible from resistivity.<br />

Further studies have since concluded that the critical Mach number M c [Woods et al.,<br />

1969] that divides subcritical shocks and supercritical shocks is a strong function of the<br />

upstream plasma beta and shock normal angle [Kennel et al., 1985]. The critical Mach<br />

number increases with decreasing plasma beta or increasing shock normal θBN (defined as<br />

the angle between the magnetic field and the shock normal) [Kennel et al., 1985]. This is<br />

because a high beta plasma has a large upstream thermal velocity to start with and thus<br />

can account for more of the net amount of dissipation by transmitted heating (like pickup<br />

ions in Chapter 2 and 3). For typical solar wind parameters, the critical Mach number is<br />

less than 2 [Kennel et al., 1985].<br />

The early magnetohydrodynamic (MHD) treatments of quasi-perpendicular shocks<br />

involved artificial collision frequencies in the fluid equations that automatically assume a<br />

picture of local and diffusive dissipation. These processes saturate when the shock is so<br />

fast (supercritical) that adiabatic heating and anomalous resistivity are not sufficient to<br />

account for energy dissipation. Kinetic effects are necessary to go beyond fluid approach.<br />

Ion reflection, a phenomenon confirmed by hybrid simulations [e.g., Quest, 1985;<br />

Gosling and Robson, 1985; Goodrich, 1985], laboratory studies [e.g., Phillips and<br />

Robson, 1972] and spacecraft observations of Earth's bow shock [e.g., Paschmann et al.,<br />

1982, Sckopke et al., 1983], has been introduced as a means to provide the extra<br />

dissipation necessary for supercritical shocks.


23<br />

The reflection process causes some ions to move back upstream of the shock. Those<br />

reflected ions are heated by the conversion of some of their ram energy into the energy of<br />

ion gyration. They are then convected downstream. Because the ion reflection process is<br />

nearly specular [Gosling and Robson, 1985], the gyrovelocities of reflected ions<br />

approximate the upstream bulk speed [Burgess, 1995; Gosling and Robson, 1985], which<br />

in this case is the upstream solar wind speed. An illustration of how the ion reflection<br />

process leads to a standard supercritical quasi-perpendicular shock structure is shown in<br />

Figure 1-6 [Wu et al., 1984]. Although Figure 1-6 is a result from a one-dimensional<br />

hybrid simulation, the structure presented in the figure is confirmed by early ISEE<br />

observations [Sckopke et al., 1983]. As Wu et al. [1984] and Kennel et al. [1985]<br />

originally discussed, the reflected ions, usually ~20%, upon reflection, enhance the<br />

upstream magnetic field as the ions compress the flow, thus creating an extended foot<br />

region as marked in Figure 1-6. As those ions gyrate around the upstream magnetic field,<br />

their Lorentz force adds to the longitudinal electric field gradually and lead to the<br />

formation of the overshoot, also marked in Figure 1-6. The overshoot effectively<br />

insulates the downstream region from the upstream region [Kennel et al., 1985]. In<br />

Chapter 2, we will re-examine specular reflection in relation to the shock structure. Our<br />

emphasis will be to provide a conceptual picture that captures the essential physical<br />

processes, which are critical to the understanding of ion kinetics. On the other hand, a<br />

complex analytic theory of ion reflection has already been developed by Leroy [1983].<br />

Leroy [1983] also offers a quantitative explanation of the magnetic field and the potential<br />

overshoot.


24<br />

Figure 1-6 The magnetic field profile of a perpendicular supercritical shock from a hybrid<br />

simulation. The foot, ramp, and overshoot are indicated. From Wu et al. [1984].<br />

In terms of the macroscopic conservation laws, since collisionless shocks largely<br />

involve magnetic forces and energies, classical hydrodynamics is insufficient to describe<br />

such shocks. Magnetodydrodynamics must be applied to derive the modified Rankine–<br />

Hugoniot relations [Burgess, 1995]. In the frame of a steady shock, both upstream and<br />

downstream plasma are assumed to satisfy the equations of MHD. Ideally, the MHD<br />

equations should be integrated from negative infinity to the shock for the upstream and<br />

from the shock to positive infinity for the downstream. Often, for a wide enough<br />

upstream and downstream region, the reasonable steady state approximations are<br />

[Burgess, 1995]:


25<br />

[ ρ u]<br />

= 0,<br />

(1.2)<br />

[ uB ] = 0,<br />

(1.3)<br />

B<br />

+ + = (1.4)<br />

2<br />

[ ρu<br />

P<br />

2<br />

]<br />

2μ0<br />

0,<br />

1<br />

γ<br />

P B<br />

2<br />

2<br />

[ ρu(<br />

u + ) + u ] = 0<br />

2 γ −1ρ<br />

μ0<br />

(1.5)<br />

for the conservation of mass, the continuity of tangential electric field, the conservation<br />

of momentum, and the conservation of energy, respectively. Here, “[X]” represents the<br />

upstream and downstream difference of the physical parameter X across the shock<br />

[X]=Xu-Xd. In order to close the system of equations (also known as the closure problem),<br />

we take the equation that describes the polytropic process P<br />

non-adiabatic situations, γ should be interpreted as a polytropic index.<br />

γ<br />

∝ ρ in Equation (1.5). For<br />

For a perpendicular shock in which the flow is perpendicular to the magnetic field,<br />

r<br />

S<br />

ρd<br />

uu Bd<br />

= = = .<br />

(1.6)<br />

ρ u B<br />

u d u<br />

Combining (1.2)–(1.5), we find [Burgess, 1995]:<br />

2−γγ2 − + + + − − + = (1.7)<br />

2<br />

( rS 1)[ rS r ( 1) ( 1)] 0,<br />

2 S<br />

γ γ<br />

2 2<br />

M A MA Mcs


where<br />

1/2<br />

A u( μρ 0 u) / u<br />

26<br />

1/2<br />

M = u B is the Alfvén Mach number and M = u ( ρ / γP<br />

) is the<br />

cs u u u<br />

sonic Mach number. It can demonstrated with algebra that the largest compression ratio<br />

one can achieve with a γ = 5 / 3 shock is rS=4 [Burgess, 1995].<br />

The Rankine–Hugoniot relation derived from MHD theory applies to shocks with a<br />

single ion component upstream. However, the termination shock has two ion components.<br />

As discussed earlier, the presence of hot (1~4 keV) pickup ions upstream changes the<br />

character of the shock by greatly increasing the effective plasma beta. The large effective<br />

plasma beta implies that the termination shock can not be treated as a classical strong<br />

shock [Wu et al, 2009], because the supercritical Mach number for such a shock is larger<br />

than a low beta shock [Kennel et al., 1985]. New models to describe as well as quantify<br />

the physical processes are thus necessary.<br />

1.5 Research Background and Goals of this Dissertation<br />

This section presents the goals of this work, in the context of previous studies of the<br />

termination shock. As discussed, the termination shock resembles some Alfvénic shock<br />

properties but is complicated by the presence of pickup ions.<br />

Various analytical models and computational simulations have reached different<br />

conclusions on the partition of dissipated energy between the solar wind ions and the<br />

pickup ions, as well as the mechanism for pickup ion energization. Zank et al. [1996] and<br />

Lee et al. [1996] independently proposed that pickup ions gain a large amount of energy<br />

through repeated reflections from a very thin–-electron Debye length scale–-cross shock


27<br />

potential layer. With a one-dimensional hybrid kinetic simulation that includes electron<br />

inertia terms (electron resistive heating), Lipatov and Zank [1999] demonstrated that<br />

pickup ions can gain more energy than the solar wind ions through multiple reflections.<br />

As a result, a few pickup ions gain substantial energy. If these ions have a sufficiently<br />

long mean free path they can, in principle, be further accelerated via diffusive shock<br />

acceleration process. However, earlier Liewer et al. [1993] reported that even with a 20%<br />

pickup ion relative density, their one-dimensional hybrid simulations showed that solar<br />

wind ions provide most of the termination shock dissipation. They also found that<br />

“pickup ions lead to the formation of an extended foot in front of the shock ramp of<br />

length approximately the gyroradius of the energetic pickup ion” [Liewer et al., 1993].<br />

The fundamental differences between the Lipatov and Zank [1999] simulations and the<br />

Liewer et al. [1993] simulations are: 1) the length scale of the shock ramp where pickup<br />

ions are most likely to be accelerated, and 2) whether the assumption of electron resistive<br />

heating is physical or not.<br />

We argue that: 1) the ramp of the termination shock scales as the ion inertial length,<br />

not the electron inertial length, and 2) the electrons are adiabatically heated. Based on<br />

early ISEE observations of strong shocks such as the Earth's bow shock (compression<br />

ratio ~4), Scudder et al. [1986a] showed that the thickness of such a shock ramp is some<br />

fraction of the upstream proton inertial length, that is, roughly the ion gyroradius. More<br />

recently, Bale et al. [2005] showed similar ramp thicknesses based on Cluster<br />

observations. Our simulations with zero percent pickup ions are consistent with these<br />

observations. Intuitively, we expect that with the addition of pickup ions at the


28<br />

termination shock, the shock will weaken (compression ratio ~2 as Voyager observed)<br />

and the ramp should widen. More directly, Burlaga et al. [2008] reported the direct<br />

observation from Voyager 2 that the ramp of the termination shock is about one ion<br />

inertial length. As to the electrons, Goodrich and Scudder [1984] demonstrate<br />

convincingly that resistive heating of electrons via various cross-field instabilities was<br />

nearly canceled due to the energy lost by electrons drifting along the shock in the ramp,<br />

so that the actual amount of energy gained by electrons was nearly adiabatic. Assuming<br />

resistive heating of electrons in a simulation causes the electrons to be overheated, this<br />

will result in a very large electron pressure gradient and hence a huge shock potential<br />

which further compresses the shock ramp to the electron scale length. This is not<br />

consistent with observations. As Lembege et al. [2004] concludes, the electron<br />

temperature jump on the order of ~50 – 100 from the Lipatov and Zank [1999] simulation<br />

is unphysical. Therefore, hybrid simulations with zero electron mass (the same model<br />

used by Liewer et al. [1993]) are a valid tool to examine the termination shock. Our study<br />

utilizes the Los Alamos hybrid simulation code to model the relative heating of the solar<br />

wind and pickup ions at the termination shock. Our simulation has fundamentally the<br />

same physics as the Liewer et al. [1993] simulation. However, we apply a more accurate<br />

set of input parameters (e.g., solar wind beta) derived from the recent Voyager<br />

observations.<br />

The presence of a substantial number of pickup ions at large heliospheric distances<br />

suggests that the environment of the termination shock is very different from that of the<br />

terrestrial bow shock. The processes by which the relatively hot pickup ions gain energy


29<br />

at the shock are fundamentally different from the specular reflection of the relatively cold<br />

solar wind ions. Using the Los Alamos hybrid simulations, we start with the simplest case<br />

of a strictly perpendicular termination shock and gradually include more realistic<br />

conditions. The physical processes involved are also examined layer by layer as we<br />

poceed.<br />

In Chapter 2, simulation methods are introduced with an emphasis on hybrid<br />

simulations—the tool we have chosen. This chapter models the termination shock in a<br />

generic sense, rather than trying to reproduce explicit features of any of the Voyager<br />

shock crossings. As such, this chapter concerns two major findings from Voyagers: that<br />

the average fraction of energy gain by the pickup ions at the shock is about 85% and that<br />

the average shock strength is about 2. The physical processes of ion energization are<br />

examined and ions are grouped according to the mechanisms by which they are energized.<br />

In Chapter 3, we develop an analytic description, the multicomponent Rankine–Hugoniot<br />

model of shock dissipation, solar wind heating and pickup ion heating. The analytic<br />

results are compared with simulations from Chapter 2. Both the analytic solutions and the<br />

simulations are consistent with Voyager observations. One puzzle reported by the<br />

Voyager team is that the measured downstream solar wind flow remains super-<br />

magnetosonic [Li et al., 2008]. We will address this issue with our analytic model as well.<br />

For both Chapter 2 and Chapter 3, most upstream parameters are fixed with Voyagers<br />

observed average values. Only the pickup ion relative density is varied to examine this<br />

most critical component’s effect on shock kinetics. Therefore, results from these two<br />

chapters are characterized as functions of the pickup ion relative density, φ , which is


30<br />

defined as the upstream pickup ion number density over the total upstream number<br />

PUI<br />

density, n / n .<br />

u u<br />

In Chapter 4, we work toward simulating a more realistic termination shock step by<br />

step. We describe the results of simulations in which we vary the pickup ion velocity<br />

distribution, the shock normal angle, and the upstream Alfvén Mach number to span a<br />

wider range of Voyager observations. Scattering at the shock is found to be so strong that<br />

different upstream pickup ion distributions result in nearly identical downstream speed<br />

distributions. The upstream Alfvén Mach number is found to be correlated with the<br />

downstream ion energies. In Chapter 5, the accuracy and convergence of our simulations<br />

are examined. In Chapter 6, we compare results from our simulations with Voyager<br />

observations and discuss the implications of our results for the IBEX mission. In Chapter<br />

7, we summarize the scientific findings, discuss the limitations of this work, and point to<br />

possible future directions of research.<br />

Throughout this work, the flow velocity and the velocity of individual ions are<br />

denoted by “u” and “v”, respectively. The subscript “u” represents upstream and the<br />

subscript “d” represents downstream. For example, the quantities u u and u d are the bulk<br />

velocities upstream and downstream, respectively. The quantity r S denotes the shock<br />

strength (or compression ratio), defined by the density jump nd / n u (downstream density<br />

over upstream density) or the jump of magnetic field transverse component. The quantity


31<br />

θ BN is the angle between the shock normal and the local magnetic field. The quantity T<br />

is the temperature. The temperature jump is denoted by τ , where τ = T / T .<br />

d u<br />

To conclude, the goal of this dissertation is to address the following questions:<br />

• How do pickup ions modify the termination shock?<br />

• How much is each ion component heated across the termination shock?<br />

• What is the partition of dissipated energy among the ion components?<br />

• How are ions (solar wind and pickup ions) energized?<br />

• What are the downstream ion velocity distributions?


2 Hybrid Simulations for an Idealized Termination Shock<br />

The termination shock is believed to be quasi-perpendicular at most heliospheric latitudes<br />

because of the Parker spiral structure of the heliospheric magnetic field and the shock's<br />

great distance from the Sun. In this chapter’s simulations, we assume an idealized<br />

termination shock: the shock is strictly perpendicular, the upstream ions are all protons;<br />

the cold solar wind ions are Maxwellian distributed, and the hot pickup ions form a shell<br />

distribution. This idealization allows us to gain physical insights of how hot pickup ions<br />

and cold solar wind ions are energized differently at the termination shock. The pickup<br />

ion relative density is varied to investigate the consequent shock response.<br />

2.1 The Hybrid Simulations<br />

In this section, the basic principles of hybrid simulations will be explained. But first, we<br />

will introduce basic plasma computational methods and review previous simulations of<br />

the termination shock. The advantages and disadvantages of each approach will be<br />

discussed. We will argue that for our goals here, the hybrid simulation is the ideal tool.<br />

2.1.1 Computer Simulation Methods<br />

When a physical system is not subject to analytic description, or is inaccessible to<br />

experimental discovery, computer simulations become necessary. The history of<br />

computer simulations can be traced back to the Manhattan project when scientists used<br />

the Monte Carlo algorithm to understand nuclear reactions. Computer simulations not<br />

only save us from performing expensive and dangerous experiments but also allow us to<br />

resolve the fast temporal and spatial variations that are difficult to track down in a


33<br />

laboratory experiment. For large scale space plasma structures, such as the termination<br />

shock, it is simply not possible to measure the whole physical domain of interest with<br />

satellites. In these circumstances, we resort to computer simulations to bridge the gap<br />

between analytic theories and experimentally unrealizable areas, hoping that this method<br />

puts the physical picture into readable form. The immediate advantage is that computer<br />

simulations are inexpensive and controllable. What makes this method attractive is that<br />

we can make as many measurements as we want. None of the measurements made<br />

perturb the system (unlike real experiments). A successful computer simulation should<br />

obey fundamental physical principles and be consistent with the limited experiments.<br />

Further, it should provide physical insight to the system studied.<br />

A typical computer simulation solves a set of mathematical equations to model the<br />

physical phenomenon of interest. The equations are cast into discrete algebraic form<br />

which is amenable to numerical solution. As the equations evolve in time with time step<br />

DT, the modeled system can be investigated as in a lab experiment. In a sense, a<br />

computer simulation is a numeric experiment—an experiment that is carried out through<br />

mathematical algorithms instead of hardware.<br />

Before we summarize previous numerical studies of the termination shock, we should<br />

briefly introduce plasma simulation methods, pioneered by Oscar Buneman (1913 –1993)<br />

and John M. Dawson (1930 – 2001). Based on the way particles are treated with respect<br />

to the electromagnetic fields, plasma simulation methods can be classified into three<br />

major categories: 1. particle orbit theory, 2. fluid simulations, and 3. kinetic simulations.


34<br />

Category 1 (particle orbit theory) is the test particle method. This method investigates<br />

plasma ion/electron behavior by tracking them as test particles in prescribed electric and<br />

magnetic fields. Hence, this method is not self-consistent, as it does not allow the<br />

plasmas to contribute back to the electric and magnetic fields. Method 2 (fluid<br />

simulations), e.g., Magnetodydrodynamic (MHD) simulation, solves a set of partial<br />

differential equations (transport equations and Maxwell’s equations) assuming that<br />

plasmas can be represented as fluids with approximate transport coefficients. This<br />

method is typically applied to large scale, low frequency phenomena where kinetic<br />

effects are not important. However, when plasma velocity distributions deviate<br />

significantly from a local Maxwellian, the fluid approximations often fail and we have to<br />

apply method 3 (kinetic simulations). Kinetic descriptions are by far the most<br />

sophisticated and complex method that allows the microscopic physical properties of<br />

plasmas to be studied.<br />

There are two distinct techniques within method 3 (kinetic description), the Vlasov<br />

method and the particle-in-cell (PIC) method. The Vlasov method treats phase-space as a<br />

continuum and integrates the distribution function f ( r , v , t) forward in time as coupled<br />

to Maxwell’s equations. It is very challenging to extend this method to multi-dimensional<br />

simulations. The PIC method represents the plasma as a large number of computational<br />

super-particles that move according to single particle equations of motion in self-<br />

consistent electromagnetic fields. In this technique: the particles are advanced in time<br />

using the equations of motion for charged particles in an electromagnetic field; and the<br />

electromagnetic fields are advanced through calculations of current densities from the


35<br />

particles. Although the particles’ motions are fully resolved, the fields are interpolated<br />

between the simulation cells. The introduction of the “cell” concept is to limit the<br />

numerical operations of this many-body system, as discussed later. Nearly all plasma<br />

physics of interest comes down to some scale lengths, within which physical quantities<br />

(density, current, etc.) can be considered/approximated as continuous. Therefore, the<br />

introduction of the simulation cell concept is not just a trick to avoid enormous amount of<br />

numerical operations, but also a physically valid and intelligent choice for describing a<br />

massively complicated system. The cell size should be carefully chosen, as will be<br />

discussed briefly later in this chapter and more extensively in Chapter 5.<br />

In hybrid simulations, fluid and particle-in-cell treatments are given to different<br />

components of a plasma.<br />

2.1.2 Previous Simulations of the Termination Shock<br />

Fluid models such as MHD simulations [e.g., Opher et al., 2006; Pogorelov et al., 2008]<br />

are the simplest way to represent the termination shock. In ideal MHD simulations, there<br />

are three forces: magnetic tension, magnetic pressure, and gas pressure. Kinetic (i.e.,<br />

particle) effects are absent. There is no physical dissipation, although there is unavoidable<br />

numerical dissipation/diffusion due to the Lax Scheme [Hockney and Eastwood, 1988]<br />

(which allows the simulation to discretize the convection equation with stability). To<br />

approximate the consequence of specular reflection, a MHD simulation has to include<br />

anomalous viscosity. Hence, MHD simulations ignore the microphysics of the shock and<br />

are not sufficient to give kinetic information on the plasmas. Without kinetic physics,


36<br />

MHD models assume that the velocity distributions of plasmas species remain near-<br />

Maxwellian, thus do not provide information about the non-thermal ion heating and<br />

thermal ion distributions which we seek. Therefore, although MHD models are useful to<br />

study the global structure as well as the shock’s relation to the solar wind, these models<br />

are too simplified to address the issues of concern to us.<br />

Kinetic Monte Carlo simulations calculate the effects of Coulomb collisions and<br />

charge-exchange processes for neutral interstellar atoms. Such calculations, when<br />

coupled directly to MHD codes, can model the interaction between the solar wind and the<br />

interstellar medium [Müller et al., 2000; Heerikhuisen et al., 2006], and provide some<br />

information about ion velocity distributions in the heliosheath. However, Monte Carlo<br />

codes typically do not include the effects of electric and magnetic fields that are critical<br />

in understanding how ions gain energy at the termination shock.<br />

The most complete representation of plasma physics is provided by PIC simulations<br />

which yield a self-consistent, kinetic description of plasma ions and electrons as well as<br />

the associated electric and magnetic fields. Although PIC simulations have been applied<br />

to the termination shock [Lee et al., 2005; Chapman et al., 2005], such computations<br />

require that the electron dynamics be followed in detail. This typically restricts the<br />

calculations to either relatively short times or to small spatial ranges and dimensions,<br />

preventing a full physical representation of the ion distributions at and downstream of<br />

this shock. Also, with a full PIC code, one typically needs to use artificial parameters,<br />

such as a small ion to electron mass ratio to reduce the difference between the electron


37<br />

and ion spatial and time scales. In addition, both the fundamental ion parameters (ion<br />

2<br />

plasma frequency ω = ne ( mε<br />

) and ion cyclotron frequency Ω = eB / m ) and the<br />

pi i i<br />

fundamental electron parameters (electron plasma frequency<br />

0<br />

i i<br />

ω = ne ( mε<br />

) ,<br />

2<br />

pe e e 0<br />

electron cyclotron frequency Ω e = eB / me<br />

, and Debye length λD) need to be included in<br />

full PIC codes. Another disadvantage of using the PIC model is that the discrete character<br />

of the super-particle electrons leads to additional numerical noise and increases the risk<br />

of the accumulation of numerical errors. One should be very careful in checking the<br />

stability as well as the energy conservation of the system. In most situations, smoothing<br />

over space and time is necessary to reduce noise. The question then, is, whether this extra<br />

effort is necessary. Both Hockney and Eastwood [1988] and Birdsall and Langdon [1985]<br />

stress that one only needs to do well enough to present the correct physics and minimize<br />

the chance of amplifying numerical noise.<br />

The middle road between MHD and PIC simulations is the hybrid simulation. Most<br />

hybrid simulations [e.g., Winske and Leroy, 1984] that have been successfully applied to<br />

shocks treat ion species as particles (as in PIC simulations) and electrons as a massless<br />

fluid (as in MHD simulations), and thus provide a fully self-consistent treatment of the<br />

electric and magnetic fields up to the ion cyclotron frequency (or gyrofrequency). It was<br />

convincingly demonstrated and validated in the past two to three decades, through<br />

comparison with observations [e.g., Winske and Leroy, 1984], that collisionless shocks<br />

are ion-inertial-length plasma phenomena which can be effectively described by hybrid<br />

simulations. While PIC codes provide a more complete representation of the physics,


38<br />

hybrid codes reduce numerical noise and extract the essential physics at frequencies up to<br />

the ion cyclotron frequency. With the hybrid code, longer run-times (in order to create a<br />

well-formed shock) and wider spatial ranges are easily achievable, neglecting the small<br />

amplitude high frequency fluctuations that are not expected to make a significant<br />

contribution to the low frequency ion physics.<br />

Liewer et al. [1993] used a one-dimensional hybrid code with a zero-mass electron<br />

fluid model to provide early, seminal results for high Mach number quasi-perpendicular<br />

termination shocks. Their simulations showed a shock transition length of about an ion<br />

inertial length and a moderate energization for pickup ions. Lipatov and Zank [1999] also<br />

used a one-dimensional hybrid code, but with a finite electron mass and a finer cell size<br />

(~0.006 c/ω pi ), to simulate the quasi-perpendicular termination shock. (Hybrid codes<br />

with finite electron mass, like PIC codes, include all the short wavelength electron scales.)<br />

In contrast to zero-electron-mass computations, Lipatov and Zank [1999] show a very<br />

narrow shock transition, on a scale less than 5% of the ion gyroradius and a pickup ion<br />

energy gain up to 0.5 MeV. It is our goal to re-address those issues, in the context of the<br />

recent Voyager 2 observations.<br />

2.1.3 Methodology of Hybrid Simulations<br />

In a hybrid model, some plasma components are treated kinetically as super-particles<br />

(like in PIC), while some plasma components are treated as fluid. First, we need to<br />

introduce the PIC concept.


39<br />

The concept of super-particle in the PIC method is constructed on the basis that the<br />

plasma is collisionless 5 . This basis is valid for the solar wind plasmas in most situations.<br />

By “super-particle”, one means that the particles are uncorrelated (also called<br />

collisionless, or weakly coupled), as their distribution function satisfies<br />

<br />

f ( xi, vi, x j, v j, t) f ( xi, vi, t) ⋅ f ( x j, v j,<br />

t)<br />

(2.1)<br />

ij i j<br />

where i and j are any two particles in the system, fij is the two-particle distribution<br />

function, and fi and fj are one-particle distribution functions. Although the super-particles<br />

do not correspond to particles of real plasmas on a one-to-one basis, they retain much of<br />

the collective behavior of the real plasma. The immediate advantage is that we can avoid<br />

solving N 2 (N is the total number of particles in the system) equations of particle-particle<br />

qq i j<br />

Coulomb force F <br />

ij = r at every time step. The forces will instead be described by the<br />

2 ij<br />

r<br />

ij<br />

continuous electric field E interpolated between mesh points (center of simulation cells,<br />

as illustrated in Figure 2-1) over simulation cells: i = i<br />

computation cycle from N 2 to N logN .<br />

<br />

F q E.<br />

This way one can reduce the<br />

Whether it is in a PIC simulation or a hybrid simulation, the fluid source terms<br />

(number density, charge density, current, electric field, and magnetic field) are specified<br />

at mesh points of computational cells. In between these mesh points at a particle’s<br />

5 Collisionless plasma satisfy nλD 3 >>1. This means that the average Coulomb interaction range a=e 2 /(4π<br />

ε 0kBT) is much smaller than the average particle spacing 1/n 1/3 (kinetic energy>>Coulomb type of<br />

potential energy).


40<br />

location, the source terms are calculated through interpolation/weighting. The most<br />

straightforward weighting method is linear weighting, which we use for our simulations<br />

of the termination shock. In the linear weighting, a source term is evenly distributed<br />

between two cells in a linear manner (one dimension: cell length; two dimensions: cell<br />

area; three dimensions: cell volume). The noise level of linear interpolation may be<br />

intolerable in some plasma problems; in such cases higher order weighting such as<br />

quadratic weighting (second order interpolation) must be applied.<br />

x= 0 H Simulation domain in one dimension<br />

XMAX<br />

1 2 3 Mesh points (cell numbers) NX NX1 NX2<br />

Figure 2-1 A one-dimensional representation of the simulation cells, the mesh points<br />

(center of each cell), and the simulation domain. Cell 1 and Cell NX2 are ghost cells that<br />

are necessary for the specification of boundary conditions. The simulation domain ranges<br />

from Cell 2 to Cell NX1.<br />

In Figure 2-1, there are two cells at the boundaries of a simulation dimension. These<br />

cells (Cell 1 and Cell NX2) are called ghost cells. They do not provide output for the<br />

simulation. The simulation domain ranges from Cell 2 to Cell NX2. The ghost cells allow<br />

easy manipulations of various boundary conditions to ensure that there is no sink/lost in<br />

the system. The particles as well as the source terms that enter Cell NX2 are either put<br />

back into cell NX1 (reflecting wall boundary) or cell 2 (periodic boundary).


41<br />

After identifying the species (some fluid, some PIC) and the cells, one can apply<br />

different numerical schemes to solve the Maxwell’s equations, the fluid equations, and<br />

the equations of particle motions (Newton’s law) under various approximations (full<br />

Maxwell, Darwin approximation, electrostatic, etc.), initial conditions, and boundary<br />

conditions that are tailored to the specific problem one intends to solve.<br />

For the simulation to run smoothly with minimum errors, natural units are often<br />

applied such that the important physical quantities are presented with values at least<br />

initially of order unity. Stability analysis should also be performed to ensure that there are<br />

no growing numerical instabilities. This can be done by applying the von Neunmann<br />

analysis to the numerical algorithm. One should check conservation laws with test runs<br />

on simple examples before going into complex physical systems. At times, smoothing is<br />

necessary in order to reduce noise.<br />

2.1.4 The Los Alamos Hybrid Simulation Model<br />

The Los Alamos hybrid code for collisionless shock was developed by Dan Winske<br />

and collaborators. In the Los Alamos Hybrid code, ions are treated kinetically as super-<br />

particles (PIC method) and electrons are regarded as a massless fluid [Winske and Omidi,<br />

1993, Winske et al., 2003]. The code is ideal for computing ion responses to plasma<br />

phenomena (such as the termination shock) at ion inertial length and time scales. The<br />

simulated plasma is quasi-neutral. The electron densities are set equal to ion densities<br />

everywhere and the electrons respond to the changes of the system instantaneously<br />

(inertia-less). The electromagnetic fields are treated in the magnetostatic condition (the


42<br />

<br />

1 ∂ E <br />

displacement current term in Ampere’s law is neglected = ∇× B− μ<br />

2<br />

0 J = 0 ) in the<br />

c ∂t<br />

low-frequency approximation. We restrict the discussion to one-dimensional simulations,<br />

which we use. This hybrid shock model has been widely used in space physics for more<br />

than two decades and evolved through several versions of numerical schemes [Winske et<br />

al., 2001]. The present form of the model has been tested to be robust and is well<br />

validated against data from quasi-perpendicular shocks, quasi-parallel shocks, and<br />

cometary shocks [Winske et al., 2001].<br />

To apply the hybrid code to the termination shock, the simulation is run in the<br />

downstream rest frame in which the shock propagates to the left, as illustrated in Figure<br />

2-2 (upper panel). (Throughtout the text, except where specified, figures are shown in the<br />

downstream rest frame.) As shown in Figure 2-2, the simulated dimension is the x<br />

direction. A perpendicular magnetic field is in the z direction. Although the simulation is<br />

one-dimensional, the ion velocities, fluid velocities and electromagnetic fields are fully<br />

three-dimensional. The one-dimensional concept implies that ∂ / ∂ y and ∂/ ∂ z are zero<br />

for all physical parameters, and that particle motions in y and z directions are not<br />

followed.


43<br />

Figure 2-2 The upper panel is a representative setup of the one-dimensional simulation<br />

for the termination shock. The bottom panel shows the time evolution of the magnetic<br />

field B z profile from a 0% pickup ion simulation. Since the simulation is run in the<br />

downstream rest frame, the shock propagates to the left. In the bottom panel, x is<br />

normalized by c / ω pi where ω pi is the ion plasma frequency and time t is normalized by<br />

Ω i where Ω i the ion cyclotron frequency based on the upstream magnetic field. From<br />

Wu et al. [2009].


and<br />

44<br />

The hybrid code solves the following equations<br />

<br />

dvi<br />

<br />

mi = e( E+ vi× B) −eηJ,<br />

dt<br />

dxi<br />

dt =<br />

<br />

(2.2)<br />

<br />

vi<br />

(2.3)<br />

for each ion super-particle. Equation (2.2) and (2.3) are solved using the leap-frog<br />

scheme: velocities are advanced at time step N+1/2; and positions are advanced at time<br />

step N. This method is second order accurate and is widely used to replace the first order<br />

Euler method.<br />

The transverse components (divergence free) of magnetic fields (By and Bz) are<br />

explicitly solved from Faraday’s law<br />

<br />

∂ B <br />

= −∇× E<br />

(2.4)<br />

∂t<br />

using the fourth order Runge–Kutta scheme and subcycling [Winske and Quest, 1988].<br />

The fourth order Runge–Kutta solver ensures simulation accuracy and the subcycling<br />

ensures the convergence of the calculation, as will be discussed in Chapter 5. We apply<br />

10 subcycles for the magnetic field solver, each performing a fourth order Runge-Kutta<br />

integration of Equation (2.4) within a time step DT/10. In contrast, solving the normal<br />

component of magnetic field (Bx) is relatively straightforward. The zero divergence of


45<br />

<br />

the magnetic field ∇⋅ B = 0 , together with this simulation’s one-dimensional setting in<br />

the x direction, imply that Bx=constant. In Equation (2.4), the electric fields are<br />

calculated from electron momentum equation<br />

∂ <br />

nmu e e e = 0 =− ene( E+ ue× B) −∇⋅ Pe + eneη⋅J, (2.5)<br />

∂t<br />

where ue is the electron bulk velocity, ui is the ion bulk velocity, η is the<br />

phenomenological resistivity (set to be a constant) to approximate the electron-ion<br />

interaction. Equation (2.5) can be rewritten as the Ohm’s law<br />

<br />

∇⋅P<br />

<br />

e E = − ue× B− + η ⋅J.<br />

(2.6)<br />

en<br />

This equation for the electric field does not require a time advance, unlike Equation (2.4)<br />

for the magnetic field. The electron pressure term Pe is advanced adiabatically. Ampere’s<br />

law<br />

is used to eliminate ue in Equation (2.5).<br />

<br />

∇× B = 4πμ J = 4 π enμ( ui −ue)<br />

0 i 0<br />

e<br />

(2.7)<br />

<br />

The last of Maxwell’s equations—the Poisson’s equation ∇⋅ E = 4 πe( n −n<br />

)/ ε0<br />

, is<br />

replaced by the quasi-neutral approximation (plasma approximation)<br />

i e<br />

ne = ni.<br />

(2.8)


46<br />

Now the simulation is self-consistent and self-contained. With appropriate choices of<br />

grid size, time step, particles per cell etc. (as will be discussed in Chapter 5) to ensure<br />

acceptable (optimal) accuracy and convergence, we are ready to study the physics of the<br />

system.<br />

In our simulations, the system is initiated with particles and fields that satisfy the<br />

Rankine–Hugoniot jump conditions and subsequently evolve into a self-sustaining shock<br />

with width of the order of an ion inertial length. As shown in Figure 2-2 (bottom panel),<br />

the shock jump is initially a step function at time zero, like in MHD simulations. As time<br />

advances, the shock gradually develops into a finite width structure that has the same<br />

features as the Leroy et al. [1982] simulation (Figure 1-6). Upstream, particles are<br />

initiated with appropriate distribution functions, depending on the particle species.<br />

Particles are continuously injected from the left wall with velocity ( uu − ud<br />

) as the<br />

simulation advances. Downstream, we have initially included a narrow region of heated<br />

plasma with plasma conditions determined by the Rankine–Hugoniot relations. As we<br />

continue to inject particles from the left hand side, the right wall boundary is set to reflect<br />

particles to ensure the balance of flux. This reflecting wall can be conveniently set up<br />

because the simulations are run in the downstream rest frame.


2.2 Simulation Results Overview<br />

47<br />

This section examines the hybrid simulation results in a general sense. We investigate the<br />

effect of changes of basic macroscopic parameters of the shock as the pickup ion relative<br />

density increases.<br />

The length of the simulated system is 400 ion inertial lengths (400 ω / ). The ion<br />

plasma frequency here is calculated from the total upstream ion density. The system is<br />

divided into 800 cells and the initial number of particles is 80,000 (about 100 particles<br />

−5<br />

per cell) for each ion component. The resistivity is set to a small value: 2× 10 (4 π / ω ) ,<br />

corresponding to a resistive length on the order 1% of the cell width. The time step is<br />

DT=0.02Ωi -1 . Here Ω i is the ion cyclotron frequency based on the upstream magnetic<br />

field. We assume that all ions are protons, and that the shock is strictly perpendicular<br />

with a shock normal θ Bn = 90 ° . The upstream parameters are chosen to be consistent with<br />

the Voyager 2 observations. Here the plasma beta β sw =0.05, and uu = 8vA<br />

(or<br />

equivalently M A = 8)<br />

in the shock frame. The solar wind's upstream velocity distribution<br />

is assumed to be Maxwellian. Furthermore, the upstream pickup ion velocity distribution<br />

is taken to be a spherical shell with the shell radius equal to the upstream bulk speed u u ,<br />

as done by Liewer et al. [1993] and Lipatov and Zank [1999]. We vary the pickup ion<br />

relative density from 0% to 40% to study different scenarios. As stated in Chapter 1, we<br />

only consider the conditions near the shock. We do not simulate the foreshock region<br />

which extends about 10–15 AU upstream of the shock, nor do we extend our simulations<br />

pi c<br />

pi


48<br />

deep into the downstream heliosheath. Both these regions are more appropriately<br />

modeled by MHD simulations, considering the scale lengths.<br />

The 0% pickup ion simulation provides a baseline computation for comparison with<br />

the more realistic simulations (with pickup ions) to follow. The bottom panel of Figure<br />

2-2 shows the spatial profiles of the magnetic field B z from the 0% pickup ion simulation<br />

at constant time intervals, stacked with the t=0 profile at the bottom, extending to the<br />

−1<br />

t=48 Ω profile on the top. This plot demonstrates a well-formed shock propagating to<br />

i<br />

the left at a nearly constant speed, u = − 1.98v<br />

. The speed of the solar wind relative to<br />

the shock is u − u + u 8v<br />

.<br />

u d sh A<br />

sh A<br />

Figure 2-3 shows the spatial profiles of the magnetic field and the ion density. The<br />

density and magnetic field profiles are similar, as expected for a perpendicular shock.<br />

The shock jump is clearly evident at x=259 c / ω pi . The upstream values are nearly<br />

constant, and the downstream average values of the density and magnetic field are ~ 4.<br />

The density and magnetic field jumps, as well as the amount of downstream heating, is<br />

consistent (to within ~ 5%) with the values based on the Rankine–Hugoniot relations<br />

[Burgess, 1995].<br />

From the simulations with varying pickup ion relative density, we calculate the<br />

downstream flow speed as a function of pickup ion relative density, as shown in Figure<br />

2-4 (top panel). The shock strength rS (bottom panel) is calculated from the density jump<br />

rS=nd/nu. In the figure, the downstream flow speed increases with increasing pickup ion


49<br />

relative density whereas the shock strength decreases. As the presence of pickup ions<br />

increase the upstream effective beta (sound speed), the magnetosonic speed is increased<br />

and the steepening (compression) of the shock is reduced. Particularly, the shock strength<br />

rS is reduced to 1.85 – 2.6 (Voyager measured range) when the pickup ion relative density<br />

falls within 20 – 30%.<br />

Figure 2-3 Density profile (upper panel) and magnetic field profile (bottom panel) from a<br />

0% pickup ion simulation. The ion density n i is normalized by the upstream ion density<br />

n u ; the magnetic field B z is normalized by the upstream magnetic field B u ; and x is<br />

normalized by c / ω pi . From Wu et al. [2009]


Figure 2-4 The shock strength (bottom panel) and the downstream flow speed (upper<br />

panel) as a function of pickup ion relative density.<br />

50


51<br />

2.3 Kinetic Structure of the Termination Shock<br />

From the previous section, we know that the scaling is monotonic. For simplicity, we<br />

compare the 0% pickup ion simulation with the 20% (Voyager inference) pickup ion<br />

simulation in this section to study the change of the kinetic structure of the shock with<br />

and without pickup ions. These kinetic structures of the shocks, as revealed by such fluid<br />

terms as E , B , n, and u , are the “sums” of the kinetic behaviors of the ions involved in<br />

the termination shock, and thus offer a first order macroscopic examination of the<br />

relevant microscopic processes.<br />

In the one-dimensional simulation, Bx=0 for perpendicular shock, as discussed before.<br />

We show in Figure 2-5 that By is trivial (nearly zero). As Ez=-uxBy, Ez is also a very<br />

small quantity (same plot). To compare the kinetic structure of the strictly perpendicular<br />

shock profiles via electric and magnetic fields, we only need to compare Bz, Ex, and Ey.<br />

Figure 2-5 By and Ez from the 0% (black) and the 20% (blue) pickup ion simulations. By<br />

is normalized to the upstream magnetic field Bu and Ez is normalized by vABu/c. The<br />

simulation dimension x is normalized by ion inertial length c/ω pi .


52<br />

The profile of Bz, Ex, and Ey from the 20% pickup ion simulation are shifted such that<br />

the shock lines up with the 0% pickup ions shock, as shown in Figure 2-6. In the top<br />

panel, one sees the well established characteristics (foot, ramp, overshoot, undershoot) of<br />

a supercritical shock from both simulations. There is a sharp peak (overshoot) of the<br />

magnetic field at the shock front (x ~262) that is much larger than the average<br />

downstream field. The magnitude of the overshoot from the 20% pickup ion simulation is<br />

significantly smaller than that of the 0% pickup ion simulation. There is an extended foot<br />

just upstream of the shock front. The length and magnitude of the foot from the 20%<br />

pickup ion simulation is larger than those of the 0% pickup ion simulation. The<br />

steepening at the shock front (where the ramp is) is weaker for the 20% pickup ion<br />

simulation than for the 0% pickup ion simulation. The upstream magnetic field Bz<br />

fluctuates as we include pickup ions, because the pickup ions increase the upstream<br />

velocity anisotropy and induce more upstream waves. In the downstream, both<br />

simulations show a number of large, quasi-periodic oscillations. The average frequency<br />

of these oscillations from the 20% pickup ion simulation is much smaller than that of the<br />

0% pickup ion simulation. Hence, the oscillating features downstream of 20% pickup ion<br />

simulation have longer wavelengths as compared with the 0% pickup ion simulation.<br />

In calculating Ex, we subtract the non-physical contributions from the hot plasma near<br />

the reflecting boundary of the wall. One sees that the average Ex upstream is<br />

approximately zero, as is the average Ex downstream. There is a decrease of Ex in the foot<br />

due to the increase in the electrostatic potential, Ex = - ∇ϕ. This Ex slows the upstream<br />

flow, which creates the compressed foot region as seen in the Bz profile. There is a sharp


53<br />

dip and peak in Ex at the shock front that is due to and<br />

∇pe<br />

E <br />

x = − uy× Bz − ⋅ x.<br />

Similarly,<br />

en<br />

the magnitude of Ex is reduced with the inclusion of pickup ions, as well as the oscillating<br />

features. And the extended foot of shock is more prounced in the 20% pickup ion<br />

simulation.<br />

Figure 2-6 Profiles of Bz (top panel), Ex (middle panel) and Ey (bottom panel) from the<br />

0% (black) and 20% (blue) pickup ions simulations. The magnetic field is normalized by<br />

the upstream magnetic field Bu; the electric fields are normalized by vABu/c, and the<br />

spatial length x is normalized to the ion inertial length c/ωpi.


54<br />

Profiles in Ey show the expected upstream motional electric field<br />

( u )<br />

E = − u B (recall that we are working in the downstream frame). This motional<br />

y u d z<br />

electric field is reduced significantly at the foot as the flow is decelerated by the reflected<br />

ions. Ey has a large negative spike at the shock front that corresponds to both the<br />

motional electric field and the induced electric field<br />

2 ∂BZ<br />

E E<br />

<br />

y = Ey−mot + y−ind =− ux× Bz −c ⋅ ⋅DT⋅y that lead to the large overshoot in Bz. In<br />

∂x<br />

the downstream region, the average value of Ey is zero, because the flow speed is zero<br />

(downstream rest frame) and therefore the motional electric field Ey-mot becomes zero.<br />

The downstream Ey oscillates at the same frequency but the opposite direction as Bz due<br />

to the induced electric field (Ey-ind).<br />

Figure 2-7 shows additional results of the kinetic structure of the two shocks (0% and<br />

20% pickup ion shocks) via the fluid densities and velocities of pickup ions, solar wind<br />

ions and all ions. Again, the z component flow velocities are not shown because these<br />

velocities have very small values for the one-dimensional simulation. Notice that Panel a<br />

(a’) and Panel b (b’) both depict the same three densities (pickup ion density: red; solar<br />

wind ion density: green; all ions density: black) however differ in the normalizations of<br />

these densities. Panel a (a’) densities are all normalized to the upstream density nu such<br />

that the magnitudes of the plotted densities can be compared. Panel b (b’) densities are<br />

normalized to each component’s upstream value such that the compression of each<br />

component is compared in the same scale. Because the unprimed panels are from the 0%


a)<br />

b)<br />

c)<br />

d)<br />

a’)<br />

b’)<br />

c’)<br />

d’)<br />

u<br />

u<br />

u<br />

u<br />

55<br />

Figure 2-7 Pickup ion (red), solar wind ion (green) and all physical ions (black) density<br />

and velocity profiles. Panel a, b, c, and d are from the 0% pickup ion simulation (where<br />

pickup ions are test particles). Panel a’, b’, c’ and d’ are from the 20% pickup ion<br />

simulation. In Panel a and a’, all densities are normalized to the upstream density (nu). In<br />

Panel b and b’, each density is normalized to its upstream value (red: nPUI / n u, PUI , green:<br />

nsw / n u sw,<br />

black: n/ n u ). The spatial length x is normalized to the ion inertial length c/ωpi.<br />

,<br />

Overshoot<br />

Dip<br />

Overshoot<br />

Dip


56<br />

pickup ion simulation where pickup ions are test particles; pickup ions make no<br />

contributions to the source terms, as can be seen in Panel a that the red line is essentially<br />

zero, and that all the green curve (solar wind ions) in the unprimed panels overlap the<br />

black curves (all ions) precisely. This is not the case in the primed panels (the 20%<br />

pickup ion simulation), where the green curves and the red curves show visible deviations<br />

and sometimes substantial local differences.<br />

In Panel a Figure 2-7, the density of the solar wind ions shows a large enhancement<br />

(overshoot) at the shock front. There is also an extended foot in the upstream shock<br />

transition region and there are downstream oscillations. These features persist in Panel a’<br />

(where there are 20% pickup ions) and are consistent with the Wu et al. [1984] specular<br />

reflection magnetic field structure for a supercritical shock (Figure 1-6). This indicates<br />

that some solar wind ions are specularly reflected to account for some of the dissipation<br />

of the shock, whether there are pickup ions or not. The difference is that with pickup ions,<br />

the length scale of the supercritical shock structure is broadened and the magnitude of the<br />

overshoot is reduced, suggesting that the solar wind reflection efficiency is reduced. Also,<br />

in Panel a’, since solar wind ions make up 80% of the density, the total density profile of<br />

the shock is determined by the solar wind density and the pickup ion density is only a<br />

small fraction of the total density. We compare pickup ion density compression at the<br />

same scale with the solar wind density compression in Panel b and b’ of Figure 2-7. In<br />

both panels, the pickup ion density rises earlier before the shock. This suggests that the<br />

extended foot (Panel a’ and b’, black curves) of the 20% pickup ion shock is caused by<br />

the pickup ions. Such an extended foot is not seen in the 0% pickup ion shock (Panel a


57<br />

and b, black curves), because as test particles pickup ions make no contribution to the<br />

source terms in this simulation. Behind the foot, the pickup ion density is nearly flat<br />

through the shock front and into the downstream, suggesting that pickup ions are not<br />

energized in the same manner as the solar wind ions for either shock [Winske et al., 2009].<br />

The small amplitude fluctuations among the nearly flat density suggest that pickup ions<br />

are mostly gyrating and drifting with almost unperturbed orbits such as those from<br />

specular reflection.<br />

The flow velocity ux (Panels c and c’) show that both the pickup ion and the solar<br />

wind ion have the same average upstream velocity (uu-ud) and the same average<br />

downstream velocity ~zero (reminder: the simulations are run in the downstream rest<br />

frame). The pickup ion upstream ux fluctuates more due to the pickup ions’ larger<br />

gyroradii. At the shock, the pickup ion ux has a rather gradual transition from (uu-ud) to<br />

zero. In contrast, the solar wind ion ux has a sharp dip (negative ux) that corresponds to<br />

the overshoot and the solar wind ion reflection. As the reflected solar wind ions are<br />

carried further downstream, after 1-2 gyroradii, the solar wind ion ux recovers from the<br />

sharp dip and displays low amplitude periodic oscillations due to the downstream solar<br />

wind ion gyrations. Again, both the dip and the oscillation magnitudes are reduced with<br />

the inclusion of the pickup ions.<br />

For the flow velocity uy (Panel d and d’), we subtract the non-physical drifting<br />

resulting from the potential created by the hot plasmas near the reflecting boundary of the<br />

wall. Therefore, in both the upstream and the downstream region, uy averages to zero.


58<br />

The fluctuations outside the shock region are caused by ion gyrations. In the shock<br />

region, there is a wide peak of positive uy that spans the entire foot region of the shock,<br />

suggesting an energization mechanism preferentially accelerating pickup ions in the y<br />

direction in the shock front. Of course the 0% pickup ion shock does not “see” this<br />

enhancement and the broadened foot, because pickup ions are test particles in this case.<br />

But pickup ions “see” the shock and display a similar response to that in the 20% pickup<br />

ion shock.<br />

The important result of Figure 2-7 is: whether we examine the 0% pickup ion<br />

simulation or the 20% pickup ion simulation, pickup ions display similar behavior that<br />

suggest a similar energization mechanism for pickup ions (will be discussed in later<br />

sections) relatively independent of the pickup ion density, as evident in the similar pickup<br />

ion features in panels b and b’, c and c’, and d and d’. Both Figure 2-6 and Figure 2-7<br />

indicate that the kinetic supercritical perpendicular shock structure is weaker and wider<br />

with the inclusion of pickup ions.<br />

2.4 Qualitative Examination of Heating via Phase-Space Plots<br />

In this section, we compare phase-space plots obtained from simulations of different<br />

pickup ion relative densities. We draw qualitative conclusions on the solar wind ion<br />

heating and reflection [Wu et al., 2009], as well as the pickup ion heating and gyration<br />

[Winske et al., 2009].


59<br />

Figure 2-8 shows a series of phase-space plots from the 0% pickup ion simulation.<br />

The shock is marked by a dashed line in Figure 2-8a and Figure 2-8b. In the figures, each<br />

dot corresponds to a simulation ion. We identify reflected solar wind ions as the particles<br />

that are streaming upstream ( v < − 0.4u<br />

and v > 0.4u<br />

) within a finite distance<br />

x u<br />

y u<br />

( 2 c / ω pi ) of the shock center. The limitation is that this method can not separate reflected<br />

ions from transmitted ions unambiguously. Because the transmitted ions can also gyrate<br />

back, a backward motion inside the shock can be due to gyration or reflection.<br />

Nevertheless, this method provides a qualitative picture of ion reflection. The transmitted<br />

solar wind ions are shown in Figure 2-8a, and the reflected solar wind ions are shown in<br />

Figure 2-8b. The v x – x and v y –x plots indicate that the gyrophases of reflected ions are<br />

∼ 180 ° off those of the transmitted solar wind ions. Further, the gyrovelocities of the<br />

reflected ions approximate the upstream flow velocity (in the shock frame). The v z – z<br />

plots show that there is no heating in the z- direction for both the transmitted and<br />

reflected solar wind ions, a consequence of the one-dimensional perpendicular shock.<br />

The vx–vy plots confirm that the upstream solar wind ions (Figure 2-8c) form a<br />

Maxwellian distribution as assumed. The velocity (vx–vy) distributions of the downstream<br />

ions (Figure 2-8d) are divided into two populations: a heated transmitted population<br />

(Figure 2-8e) and a suprathermal tail formed by the reflected ions (Figure 2-8f). Both<br />

downstream populations are gyrotropic.


60<br />

Figure 2-8 The phase-space plots of solar wind ions in the 0% pickup ion simulation. The<br />

panels on the top are the v x – x , y v – x and v z – x plots for a) transmitted solar wind ions<br />

and b) reflected solar wind ions. The panels on the bottom are v x velocity distributions<br />

and x v – v y phase-space plots for c) upstream solar wind ions, d) downstream solar wind<br />

ions, e) downstream transmitted solar wind ions, and f) downstream reflected solar wind<br />

ions. The shock is marked by a dashed line in both Panel a) and b). The solar wind phasespace<br />

plot evolves from the upstream Panel c) into the downstream Panel d) passing the<br />

shock. Empirically, we can separate the downstream population d) into the transmitted<br />

ions in Panel e) and the reflected ions in Panel f). The velocities are all normalized by the<br />

upstream Alfvén speed v A and x is normalized by c / ω pi . From Wu et al. [2009].


61<br />

Figure 2-9 presents the phase-space plots of both the solar wind ions and the pickup<br />

ions for the 0% pickup ion simulation. The pickup ions are test particles here—they are<br />

subject to the electromagnetic fields as usual, but they make no contribution to the source<br />

terms such as density and current. In other words, the test particles are acted on by the<br />

plasma but they do not react back to modify the plasma flow or shock in any way. In the<br />

figure, solar wind ions are plotted in the panels on the left: a), c) (upstream), and d)<br />

(downstream). Pickup ions are plotted in the panels on the right: b), e) (upstream), and f)<br />

(downstream). The upper panels of Figure 2-9a and Figure 2-9b display the same<br />

magnetic field profile B z from the simulation. The profile is plotted on top of both the<br />

solar wind ion panels and the pickup ion panels so that we can compare the scale length<br />

of B z with the scale lengths of both populations. As Figure 2-9f illustrates, some pickup<br />

ions are energized. However, the average heating of the pickup ions is much weaker than<br />

that of the solar wind ions. In Figure 2-9b, the energized pickup ions upstream of the<br />

shock form an extended region in the v x –x and the v y –x spaces (middle and bottom<br />

panels) as marked by the two vertical grey lines. However, as test particles, pickup ions<br />

do not affect the foot of the magnetic profile (top panel of the same figure).<br />

Figure 2-10 presents the same plot as Figure 2-9 for the 20% pickup ion simulation.<br />

In this simulation, the shock is much weaker ( r S ~ 2.5 ). There is a broader foot in front of<br />

the shock front in the B z profile, which is roughly equal to the width ( ∼ 8 pi / c ω ) of the<br />

upstream pickup ion gyroradius [Liewer et al., 1993]. This wider shock foot, due to the<br />

presence of pickup ions in front of the shock, is marked by the two vertical grey lines in


62<br />

Figure 2-9 The phase-space plots of the 0% pickup ion simulation where pickup ions are<br />

treated as test particles. Solar wind ions are plotted in the panels (a,c,d) on the left and<br />

pickup ions are plotted in the panels (b,e,f) on the right. Upper panels are magnetic field<br />

profile z B –x, x v –x phase-space plots, and v y –x phase-space plots for a) solar wind ions<br />

and b) pickup ions. Lower panels are v x velocity distributions and v x – v y phase-space<br />

plots for c) upstream solar wind ions, d) downstream solar wind ions, e) upstream pickup<br />

ions, and f) downstream pickup ions. The velocities are all normalized by the upstream<br />

Alfvén speed v A and x is normalized by c / ω pi . From Wu et al. [2009].


63<br />

Figure 2-10b. Recall that in the top panel of Figure 2-9b (the 0% pickup ion simulation),<br />

the foot is much shorter and not as visible.<br />

Figure 2-10a shows that for the 20% pickup ion simulation, solar wind ions are<br />

heated less than in the 0% pickup ion simulation (Figure 2-9a). When the pickup ion<br />

relative density is 20%, fewer solar wind ions are reflected. Assuming that the wings of<br />

the downstream solar wind distributions correspond to the more energetic reflected ions,<br />

the reflection efficiency of the solar wind in the simulation with 20% pickup ions (Figure<br />

2-10d) is significantly reduced in comparison with the reflection efficiency in the 0%<br />

pickup ion (Figure 2-9d) simulation. The average heating of the solar wind ions,<br />

however, is still stronger than that of the pickup ions due to reflection. The downstream<br />

thermal velocity of the solar wind (Figure 2-10d) is roughly 3–4 times its upstream<br />

velocity (Figure 2-10c). Some pickup ions are energized to form a heated ring<br />

distribution (Figure 2-10f), as in Figure 2-9f. The downstream pickup ions (Figure 2-10f)<br />

are heated only ~ 2 times the upstream pickup ion velocity (Figure 2-10e) because the<br />

shock is much weaker [Wu et al., 2009]. More precisely quantified heating ratios will be<br />

discussed in the next section.


64<br />

Figure 2-10 The phase-space plots of the 20% pickup ions simulation. Solar wind ions<br />

are plotted in the panels (a,c,d) on the left and pickup ions are plotted in the panels (b,e,f)<br />

on the right. Upper panels are magnetic field profile z B –x, v x –x phase-space plots, and<br />

v y –x phase-space plots for a) solar wind ions and b) pickup ions. Lower panels are v x<br />

velocity distributions and v x – v y phase-space plots for c) upstream solar wind ions, d)<br />

downstream solar wind ions, e) upstream pickup ions, and f) downstream pickup ions.<br />

The velocities are all normalized by the upstream Alfvén speed v A and x is normalized<br />

by c / ω pi . From Wu et al. [2009].


2.5 Heating and Energy Partition<br />

65<br />

In this section, we use temperature to quantify the average microscopic particle<br />

energization/heating, and pressure to quantify the macroscopic partition of energy. We<br />

conclude that the temperature of solar wind ions is raised by a larger factor than the<br />

temperature of the pickup ions; however, the majority of the dissipated energy is<br />

deposited on the pickup ions [Wu et al., 2009].<br />

Generally, the thermal pressure can be calculated from P= nkBT . The temperature, T,<br />

has the dimension of energy per unit particle and represents the average kinetic energy of<br />

these particles. The pressure, P, has the dimension of energy per unit volume. For our<br />

simulations, the temperatures are defined as follows<br />

m<br />

T = < v −< v > >+< v −< v > >+< v −< v > > (2.9)<br />

j 2 2 2 2 2 2<br />

j ( x x y y z z ).<br />

3kB<br />

Here the angular brackets denote the average value (ensemble average) of the<br />

enclosed parameter over all the ions of component j (j could be the pickup ion or the solar<br />

wind ion) and over some finite spatial region within the simulation domain; m j is the<br />

mass of ions of component j; kB is the Boltzmann constant. In the one-dimensional<br />

simulation of perpendicular shocks, as shown in Figure 2-8, there is nearly no heating in<br />

v z (the velocity component parallel to the background magnetic field). Therefore, it is<br />

more accurate to calculate the temperatures inferred from the simulation using only the<br />

v x and y<br />

v terms (velocity components perpendicular to the background magnetic field)


66<br />

in Equation (2.9). The downstream temperature calculated in this manner from the 0%<br />

pickup ion simulation satisfies the Rankine-Hugoniot relations [Wu et al., 2009].<br />

Applying the temperatures computed over a spatial extent of 60 c / ω pi upstream and<br />

downstream of the shock, we calculate the temperature jump of the solar wind ions τ sw<br />

and the temperature jump of the pickup ions τ PUI , both of which are listed in Table 2-1<br />

and plotted in Figure 2-11. In the table, nPUI / n (input column) is the relative density of<br />

upstream pickup ions. Ideally, if all the particles are transmitted, a polytropic energy<br />

equation predicts that the temperature jump across a shock should be given by (see<br />

Section 3.1.1 for derivation)<br />

τ<br />

T<br />

r<br />

d γ −1<br />

= = S .<br />

(2.10)<br />

Tu<br />

Here the polytropic index is to be decided and should not be mistaken as the adiabatic<br />

value of 5/3.<br />

In Table 2-1 and Figure 2-11, an adiabatic temperature jump corresponding to γ =5/3<br />

is presented by the τ adiabat column. An adiabatic transition involves no change in entropy<br />

across the shock. However, by definition, a shock is a transition across which there is an<br />

entropy increase. Because of this non-adiabatic nature of shock crossings, both the solar<br />

wind temperature jump τ sw and the pickup ion temperature jump τ PUI are larger<br />

thanτ adiabat . In particular, the solar wind ions go through a much larger temperature jump


67<br />

Table 2-1 Results Calculated from the Hybrid Simulations (MA = 8, β sw = 0.05).<br />

PUI<br />

nu<br />

nu<br />

τ adiabat τ sw τ PUI γ PUI η PUI<br />

Pd<br />

P u<br />

PUI<br />

Pd<br />

χ d =<br />

Pd<br />

0% 2.47 433.64 6.91 2.43 - 1632.68 -<br />

5% 2.18 271.24 5.86 2.51 0.42 43.32 0.46<br />

10% 2.04 181.30 4.75 2.45 0.59 23.82 0.65<br />

15% 1.95 88.79 4.42 2.49 0.80 15.71 0.84<br />

20% 1.86 49.07 3.85 2.44 0.87 12.13 0.90<br />

25% 1.75 48.85 3.43 2.45 0.87 9.97 0.90<br />

30% 1.67 42.69 3.04 2.44 0.84 8.14 0.89<br />

35% 1.59 35.39 2.87 2.52 0.88 7.05 0.92<br />

40% 1.50 31.51 2.59 2.56 0.89 5.76 0.93<br />

Here energy partition η j =<br />

j j<br />

j<br />

φjTd −φjTu<br />

Td<br />

[see Equation (2.12)] and temperature jump τ i i<br />

j = ,<br />

j<br />

( φT −φT<br />

)<br />

T<br />

∑<br />

i= sw, PUI<br />

i d i u<br />

where j denotes an ion component. The polytropic index γ PUI is derived from columns τ PUI and S<br />

u<br />

r using<br />

PUI 1<br />

τ PUI = rS γ − . In the 0% pickup ion simulation, pickup ions are treated as test particles. They are subject<br />

to the electricmagnetic fields but make no contributions to the source terms. Adapted from Wu et al. [2009].<br />

γ PUI =2.48<br />

γ =5/3<br />

adiabatic<br />

Figure 2-11 The simulated temperature jumps (blue: solar wind; red: pickup ion) as<br />

compared with adiabatic temperature jumps. The polytropic indexes are calculated from<br />

Equation (2.10). In the 0% pickup ion simulation, pickup ions are treated as test particles.<br />

They are subject to the electricmagnetic fields but make no contributions to the source<br />

terms.


68<br />

than the pickup ions: τ sw >> τPUI > τadiabat<br />

(Figure 2-11). Recall that the Rankine–<br />

Hugoniot relations determine the temperature jump across the shock, but do not indicate<br />

the nature of the heating, nor the partition of the heating if there is more than one plasma<br />

component. A polytropic law of the form of Equation (2.10) has often been used to<br />

characterize the heating at shocks [e.g., Feldman, 1985], where γ becomes a parameter<br />

determined from observations or simulations. We thus define γ PUI as such a parameter<br />

for characterizing pickup ion heating. The values of γ PUI from the simulations can be<br />

obtained from Equation (2.10), as shown in Table 2-1. It is apparent that γ PUI varies<br />

between 2.4 and 2.6 and does not depend on the pickup ion relative density [Wu et al.,<br />

2009]. This result will be applied in Chapter 3 to analytically describe heating and energy<br />

partition.<br />

Compared with the observed solar wind temperature jump in Table 1-3, the simulated<br />

values of τ sw in Table 2-1 are larger than the observed ones. This is because the plasma<br />

instrument on Voyager 2 does not measure reflected solar ions in the suprathermal range<br />

(~keV) of the velocity distributions and therefore underestimatesτ sw , in comparison with<br />

the simulation values [Wu et al., 2009].<br />

We define η j as the percentage of thermal energy gained by each ion component<br />

η<br />

PV −P( rV) P −Pr<br />

,<br />

( ) ( )<br />

j j j j<br />

j = d d u S d<br />

PV d d −Pu rV S d<br />

= d u S<br />

i i<br />

∑ Pd −Pr<br />

u S<br />

i= sw, PUI<br />

(2.11)


69<br />

where j denotes the ion component (the pickup ion or the solar wind ion); V d is the<br />

volume of the downstream ions; and so ( rV S d)<br />

is the ions’ upstream volume before they<br />

j j j<br />

cross the shock. As P = n kT , Equation (2.11) can be re-written as<br />

η<br />

φ T −φ<br />

T<br />

j j<br />

j =<br />

j d<br />

∑<br />

i= sw, PUI<br />

i<br />

j<br />

i<br />

d<br />

u<br />

i<br />

i<br />

u<br />

( φT −φT<br />

)<br />

(2.12)<br />

for a simulation. Here φ j is the relative density of the ion component j. The calculated<br />

η PUI (Table 2-1) from simulations shows that the net energy gain of pickup ions<br />

increases with the increasing pickup ion relative density. When the pickup ion relative<br />

density is greater than 15%, pickup ions account for more than 80% of the dissipation.<br />

This is consistent with the postulation that most energy goes to pickup ions, which<br />

Richardson et al. [2008] draws from the Voyager 2 observations [Wu et al., 2009]. This<br />

does not contradict the results that τ sw >> τ PUI.<br />

Because pickup ion starts with an initial<br />

energy that is ~1000 times that of the solar wind energy, even a relatively small jump in<br />

the temperature implies a large value of net energy gain, as illustrated in the following<br />

algebraic (unitless) analogy for a single representative pickup ion and a single<br />

representative solar wind ion:<br />

τ sw 40 >> τ pui <br />

4<br />

Δ Esw = Esw(<br />

τ sw − 1) = 1 × (40− 1) = 39,<br />

Δ EPUI = Epui(<br />

τ pui − 1) = 1000 × (4 − 1) = 4000,<br />

Δ E = 4000 >>Δ E = 39<br />

PUI sw


70<br />

The shock has more than a single pickup ion and a single solar wind ion. The complete<br />

calculation is more complex than the simple analogy, as will be analytically formulated in<br />

Chapter 3.<br />

Another way of characterizing the energy partition, for easy comparison with<br />

Voyager observations, is through the downstream thermal pressure ratio defined as<br />

PUI<br />

Pd<br />

χ d = .<br />

(2.13)<br />

P<br />

The pressure ratio χ d is displayed in Table 2-1. For the most cases, χ d<br />

approximates η PUI . Both χ d and η PUI increase as the pickup ion relative density<br />

increases. With a pickup ion ratio of 15%, χ d is about 84%—the fraction gain of<br />

dissipated energy for pickup ions, as inferred from the Voyager 2 observations by<br />

Richardson et al. [2008].<br />

2.6 Pickup Ion Energization<br />

The primary mechanisms for heating the solar wind ions are compression of the<br />

transmitted ions and shock reflection of the relatively small number of ions which are<br />

turned back upstream. Both processes have been analyzed in great detail elsewhere [e.g.,<br />

Leroy et al., 1982; Leroy, 1983; Burgess et al., 1989] and are not the emphasis here. The<br />

pickup ions, however, are much more energetic than the cool solar wind ions upstream;<br />

as a result, they gain energy at the shock in a completely different manner. Several<br />

different studies have suggested energy gains by orders of magnitude larger than the<br />

d


71<br />

upstream energy [e.g., Zank et al., 1996; Lee et al., 1996], e.g., 0.5 MeV. In contrast, we<br />

address the initial heating of the pickup ions.<br />

2.6.1 Simulated Pickup Ion Kinetics<br />

Early on in Section 2.3, we demonstrated that at the level of a fluid description, the<br />

supercritical characteristics of the termination shock are a result of solar wind ion<br />

specular reflection. We have also shown that the collective behavior of pickup ions<br />

extends the length of the termination shock foot and the spatial scales of the downstream<br />

oscillations (Figure 2-6 and Figure 2-7). This section goes beyond a fluid parameter<br />

analysis and examines the response of individual pickup ions.<br />

The simplist case is one in which that pickup ions are treated as test particles, the 0%<br />

pickup ion simulations. Figure 2-12 shows individual ion’s vx – vy phase (pickup ions:<br />

left panel; solar wind ions: right panel) from such a simulation. In the figures, each dot<br />

corresponds to a simulation ion. These ions are the ions from the vicinity of 20 c/ω pi that<br />

covers the range of the foot, the ramp, the overshoot and the undershoot of the shock as<br />

well as some partial upstream adjacent to the foot.<br />

Solar wind ion reflection is apparent in Figure 2-12, where some solar wind ions<br />

stream upstream from a collection of cold upstream ions concentrated at vx/vA ~ 5. These<br />

ions acquire significant energy through vy as they return toward the shock and develop a<br />

ring-like character. In contrast, there seem to be two streams of ions that are pulled out<br />

from the pickup ion shell at two particular gyro-phase angles, 1. vx< 0 and vy>0 (later


72<br />

Figure 2-12 Velocity phase-space plots (vx–vy) for pickup ions (left panels) and solar<br />

wind ions (right panels) within 20 c/ωi of the shock front from the 0% pickup ion<br />

simulation. All velocities are normalized by vA.<br />

will be referred to as “reversed” ions as the backward gyration that leads to energization<br />

starts in the foot region upstream of the shock, therefore the ions encounter the shock<br />

front from upstream), 2. vx>0 and vy


73<br />

If we examine a spatial series of the above phase space plots, we obtain the evolution<br />

of the ions’ velocities as they cross the shock. Such an analysis for the 20% pickup ion<br />

simulation is shown in Figure 2-13. Again, we include both pickup ions and solar wind<br />

ions, respectively. Because the shock compression is reduced with the inclusion of pickup<br />

ions in this simulation, the energization level is reduced compared with Figure 2-12.<br />

Nevertheless, the same features of the energization processes persist, indicating that the<br />

density of pickup ions does not affect the basic energization mechanism.<br />

In Figure 2-13, the red labeled pickup ion phase space corresponds to the red labeled<br />

upstream and partial foot region of the kinetic shock profile (detail description and<br />

explanation of the profile can be found in Section 2.3 and thus will not be repeated here).<br />

There is a small stream of “reversed” pickup ions in the vx0 region. As the<br />

pickup ions enter further into the foot, more “reversed” pickup ions are seen (as shown in<br />

the orange labeled phase space plot), indicating that the “reversed” ions originate from<br />

the foot region upstream of the shock. Up to this point, there is no evidence of the<br />

“crossing” ions in both the red and the orange labeled regions which include the foot but<br />

exclude the other shock structures. Interacting with the negative Ex in this upstream foot<br />

region, the “reversed” pickup ions gain energy because vxEx>0. As we move to the<br />

yellow labeled region which covers the ramp, the overshoot, and the undershoot as well<br />

as the foot, the “crossing” ions appear in the vx>0 and vy0. Both groups of pickup ions continue to gain energy as they<br />

gyrate around the local magnetic fields, as shown in the green labeled phase space. As we


74<br />

Figure 2-13 Phase space evolution of the pickup ions (first row of the vx–vy panels) and<br />

of the solar wind ions (second row of the vx–vy panels) as they cross the 20% pickup ion<br />

shock. The kinetic shock structure from Section 2.3 is zoomed in for comparison. Each<br />

panel of the evolution series corresponds to the color marked spatial region of 20 c/ωpi.<br />

The center of each adjacent panel is 5 c/ωpi apart.


75<br />

continue to move downstream, the energized pickup ions form a gyrotropic distribution<br />

that is nearly uniform in the covered velocity range, suggesting that both energization<br />

processes are continuous, just as indicated by the constant value of the pickup ion density<br />

across the shock.<br />

In comparison, some solar wind ions show an initial reversal of vx that later develops<br />

into the low density halo around the ions that are heated through transmission. The<br />

downstream gyrotropic distribution of the solar wind (purple labeled phase space plot) is<br />

rather non-uniform with the reflected ion ring bounding the denser core. This is<br />

consistent with the fact that specular reflection creates an energization gap between<br />

transmitted ions and reflected ions.<br />

Phase-space evolutions of pickup ions in a similar format as Figure 2-13 also appear<br />

in Lipatov and Zank [1999] and Chapman et al. [2005], but look significantly different.<br />

In the former case, the simulation contains strong resistive electron heating and very<br />

small spatial scales (~ electron Debye length) that generating a very large spike in Ex that<br />

accelerates pickup ions significantly through multiple reflections (surfing). In the latter<br />

case, a reformatting shock that involves large, time-varying electric fields is quoted as the<br />

reason for the acceleration of the pickup ions. Our hybrid simulations, instead, gives<br />

steady state solutions.<br />

Figure 2-13 suggests, and Figure 2-14 confirms, that there is a gyrophase-dependent<br />

energy gain process initially. The panels of Figure 2-14 illustrate the temporal evolution<br />

of pickup ions that originate from the same upstream location at the start of the


76<br />

simulation, with one ion per panel. The start time in each panel is marked with an asterisk,<br />

and the color changes from black through blue, green, yellow, orange and red; the last<br />

color corresponds to late-time downstream conditions. Except for the last panel, these<br />

trajectories are characteristic of the fraction of pickup ions that gain substantial energy at<br />

the shock. The black to green circles with relatively large average v x (~uu-ud) are<br />

upstream velocity trajectories; the transition region from green to yellow corresponds to<br />

parts of the trajectories near the shock; and the large red circles with relatively small<br />

average v x (~0, recall that the simulations are run in the downstream frame) are from<br />

downstream. These trajectories suggest that some pickup ions can achieve a significant<br />

energy gain at the shock, and the process involves both v x and v y . The mechanism by<br />

which these ions are energized [Winske et al., 2009] is the gyrophase-dependent<br />

acceleration (GPDA), discussed in the next section.<br />

2.6.2 Gyrophase-Dependent Acceleration Model for Pickup Ions<br />

To explain how the two streams of pickup ions are energized across the shock, consider<br />

the Figure 2-15 diagram. The dashed circles mark the upstream pickup ion trajectories in<br />

x–y space and vx–vy space. In the context of this figure, we consider two types of pickup<br />

ions.<br />

In the downstream rest frame, consider the first type of pickup ion well upstream of<br />

the shock (in the left side of the figure), as it gyrates in the upper right quarter of its<br />

upstream orbit in x–y space (from Point P to Point B). By nature of gyration, the ion has a


77<br />

Figure 2-14 Pickup ion trajectories from the 20% pickup ions simulation in velocity<br />

space perpendicular to the background magnetic field. Both the x v and v y are normalized<br />

to the upstream Alfvén speed v A . Each panel shows a representative ion trajectory from<br />

time zero (indicated by an asterisk) to the end of the simulation. The arrows indicate the<br />

sense of temporal progress of each trajectory. From time zero, the color of each trajectory<br />

changes from black through blue, green, yellow, orange and, near the end of the<br />

simulation, to red. From Wu et al. [2009].


78<br />

positive vx and negative vy. The ion loses energy in the foot region because ∫ vE x x 0 and Ex


“Reversed” ion<br />

B<br />

D x<br />

Ex<br />

0<br />

Ey<br />

0<br />

C<br />

vy<br />

y<br />

C<br />

D<br />

Shock front<br />

P<br />

vx<br />

79<br />

“Crossing” ion<br />

B<br />

P<br />

P<br />

B<br />

Figure 2-15 Schematic interpretation of pickup ion energy gain (downstream rest frame).<br />

Left side of the figure corresponds to a representative trajectory of a “reversed” ion.<br />

Right side of the figure corresponds to a representative trajectory of a “crossing” ion.<br />

Both curves correspond to gyromotions with increasing Lamour radii. The electric field<br />

Ex and the electric field Ey in the shock region (foot, overshoot) are also shown. The<br />

downstream oscillation these fields are not shown. Instread, thick dash lines mark the<br />

average downstream fields.<br />

Ex<br />

0<br />

Ey<br />

0<br />

D<br />

C<br />

vy<br />

Shock front<br />

y<br />

P<br />

C<br />

D<br />

B<br />

x<br />

vx


80<br />

The ion gains some energy initially by∫ vE y y >0 (for vy0) and vE x x<br />

∫ >0 (for vx


81<br />

loss), while the “crossing” ions gain some initial energy as they cross back over the shock<br />

front in a narrow region (the width of the overshoot, ~ 1–2 c/ωpi) with a large |Ey|.<br />

For both types of ion, the major share of the energy gain occurs when the ions have<br />

returned upstream of the shock front (from Point C to Point D), through the usual shock<br />

drift acceleration (SDA) mechanism (vy>0, Ey>0). That is, the drift of the ions along the<br />

motional electric field upstream of the shock. The important difference is, however, that<br />

the “crossing” ions receive an initial kick for a minor energy gain before this major<br />

energy gain, while the “reversed” ions receive an initial kick for a minor energy loss<br />

before the major energy gain. This initial energy gain/loss is a result of the ion’s first<br />

interaction with the shock at different physical locations (in x–y space), or the shock<br />

front’s first interaction with the ions at different phases (in vx–vy space). Therefore, we<br />

name our model the gyrophase-dependent acceleration (GPDA), to distinguish it from<br />

SDA in our emphasis of the initial minor energy gain/loss, as seen in velocity phase space<br />

(recall Figure 2-14).<br />

In order to examine the relative efficiency of these two processes, we have plotted<br />

Figure 2-16 from a simulation with 20% pickup ions. The “reversed” ions and the<br />

“crossing” ions are tagged and followed, respectively, as they cross the shock in the<br />

simulation. Their “final” energy is then computed some distance (~ 50 c/ ω pi )<br />

downstream of the shock front. The maximum energy achieved of the “crossing” ions is<br />

clearly larger than that gained by the “reversed” ions, as expected. More importantly,<br />

there are many more “crossing” ions with higher energy gain than “reversed” ions, due to


82<br />

the different initial kicks the ions receive from the shock that result in minor energy<br />

gain/loss as predicted by GPDA.<br />

Figure 2-16 Energy distribution histogram of the downstream pickup ions. Blue line: the<br />

distribution for “reversed” ions. Black line: the distribution for “crossing” ions. The<br />

pickup ion energy is normalized to the upstream solar wind dynamic energy.<br />

2.7 Discussion<br />

“Reversed” ions<br />

“Crossing” ions<br />

We have used the one-dimensional Los Alamos hybrid code to carry out a series of<br />

simulations of the idealized perpedicular termination shock in the presence of solar wind<br />

ions and pickup ions. Our simulations show that, although the presence of the pickup ions<br />

weakens the shock, some of the upstream solar wind ions are still reflected, helping to<br />

dissipate the energy required for the shock transition. The reflected solar wind ions gain a


83<br />

gyrotropic speed of the order of the upstream flow speed, and are swept downstream.<br />

This process gives some solar wind ions the same order of magnitude of kinetic energy as<br />

the pickup ions, so that it may be difficult to observationally separate reflected solar wind<br />

ions from pickup ions downstream. The simulations also show that the presence of the<br />

pickup ions significantly raises the magnetosonic wave speed so that the termination<br />

shock is significantly weakened. The shock strength reduces to between 2 and 2.5 when<br />

the pickup ion relative density is 20–30%. Recall that the compression ratio is about 4 for<br />

strong supercritical shocks.<br />

The simulations also show that the temperature jump of solar wind ions across the<br />

shock is significantly larger than that of the pickup ions. This does not contradict the<br />

finding (also from the simulations) that pickup ions account for more than ~85% of the<br />

dissipated energy when the pickup ion relative density is larger than 15%. The reason is<br />

that the pickup ions start with initial energies that are ~1000 times those of the solar wind<br />

ions. A small fraction of energy gain results in a large portion of the dissipated energy<br />

being transferred to pickup ions.<br />

We demonstrate that there are two types of energized pickup ions, consistent with the<br />

Lever et al. [2001] test particle calculation. Both types gain moderate energies but on<br />

average the “crossing” ions gain more energy than the “reversed” ions, consistent with<br />

the finding of Lever et al. [2001]. The acceleration efficiency of the “crossing” ions is<br />

independent of the pickup ion relative density, or the thickness of the shock ramp. Most<br />

importantly, we propose the gyrophase-dependent acceleration (GPDA) to explain the


84<br />

physical difference in the initial kick of the minor energy gain/loss for the<br />

“crossing”/”reversed” ions, a mechanism not included in the Lever et al. [2001] model.<br />

The primary method for the initial energy gain by the pickup ions at the termination<br />

shock is through the gyrophase-dependent acceleration (GPDA) process described by<br />

Winske et al. [2009]. This GPDA mechanism demonstrates that, rather than a simple<br />

reversal of the radial velocity ( v x ) as in reflection, both the radial and the transverse ( v y )<br />

velocity components play a role in the transfer of energy to the pickup ions [Winske et al.,<br />

2009]. GPDA also explains why the pickup ion energy gain is greater than the value<br />

predicted by simple compression. Only pickup ions with sizeable components of vx and<br />

vy can sample the region near the shock front where the motional electric field (Ey) has a<br />

large negative spike on the spatial scale of the width of the magnetic field overshoot.<br />

When the pick-up ions dominate the downstream pressure, the upstream foot and<br />

downstream oscillations of Bz have a longer spatial scale length, characterized in terms of<br />

the gyro-radius of the energized pickup ions. Thus, The GPDA includes SDA but<br />

emphasizes the pickup ion’s very first moment of encounter with the shock.<br />

In conclusion, from the ion energization point of view, there are three primary<br />

mechanisms for ion heating at the termination shock: 1) the compression of transmitted<br />

solar wind ions; 2) the specular reflection of some solar wind ions; and 3) the GPDA<br />

mechanism operating on the pickup ions. We can use the polytropic index γ PUI (as<br />

shown in Table 2-1) to characterize the pickup ions. The next chapter will thus be built<br />

upon the above findings.


3 Analytic Model for an Idealized Termination Shock<br />

This chapter develops an analytic multicomponent Rankine–Hugoniot model to address<br />

the partition of dissipated energy at the perpendicular termination shock. A polytropic<br />

index is introduced to characterize the shock. Based on the model, we analytically<br />

formulate the downstream multi-ion speed distribution as a superposition of Maxwellian<br />

distributions.<br />

3.1 Baseline Analytic Formula<br />

3.1.1 Heating of a Transmitted Ion<br />

To derive the heating of an ion for a single transmission across the shock, we start with<br />

the polytropic energy equation<br />

where<br />

or<br />

P nk T nk v<br />

V ∝ n,<br />

so<br />

2<br />

= B ∝ B and 1/<br />

γ<br />

PV = constant,<br />

(3.1)<br />

2 γ 2 γ−1<br />

nv n v n constant<br />

(1 / ) = / = ,<br />

(3.2)<br />

v / n = v / n .<br />

(3.3)<br />

2 γ−1 2 γ−1<br />

u u d d<br />

If we substitute shock strength r = n / n into Equation (3.3), we arrive at<br />

S d u<br />

v = v r<br />

(3.4)<br />

( γ −1)/2<br />

d u( S)<br />

.


Defining the temperature jumpτ ,<br />

τ<br />

T<br />

86<br />

v v r<br />

d<br />

2 γ −1<br />

= = ( d / u) = ( S)<br />

.<br />

(3.5)<br />

Tu<br />

Here the derivation is very general and the polytropic index should not be mistaken for<br />

the adiabatic value of 5/3. It should be determined by the shock’s characteristics, as we<br />

discuss below.<br />

3.1.2 Analysis of Upstream Mach Numbers<br />

We analyze the upstream Mach numbers to gain preliminary insights into the termination<br />

shock. The magnetosonic Mach number M MS , as the coupled Mach number of the Alfvén<br />

Mach number A M and the sonic Mach number M cs , can be written as<br />

M<br />

MS<br />

uuM AMcs M A<br />

Mcs<br />

= = = =<br />

2 2 2 2<br />

v + v M + M β γ /2+ 1 2/( β γ)<br />

+ 1<br />

A cs A cs<br />

eff eff<br />

(3.6)<br />

where β eff is the effective plasma beta (calculated from the sum of solar wind beta and<br />

the pickup ion beta). Assuming γ =5/3, MA=8 (Voyager observations), β sw =0.05<br />

(Voyager observations) and β PUI =8.53 6 , we find that MMS=2.80~Mcs=2.99. Both Mach<br />

numbers are much smaller than the Alfvén Mach number (MMS~Mcs


87<br />

the effective beta dramatically and change the shock’s characteristic from the MA-<br />

dominant to the Mcs-dominant form.<br />

3.2 Multicomponent Rankine–Hugoniot Model<br />

Because of the importance of pickup ions, we here develop an analytic model for the<br />

perpendicular termination shock, based on the Rankine–Hugoniot jump conditions, a<br />

polytropic energy equation, and specular reflection (for solar wind ions). The model<br />

computes the relative energy gain at a perpendicular shock for three proton components:<br />

transmitted solar wind ions, reflected solar wind ions, and pickup ions.<br />

3.2.1 Modified Compression Ratio<br />

To apply the Rankine–Hugoniot relations, we first need to define the upstream conditions.<br />

As noted earlier, we consider a perpendicular shock, with an upstream velocity u u and a<br />

corresponding Alfvén Mach number M A = uu / uA<br />

, where u A is the upstream Alfvén<br />

speed<br />

u = B /( μ ρ ) . We also specify the upstream solar wind temperature sw<br />

T<br />

2 2<br />

A u 0 u<br />

sw sw 2<br />

using β = 2 μ kT ρ /( mB)<br />

. For the pickup ions, we assume the idealized spherical<br />

sw 0 B u u i u<br />

shell distribution (as done in Chapter 2), with speed vshell = uu.<br />

It then follows from<br />

Equation (2.9) that<br />

PUI<br />

2 2<br />

Pu = ρPUIvPUI, u /3 = φρuuu<br />

/3,<br />

(3.7)<br />

where φ is the upstream pickup ion relative density. Therefore, the upstream thermal<br />

pressure is given by<br />

u


It is then useful to define<br />

88<br />

β φ<br />

P = − P + P = − + u<br />

(3.8)<br />

sw PUI sw<br />

2<br />

u (1 φ) u φ u [(1 φ) ] ρ .<br />

2 u u<br />

2MA3 P β φ<br />

δ = = (1 − φ)<br />

+ .<br />

(3.9)<br />

ρ<br />

2 3<br />

u sw<br />

2 2<br />

uuu MA<br />

With given upstream solar wind beta β sw , Alfvén Mach number M A and heating<br />

parameters (γ , γ PUI ), for any chosen pickup ion relative density φ we can calculate δ<br />

and the sonic Mach number<br />

Then Equation (1.7) can be rewritten as<br />

M = 1/( γδ ).<br />

(3.10)<br />

2<br />

cs<br />

2 −γ<br />

γ<br />

r − r + r + + − − + = (3.11)<br />

2<br />

( S 1)[ S ( 2 1) ( 1)] 0<br />

2 S γδ γ γ<br />

2<br />

MA MA<br />

and solved for r S . Hence, we modified the original Rankine-Hugoniot compression ratio<br />

equation with the inclusion of the pickup ion relative density [Wu et al., 2009].<br />

3.2.2 Component Pressures<br />

Both the Voyager 2 observations and our hybrid simulations illustrate that downstream<br />

pickup ions gain more energy than would be described by γ = 5 / 3 in Equation (2.10). To<br />

represent this concept in the analytic model, we assume that the thermalization of the


89<br />

pickup ions follows a polytropic energy equation on the average sense, as characterized<br />

in Chapter 2. Using γ PUI to represent the amount of pickup ion heating,<br />

Equation (3.7) and (3.12) together give<br />

PUI γ PUI PUI<br />

P r P .<br />

(3.12)<br />

d S u<br />

P = r u<br />

(3.13)<br />

PUI γ PUI 2<br />

d S φρu<br />

u /3.<br />

Our simulations further show that, even with the inclusion of pickup ions, some solar<br />

wind ions are specularly reflected at the shock. Therefore, we assume that the solar wind<br />

ions are divided into two components: a transmitted component and a reflected<br />

component. Let ε ref be the reflection efficiency of the solar wind ions: the number<br />

density of reflected solar wind ions divided by the number density of the solar wind ions.<br />

We further assume that the heating of transmitted solar wind ions obeys the polytropic<br />

energy equation, with the same γ as used in Equation (3.11). Then the transmitted solar<br />

wind population has a downstream pressure of<br />

sw−trans γ sw−transγ sw<br />

P = r P = r (1 −ε ) P ,<br />

(3.14)<br />

d S u S ref u<br />

where the superscripts “ref” and “sw-trans" represent “reflected solar wind ions” and<br />

“transmitted solar wind ions”, respectively. In addition, we can express the upstream<br />

solar wind pressure as


90<br />

P = P − P = P − u<br />

(3.15)<br />

sw PUI<br />

2<br />

u u u u φρu<br />

u /3,<br />

where we made use of equation (3.7). Equations (3.14) and (3.15) give<br />

P r ε P u<br />

(3.16)<br />

sw−trans γ<br />

2<br />

d = S (1 − ref )( u −φρu<br />

u / 3).<br />

For the reflected solar wind ions, because of specular reflection and the subsequent<br />

sw−ref energy gain in the upstream region, their downstream thermal speed is v ~ u<br />

d u<br />

sw−ref [Gosling and Robson, 1985]. In fact, our 0% pickup ion simulation gives v = 1.5u<br />

.<br />

d u<br />

The additional energy gain comes from the shock compression and other non-linear<br />

processes in the shock. To simplify the calculation, we adapt a value of 2 and rewrite<br />

sw−ref v 2u<br />

(3.17)<br />

d u<br />

for the reflected solar wind ions. It then follows that<br />

P = ρ v = ε − φ ρ u = r ε −φ<br />

ρ u (3.18)<br />

sw−ref sw−ref sw−ref 2 2 2<br />

d d ( d ) /3 ( ref (1 ) d )2 u /3 2 S ref (1 ) u u /3.<br />

The total downstream thermal pressure can then be expressed as the sum of the<br />

transmitted solar wind thermal pressure, the reflected solar wind thermal pressure and the<br />

transmitted pickup ion thermal pressure<br />

sw−trans sw−ref PUI<br />

P = P + P + P .<br />

(3.19)<br />

d d d d


91<br />

Using Equation (3.14), (3.16) and (3.18), we can rewrite the above equation as [Wu et al.,<br />

2009]<br />

P = r −ε P − φρ u + r ε − φ ρ u + r φρ u (3.20)<br />

γ<br />

2 2 γ PUI 2<br />

d S (1 ref )( u u u / 3) 2 S ref (1 ) u u / 3 S u u / 3.<br />

3.2.3 Reflection Efficiency<br />

The Rankine–Hugoniot conservation of momentum requires that<br />

B B<br />

ρ + + = ρ + + .<br />

(3.21)<br />

2 2<br />

2<br />

uuu Pu u<br />

2μ0 2<br />

dud Pd<br />

d<br />

2μ0<br />

Substituting d P with equation (3.20) and B d with Equation (1.6), we obtain<br />

B rB<br />

ρ ρ (1 )( φρ / 3) 2 (1 φ) ρ / 3 φρ / 3 .<br />

2 2 2<br />

2 u<br />

uuu + Pu + =<br />

2μ0 2 γ<br />

dud + rS −εref Pu −<br />

2<br />

uuu + rSε ref −<br />

2<br />

uuu γ PUI + rS 2<br />

uuu S u +<br />

2μ0<br />

2<br />

Dividing Equation (3.22) by u u u ρ and solving forε ref , we find<br />

(3.22)<br />

2<br />

1 1−<br />

rS<br />

φ γ γ PUI<br />

1 − + + (1 − r ) ( )<br />

2<br />

S δ + rS −rS<br />

rS 2M A<br />

3<br />

ε ref =<br />

.<br />

(3.23)<br />

γ<br />

2 rS(1 −φ)<br />

rSφ<br />

γ<br />

+ −rSδ<br />

3 3


92<br />

Because the compression ratio r S has been obtained from Equation (3.11) in Section<br />

3.2.1, we can solve for the solar wind reflection efficiency ε ref from Equation (3.23). In<br />

addition, the thermal pressure jump can be derived as<br />

γ<br />

2 (1 ) PUI<br />

P r<br />

d γ φ Sεref−φ rS<br />

φ<br />

= rS<br />

(1 −ε ref )(1 − ) + +<br />

(3.24)<br />

P<br />

3δ3δ 3δ<br />

u<br />

and the downstream pickup ion thermal pressure ratio [defined in Equation (2.13)] is [Wu<br />

et al., 2009]<br />

PUI<br />

γ PUI<br />

Pd rS<br />

φ<br />

χ = =<br />

.<br />

γ<br />

γ PUI<br />

P r (1 −ε )(3 δ − φ) + 2 r ε (1 − φ) + r φ<br />

d S ref S ref S<br />

(3.25)<br />

In order to compare model results with Voyager observations and hybrid simulations,<br />

we choose the same input values of M A = 8 and β sw = 0.05 (as in our simulations) for our<br />

evaluation of the analytic model. In addition, Table 2-1 shows that γ PUI varies between<br />

2.4 and 2.6. We assume an average value, γ PUI =2.48, independent of the pickup ion<br />

relative density, in our analytic formula. Figure 3-1 then shows model results for<br />

PUI<br />

r , P / P , and P / P as functions of the pickup ion relative density for the choice of<br />

S<br />

d u<br />

d d<br />

two different values of γ : the usual adiabatic value γ =5/3 and an empirical choice of<br />

γ =2.2; the simulation results are indicated by the diamond symbols [Wu et al., 2009].<br />

Our multicomponent Rankine–Hugoniot analysis with these input conditions shows<br />

that, when the pickup ion relative density φ increases, the compression ratio r S (see the


93<br />

top panel of Figure 3-1) and the pressure jump Pd/ P u (as in the middle panel of Figure<br />

3-1) drop as the shock weakens. In contrast, the downstream partition of pickup ion<br />

PUI<br />

energy density P / P (bottom panel) increases as the pickup ion number density<br />

d d<br />

increases. The reason is that the increase of upstream pickup ion number density raises<br />

the partition of upstream pickup ion energy density significantly. For a fixed value of the<br />

pickup ion relative density, an increase in the polytropic index γ corresponds to a<br />

PUI<br />

decrease in the pressure jump and a smaller compression ratio. Both the r S and Pd / P d<br />

panels demonstrate that with the increase of the pickup ion relative density, the simulated<br />

values trend from the γ =5/3 curve toward the γ =2.2 curve [Wu et al., 2009].<br />

Overall, as seen from the r S panel, above ~15% pickup ion relative density, the shock<br />

tends toward the γ =2.2 solution. With the γ =5/3 solution, the downstream partition of<br />

PUI<br />

pickup ion energy density P / P reaches ~100% abruptly when the pickup ion relative<br />

d d<br />

density reaches 11%. This solution is not consistent with Voyager result. So a γ =5/3<br />

solution is unlikely to be true even when pickup ion density is below 15% but large<br />

enough to change the shock’s character. Approximately when pickup ion relative density<br />

approaches 5%, a different γ should be applied. In fact, the transition from γ =5/3 to<br />

γ =2.2 should be gradual and monotonic when we raise pickup ion relative density from<br />

0% to 40%. There are many intermedium values of γ between 5/3 and 2.2. In particular,<br />

the solution of γ = 1.95 fits the simulation with 20% pickup ions.


94<br />

Figure 3-1 Compression ratio r S , pressure jump Pd / P u and downstream pickup ions<br />

PUI<br />

thermal pressure ratio Pd / P d as a function of the pickup ion relative density φ .<br />

Dashed lines are theoretical predictions when γ is set to be 5/3, the solid lines are<br />

theoretical predictions when γ is 2.2. The diamond symbols mark values from<br />

simulations. From Wu et al. [ 2009].


95<br />

PUI<br />

The downstream fraction of pickup ion energy density ( P / P ) inferred from<br />

d d<br />

Voyager 2 observations [Richardson, 2008] is approximately 85%. The values from our<br />

PUI<br />

simulations in Chapter 2 show that, for φ > 15% , P / P approaches a constant value<br />

d d<br />

of approximately 90%, quite close to the Richardson [2008] inference. If we combine this<br />

with the results of the top panel of Figure 3-1 and apply the observed average<br />

compression ratio of ~ 2, the simulations and the model at γ = 2.2 both indicate that the<br />

pickup ion relative density near the nose of the termination shock is around 25%, which<br />

is consistent with the Richardson [2008] estimate of 20% [Wu et al., 2009].<br />

The solar wind reflection efficiency ε ref is plotted in black as a function of the pickup<br />

ion relative density φ in Figure 3-2. The dashed line corresponds to γ =5/3 and the solid<br />

line corresponds to γ =2.2. The modeled reflection efficiency for the 0% pickup ion case<br />

is 25% and is consistent with earlier hybrid simulations [e.g., Leroy et al.; 1982]. The<br />

value of the reflection efficiency drops dramatically to zero when the pickup ion relative<br />

density arrives at φ =11% for the γ =5/3 solution. For the γ =2.2 solution, the reflection<br />

efficiency of solar wind ions remains notable for all values of the pickup ion relative<br />

density. When the pickup ion relative density is 20%, the model predicts that there are<br />

still a significant number of reflected solar wind ions, consistent with the phase-space<br />

plots in Capter 2 (Figure 2-10) [Wu et al., 2009].


96<br />

Figure 3-2 Solar wind reflection efficiency ε ref (%) as a function of the pickup ion<br />

relative density φ . The dashed lines correspond to γ =5/3 and the solid lines correspond<br />

to γ =2.2. Adapted from Wu et al. [2009].<br />

3.2.4 Energy Partition<br />

Applying the results from the last section, the percentage of thermal energy gain by the<br />

transmitted solar wind ions can be derived from Equation (2.11)<br />

η<br />

sw−trans γ<br />

(1 −εref )( rS−rS)( δ −φ<br />

/ 3)<br />

=<br />

.<br />

γ PUI<br />

γ<br />

φ( r − r )/3 + ( δ −φ /3)[ r (1 −ε ) − r ] + 2 ε r (1 −φ)/3<br />

S S S ref S ref S<br />

(3.26)


97<br />

Similarly, the percentages of thermal energy gain by reflected solar wind ions and pickup<br />

ions are<br />

and<br />

η<br />

sw−ref η<br />

PUI<br />

2 εref rS(1 −φ )/3 −εref rS(<br />

δ −φ<br />

/3)<br />

=<br />

,<br />

γ PUI<br />

γ<br />

φ( r − r )/3 + ( δ −φ /3)[ r (1 −ε ) − r ] + 2 ε r (1 −φ)/3<br />

S S S ref S ref S<br />

γ PUI φ(<br />

rS − rS)/3<br />

=<br />

,<br />

γ PUI<br />

γ<br />

φ( r − r )/3 + ( δ −φ /3)[ r (1 −ε ) − r ] + 2 ε r (1 −φ)/3<br />

S S S ref S ref S<br />

(3.27)<br />

(3.28)<br />

respectively. The fractions of dissipation energies acquired by the reflected solar wind<br />

( ηsw− ref<br />

, blue) and the pickup ions ( η PUI , in red) are plotted in Figure 3-3. The dashed<br />

lines correspond to γ =5/3, the solid line corresponds to γ =2.2, and the simulation results<br />

are indicated with diamond symbols. The percentage of thermal energy gain by the<br />

reflected solar wind ions decreases with increasing pickup ion relative density φ . The<br />

percentage of thermal energy gain by the transmitted solar wind ions ( ηsw− trans<br />

negligibly small and is not shown. The percentage of thermal energy gain by the pickup<br />

ions η PUI increases with an increasing pickup ion relative density. Above φ ~ 15% , the<br />

simulated η PUI is nearly constant. At relative small values of pickup ion relative density,<br />

the simulated values are more in agreement with the Multicomponent Rankine–Hugoniot<br />

model based on a γ = 5/3 shock. These values tend toward the γ = 2.2 shock solution at<br />

higher pickup ion relative densities as Figure 3-2 shown [Wu et al., 2009]. The solutions<br />

) is


98<br />

with intermedium values of γ are not shown but can be infered to fall monotonically<br />

within the range of the γ = 5 / 3 solutions and the γ = 2.2 solutions.<br />

Figure 3-3 Termination shock energy partition η (%) as a function of the pickup ion<br />

relative density φ . The dashed lines correspond to γ =5/3 and the solid lines correspond<br />

to γ =2.2. Red diamond symbols mark the percentage of heating the pickup ions gain<br />

from our simulations. Adapted from Wu et al. [2009].


3.2.5 Downstream Mach numbers<br />

The downstream Alfvén Mach number is obtained with the aid of Equation (1.6)<br />

99<br />

u μ ρ μ ρ<br />

M = = u ( ) = r u ( ) = r M . (3.29)<br />

d 0 d 1/2 −1.50 u 1/2 −1.5<br />

A, d<br />

vAd ,<br />

d 2<br />

Bd S u 2<br />

Bu<br />

S A<br />

Let the upstream Alfvén Mach number be M A = 8,<br />

for the pickup ion relative density<br />

range φ =[0, 40%]. Then, M A, d =[1.11, 2.26] (for γ =5/3) or A, d<br />

M =[1.90, 3.43] (for<br />

γ =2.2). This indicates that downstream of the termination shock, the flow is still super-<br />

Alfvénic.<br />

The downstream sonic Mach number can also be obtained analytically<br />

u ρ M<br />

M = = u = (3.30)<br />

d d 1/2<br />

cs<br />

cs, d d ( ) ,<br />

vcs γ PdrP S d / Pu<br />

where the upstream sonic Mach number M = 1/ ( γδ ) (as discussed in Section 3.2.1). The<br />

cs<br />

pressure jump Pd / P u is known from Equation (3.24). For the pickup ion relative density<br />

range φ = [0, 40%], we obtain that the downstream sonic Mach number M cs, d = [0.50,<br />

0.62] (γ =5/3) or M cs, d = [0.55, 0.69] (for γ =2.2).


100<br />

Figure 3-4 Upstream and downstream Mach numbers as functions of pickup ion relative<br />

density. The three panels on the right are subsets of the same quantities of the three<br />

panels on the left: Alfvén Mach number M A (top panel), sonic Mach number M cs<br />

(middle panel), Magnetosonic Mach number M MS (bottom panel). In the right panels, we<br />

have a better view of how the solid red lines differ from the dashed lines. The black lines<br />

are the upstream Mach numbers as a function of the pickup ion relative density φ ; the red<br />

lines are the analytically calculated downstream Mach numbers at γ =2.2 and the red<br />

dashed lines are the analytically calculated downstream Mach numbers at γ =5/3. From<br />

Wu et al. [2009].


101<br />

The magnetosonic Mach number M MS can be calculated from<br />

M = M M M + M For all of our simulations, the downstream magnetosonic<br />

2 2<br />

MS A cs A cs .<br />

Mach numbers fall within the range of [0.46, 0.60] (for γ =5/3) or [0.53, 0.67] (for<br />

γ =2.2).<br />

The analytically calculated Mach numbers are plotted in Figure 3-4 as a function of<br />

the pickup ion relative density φ . The polytropic index of γ =2.2 predicts slightly larger<br />

downstream Mach numbers than the Mach numbers predicted by γ =5/3. This is because<br />

a larger γ increases the heating, and thus lowers the compression ratio, which results in a<br />

faster downstream flow with larger downstream Mach numbers. Although the<br />

downstream flow is super-Alfvénic (or measured to be super-magnetosonic due to<br />

instrument limitations), it is still subsonic and strictly sub-magnetosonic. The termination<br />

shock has the character resembling a gas kinetic shock as opposed to the terrestrial bow<br />

shocks. The participation of pickup ions in the shock dynamics give rise to the gas kinetic<br />

nature of the termination shock [Fisk, 1996]. As indicated in Section 3.2.1, the<br />

termination shock is a sonic Mach number dominant shock.<br />

3.3 Downstream Multi-Ion Speed Distributions<br />

In this section, we build on the Multicomponent Rankine–Hugoniot model to derive<br />

functions that describe the ion speed distributions downstream of the termination shock.<br />

Recall from Section 2.5 and Figure 2-8 that the parallel velocity of a perpendicular shock


102<br />

d u<br />

from a one-dimensional simulation remains nearly unchanged f ( v ) = f ( v )<br />

consider the perpendicular speed distribution f ( v⊥ ) in this work.<br />

3.3.1 Dimensional Analysis<br />

, we only<br />

In this section, we examine if the one-dimensional Maxwellian can be generalized to a<br />

higher dimension.<br />

For the purposes of comparing with our one-dimensional simulations (where only<br />

perpendicular heating is achieved), we restrict the discussion to the perpendicular<br />

velocities vx and vy. Assuming the distribution is gyrotropic, we write<br />

where the perpendicular speed<br />

f ( v ) dv ⋅ f( v ) dv = f( v , θ ) ⋅ v dv dθ,<br />

(3.31)<br />

x x y y<br />

v = v + v .<br />

⊥<br />

2 2<br />

x y<br />

Consider the Maxwellian distribution,<br />

⊥ ⊥ ⊥<br />

2<br />

2 2 2<br />

v vy vx+ v<br />

x<br />

y<br />

− − −<br />

2 2 2<br />

vthvth vth<br />

x θ x⋅ y y = 1 ⋅ 2 x y = 1 2<br />

x y<br />

f ( v , ) dv f( v ) dv ae a e dv dv aa e dv dv<br />

So ( , )<br />

f v θ<br />

⊥ is still Maxwellian in form.<br />

2<br />

v⊥<br />

−<br />

2<br />

vth<br />

= ae vdvdθ = f( v) ⋅vdvdθ.<br />

⊥ ⊥ ⊥ ⊥ ⊥<br />

(3.32)


103<br />

3.3.2 Downstream Two-Maxwellian Speed Distribution<br />

In this section, we characterize the downstream ion speed distribution as a superposition<br />

of two Maxwellian distributions.<br />

Although the reflected solar wind ions are heated a lot more than the transmitted solar<br />

wind ions, their energies still fall to the lower end of the downstream pickup ion energy<br />

range. We make an assumption (to be tested in Chapter 4 simulations) that downstream<br />

of the shock, the solar wind ions form one Maxwellian and the pickup ions form another<br />

Maxwellian. This assumption is only valid when there are not too many reflected solar<br />

wind ions, i.e., when the pickup ion relative density is sufficiently large (as will be<br />

discussed in Section 3.4). So the downstream multi-ion speed distribution can be written<br />

as<br />

v v<br />

f( v ) = a exp[ − ( ) ] + a exp[ − ( ) ]<br />

(3.33)<br />

⊥<br />

2 2<br />

⊥ 2 ⊥ 2<br />

1 sw<br />

vth, d<br />

2<br />

PUI<br />

vth,<br />

d<br />

in the plasma frame. Here a1 and a2 are normalization constants and vary from case to<br />

sw<br />

PUI<br />

case, v th, d is the downstream thermal speed of the solar wind Maxwellian, and th, d<br />

sw<br />

downstream thermal speed of the pickup ion Maxwellian. To specify ,<br />

have to apply the results from the multicomponent Rankine-Hugoniot model.<br />

sw<br />

First, we calculate the upstream solar wind thermal speed v th, u from<br />

v is the<br />

PUI<br />

v th d and th, d<br />

v , we


β<br />

sw<br />

104<br />

2kT 2kT<br />

n k T<br />

= =<br />

m<br />

=<br />

m<br />

=<br />

B sw B sw<br />

sw B<br />

2<br />

B<br />

sw i<br />

2<br />

B<br />

i<br />

2<br />

B<br />

sw 2<br />

(1 −φ<br />

)( vth,<br />

u )<br />

, 2<br />

vA<br />

2 μ μ n m μ (1 −φ)<br />

n m<br />

0 0 sw i 0<br />

u i<br />

(3.34)<br />

where φ , as defined before, is the pickup ion relative density. Equation (3.34) can be<br />

rewritten as<br />

β<br />

v v<br />

(1 −φ<br />

)<br />

sw sw 1/2<br />

th, u = [ ] ⋅ A.<br />

PUI<br />

Similarly, the upstream pickup ion effective thermal speed v , is<br />

Then the downstream pickup ion thermal speed is<br />

th u<br />

(3.35)<br />

PUI βPUI<br />

1/2<br />

vth, u = ( ) vA.<br />

(3.36)<br />

φ<br />

γ PUI −1<br />

PUI PUI 2<br />

th, d th, u S ,<br />

v = v ⋅ r<br />

(3.37)<br />

where we apply Equation (3.4). Equations (3.36) and (3.37) together specify<br />

For the downstream solar wind thermal speed,<br />

γ PUI −1<br />

PUI βPUI<br />

1/2 2<br />

vth, d = ( ) ⋅rS⋅ vA.<br />

(3.38)<br />

φ


105<br />

v = [ ε ( v ) + (1 −ε<br />

)( v ) ]<br />

sw sw ref 2 sw trans 2 1/2<br />

th, d ref th, d ref th, d<br />

ref<br />

2 2<br />

M AvA ref<br />

sw<br />

vth, u<br />

2 1 1/2<br />

rS γ<br />

− −<br />

= ε + −ε<br />

−<br />

[ (2 ) (1 )( ) ] ,<br />

(3.39)<br />

where we apply Equation (3.17) and (3.4). Equations (3.36) and (3.37) together specify<br />

β<br />

v ε M ε r v<br />

(1 −φ<br />

)<br />

sw 2 sw γ −1<br />

1/2<br />

th, d = [2 ref ⋅ A + (1 − ref ) ⋅ ⋅ S ] A.<br />

(3.40)<br />

In Equation (3.38) and Equation (3.40), the compression ratio rS and the solar wind<br />

reflection efficiencyεref can be obtained analytically from Equations (3.11) and (3.23).<br />

The pickup ion polytropic index γ PUI is set to be 2.48 as done before. The other<br />

polytropic index γ is varied from 5/3 to 2.2 with an intermediate value γ =1.95 which<br />

corresponds to the 20% pickup ion simulation (Section 3.2.2). With given upstream<br />

parameters β sw , β PUI , and MA, the downstream thermal velocities of the two<br />

sw<br />

Maxwellians ( ,<br />

PUI<br />

v th d and th, d<br />

v ) can be analytically solved as functions of the pickup ion<br />

relative density and the polytropic index γ . Figure 3-5 shows such solutions of the two<br />

downstream thermal speeds for the upstream parameters: β sw =0.05, β PUI =8.53, and<br />

MA=8. One should be very careful in applying this figure. As stated, the assumptions of<br />

this section’s derivation only apply when the pickup ion relative density is sufficiently<br />

large that the reflected solar wind ions are a very small percentage of total density.<br />

Figure 3-5 shows that as the pickup ion relative density φ increases, both the<br />

downstream solar wind thermal speed (top panel) and the downstream pickup ion thermal


106<br />

Figure 3-5 Downstream solar wind (top panel) and pickup ion (bottom panel) thermal<br />

speeds as functions of the pickup ion relative density and the polytropic index γ for a<br />

MA=8, sw<br />

β =0.05 perpendicular termination shock. The vertical line marks the 20% shell<br />

distributed pickup ion case, in which the γ =1.95 solutions predict that the downstream<br />

solar wind thermal speed is 1.93 vA and the downstream pickup ion thermal speed is<br />

12.20 vA, consistent with the simulation fitted thermal speeds (Chapter 4).


107<br />

speed (bottom panel) decrease, in response to the weakening of the shock. The decreasing<br />

of the solar wind thermal velocity is so fast that the γ =5/3 solution is incapable of<br />

capturing the right physics at φ >11% (consistent with Section 3.2 result). The γ =2.2<br />

solution of the downstream solar wind thermal speed is very different from the γ =5/3<br />

solution. It is then necessary to find the exact value of γ (which is 1.95) for the 20%<br />

pickup ion case that we are particularly interested in. The γ =1.95 solution (dotted<br />

curves) indicates that with 20% pickup ions (marked by the grey vertical line), the solar<br />

wind thermal speed is 1.93 vA and the pickup ion thermal speed is 12.20 vA, which imply<br />

that downstream the multi-ion speed distribution for the 20% shell distributed pickup ion<br />

case is<br />

2 2<br />

v⊥ v⊥<br />

f( v⊥ , 20%, M A = 8) = 0.54nd exp( − ) + 0.0027n exp( ).<br />

2 d − (3.41)<br />

2<br />

1.93 12.20<br />

where nd is the downstream density, the normalization factors 0.54 and 0.0027 are<br />

derived from the integration of the distribution functions and the relative densities (0.8<br />

0.8<br />

and 0.2) of the two Maxwellian components: = 0.54<br />

2<br />

v⊥ exp( − ) 2<br />

1.93<br />

0.2<br />

= 0.0027 . This relationship will be tested in Chapter 4 in simulations.<br />

2<br />

v⊥ exp( − ) 2<br />

12.20<br />

In addition, the downstream pickup ion thermal speed (bottom panel of Figure 3-5) is<br />

not as a sensitive function of γ as the downstream solar wind thermal speed. The reason<br />

and


108<br />

is that the gyrophase-dependent acceleration mechanism [Winske et al., 2009] for pickup<br />

ions is not a sensitive function of γ .<br />

3.3.3 Refined Downstream Two-Maxwellian Speed Distribution<br />

In the previous section, we make the assumption to characterize the downstream into a<br />

solar wind Maxwellian and a pickup ion Maxwellian. However, as the simulations show,<br />

the reflected solar wind ions have gained slightly more energy than those pickup ions<br />

which gain no energy (as shown in the last panel of Figure 2-14). A better description of<br />

the downstream components would be to include the low energy pickup ions into the<br />

relatively cold Maxwellian (which we refer to as Maxwellian-1) and to include the<br />

reflected solar wind ions into the relatively hot Maxwellian (which we refer to as<br />

Maxwellian-2). This method, again, only applies when the pickup ion relative density is<br />

large enough so that there are sufficient low energy pickup ions to thermalize fully with<br />

the transmitted solar wind ions.<br />

Using this prescription with Equations (3.37) and (3.39), the thermal speeds for<br />

Maxwellian-1 and Maxwellian-2 are<br />

and<br />

γ PUI −1<br />

PUI 2<br />

2 2<br />

th d th u S εrefφ εref φ<br />

v 1, = v , ⋅r − ⋅(1 − ) ⋅ M + ⋅(1 − ) ⋅ 2M<br />

(3.42)<br />

A A<br />

v [ ( M v ) (1 )( v ) r ] ,<br />

γ<br />

= ε + − ε<br />

(3.43)<br />

2 2 sw 2 −1<br />

1/2<br />

th2, d ref A A ref th, u S


109<br />

respectively. For the same parameters as used in the previous section, the modified<br />

and<br />

Max2<br />

th, d<br />

v<br />

Max1<br />

th, d<br />

v are plotted in Figure 3-6. This modified method of decomposing the<br />

Figure 3-6 Downstream Maxwellian-1 (top panel) and Maxwellian-2 (bottom panel)<br />

thermals speeds as functions of the shell distributed pickup ion relative density and the<br />

polytropic index γ for a MA=8, sw<br />

β =0.05 perpendicular termination shock. For the 20%<br />

pickup ion case, the γ =1.95 solutions predict that the downstream Maxwellian-1 thermal<br />

speed is 1.39 vA and the downstream Maxwellian-2 thermal speed is 12.26 vA, consistent<br />

with the simulation fitted thermal speeds (Chapter 4).


110<br />

distribution into two Maxwellian components reduces the temperature of the relatively<br />

cold Maxwellian notably and increases the temperature of the relatively hot Maxwellian<br />

slightly. The method also increases the normalization factor of the relatively cold<br />

Maxwellian. So the downstream distribution for the MA=8, φ =20% shell distributed<br />

pickup ion case is predicted to be<br />

2 2<br />

v⊥ v⊥<br />

f( v⊥ ,20%, MA = 8) = 0.83nd exp( − ) + 0.0027n exp( ).<br />

2 d − (3.44)<br />

2<br />

1.39 12.26<br />

This relation will also be tested with the simulations, as compared to the relation in<br />

Equation (3.41).<br />

3.4 Discussion<br />

Our Multi-Component Rankine–Hugoniot model is based on the Rankine–Hugoniot jump<br />

conditions, a polytropic energy equation, and specular reflection (for a few solar wind<br />

ions). The model computes the relative energy gain at a perpendicular shock for three<br />

proton components: transmitted solar wind ions, reflected solar wind ions, and pickup<br />

ions. Applying fitting parameters derived from our hybrid simulations to the polytropic<br />

energy equation, we can characterize the energy gains for transmitted solar wind ions and<br />

pickup ions. The reflected solar wind ion energy gain is characterized by a different<br />

parameter (which represents specular reflection) derived from the 0% pickup ion<br />

simulation. The model results are in good agreement with our simulations and are<br />

consistent with the plasma observations made by Voyager 2 at the termination shock [Wu<br />

et al., 2009].


111<br />

Our model and our simulations predict that the pickup ion relative density at the<br />

Voyager 2 crossing of the termination shock is about 25%, similar to the value of ~ 20%<br />

inferred from the Voyager data by Richardson [2008]. Comparison between the analytic<br />

results and the simulations shows that, as pickup ion relative density increases, the<br />

polytropic index increases. As the pickup ion relative density approaches 40% (from 0%),<br />

the polytropic index approaches 2.2 (from 5/3). The reason is that the pickup ions widen<br />

the shock transition region (as shown in Chapter 2 Figure 2-10). A wider shock has a<br />

longer compression region and thus enhanced heating. This increased heating,<br />

represented by γ >5/3 and γ PUI >5/3, reduces the shock strength as can be sensed from<br />

the polytropic equation that the temperature jumps scale with<br />

1<br />

rS γ − . When γ increases, rS<br />

is allowed to “relax” without having to sacrifice too much on the temperature jump. For<br />

example, the strongest shock with γ =5/3 has a temperature jump of 4 5/3-1 =2.52 for<br />

transmitted solar wind ions. To achieve the same temperature jump at γ =2.2, the shock<br />

strength only has to be rS=2.2 (2.2 2.2-1 =2.52). Essentially this change of shock<br />

characteristic is an unavoidable result of the Winske et al [2009]’s GPDA mechanism. It<br />

is the GPDA gyrating motion of pickup ion that widens the shock transition region (e.g.,<br />

the foot of the shock is determined by pickup ion gyroradius) and relaxes the shock to a<br />

state that the compression ratio can be lowered to a value ~2.<br />

The simulations and the model further show that the relative energy gains of the two<br />

upstream components are sensitive functions of the relative pickup ion density. Only<br />

when the pickup ion relative density is less than ~ 5% does more energy go into the solar


112<br />

wind reflection than the pickup ion heating; in this case, the solar wind component<br />

dominates the shock dissipation [Liewer et al., 1993]. However, for larger pickup ion<br />

relative densities, the energy gain of pickup ions becomes much greater than that of the<br />

solar wind ions, which is the case of the termination shock. Our results support the<br />

inference by Richardson [2008] that most ( ~ 90%) of the dissipation is caused by pickup<br />

ions.<br />

Further, our multicomponent Rankine–Hugoniot model allows the downstream Mach<br />

number to be formulated analytically. We find that downstream of the shock the flow is<br />

still super-Alfvénic; however, it is subsonic and sub-Magnetosonic. There is no mystery<br />

to the observation that the downstream appears to be super-magnetosonic, because the<br />

observation inferred magnetosonic Mach number excludes pickup ions and possibly<br />

reflected solar wind ions. We infer that instead of a strong Alfvénic shock such as the<br />

familiar terrestrial bow shock, the presence of the energetic pickup ions raises the<br />

magnetosonic wave speed so that the termination shock has the dual identity of a<br />

Alfvénic shock and a gas kinetic shock [Wu et al., 2009].<br />

Using the multicomponent Rankine–Hugoniot model and the general notion that the<br />

downstream distributions are better described by the superposition of at least two<br />

Maxwellians, we formulate the analytic two-Maxwellian speed distribution for the<br />

downstream. The two components approximately correspond to heated solar wind ions<br />

and heated pickup ions, assuming that each ion population is fully thermalized within<br />

itself by the shock (as will be tested by simulations in the next Chapter). The thermal


113<br />

speeds of the two groups are described as functions of the pickup ion relative density and<br />

the polytropic index γ . It is found that the downstream solar wind thermal velocity has a<br />

stronger dependence on γ while the downstream pickup ion thermal velocity is weakly<br />

dependent on γ . The reason lies in the pickup ion energization mechanism proposed by<br />

Winske et al. [2009] (Chapter 2). Because the solar wind ions are energized through the<br />

combination of specular reflection and transmission, and because specular reflection<br />

(which allows significantly more energy gain than the simple transmission) strongly<br />

depends on γ (Figure 3-2), a small change in γ greatly alters the downstream thermal<br />

velocity. For the pickup ions, the mechanism is rather different. The gyrophase-<br />

dependent acceleration (Winske et al. [2009]) does not show significant dependence on<br />

either γ , or γ PUI .<br />

As reflected solar wind ions have a speed that is slightly higher than un-energized<br />

pickup ions, the real situation is more complex. A refined two-Maxwellian speed<br />

distribution is formulated to account for the effect. The relatively cold Maxwellian is<br />

assumed to be a mixture of transmitted solar wind ions and low energy pickup ions. Both<br />

the two-Maxwellian model and the refined version only apply to situations where the<br />

pickup ion relative density is large (>=20%). When the pickup ion relative density is very<br />

small (


114<br />

observation and simulation. Our analytic approximation will still hold for the range of<br />

parameters we are concerned with.


4 Hybrid Simulations for a More Realistic Termination Shock<br />

Our simulations of an idealized termination shock (Chapter 2) use most of Voyager 2<br />

observed upstream values as inputs, but assume a perpendicular shock and vary pickup<br />

ion relative density to gain physical insight into the ion energization processes and the<br />

energy partition. In this chapter’s simulations, the pickup ion relative density will be<br />

fixed at 20%. First, we vary the upstream pickup ion velocity distribution to understand<br />

the physical consequences of such variations. Second, we vary the upstream Alfvén Mach<br />

number to investigate the associated downstream ion energy dependence. Third, we will<br />

vary the shock normal in our simulations to span the observed values, thereby<br />

investigating the physics of quasi-perpendicular shocks.<br />

4.1 Variation of Pickup Ion Velocity Distributions<br />

The pickup ions initially form a ring velocity distribution initially by virtue of the pickup<br />

process. Subsequent scattering produces broader velocity distributions such as the<br />

Vasyliunas–Siscoe distribution. In the simulations described here, we assume that the<br />

upstream pickup ion distributions may be presented as a velocity shell, a velocity sphere,<br />

a Vasyliunas–Siscoe distribution and a Maxwellian distribution, corresponding to<br />

successive levels of thermalization the pickup ions go through as they flow away from<br />

the Sun. For each of these simulations, we demonstrate that scattering near the shock is<br />

strong enough to yield almost identical distributions downstream.


116<br />

4.1.1 Analytic Derivation of Pickup ion Beta and Speed Range<br />

This section provides an analytic description of the assumed velocity specifications for<br />

the hybrid simulations of a more realistic termination shock. We assume that pickup ions<br />

gain an initial gyrovelocity that is equal to the solar wind upstream speed uu. Hence,<br />

newly born pickup ions form a ring-shaped velocity distribution with an average flow<br />

speed equal to that of the solar wind. As they flow outward toward the termination shock,<br />

the pickup ions are scattered by both particle-particle and wave-particle interactions,<br />

broadening this distribution in both pitch-angle and in speed. The upstream pickup ion<br />

distribution also depends on the interstellar neutral ion density, the charge-exchange rate,<br />

and the ionization rate along the ion trajectories as they flow away from the Sun<br />

[Vasyliunas and Siscoe, 1976].<br />

Here we consider five distinct velocity distributions corresponding to successive level<br />

of scattering which follow the initial pickup:<br />

• A ring distribution with perpendicular speed equal to the upstream flow speed,<br />

• A shell distribution with speed equal to the upstream flow speed;<br />

• A velocity sphere with maximum speed determined by the pickup ion beta;<br />

• The Vasyliunas–Siscoe distribution with maximum speed determined by the<br />

pickup ion beta;<br />

• A fully thermalized Maxwellian distribution.


117<br />

For each of these upstream distributions, we derive and calculate the effective pickup ion<br />

beta,<br />

ring<br />

β PUI ,<br />

shell<br />

β PUI ,<br />

sphere<br />

β PUI ,<br />

β , and Max<br />

β , respectively. Here VS denotes the Vasyliunas–<br />

VS<br />

PUI<br />

Siscoe distribution and “Max” denotes the Maxwellian.<br />

PUI<br />

We assume that for each distribution, the pickup ion relative density is φ and an<br />

upstream Alfvén Mach number MA is maintained constant. For a pickup ion distribution<br />

f(v), the corresponding β PUI in the solar wind frame is<br />

β<br />

PUI<br />

φ<br />

m<br />

∫<br />

<br />

∫<br />

<br />

= = ⋅ ,<br />

i nk u B<br />

3kB 2<br />

f( vvdv )<br />

2φ<br />

2<br />

f( vvdv )<br />

2<br />

Bu /(2 μ0)<br />

<br />

f( v) dv<br />

2<br />

3vA<br />

<br />

f( v) dv<br />

∫ ∫<br />

(4.1)<br />

where kB is the Boltzmann constant; μ0 is the permeability; nPUI,u is the upstream<br />

pickup ion number density, nu is the total upstream number density; Bu is the upstream<br />

magnetic field; mi is the pickup ion (here proton) mass.<br />

ring<br />

Starting with the ring distribution, f ( v , v ) = δ( v −u<br />

) δ(<br />

v )<br />

PUI ⊥ ⊥ u ,<br />

∞<br />

2 2<br />

δ( v uu ) δ(<br />

v )( v v ) v dv dv<br />

ring 2φ ∫ ⊥ − + ⊥ ⊥ ⊥<br />

0<br />

2φ 2 2 2<br />

PUI = ⋅ = u 2 2 u = M<br />

∞<br />

A<br />

3vA 3vA 3<br />

∫δ(<br />

v⊥ − uu ) δ(<br />

v) v⊥dvdv⊥ 0<br />

β φ<br />

shell<br />

For a shell distribution, f ( v) = δ ( v − u ) ,<br />

PUI u<br />

,<br />

(4.2)


118<br />

∞<br />

4<br />

δ ( v uu) v dv<br />

shell 2φ ∫ −<br />

0<br />

2φ<br />

2 ring<br />

PUI = ⋅ = ⋅ u<br />

2 2 u =<br />

∞<br />

PUI<br />

3vA 2 3vA<br />

∫δ<br />

( v − uu) v dv<br />

0<br />

β β<br />

(4.3)<br />

For the other three distributions, our procedure is to choose a velocity cutoff or a<br />

thermal speed so that the corresponding pickup ion beta matches those of Equations (4.2)<br />

and (4.3). This will provide us with pickup ion distributions with the same upstream<br />

kinetic energy in the simulations described in the next section.<br />

sphere<br />

sphere<br />

For a velocity sphere, f ( v ) = 1 at 0 < v< v and zero otherwise,<br />

β<br />

PUI<br />

sphere<br />

vu<br />

4<br />

1 vdv<br />

sphere 2φ ∫ ⋅<br />

0 2φ 3 sphere 2<br />

PUI = ⋅ ( ) ,<br />

2 sphere = ⋅ v<br />

v<br />

2 u<br />

3v u<br />

A 3v 5<br />

2<br />

A<br />

∫<br />

0<br />

1⋅vdv<br />

In order for Equation (4.4) to be equal to ring<br />

β PUI , the upper limit of the sphere has to be<br />

u<br />

(4.4)<br />

sphere 5<br />

vu = uu.<br />

(4.5)<br />

3<br />

For a velocity distribution described by Vasyliunas and Siscoe [1976],<br />

⎧ 3n<br />

β r<br />

λ<br />

f v v v<br />

2<br />

PUI H 11<br />

VS ( ) = ⎪<br />

exp[ − ],0 ≤ 4 3/2 3/2<br />

⎨8<br />

πur<br />

u TS( v/ uu) rTS( v/ uu)<br />

⎪<br />

VS<br />

0, v> vu<br />

≤<br />

VS<br />

u<br />

⎩<br />

(4.6)


119<br />

where nH is the density of the interstellar neutrals (~0.1cm -3 ); β 1 is the ionization rate at 1<br />

astronomical unit (AU); r1 is the length of 1 AU, rts is the location of the termination<br />

shock (~94 AU); uu is the upstream solar wind speed (~300 km/s); λ is where neutrals get<br />

PUI<br />

fully ionized and fully picked up (~4 AU). The complete integration of f ( v ) in three<br />

dimensions (v 2 λ uu<br />

3/2<br />

dv) involves the exponential integral function Ei<br />

( − ( ) ) , where<br />

r v<br />

∞ −<br />

t<br />

e dt<br />

Ei( x)<br />

=−∫ (van Heemert, 1957). The exponential integral function is not<br />

t<br />

− x<br />

PUI<br />

analytically solvable. Notice that because the exponential term in f () v does not affect<br />

f () v much when v is larger than 0.1uu, we simplify Equation (4.7) to<br />

PUI<br />

VS<br />

3n<br />

β r<br />

f () v = ,0.01 ≤v≤ v .<br />

(4.7)<br />

PUI H<br />

2<br />

11<br />

VS<br />

VS 4 3/2<br />

8 πur<br />

u TS( v/ uu)<br />

u<br />

Therefore, the pickup ion beta can be analytically integrated,<br />

β<br />

VS VS<br />

vu vu<br />

PUI 4 5/2<br />

∫ fVS () v ⋅v<br />

dv ∫ v dv<br />

VS φ 0.1u φ u 0.1uu PUI = ⋅ 2 VS = ⋅ 2 VS<br />

v 3 u v<br />

v 3 u<br />

A v<br />

PUI 2 A 1/2<br />

∫ fVS () v ⋅v<br />

dv ∫ v dv<br />

φ<br />

⋅ 2<br />

3vA 7<br />

VS<br />

vu<br />

2<br />

0.1u 0.1u<br />

2 2 2 3<br />

( ) . (4.8)<br />

u u<br />

In order for Equation (4.8) to be equal to ring<br />

β PUI , the upper limit of the Vasyliunas–Siscoe<br />

distribution has to be<br />

VS<br />

TS<br />

VS


120<br />

VS 7<br />

vu = uu.<br />

(4.9)<br />

3<br />

v / vth<br />

For a Maxwellian distribution, f () v ce −<br />

= , where c is a constant. Hence,<br />

β<br />

Max<br />

PUI<br />

∞<br />

∫<br />

2 2<br />

−v<br />

/ vth<br />

4<br />

2 2<br />

2φ = ⋅<br />

e ⋅v ⋅dv<br />

2φ 5<br />

= ⋅ v ,<br />

e ⋅v ⋅dv<br />

Max<br />

0<br />

2<br />

PUI 2<br />

3vA ∞<br />

2 2<br />

−v<br />

/ vth<br />

2<br />

2<br />

3vA 2<br />

th<br />

In order for Equation (4.10) to be equal to ring<br />

β PUI , vth has to satisfy<br />

∫<br />

0<br />

(4.10)<br />

2<br />

vth = uu.<br />

(4.11)<br />

5<br />

If we set φ =20% and MA=8, then β PUI =8.53 for all the above five pickup ion<br />

distributions.<br />

Below, we show the analytic solutions of the perpendicular distribution function<br />

v<br />

u<br />

( fPUI ( v⊥) = ∫ f( v) ⋅dv<br />

) for β PUI =8.53 pickup ions, applying the above prescriptions for<br />

v<br />

− u<br />

four cases we will examine soon in the simulations (next section).


121<br />

Figure 4-1 The distribution function fPUI ( v⊥ ) versus v⊥ for the pickup ion shell (black),<br />

the pickup ion sphere (green), the Vasyliunas–Siscoe distributed pickup ions (orange),<br />

and the Maxwellian pickup ions (red) at the same β PUI =8.53. The panel on the left is<br />

ploted in linear-linear scale and the panel on the right is the same plot in linear-log scale.<br />

The pickup ion’s perpendicular velocity vperp is normalized to the upstream velocity uu .<br />

From Wu et al. [in preparation].<br />

For the pickup ion shell case,<br />

∞<br />

∫ δ (<br />

2 2<br />

v + v⊥−uu) ⋅dv<br />

PUI −shell ⊥<br />

−∞ = ∞<br />

2<br />

∫ v ⋅δ( v−uu) ⋅dv<br />

−∞<br />

f ( v ) .<br />

(4.12)<br />

To simplify the analytic solution, we normalized all velocities to uu, the integration<br />

becomes<br />

∞<br />

v<br />

2<br />

fPUI −shell ( v⊥) = ∫ δ ( v−1) ⋅ ⋅ dv=<br />

.<br />

(4.13)<br />

v −v 1−v<br />

2 2 2<br />

−∞ ⊥ ⊥


122<br />

As the black line (Figure 4-1) shows, the shell pickup ion distribution peaks at v ⊥ =uu,<br />

drops sharply below uu, has a small value at smaller v ⊥ , and is zero when v ⊥ is larger<br />

than uu.<br />

For the pickup ion sphere<br />

0<br />

5/3<br />

dv<br />

fPUI −sphere( v⊥<br />

) =<br />

− 5/3<br />

5/3<br />

2<br />

v ⋅ dv<br />

= 18/5.<br />

∫<br />

∫<br />

(4.14)<br />

As the green line (Figure 4-1) shows, the sphere pickup ion distribution is flat and has a<br />

wider range with the v ⊥ upper limit 5/3uu.<br />

For the Vasyliunas–Siscoe distributed pickup ions,<br />

f ( v ) =<br />

PUI −VS ⊥<br />

2<br />

7/3<br />

∫<br />

0.1<br />

2 2 3/2<br />

⊥<br />

7/3 2<br />

0.1<br />

1<br />

( v + v )<br />

∫<br />

v<br />

v<br />

3/2<br />

⋅dv<br />

⋅dv<br />

<br />

1 3 5 2 1 3 5<br />

2<br />

= 10.222 2F1[ , , , −100 v⊥] −2.615 2F1[ , , , −0.429<br />

v⊥].<br />

4 4 4 4 4 4<br />

(4.15)<br />

The integration of Equation (4.15) results in Gauss’s hypergeometric function 2F1, which<br />

can be solved with the Mathematica program numerically. As the orange line (Figure 4-1)<br />

shows, the Vasyliunas–Siscoe pickup ion distribution has a wider range than the pick up<br />

ion sphere, with the v ⊥ upper limit 7/3uu.


For the Maxwellian pickup ions,<br />

123<br />

∞ 5 2 2<br />

− ( v⊥+ v)<br />

2<br />

∫<br />

e ⋅dv<br />

5 2<br />

− v⊥<br />

2<br />

−∞<br />

fPUI −Max ( v⊥) = = 10 e .<br />

∞ 5 2<br />

− ( v )<br />

2 2 v ⋅e ⋅dv<br />

∫<br />

−∞<br />

<br />

(4.16)<br />

As the red line (Figure 4-1) shows, the Maxwellian pickup ion distribution has the widest<br />

range but as it drops off very fast it is not obvious at large v ⊥ .<br />

4.1.2 Simulated Downstream Speed Distribution<br />

The hybrid simulations described in this section utilize the isotropic shell, sphere,<br />

uniform sphere and the Vasyliunas–Siscoe pickup ion distributions as well as the velocity<br />

range discussed in the last section as input to study the ion speed distribution downstream<br />

of the perpendicular termination shock. The ring pickup ion distribution is not simulated<br />

here because the distribution is highly anisotropic and should be simulated using a three-<br />

dimensional hybrid code as will be discussed in Chapter 7.<br />

The Los Alamos hybrid code used here is the same as that used in Chapter 2 with the<br />

ions represented as superparticles and the electrons described by a zero-mass adiabatic<br />

fluid model with γ =5/3. The simulations are run on a system of length 400 ω / and<br />

−1<br />

800 cells in the x-direction; the integration time step is DT=0.02 Ω . The magnetic field<br />

is in the z-direction; spatial variations are allowed only in the x-direction, but the full<br />

three-dimensional variations of the fields and the superparticle velocities are included.<br />

ci<br />

pi c


124<br />

The simulation parameters used here are similar to the ones described in Chapter 2. The<br />

initial number of particles is 80,000 (about 100 particles per cell) for each ion component.<br />

The resistivity is set to a small value:<br />

−5<br />

2 10 (4 π / ω pi )<br />

× . We assume that all ions are<br />

protons, and that the shock is perpendicular (shock normal θ Bn = 90 ° ).<br />

The upstream Alfvén Mach number is set to be MA=8, the upstream pickup ion<br />

relative density is set to be 20%; and the upstream solar wind velocity distribution is<br />

assumed to be a Maxwellian with β sw =0.05 as inferred from Voyager 2 observations<br />

[e.g., Richardson et al. 2008; Li et al., 2008]. The upstream β PUI pickup ion beta is set to<br />

be 8.53 for all four of the upstream distributions used here.<br />

The results of the simulations are listed in Table 4-1. Both the compression ratio rS<br />

and the downstream bulk speeds are relatively independent of the upstream pickup ion<br />

velocity distribution. The magnetic field profiles, as shown in Figure 4-2, illustrate that<br />

the simulations show very similar shock profiles in magnitude and oscillations, also<br />

independent of the upstream pickup ion distributions.


125<br />

Table 4-1 Results Calculated from the Hybrid Simulations (MA=8, β sw =0.05, β PUI =8.53)<br />

Pickup ion<br />

velocity distribution<br />

Shell Sphere Vasyliunas–Siscoe Maxwellian<br />

compression ratio rS 2.47 2.38 2.38 2.39<br />

ud/vA 2.93 2.97 3.02 3.04<br />

nd,sw /nu 1.98 1.95 1.99 1.97<br />

nd,PUI /nu 0.50 0.49 0.49 0.50<br />

Figure 4-2 The magnetic field profile (Bz) of the four simulations. Black: the simulation<br />

with a pickup ion velocity shell. Green: the simulation with a pickup ion velocity sphere.<br />

Orange: the simulation with the Vasyliunas–Siscoe distributed pickup ions. Red: the<br />

simulation with the Maxwellian distributed pickup ions. From Wu et al. [in preparation].


126<br />

The perpendicular speed distributions (un-normalized) upstream and downstream of<br />

the shock are plotted in their respective plasma frames, as shown in Figure 4-3. Both<br />

distributions are displayed in the log scale and the linear scale for the x axis. For the y<br />

axis, all the four panels are plotted in log scale. Upstream of the shock, the solar wind<br />

forms a Maxwellian close to the y axis for all four simulations. The velocity shell (black)<br />

have a speed peak at ~8 vA while the velocity sphere (green) forms a flat plateau beyond<br />

the solar wind Maxwellian. Despite these differences, all four downstream ion speed<br />

distributions have a two-Maxwellian character: a relatively cold Maxwellian (which we<br />

refer to as Maxwellian-1) and a relatively hot Maxwellian (which we refer to as<br />

Maxwellian-2). Maxwellian-1 resulted from all the four simulations are nearly identical,<br />

while Maxwellian-2 differentiates slightly at the higher energy end for the four<br />

simulations. Although the macroscopic electric fields and magnetic fields in a one-<br />

dimensional simulation are unable to isotropize downstream ions, they strongly scatter<br />

the ions into a downstream distribution that is relatively independent of the upstream<br />

pickup ion distribution.


127<br />

Figure 4-3 The upstream and the downstream speed distributions in their respective<br />

plasma frames. Black: the simulation with a pickup ion shell. Green: the simulation with<br />

a pickup ion sphere. Orange: the simulation with the Vasyliunas–Siscoe distributed<br />

pickup ions. Red: the simulation with the Maxwellian distributed pickup ions. Upper left<br />

panel: upstream speed distribution in log-log scale. Upper right panel: upstream speed<br />

distribution in log-linear scale. Lower left panel: downstream speed distribution in loglog<br />

scale. Lower right panel: downstream speed distribution in log-linear scale. From Wu<br />

et al. [in preparation].


128<br />

Table 4-2 Properties of the two downstream Maxwellians from the four simulations<br />

(MA=8, β sw =0.05, β PUI =8.53)<br />

Pickup ion<br />

velocity distribution<br />

Shell Sphere Vasyliunas– Siscoe Maxwellian<br />

nmax1/nd 0.79 0.80 0.80 0.79<br />

nmax2/nd 0.21 0.21 0.20 0.21<br />

nmax1 /nu 1.95 1.94 1.99 1.96<br />

nmax2 /nu 0.53 0.50 0.49 0.51<br />

T 45.11 41.59 43.25 49.00<br />

Tmax1 /<br />

Tmax2 /<br />

sw<br />

u<br />

T 2266.56 2394.77 2407.63 2252.26<br />

sw<br />

u<br />

vth1 /vA 1.68 1.61 1.64 1.75<br />

vth2 /vA 11.90 12.23 12.27 11.86<br />

T 61.52 53.56 56.85 53.56<br />

Tsw,d /<br />

TPUI,d /<br />

sw<br />

th, d<br />

sw<br />

u<br />

T 2239.18 2394.41 2353.45 2394.42<br />

sw<br />

u<br />

v /vA 1.96 1.83 1.89 1.83<br />

v /vA 11.83 12.23 12.13 12.23<br />

PUI<br />

th, d<br />

Td/ T 497.05 521.73 516.17 521.73<br />

sw<br />

u<br />

Note: As discussed in Chapter 2, we calculate the downstream temperature using only the perpendicular<br />

velocity components for the two Maxwellians.


129<br />

We divide the distributions into two Maxwellian components at '<br />

v =4.4 vA. This<br />

⊥<br />

value of<br />

'<br />

v is empirically chosen such that the relatively cold Maxwellian has 20% of the<br />

⊥<br />

downstream density, and such that we can test Sections 3.3.2 and 3.3.3 predictions. The<br />

relative densities of the two simulated Maxwellian components are listed in Table 2 for<br />

the MA=8 simulations: nmax1 for Maxwellian-1 and nmax2 for Maxwellian-2, both in the<br />

units of the downstream density nd and the upstream density nu. Compared with the<br />

downstream solar wind and pickup ion densities nd,sw and nd,PUI in Table 4-1, we find that<br />

nmax1 ≈ nd,sw and nmax2 ≈ nd,PUI as expected, suggesting that most of Maxwellian-1 ions are<br />

heated solar wind ions and most of Maxwellian-2 ions are heated pickup ions. However,<br />

an examination of the temperatures (or effective thermal speed) shows that although<br />

Tmax2 ≈ TPUI,d (or vth2 ≈ ,<br />

PUI<br />

sw<br />

v th d ) for each simulation, Tmax1 < Tsw, d (or vth1 < th, d<br />

v ) notably.<br />

This is because the relatively low energy pickup ions (as suggested by Section 3.3.3)<br />

instead of the reflected solar wind ions (as suggested by Section 3.3.2) participate in the<br />

thermalization process with the transmitted solar wind ions to form Maxwellian-1. The<br />

reflected solar wind ions (which have more energy than the low energy pickup ions)<br />

instead fall into Maxwellian-2. In addition, the four simulated speed distributions have<br />

nearly the same effective downstream temperatures, as required by conservation of<br />

energy. Given that Maxwellian-1 has nearly the same density as the solar wind ions, we<br />

can apply section 3.3.3 result (refined two-Maxwellian) multiplied by a normalization<br />

factor to fit the distributions (Figure 4-4). We also apply the two-Maxwellian result of


130<br />

section 3.3.2 multiplied by a normalization factor to fit the distributions as a comparison<br />

(Figure 4-5).<br />

Section 3.3.4 fit Section 3.3.4 fit<br />

Figure 4-4 The fitting (orange line) of the downstream distributions (in their respective<br />

plasma frames) with the two maxwellian components as derived from Section 3.3.3. The<br />

left panel is plotted in log-log scale and the right panel is plotted in the log-linear scale.<br />

The dotted lines are downstream speed distributions from the simulations. Black dots: the<br />

simulation with a pickup ion ring shell. Green: the simulation with a pickup ion sphere.<br />

Orange: the simulation with the Vasyliunas–Siscoe distributed pickup ions. Red: the<br />

simulation with the Maxwellian distributed pickup ions. Right panel: log-linear scale.<br />

Left panel: log-log scale. The perpendicular speed vperp is in the unit of the upstream<br />

Alfvén speed vA. From Wu et al. [in preparation].


Section 3.3.3 fit<br />

131<br />

Section 3.3.3 fit<br />

Figure 4-5 The fitting of the downstream distributions (in their respective plasma frames)<br />

with the two maxwellian components as derived from Section 3.3.2. The left panel is<br />

plotted in log-log scale and the right panel is plotted in the log-linear scale. The dotted<br />

lines are downstream speed distributions from the simulations. Black dots: the simulation<br />

with a pickup ion ring shell. Green: the simulation with a pickup ion sphere. Orange: the<br />

simulation with the Vasyliunas–Siscoe distributed pickup ions. Red: the simulation with<br />

the Maxwellian distributed pickup ions. Right panel: log-linear scale. Left panel: log-log<br />

scale. The perpendicular speed vperp is in the unit of the upstream Alfvén speed vA. From<br />

Wu et al. [in preparation].<br />

Table 4-3 The goodness (R) of the two-Maxwellian and the refined two-Maxwellian fits<br />

for the four simulations (MA=8, β sw =0.05, β PUI =8.53)<br />

Pickup ion<br />

velocity distribution<br />

Shell Sphere Vasyliunas– Siscoe Maxwellian<br />

Refined two-Maxwellian 0.988 0.993 0.994 0.988<br />

Two-Maxwellian 0.954 0.961 0.950 0.958


132<br />

2<br />

The goodnesses ( κ =R) of the fits to the simulations are listed in Table 4-3; the<br />

refined two-Maxwellian formula fits the MA=8 and φ =20% simulations much better as<br />

can be inferred from the goodness of the fits and the figures. The refined formula<br />

describes the mixing region of the two Maxwellian components rather poorly, as can be<br />

seen from the figure. This will be discussed at the end of this chapter’s discussion section.<br />

4.2 Variation of Alfvén Mach Number<br />

In this section, the upstream pickup ion velocity distribution is chosen to be a Vasyliunas-<br />

Siscoe population with 20% relative density. The upstream Alfvén Mach number is<br />

varied to investigate how this parameter changes the downstream ion properties.<br />

Assuming that the upstream magnetic field variation is very small, the variation of the<br />

upstream Alfvén Mach number corresponds to the variation of the upstream solar wind<br />

speed as well as the variation of the upstream pickup ion beta (as implied by Equation<br />

(4.2)). As shown in Figure 4-6, when the Alfvén Mach number increases, the speed<br />

distribution broadens because the upstream flow has more energy.<br />

The temperatures of the solar wind ions and pickup ions are listed in Table 4-4. As<br />

the Alfvén Mach number ranges from MA=[6,16], the pickup ion effective thermal energy<br />

falls within the range=[600 eV, 6 keV], the energy range of interest to IBEX. We will<br />

return to these results in Chapter 6.


MA=4<br />

MA=6<br />

MA=8<br />

MA=12<br />

MA=16<br />

133<br />

Figure 4-6 The downstream ion speed distributions (in their respective plasma frames)<br />

from the simulations: MA=4 (blue), MA=6 (green), MA=8 (black), MA=12 (orange) and<br />

MA=16 (red). The left panel is in the log-log scale and the right panel is in the log-linear<br />

scale. From Wu et al. [in preparation].<br />

Table 4-4 The downstream temperatures of the simulations ( β sw =0.05, φ =20%<br />

Vasyliunas-Siscoe distributed upstream pickup ions)<br />

sw<br />

Tsw,d/ T Tsw,d (eV) sw<br />

u<br />

TPUI,d/ T u<br />

TPUI,d (keV)<br />

MA=4 5.42 2.33 467.56 0.20<br />

MA=6 20.54 8.83 1256.07 0.54<br />

MA=8 56.85 24.45 2353.45 1.01<br />

MA=12 196.20 84.366 5093.85 2.19<br />

MA=16 460.38 197.96 8680.65 3.73<br />

As calculated in Section 1.3, at the Voyager 2 observed average upstream conditions<br />

(vA=38 km/s), the upstream solar wind temperature is about T ≅ 0.43 eV.<br />

sw<br />

u<br />

MA=4<br />

MA=6<br />

MA=8<br />

MA=12<br />

MA=16


4.3 Variation of Shock Normals<br />

134<br />

Due to the Parker spiral (Chapter 1), the idealistic termination shock has been assumed to<br />

have a shock normal of 90°. In reality, Voyager 2 observed termination shock normals (to<br />

the magnetic field) to be 74° and 82°. It is safe to assume that the realistic quasi-<br />

perpendicular termination shock near the ecliptic plane has a shock normal in the range of<br />

[70°, 90°].<br />

For the simulations in this section, we chose the upstream pickup ion distribution to<br />

be of the Vasyliunas–Siscoe type with a relative density of 20%, the Alfvén Mach<br />

number MA=8, β sw =0.05 and β PUI =0.05. The shock normal θBN is set to be 70°, 80°, 90°<br />

in successive simulations. Figure 4-7 shows the magnetic field profiles of the simulations<br />

after the same run time. The Bz components of the simulations have the similar<br />

oscillating features. As the shock normal decreases, it appears that the shock propagates<br />

slightly slowly. This is caused by the fact that it takes slightly longer simulation time to<br />

form a self-sustaining shock when the shock normal decreases. Actually, the shock speed<br />

increases slightly when the shock normal reduces. The shock is slightly weaker than the<br />

strictly perpendicular shock. The magnitude of Bz also decreases slightly with decreasing<br />

shock normal because Bz = Bu⋅ CosθBN.<br />

The other transverse component By is found to<br />

increase with decreasing shock normal. In the quasi-perpendicular shock, the non-zero<br />

parallel component of magnetic field Bx introduces Ez uy x B = − × . This oscillating (uy


135<br />

oscillates due to gyration) Ez gives rise to the oscillating By component, as required by<br />

z By= E ∂<br />

⋅ DT<br />

∂x<br />

. When θBN = 90°, Bx is zero and so is By, as seen in Chapter 2.<br />

θBN = 70°<br />

θBN = 80°<br />

θBN = 90°<br />

θBN = 70°<br />

θBN = 80°<br />

θBN = 90°<br />

Figure 4-7 The magnetic field profile (Bz, Bx) of the three simulations with different<br />

shock nomal. Black: the simulation with θBN = 90°. Orange: the simulation with θBN = 80°.<br />

Blue: the simulation with θBN = 70°.


136<br />

The downstream ion speed distributions, as shown in Figure 4-8, tend toward slightly<br />

lower energy when the shock normal decreases. However, the change, again, is very<br />

small, unlike the changes caused by the variation of Alfvén Mach numbers shown in the<br />

previous section.<br />

Figure 4-8 The downstream ion speed distributions (in their respective plasma frames)<br />

from the three simulations: θBN = 90° (black), θBN = 80° (orange) and θBN = 70° (blue).<br />

The left panel is in the log-log scale and the right panel is in the log-linear scale.<br />

4.4 Discussion<br />

θBN = 70°<br />

θBN = 80°<br />

θBN = 90°<br />

θBN = 70°<br />

θBN = 80°<br />

θBN = 90°<br />

We first carry out four different simulations with four distinct upstream pickup ion<br />

distributions (shell, sphere, Vasyliunas–Siscoe, Maxwellian) with the same temperature.<br />

The simulated compression ratio, the magnetic field profile and the speed profile all<br />

indicate that all four shocks are nearly identical, despite the different prescription for the<br />

upstream pickup ions. The pickup ions’ interaction with the electric and magnetic fields


137<br />

inside the shock allows them to thermalize into nearly identical downstream speed<br />

distributions, independent of the upstream distributions. This result is not unexpected,<br />

given the pickup ion heating mechanism (GPDA) described in Section 2.6. Therefore, we<br />

can not use the downstream information to infer the upstream pickup ion velocity<br />

distribution.<br />

The two-Maxwellian distribution and refined two-Maxwellian distribution provide<br />

good fits downstream in the MA=8 simulations. Particularly, the refined formula is more<br />

successful with an average goodness of fit equal to 0.99. However, the two-Maxwellian<br />

fit describes the mixing region of the two Maxwellian components poorly. This implies<br />

that an analytic formula derived from the simple assumption of two Maxwellian<br />

components does not capture all the correct physics of the shock. Particularly, the<br />

thermalization processes inside the shock mixes the three ion components (transmitted<br />

solar wind ions, reflected solar wind ions, and pickup ions) in a non-linear manner. This<br />

mixing should be a function of the shock strength, the upstream Mach numbers, the<br />

pickup ion relative density etc. An analytic model is unlikely to capture the full<br />

complexity of the processes. A rigorous multi-Maxwellian decomposition of the speed<br />

distribution is unlikely to be useful for practical purposes. In data analysis, it is better to<br />

fit the tail part of the distribution with a power law, as will be done in Chapter 6.<br />

To summarize, this chapter emphasizes the construction of the various downstream<br />

heated ion speed distributions from the more realistic simulations. The upstream pickup<br />

ion velocity distributions, the upstream Alfvén Mach numbers, and the upstream shock


138<br />

normals are varied independently to examine the relative importance of these three<br />

factors. We found that the upstream pickup ion velocity distribution and the shock normal<br />

have very little influence on the downstream speed distributions and the downstream<br />

temperatures. In contrast, the upstream Alfvén Mach number affects the downstream<br />

speed distribution and the downstream temperatures significantly. A large upstream<br />

Alfvén Mach number implies a large ion energy input to the shock. The downstream<br />

temperatures calculated from the MA = [6, 16] simulations fall within the range [600 eV –<br />

6 keV], the range of interest to IBEX observations, as discussed later in Chapter 6.


5 Accuracy and Convergence<br />

In Chapter 4, we studied the sensitivity of the simulation results to variations in physical<br />

parameters; in this chapter, we study the sensitivity of the simulation results to variations<br />

in numerical parameters. Therefore, the goal here is to explore the validity of our<br />

simulations. We vary the cell size and the number of super-particles per simulation cell to<br />

investigate how these parameters affect the stability as well as the physical interpretations<br />

of the simulations.<br />

5.1 Cell Size<br />

A van Neumann stability analysis shows that a simulation is stable if it satisfies the<br />

Courant condition (also called the Courant–Friedrichs–Levy condition)<br />

v⋅ DT < H<br />

(5.1)<br />

where v is the phase speed of the fastest wave in the system, DT is the computational<br />

time step and H is the cell size. This relation means that the cell size must be large<br />

enough so no signal can travel beyond one cell within one computational time step DT. If<br />

this condition is not satisfied, causality is violated: the source terms in the cell change<br />

faster than the physics such that the simulation either crashes or produces unphysical<br />

results. For a hybrid simulation, we further require that the fastest ion should not move<br />

more than one cell length in one time step. Actually it is better if the fastest ion does not<br />

move more than ½ of a one cell length in one time step. This is because the<br />

electromagnetic forces that the ion experiences are interpolated from 2 adjacent grid


140<br />

points. At the same time, the hybrid simulation time step is limited by the condition that<br />

the ion gyro-motion must be well resolved. This means that DT should be small<br />

compared with the inverse ion cyclotron frequency. A typical rule of thumb is to set<br />

DT0.1 c/ω pi [Winske and Omidi, 1993].<br />

Even when the Courant condition is satisfied, the cell size still needs to be carefully<br />

chosen. If the cell size is too big, the resolution is too low to resolve the physical<br />

processes. If the cell size is too small, DT has to be reduced to satisfied Courant condition,<br />

and the total number of super-particles has to be increased to keep the same thermal noise<br />

level (discussed in the next section). Both changes result in longer run times. In addition,<br />

numerical whistler waves can be introduced. Recall that the simulations are zero electron<br />

mass simulations. Under such an approximation, whistler waves have a strong dispersion<br />

ω ~ k 2 at short wavelengths. These fast high frequency whistler waves are artificial in a<br />

sense that when the electron mass is taken into account, as is the case in reality, the<br />

speeds of high frequency whistler waves are reduced. To suppress the artificial fast<br />

whistler waves in zero electron mass models such as the hybrid simulations, a medium<br />

cell size is preferred, as can be seen in the following examples.<br />

5.1.1 Cases without Pickup Ions<br />

To study the effect of simulation cell size, we first show results from simulations without<br />

pickup ions. The shock is therefore similar to the Earth’s bow shock, with structure that<br />

has been observationally verified. Three simulations with different cell sizes are studied:<br />

a) H= c/ω pi , b) H=0.5 c/ω pi (the cell size generally used in other chapters of this


141<br />

dissertation), c) H=0.25 c/ω pi . For case c), the time step DT is also reduced. The physical<br />

input parameters are all the same for the three cases: MA=8, θBN=90°, and β sw =0.05. This<br />

is evidently a supercritical shock that should yield a compression ratio of 4.<br />

As shown in the upper panel of Figure 5-1, a coarse spatial resolution of c/ω pi gives<br />

an incorrect downstream magnetic field profile. There is no clearly defined overshoot,<br />

because some of the downstream peaks are larger than the first peak. This simulation<br />

does not represent the downstream situation well. The compression ratio of this case<br />

fluctuates and is much larger than 4 (Figure 5-2 blue curve); this is unphysical because<br />

the maximum compression value for a supercritical perpendicular shock is rS=4 if γ =5/3<br />

[Burgess, 1995].<br />

When the cell size is reduced to 0.5 c/ω pi (middle panel of Figure 5-1), a steady<br />

shock forms. There is an overshoot at the shock front and there are downstream<br />

oscillations, in agreement with typical bow shock observations [e.g., Sckopke et al., 1983,<br />

Leroy et al., 1982]. The compression ratio is ~4 (Figure 5-2 black curve) as expected. If<br />

the cell size is further reduced to 0.25 c/ω pi (bottom panel of Figure 5-1), the<br />

downstream oscillation near the overshoot is very regular, which is not typically observed.<br />

In addition, the compression ratio is a bit small (Figure 5-2 red curve), because of more<br />

time variation of the shock structure that gives rise to slightly more heating. At even<br />

smaller cell size, artificial fast whistler waves set in. These waves increase heating and<br />

reduce the compression ratio even further. Our choice is either to adapt the medium cell


142<br />

size or to increase resistivity to suppress these unwanted numerical waves. We have<br />

chosen the former in the thesis, because it is better not to include the unphysical waves<br />

and artificially large resistivity for compensation.<br />

a)<br />

b)<br />

c)<br />

Figure 5-1 Magnetic field profiles of the three simulations without pickup ions: a) H=<br />

c/ω pi (top panel), b) H=0.5 c/ω pi (middle panel), and c) H=0.25 c/ω pi (bottom panel).


143<br />

Figure 5-2 The compression ratios as a function of time calculated from the average<br />

density of a 50 c/ω pi wide region that is 10 c/ω pi away from the overshoot in the<br />

downstream. The three simulations are presented: a) H= c/ω pi (blue), b) H=0.5 c/ω pi<br />

(black), and c) H=0.25 c/ω pi ( red curve).<br />

5.1.2 Cases with Pickup Ions<br />

In this section, shell distributed pickup ions with 20% relative density are added to the<br />

simulations described in the previous section: a). H= c/ω pi , b). H= 0.5 c/ω pi (our<br />

choice), and c). H= 0.25 c/ω pi . We add case c’) (H= 0.25 c/ω pi with higher resistivity)<br />

to show that a higher resistivity helps to suppress some numerical waves. The physical<br />

input parameters are all the same for the four cases: MA=8, θBN=90°, β sw =0.05, and<br />

β PUI =8.32.


a)<br />

b)<br />

c)<br />

c’)<br />

144<br />

Resistivity<br />

Resistivity<br />

Resistivity<br />

Resistivity<br />

η = ×<br />

π ω<br />

−5<br />

2 10 (4 / pi )<br />

η = × π ω<br />

−5<br />

2 10 (4 / pi )<br />

η = × π ω<br />

−5<br />

2 10 (4 / pi )<br />

η = π ω<br />

−4<br />

10 (4 / pi )<br />

Figure 5-3 Magnetic field profiles of the three simulations at the same resistivity but<br />

different cell sizes: a) H= c/ω pi , b) H=0.5 c/ω pi , and c) H=0.25 c/ω pi . Panel c) shows<br />

an unphysical fast moving shock caused by the small cell size. This effect can be<br />

eliminated with a larger resistivity as shown in Panel c’), where the cell size is as small as<br />

Panel c) but the resistivity is 5 times bigger the previous 3 panels.


145<br />

Figure 5-3 illustrates the magnetic field profiles of the three runs at the same real time<br />

(50 Ωi -1 ) when the shock is well formed. For the coarse cell size case (H= c/ω pi ) in Panel<br />

a), the downstream is not well resolved, as is the case a) in the previous section. For the<br />

medium cell size case (H=0.5 c/ω pi , our choice in this work) in Panel b), the supercritical<br />

shock features are well presented. For the smaller cell size case (H=0.25 c/ω pi ) in Panel<br />

c), the magnitude of the overshoot is too large and the downstream oscillation is too<br />

regular, even after a few pickup ion gyrating periods. (Here, the oscillation is associated<br />

with pickup ions, as discussed in Chapter 2.) Most disturbingly, the shock propagates too<br />

fast toward upstream, which indicates that there are unphysical (numerical) waves that<br />

over-heat the ions. This shock speed can be corrected if we increase the resistivity as<br />

shown in Panel c’). Still, the downstream oscillation is too regular and unlikely to be real.<br />

The compression ratios of the four simulations are plotted in Figure 5-4. Case a) and<br />

case b) have the same overall compression ratios. Case c) (dashed red curve) predicts a<br />

compression ratio that is too small. However, with a larger resistivity, case c’) (solid red<br />

curve) achieves the correct compression ratio as the other cases. In terms of resolution,<br />

the noise is smaller when we reduce the cell size.


146<br />

Figure 5-4 The compression ratios as a function of time calculated from the average<br />

density of a 50 c/ω pi wide region that is 10 c/ω pi away from the overshoot in the<br />

downstream. The four simulations are presented: a) H= c/ω pi (blue), b) H=0.5 c/ω pi<br />

(black), c) H=0.25 c/ω pi (dashed red curve) and c’) H=0.25 c/ω pi with higher resistivity<br />

−4<br />

(red curve) η = 10 (4 π / ω ) .<br />

5.2 Super-Particles per Cell<br />

pi<br />

The number of ions per simulation cell is varied in the section. For a simulation, the<br />

thermal noise is proportional to 1/ N cell , where Ncell is the number of super-particles per<br />

cell. Ideally if we increase Ncell, the thermal noise will decrease and the result will be<br />

more accurate. The tradeoff is that the computing time will increase significantly. For this<br />

thesis, we adopt a value of Ncell=100, as typically chosen in most codes that apply the PIC<br />

techniques.<br />

Black: H=0.5 c/ω pi<br />

Blue: H= c/ω pi<br />

Dashed Red: H=0.25 c/ω pi<br />

Solid Red: H=0.25 c/ω pi with higher resistivity<br />

η =<br />

π ω<br />

−4<br />

10 (4 / pi )


147<br />

We compare our simulations to simulations with Ncell=25, and Ncell=400. This small<br />

variation of noise level (a factor of 1/2 and of 2) hardly affects the results or the overall<br />

conclusions for the shock simulation. The shock speed and the compression ratios are<br />

nearly identical.<br />

Figure 5-5 show the magnetic field profile of the simulations. In the upstream, it is<br />

apparent that with more particles per cell, the simulation is less noisy. The magnitudes of<br />

the overshoots from the three simulations are very similar. However, for the Ncell=25<br />

simulation, the overshoot bifurcates and is not well resolved. This bifurcation is reduced<br />

when the particles per cell is increased to 100. When there are 400 particles per cell, the<br />

bifurcation disappear and there is a well resolved supercritical shock structure. The<br />

Ncell=100 solution, although not as perfect as the Ncell=400 simulation, is still appropriate<br />

for our problem, because it preserves the physics without requiring a longer run time the<br />

Ncell=400 simulation demands. In fact, even the 25 particles per cell simulation gives<br />

correct shock jump condition, despite the noisy magnetic field profile. Recall that the<br />

simulated super-particles do not correspond to real particles on a one-on-one basis. In fact,<br />

all published hybrid or PIC simulations contain orders of magnitude fewer particles than<br />

are found in real plasmas. What is preserved in those simulations is the general character<br />

of the plasma behavior, the overall physics and the scaling of the systems. Simulations<br />

with realistic particle density will not be available any time in the near future.


148<br />

Figure 5-5 Magnetic field profiles of three simulations with different numbers of particles<br />

per cell. Top panel: simulation with 25 particles per cell; middle panel: simulation with<br />

100 particles per cell; bottom panel: simulation with 400 particles per cell.


5.3 Discussion<br />

149<br />

In this chapter, we have demonstrated that the choice of cell size is crucial for the proper<br />

representation of shock physics. If the cell size is too big, the simulation resolution is<br />

inadequate and the small scale physics is not properly represented. However, if the cell<br />

size is too small, unwanted numerical errors increase. The ideal cell size should optimize<br />

the system, demonstrate the important physical processes, and produce results that are in<br />

agreement with observations. The other issue with small cell sizes is that the total amount<br />

of cells in the simulation becomes very large, thus increasing the simulation time. Often,<br />

we also need to reduce the time step. The larger number of particles and smaller time<br />

steps result in very time-consuming and costly simulations. The important issue is that in<br />

the zero electron mass hybrid simulations, whistler waves become unphysically fast if the<br />

cell size is too small. Increasing the resistivity tends to suppress short wavelength modes<br />

and makes the downstream look much too regular. It’s still a solution of the physical<br />

equations, but may not look as “real”. The cell size we chose for this thesis is<br />

H=0.5 c/ω pi .<br />

Unlike the cell size, we find that the number of super-particles per cell Ncell does not<br />

significantly affect the physical results, e.g., the compression ratio stays unchanged.<br />

However, the detail magnetic field structure is less noisy when there are more particles<br />

per cell due to the decrease of thermal noise. A typical thumb of rule is to choose 100<br />

particles per cell, as we did in this work.


150<br />

Numerical algorithms are also directly related to the accuracy and convergence of the<br />

simulation. Particularly, the solver for magnetic field has gone through several<br />

generations of change. Originally, in the early 1980’s, the magnetic field solver was<br />

implicit [Winske and Leroy, 1984]. One solves the inverse matrix for the vector potential<br />

at a time step. This method is not as accurate as explicit methods, which are more<br />

commonly used nowadays. The fourth order Runge–Kutta method is a fairly accurate (to<br />

the fourth order) explicit method. However, this method lacks convergence at large time<br />

steps. Winske and Quest [1988] presented a new method for solving the magnetic field in<br />

perpendicular shock simulations. This method divides the simulation time step DT into<br />

several sub time steps DDT=DT/Ncycle. Here Ncycle is the number of sub-cycles within a<br />

time step. Within each of these sub-cycles, the magnetic field is calculated using the<br />

fourth order Runge–Kutta algorithm. This method has achieved much success as a robust<br />

solver and is applied in our simulations. Of course, there is still room for further<br />

numerical improvement. One may try different weighting algorithms for interpolating the<br />

source terms, or include several smoothing cycles using the Hamming filter [Hamming,<br />

1977].<br />

In addition, we have checked conservation laws (the Rankine–Hugoniot relations, as<br />

discussed in detail in Chapter 3) to ensure the validity of the code before analyzing the<br />

simulations. Finally, we compute the entropy of the simulation system to make a<br />

connection with gas-dynamic and MHD shocks. Take the 20% pickup ion (Vasyliunas–


151<br />

Siscoe distribution) simulation with H=0.5 c/ω pi and 100 particles per cell. The values of<br />

entropy are calculated in each cell using<br />

S =−kB∫ dv⊥⋅v⊥f( v⊥)ln f( v⊥)<br />

(5.2)<br />

for all ions at the end of the simulation. That is, we computer an entropy value every half<br />

ion inertial length for ~100 ions. In calculating each entropy value, the velocity bin is<br />

0.125 vA wide and apart. A different bin size will not change the overall result much but<br />

will alter the details of the values. The entropy values are very noisy, calculated from<br />

binning only ~100 ions in each cell. Therefore, we apply the Hamming filter [Hamming,<br />

1977] once to smooth the values.<br />

As shown in Figure 5-6, entropy increases across the shock. There are localized<br />

decrease due to fluctuactions, similar to the local decrease shown in Gombosi [1994].<br />

This does not violate the second law of thermaldynamics, because the system’s total<br />

entropy still increases with time as ions move from the low average entropy upstream<br />

region to the high average entropy downstream region. Physically, this is what we<br />

expected because the shock crossing is an irreversible process that thermalizes these ions.<br />

The thermalization turns the ion’s distributions from the relatively cold distributions<br />

(upstream) into the relatively hot distributions (downstream). A cold distribution is<br />

perfectly ordered; all particles have similar velocities. As temperature increases, there is<br />

a greater thermal speed with greater thermal spread, which corresponds to more disorder<br />

and larger entropy. Put simply, the number of microstates increases across the shock.


152<br />

Notice that if the particle distribution is a single Maxwellian, Equation (5.2) becomes<br />

S =−k dv ⋅v<br />

f ( v )ln f ( v )<br />

Max B ⊥ ⊥ Max ⊥ Max ⊥<br />

2 2 2 2<br />

⊥/ th ⊥/<br />

th<br />

∝−k dv ⋅v<br />

e ln e<br />

B<br />

B<br />

v<br />

= k dv ⋅v e ⋅<br />

∝ T<br />

∫<br />

Max<br />

∫<br />

∫<br />

.<br />

−v v −v<br />

v<br />

⊥ ⊥<br />

2<br />

⊥ ⊥<br />

2 2<br />

−v⊥/ vth<br />

⊥<br />

2<br />

vth<br />

(5.3)<br />

Figure 5-6 Entropy (black) of the system at the end of the simulation run: 20% pickup ion<br />

shell, MA=8,<br />

overplotted.<br />

β sw =0.05, β PUI = 8.53. The temperature profile (blue) is scaled and


153<br />

Equation (5.3) says that the entropy of a Maxwellian distribution is proportional to its<br />

temperature. In Figure 5-6, the temperature (blue) calculated from each simulation cell is<br />

re-scaled such that the average temperatures upstream and downstream overlap the<br />

average entropies upstream and downstream, respectively. Like for the entropy, we also<br />

smooth the temperature by applying the Hamming filter once. In the upstream, there are<br />

80% of the ions (solar wind ions) in a Maxwellian that make little contribution to the<br />

upstream temperature. The upstream temperature shape deviates from the upstream<br />

entropy shape significantly, as the upstream temperature is dominant by the non-<br />

Maxwellian pickup ions. Therefore, in the figure the upstream temperature is nearly flat<br />

while the upstream entropy fluctuates. In contrast, the shape of the downstream<br />

temperature shows more agreement with the shape of the downstream entropy. This is<br />

due to Maxwellian-1 (discussed in Chapter 4) that includes both the transmitted solar<br />

wind ions and some pickup ions.<br />

I want to end this Chapter by quoting Charles K. Birdsall and A. Bruce Langdon who<br />

have contributing immeasurably to the subject of numerical simulations of plasmas and<br />

whom I greatly admire, “Practitioners must exercise a great deal of care, enough to obtain<br />

the essence of the problem, but not so much as to inhibit achieving any result.”


6 Theory/Model Connection to Observations<br />

6.1 Comparison with Voyager 2<br />

In Chapter 2, we found agreement between Voyager observations and several simulation<br />

results including the energy partition among several downstream ion components, and the<br />

pickup ion relative density. This section further compares observed termination shock<br />

properties with our simulated shock properties.<br />

6.1.1 Structure of the Termination Shock<br />

Figure 6-1 shows a magnetic field profile from a simulation with the following upstream<br />

inputs that are consistent with Voyager TS-3 crossing: θBN=74°, MA=8.8 and pickup ion<br />

relative density φ =20%. The pickup ion velocity distribution upstream follows the<br />

Vasyliunas–Siscoe prescription, which is widely believed to be the pickup ions velocity<br />

distribution at the termination shock [e.g., Schwadron, 1998]. In addition, our analysis in<br />

Chapter 4 shows that the pickup ion velocity assumption does not affect the magnetic<br />

field profile of the shock much. The choice of the Vasyliunas–Siscoe pickup ion for the<br />

purpose of comparison with Voyager 2 observation is therefore a valid choice.<br />

The structure (foot, ramp, overshoot, undershoot) of the termination shock observed<br />

by Burlaga et al. [2008] is compared side by side with our simulated shock structure, as<br />

shown in the figure. The two magnetic field profiles are superimposed in a way that 1. x-<br />

axis: the ramps from both profiles overlap and the scale length of the feet of both profiles<br />

overlap; 2. y-axis: the average upstream magnetic fields from both profiles overlap and<br />

the average downstream magnetic field from both profile overlap.


Ramp<br />

Foot<br />

155<br />

Overshoot<br />

Undershoot<br />

Figure 6-1 Comparison of our simulated magnetic field profile (orange) with the Voyager<br />

observed TS-3 magnetic field profile (black). Voyager 2 observing time is label on the x<br />

axis on the bottom and the simulation length is marked on the x axis on the top. Voyager<br />

2 Data are provided by Leonard F. Burlaga.<br />

Figure 6-1 demonstrates that the scaling of the simulation is very similar to the<br />

scaling of the observation. In particular, the simulated ramp is about one ion inertial<br />

length, just as Burlaga et al. [2008] conclude from Voyager 2 observations. The<br />

simulated lengths of the foot is about 8.8 ion inertial length, that is, the pickup ion<br />

gyroradius, also consistent with observations. The magnitudes and scale lengths of the<br />

overshoot and the undershoot are also in good agreement with the observations. The<br />

simulated downstream oscillation has similar magnitude as the observed oscillations.


156<br />

However, the exact phases of the oscillations are not in sync. In addition, the magnitude<br />

of the simulated foot is slightly smaller than the observed foot; the magnitude of the<br />

simulated overshoot is slightly larger than the observed overshoot; and the simulated<br />

overshoot bifurcates (note that as shown in Figure 5-5, the 400 particles per cell<br />

simulation has a better resolved shock structure without bifurcation). These differences<br />

are unavoidable results of comparing space plasma observations which have substantial<br />

temporal and spatial fluctuations as well as variations in the upstream solar wind<br />

conditions with a limited one-dimensional simulation with limited particles per cell (100<br />

particles per cell). Nevertheless, the qualitative agreement between the two magnetic<br />

field profiles provides convincing evidence that the simulation has captured the basic<br />

properties (foot, ramp, overshoot, undershoot) of the observed supercritical shock.<br />

6.1.2 Core Ions<br />

In Chapters 3 and 4, we fit the downstream ions with two Maxwellian components for<br />

theoretical purposes. In this section, for comparison with Voyager 2 observations, we fit<br />

the downstream ions with a single Maxwellian and ignore the high velocity “tail” of the<br />

distribution. As Bridge et al. [1977] originally presented, Voyager 2’s plasma instrument<br />

measures from 10 eV to 5950 eV. However, the instrument’s minimum measurable flux<br />

is above the flux of the suprathermal ions. Therefore, Voyager 2 can not distinguish<br />

suprathermal ion flux from the background. In fitting a single Maxwellian to the<br />

simulated downstream ions, we ignore the low flux suprathermal ions and bring out the<br />

core ions measured by Voyager 2 measures.


157<br />

Consider the 20% pickup ion simulation of a perpendicular shock with a Vasyliunas–<br />

Siscoe distribution upstream, MA=8, β sw =0.05, and β PUI =8.53. Figure 6-2 shows the<br />

single Maxwellian fit to the downstream ions in log and linear scales. The goodness of<br />

this fit is R=0.995, despite the fact that we ignore the tenuous suprathermal part of the<br />

distribution (as clearly evident in the log scale figure). The temperature of this fitted<br />

Maxwellian is Tcore 19 Tu,sw. Similarly, if we fit the MA=6 simulation with a single<br />

Maxwellian, the temperature is Tcore 10 Tu,sw. Both temperature jumps with respect to<br />

the upstream solar wind temperature fall within Voyager 2 observed temperature jump<br />

range ~10–20.<br />

These results suggest that the Voyager observed downstream core ion is a thermalized<br />

population that includes both the transmitted solar wind ions (whose temperature jump is<br />

1<br />

rS γ − ~2 as describe in Chapter 3) and some low energy pickup ions whose effective<br />

temperature is more than 10 times of the upstream solar wind temperature. The mixing of<br />

these two populations generate the core Maxwellian with a temperature that is ~10–20<br />

times of the upstream solar wind temperature when MA= 6–8, consistent with Voyager 2<br />

measurements.


( )<br />

f v ⊥<br />

10 4<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

y = m1*exp(-x^ 2/1.1^ 2)<br />

m1<br />

Chisq<br />

Value<br />

3723.7<br />

7.1089e+5<br />

Error<br />

24.835<br />

NA<br />

158<br />

R 0.99506 NA<br />

0.01<br />

0.01 0.1 1 10 100<br />

4000<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

y = m1*exp(-x^ 2/1.1^ 2)<br />

m1<br />

Chisq<br />

Value<br />

3723.7<br />

7.1089e+5<br />

Error<br />

24.835<br />

NA<br />

R 0.99506 NA<br />

0<br />

0.01 0.1 1 10 100<br />

Figure 6-2 The fitting of the simulated downstream distribution (MA=8, β sw =0.05, and<br />

β PUI =8.53) with one maxwellian component. The left panel is plotted in log-log scale<br />

and the right panel is plotted in the log-linear scale. The circles are downstream speed<br />

distributions from the simulations. The lines are the fitted function. The perpendicular<br />

speed vperp is in the unit of the upstream Alfvén speed vA.<br />

6.2 Implication for IBEX<br />

In this section, we present the x direction (r direction in the RTN coordinate system)<br />

energy spectra of the sunward moving downstream ions (in the shock frame) and discuss<br />

implications for IBEX.<br />

v⊥ / vA<br />

v⊥ / vA<br />

Figure 6-3 shows the complete spectra of three simulations with MA=4, 8, 16. These<br />

Mach numbers correspond to upstream solar wind speeds uu=128 km/s, 304 km/s and 608<br />

km/s, respectively, assuming that the upstream Alfvén speed is 38km/s (Chapter 1). The<br />

horizontal axis represents the x direction energy in the shock frame. The vertical axis<br />

represents the counts of ions as a function of the x direction energy. As expected from the


159<br />

discussion in Chapter 4, when the Alfvén Mach number increases, the spectra broaden<br />

and more suprathermal ions are produced. Using the Alfvén speed observed by Voyager<br />

2 vA=38 km/s (Chapter 1), IBEX-Hi detector’s energy range (~0.6–6 keV) [Funsten et al.,<br />

2009] corresponds to vx 2 /vA 2 = [90, 804], as shown in the right panel. If we fit these<br />

spectra with power law functions, the corresponding power law indexes are κ =5.89<br />

(MA=4), κ =2.76 (MA=8) and κ =1.22 (MA=16). A higher Alfvén Mach number leads to<br />

a gentler slope for the suprathermal ions. The reason is that a higher Mach number<br />

implies a hotter upstream population (larger pickup ion speeds) as well as more strongly<br />

heated downstream ions.<br />

10 4<br />

1000<br />

100<br />

10<br />

1<br />

1 10 100 1000 10 4<br />

MA=4<br />

MA=8<br />

MA=16<br />

2 2<br />

v /vA<br />

x<br />

1000<br />

100<br />

10<br />

y = 6.4518e+12 * x^(-5.8907) R= 0.99048<br />

y = 8.7629e+7 * x^(-2.7604) R= 0.94592<br />

y = 49504 * x^(-1.222) R= 0.97503<br />

MA=4<br />

MA=8<br />

MA=16<br />

1<br />

10 100 1000<br />

2 2<br />

v /vA<br />

x<br />

Figure 6-3 The modeled x direction (r direction in the RTN coordinate system) energy<br />

spectra of the sunward directing ions in the shock frame (left). Partial spectra of Figure<br />

6-3 that fall into the IBEX-Hi detector limit (right). We fit the spectra with power law<br />

functions (solid lines). Blue: MA=4, black: MA=8, and red: MA=16.


160<br />

Ions heated at the termination shock that undergo charge exchange with interstellar<br />

neutrals are converted into neutrals, which may subsequently be detected by IBEX. Our<br />

result implies that for IBEX high energy detector observed spectra, the decreases in the<br />

power law index scales with the increasing solar wind speeds. This scaling is consistent<br />

with the new results returned from the IBEX measurements that the spectral index is<br />

relatively large ( κ =1.95±0.09 at 1.7 keV and κ =1.91±0.07 at 2.7 keV) at low<br />

heliospheric latitudes where the solar wind is relatively slow and relatively small<br />

(κ =1.49±0.05 at 1.7 keV and κ =1.39±0.08 at 2.7 keV) at high heliospheric latitudes<br />

where the solar wind is relatively fast [Funsten et al., 2009].<br />

The predicted κ range of the downstream ions is however slightly larger than the<br />

IBEX measured κ range of the energetic neutral atoms (ENAs), as McComas et al.<br />

[2009]’s Figure 2 illustrated sky map of κ shows that κ has a range of approximately [1,<br />

4]. A more complete comparison of IBEX observations and simulation results is beyond<br />

the scope of this work. Recall that IBEX globally observes the line-of-sight integration of<br />

the sunward propagating heliosheath neutrals. Heating processes beyond the termination<br />

shock and into the heliosheath [e.g., Schwadron and McComas, 2006; Fisk and Gloeckler,<br />

2006; Drake et al., 2009] as well as the dynamical properties of the heliopause must be<br />

included. The deflection of flow in the heliosheath can be described by MHD models<br />

[e.g., Opher et al., 2006; Pogorelov et al., 2008]. The charge exchange processes to<br />

convert termination shock heated ions into ENAs can be modeled with a code developed<br />

in Boston University (BU) [Prested et al., 2008, Schwadron et al., 2009]. In addition, the<br />

BU model can take into account the ENA spectral’s energy shift as they are deflected and


161<br />

by the Sun’s gravitational field and radiation pressure [Schwadron et al., 2009, Möbius et<br />

al., 2009]. Nevertheless, our result provides physical insights for the scaling of IBEX<br />

measurements, as well as some quantitative estimates of how much heating the<br />

termination shock gives to the ions and how that amount of heating compares to the range<br />

that IBEX measures.<br />

Finally, it is worth mentioning that the newly returned IBEX measurements discover<br />

a “ribbon” of higher flux ENAs across the global sky map, with spectral indexes not<br />

significantly different from adjacent areas at the same heliospheric latitude [McComas et<br />

al., 2009, Funsten et al., 2009, Schwadron et al., 2009, Fuselier et al., 2009]. This<br />

suggests that the ribbon is caused by merely local confinement of higher density plasmas,<br />

other than additional dynamic heating at the termination shock [McComas et al., 2009,<br />

Funsten et al., 2009, Schwadron et al., 2009]. Many ideas are suggested to account for<br />

the ribbon, among them there is the concept of a strong interstellar magnetic field that<br />

distorts the heliopause and possibly the heliosheath [McComas et al., 2009, Funsten et al.,<br />

2009, Schwadron et al., 2009]. Given the interstellar magnetic field is inferred to be more<br />

important than the community has previously anticipated, it may be important to include<br />

this magnetic field in future modeling efforts of the heliopause, which can also be done<br />

with hybrid simulations.


7 Summary<br />

7.1 General Overview<br />

The study of ion kinetics at the helispheric termination shock is driven by scientific<br />

curiosity and practical necessity. As the Voyager spacecraft return in situ measurements<br />

from their crossings of the termination shock and traversal of the heliosheath, and as the<br />

IBEX spacecraft starts to make remote measurements of this region, space scientists are<br />

using theoretical studies and numerical simulations to interpret these observations and<br />

increase our physical understanding of this most remote region of our solar system. This<br />

dissertation describes our research using hybrid simulations to self-consistently model the<br />

ion scattering and energization processes at the shock. We quantify ion heating, ion<br />

energy dissipation and the ion speed distribution with the aid of analytic methods.<br />

The organization of the thesis is summarized as follows. Chapter 1 describes the<br />

motivation, summarizes what known and unknown about the termination shock in both<br />

theory and observations, and sets the research background of the topic. Chapter 2 focuses<br />

on understanding the fundamental ion energization processes at the termination shock,<br />

simulating an idealized termination shock with varying pickup ion relative density.<br />

Chapter 3 builds on the knowledge from Chapter 2, providing an analytic multi-<br />

component Rankine–Hugoniot model to address the heating, energy partition and<br />

velocity distributions. Chapter 4 examines the sensitivity of the resulting downstream ion<br />

speed distribution to the upstream pickup ion velocity distribution, the Alfvén Mach<br />

number, and the shock normal angle relative to the magnetic field. Chapter 5 investigates


163<br />

the sensitivity of the simulations to simulation cell size and numbers of particles per cell.<br />

Chapter 6 further validates the simulation model with direct comparison with Voyager<br />

observations and discusses the implication for the IBEX mission.<br />

It is important to clarify two important aspects of this work. First, we do not model<br />

the foreshock region, which is on the scale of ~10 AU, nor do we model the inner<br />

heliosheath, which is on the scale of ~30 AU. Both regions represent large scale<br />

structures beyond the scope we consider. The foreshock involves instabilities that scatter<br />

the upstream ion distributions and the heliosheath involves instabilities that isotropize the<br />

downstream ions. Those physical processes are not addressed here. Second, we do not<br />

model the energetic particles at energies greater than 10 keV or their acceleration<br />

mechanisms. Instead, we address the acceleration of thermal (eV) and suprathermal (few<br />

keV) ions within IBEX’s range of measurements, as discussed in Chapter 1 (end of<br />

Section 1.1) and Chapter 6 (Section 6.2). The issue with energetic particles is future work<br />

and should be carefully modeled with three-dimensional simulations in large simulation<br />

domains.<br />

7.2 Review of Scientific Context<br />

To summarize the scientific content, we re-visit the goals of this dissertation (end of<br />

Chapter 1), answering the following essential questions.<br />

How do pickup ions modify the termination shock? Pickup ions, typically ~1000<br />

times hotter than the solar wind ions, significantly increase the effective beta of the


164<br />

upstream plasma. As a result, both the upstream sound speed and the fast wave speed are<br />

increased with increasing pickup ion relative density. When 20% of the ions are pickup<br />

ions, the upstream magnetosonic Mach number is reduced. The shock becomes much<br />

weaker than the terrestrial bow shock. Still, the shock remains supercritical, as is clearly<br />

evident in the simulations with the formation of the overshoot in the magnetic field<br />

profile and through the solar wind ion specular reflection. The difference here is that the<br />

foot of the shock is determined by the pickup ion gyroradius, as a result of the gyro-phase<br />

dependent acceleration. This foot is notably broader than that resulted from solar wind<br />

ion gyrations.<br />

How much is each ion component heated across the termination shock? The<br />

heating rate of the ions depends on the pickup ion relative density, which modifies the<br />

compression ratio (strength) of the shock. The heating is quantified by our analytic multi-<br />

component Rankine–Hugoniot model. For transmitted solar wind ions, the heating is very<br />

1<br />

small and the temperature jump is about τ = r γ − , which is about τ ~2 for a compression<br />

ratio rS~2 shock. This polytropic index increases with pickup ion relative density. When<br />

there are no pickup ions, γ ≈5/3; when there are 20% pickup ions, γ ≈2; and when there<br />

are 40% pickup ions, γ ≈2.2. For reflected solar wind ions, their dynamic energy is<br />

converted into thermal energy. The temperature jump of reflection solar wind ions is on<br />

the order of a thousand. Because the reflected solar wind ions are a small portion of the<br />

solar wind population (reflection efficiency is less than 10% when pickup ion relative<br />

density is larger than 20%), the temperature jump of all solar wind ions are on the order<br />

S


165<br />

of ten. For pickup ions, the average temperature jump is described by<br />

is roughly a constant (~2.5), independent of the pickup ion relative density.<br />

γ<br />

1<br />

rS PUI γ −<br />

, where PUI<br />

What is the partition of dissipated energy among the ion components? The<br />

partition of dissipated energy is also a function of the pickup ion relative density. When<br />

this density is smaller than 5%, solar wind ion reflection is the dominant dissipation<br />

process. Most of the dissipated energy is deposited into solar wind ions. When the pickup<br />

ion relative density is larger than 5%, pickup ions account for most of the shock<br />

dissipation. Particularly when pickup ion relative density is equal or larger than 20%,<br />

more than 85% of the dissipated energy is converted into pickup ion heating. This result<br />

is consistent with inference from Voyager 2 measurements. In addition, this result does<br />

not contradict the finding that the temperature jump of the pickup ion is much smaller<br />

than that of the solar wind ions. Because pickup ions started with an initial temperature<br />

that is three orders of magnitude larger the initial temperature of the solar wind ions, a<br />

small temperature jump still enables a much larger net dissipated energy gain for the<br />

pickup ions than for the solar wind ions.<br />

How are ions (solar wind and pickup ions) energized? The ion components are<br />

heated through different mechanisms. The transmitted solar wind ions are heated by<br />

shock compression. The reflected solar wind ions are heated through specular reflection,<br />

which convert these ions’ incident velocity into gyrovelocity. These mechanisms are the<br />

same as the ion energization mechanisms of the terrestrial bow shocks. What is different<br />

is the energization process of the pickup ions. The pickup ions already have very high


166<br />

upstream temperatures. Therefore, they encounter the termination shock in a completely<br />

different manner compared with the cold solar wind ions. This encounter involves both<br />

the guiding center motions and the gyro motions of the pickup ions. We found that<br />

pickup ions’ energization is a gyrophase-dependent acceleration (GPDA) process, with<br />

the “crossing” ions gaining more energy than the “reversed” ions. This is because the<br />

“crossing” ions’ initial encounter with the shock results in some initial minor energy gain,<br />

while the “reversed” ions’ initial encounter with the shock results in some initial minor<br />

energy loss. While both types of pickup ions’ major energy gain is due to the well-known<br />

shock drift acceleration (SDA) during a single encounter with the shock, GPDA causes<br />

differences in the ions’ final energy, according to the gyrophases of the ions as they<br />

encounter the shock. Those pickup ions with favorable phase angles (positive vx, negative<br />

vy) will be preferentially accelerated by the inhomogeneous electric field Ey (


167<br />

In addition, results from our one-dimensional hybrid simulation are in agreement with<br />

Voyager 2 observed termination shock magnetic field profile, Voyager 2 inferred pickup<br />

ion relative density, and Voyager 2 inferred energy partition of dissipated energy. We<br />

also validate our simulations by checking accuracy and convergence.<br />

7.3 Future Work<br />

This dissertation makes a fundamental contribution to the understanding of the physical<br />

processes associated with scattering and heating of thermal and suprathermal ions at the<br />

termination shock. Along the way, we identify directions for future work:<br />

1. From the perspective of hybrid simulations, the following projects can be<br />

considered in the future:<br />

I. Dimension. An important limitation of our work is that ions are tied to magnetic<br />

fields in the one-dimensional simulation. The one-dimensional code precludes the<br />

excitation of some downstream waves as well as cross-field diffusion. For example,<br />

quasi-perpendicular shocks often generate strong T⊥ > T|| anisotropies which can lead to<br />

ion cyclotron instabilities [e.g., Scholer et al., 2000]. These waves scatter the heated solar<br />

wind and pickup ions in the direction of the magnetic field in two-dimensional or three-<br />

dimensional simulations. The scattering relaxes the downstream plasma [e.g., Winske and<br />

Quest, 1986; Thomas, 1989; McKean et al. JGR, 1995], which in one-dimensional<br />

simlations remains in a highly anisotropic state. With three-dimensional simulations, one


168<br />

can fully resolve the downstream ion velocity distribution as well as investigating the<br />

additional accelerations (if any) caused by various instabilities.<br />

II. Energetic Particles. To describe energetic particles, one can include the energetic<br />

particle population into the upstream input, along with the solar wind population and the<br />

pickup ion population. The ubiquitous co-existence of these three populations is observed<br />

in the heliosphere all the time, even during quiet-time solar wind conditions [Fisk and<br />

Gloeckler, 2006]. Assuming that those ions are isotropic protons, Figure 7-1 depicts our<br />

illustration of the three populations in the upstream region. In the figure, the solar wind is<br />

a Maxwellian, the pickup ions form a Vasyliunas–Siscoe distribution, and the energetic<br />

particles form the power law tail of the spectra.<br />

In addition, the simulation box should be extended significantly when energetic<br />

particles are included. A MeV energetic particle will have a gyrospeed that is more than<br />

1000 times of the Alfvén speed. Then this particle’s gyroradius is more than 1000 times<br />

the ion inertial length 7 . To describe energetic particles properly, one should extend the<br />

simulation box to resolve energetic particle gyrations. The appropriate length of the<br />

simulation box is expensive to achieve.<br />

Further, Giacalone and Ellison [2000] argue that energetic particles can be generated<br />

directly from thermal particles. Treating pickup ions as test particles, they apply<br />

7 Energetic particle larmor radius (gyroradius) ,<br />

here λi is the ion inertial length.<br />

v v v c ω v c v<br />

EP EP A pi EP EP<br />

LEP= = ⋅ ⋅ ⋅ = ⋅ = i<br />

Ωc vA c ωpi Ωc<br />

vA ωpi<br />

vA<br />

r<br />

λ ,


169<br />

Figure 7-1 The three upstream components in the solar wind frame. The solid curve<br />

represents the Maxwellian solar wind; the dotted curve represents the Vasyliunas–Siscoe<br />

pickup ions; the dashed curve represents the powerlaw tail.<br />

prescribed Alfvén waves to their three-dimensional hybrid simulations. They find that the<br />

downstream spectra broaden when the simulation box size increases, thereby suggesting<br />

that the long wavelength waves accelerate particles by enabling cross-field diffusion and<br />

magnetic field random walks. To test their acceleration mechanism, one needs the system<br />

length to be on the order of 20,000 c/ω pi (~1 AU) in the direction of solar wind<br />

propagation (x direction in our simulation). The requirement is unattainable in present<br />

three-dimensional simulations but may be possible with more powerful parallel<br />

computatation.<br />

Solar wind<br />

Pickup ions<br />

Energetic particles


170<br />

III. Heavy Ions. Although to first order, heavy ions are not included in this work, it<br />

would be a useful addition to carry out more realistic simulations in the future. Helium<br />

ions, which usually make up


171<br />

solar system boundaries. Future projects can be built upon this work to continue the<br />

exploration of the termination shock, other astrophysical shocks, and the universe.


List of Journal Abbreviations<br />

Adv. Space Res. Advance in Space Research<br />

Astrophys. J. Astrophysical Journal<br />

Geophys. Monogr. Ser. Geophysical Monograph Series<br />

Geophys. Res. Lett. Geophysical Research Letters<br />

J. Geophys. Res. Journal of Geophysical Research<br />

J. Phys. G: Nucl. Part. Phys. Journal of Physics G: Nuclear and Particle<br />

Physics<br />

J. reine angew. Math. Journal für die reine und angewandte<br />

Mathematik<br />

Mon. Not. R. Astron. Soc. Monthly Notices of the Royal Astronomical<br />

Society<br />

Phys. Fluids Physics of Fluids<br />

Phys. Rev. Lett. Physical Review Letters<br />

Phys. Plasmas Physics of Plasmas<br />

Proc. R. Soc. Proceedings of the Royal Society<br />

Space Sci. Rev. Space Science Reviews


References<br />

Aellig, M. R. and A. J. Lazarus and J. T. Steinberg (2001), The solar wind Helium<br />

abundance: variation with wind speed and the solar cycle, Geophys. Res. Lett., 28(14),<br />

2767-2770.<br />

Bale, S. D., M. A. Balikhin, T. S. Horbury, V. V. Krasnoselskikh, H. Kucharek, E.<br />

Möbius, S. N. Walker, A. Balogh, D. Burgess, B. Lembege, E. A. Lucek, M. Scholer,<br />

S. J. Schwartz, and M. F. Thomsen (2005), Quasi-perpendicular shock structure and<br />

processes, Space Sci. Rev., 118, 161–203, doi:10.1007/s11214-005-3827-0.<br />

Birdsall, C. K. and A. B. Langdon (1985), Plasma Physics Via Computer Simulation,<br />

McGraw-Hill Book Company, New York.<br />

Bridge, H. S., J. W. Belcher, R. J. Butler, A. J. Lazarus, A. M. Mavretic, J. D. Sullivan, G.<br />

L. Siscoe, and V. M. Vasyliunas (1977), The plasma experiment on the 1977 Voyager<br />

mission, Space Sci. Rev., 21, 259–287.<br />

Burgess, D. (1989), Alpha particles in field-aligned beams upstream of the bow shock:<br />

Simulations, Geophys. Res. Lett., 16(2), 163–166.<br />

Burgess, D. (1995), Collisionless shocks, in Introduction to Space Physics, edited by M.<br />

G. Kivelson and C. T. Russell, 129–163, Cambridge University Press, New York.


174<br />

Burgess, D. (1997), Solar wind and interstellar medium coupling, in Lecture Notes in<br />

Physics, 489, 1616–6361, Springer, Berlin/Heidelberg.<br />

Burlaga, L. F., N. F. Ness, M. H. Acuna, R. P. Lepping, J. E. Connerney, E. C. Stone, and<br />

F. B. McDonald (2005), Crossing the termination shock into the heliosheath: Magnetic<br />

fields, Science, 309, 2027–2029, doi:10.1126/science.1117542.<br />

Burlaga, L. F., N. F. Ness, M. H. Acuna, R. P. Lepping, J. E. P. Connerney and J. D.<br />

Richardson [2008], Magnetic Fields at the Solar Wind Termination Shock, Nature,<br />

doi:10.1038/nature07029.<br />

Chapman, S. C., R. E. Lee, and R. O. Dendy (2005), Perpendicular shock reformation<br />

and ion acceleration, Space Sci. Rev., 121, 5.<br />

Cravens, T. E. (1997), Physics of Solar System Plasmas, 134, Cambridge University<br />

Press, New York.<br />

Decker, R. B., S. M. Krimigis, E. C. Roelof, M. E. Hill, T. P. Armstrong, G. Gloeckler, D.<br />

C. Hamilton, and L. J. Lanzerotti (2005), Voyager 1 in the foreshock, termination<br />

shock, and heliosheath, Science, 309 (5743), 2020–2024, doi:10.1126/science.1117569.<br />

Decker, R. B., S. M. Krimigis, E. C. Roelof, M. E. Hill, T. P. Armstrong, G. Gloeckler, D.<br />

C. Hamilton, and L. J. Lanzerotti (2008), Mediation of the solar wind termination<br />

shock by non-thermal ions, Nature, 454, 67–70, doi:10.1038/nature07030.


175<br />

Drake, J, Reconnection of the sectored heliospheric magnetic field near the heliopause: a<br />

mechanism for the generation of anomalous cosmic rays, the Solar Heliospheric<br />

INterplanetary Enviroment (SHINE) workshop, August, 2009, Nova Scotia, Canada.<br />

Feldman, W. C. (1985), Electron velocity distributions near collisionless shocks, in<br />

Collisionless Shocks in the Heliosphere: Reviews of Current Research, Geophys.<br />

Monogr. Ser., vol. 35, edited by B. T. Tsurutani and R. G. Stone, pp. 195–205, AGU,<br />

Washington, D.C.<br />

Fisk, L. A. (1996), Implication of a weak termination shock, Space Sci. Rev., 78, 129–<br />

136, doi:10.1007/BF00170799.<br />

Fisk, L. A., G. Gloeckler, and T. H. Zurbuchen (2006), Acceleration of low-energy ions<br />

at the termination shock of the solar wind, Astrophy. J., 644, 631–637.<br />

Fisk, L. A., and G. Gloeckler (2006), The common spectrum for accelerated ions in the<br />

quiet-time solar wind, Astrophy. J., 640 (1), L79–L82, doi:10.1086/503293.<br />

Funsten, H. O., F. Allegrini, G. B. Crew, R. DeMajistre, P. C. Frisch, S. A. Fuselier, M.<br />

Gruntman, P. Janzen, D. J. McComas, E. Möbius, B. Randol, D. B. Reisenfeld, E. C.<br />

Roelof, and N. A. Schwadron (2009), Structures and spectral variations of the outer<br />

heliosphere in IBEX energetic neutral atom maps, Science, doi:10.1126/science.<br />

1180927.


176<br />

Fuselier, S. A., and W. K. H. Schmidt (1994), H + and He 2+ Heating at the Earth's bow<br />

shock, J. Geophys. Res., 99(A6), 11,539–11,546.<br />

Fuselier, S. A., F. Allegrini, H. O. Funsten, A. G. Ghielmetti, D. Heirtzler, H. Kucharek,<br />

O. W. Lennartsson, D. J. McComas, E. Möbius, T. E. Moore, S. M. Petrinec, L. A.<br />

Saul, J. A. Scheer, N. Schwadron, and P. Wurz (2009), Width and variation of the<br />

ENA flux ribbon observed by the interstellar boundary explorer, Science,<br />

doi:10.1126/science.1180981.<br />

Giacalone, J., and D. C. Ellison (2000), Three-dimensional numerical simulations of<br />

particle injection and acceleration at quasi-perpendicular shocks, J. Geophys. Res.,<br />

105(A6), 12,541–12,556.<br />

Gloeckler, G., and J. Geiss (1996), Abundance of 3 He in the local interstellar cloud,<br />

Nature, 381, 210–212, doi:10.1038/381210a0.<br />

Gloeckler, G (1998), Interstellar and inner source pickup ions observed with SWICS on<br />

Ulysses, Space Sci. Rev., 86, 127–159.<br />

Goodrich, C. C., and J. D. Scudder (1984), The adiabatic energy change of plasma<br />

electrons and the frame dependence of the cross-shock potential at collisionless<br />

magnetosonic shock waves, J. Geophys. Res., 89(A8), 6654–6662.


177<br />

Goodrich, C. C. (1985), Numerical simulations of quasi-perpendicular collisionless shock,<br />

in Collisionless Shocks in the Heliosphere: Reviews of Current Research, Geophys.<br />

Monogr. Ser., vol. 35, edited by B. T. Tsurutani and R. G. Stone, pp. 153–168.<br />

Gosling, J. T., and A. E. Robson (1985), Ion reflection, gyration, and dissipation at<br />

supercritical shocks, in Collisionless Shocks in the Heliosphere: Reviews of Current<br />

Research, Geophys. Monogr. Ser., vol. 35, edited by B. T. Tsurutani and R. G. Stone,<br />

pp. 141–152<br />

Gombosi, T. I. (1994), Gaskinetic Theory, Cambridge University Press, 262–266.<br />

Hamming, R. W. (1977), Digital Filters, Prentice Hall, Englewood Cliffs, NJ.<br />

Hockney, R. W. and J. W. Eastwood (1988), Computer Simulation Using Particles, IOP<br />

Publishing Ltd, London.<br />

Heerikhuisen, J., V. Florinski, and G. P. Zank (2006), Interaction between the solar wind<br />

and interstellar gas: A comparison between Monte Carlo and fluid approaches, J.<br />

Geophys. Res., 111, A06110.<br />

Kasper, J. C., M. L. Stevens, A. J. Lazarus, J. T. Steinberg and K. W. Ogilvie (2007),<br />

Solar wind Helium abundance as a function of speed and heliographic latitude:<br />

variation through a solar cycle, Astrophy. J., 660, 901-910.


178<br />

Intriligator D. S., Siscoe G. L., Miller W. D. (1996): Interstellar pickup H+ ions at 8.3-<br />

AU - Pioneer-10 plasma and magnetic-field analyses. Geophys. Res. Lett. 23, 2181-<br />

2184.<br />

Jokippi, J. R. (1986), Particle acceleration at a termination shock. I - Application to the<br />

solar wind and the anomalous component, J. Geophys. Res., 91, p. 2929-2932, doi:<br />

10.1029/JA091iA03p02929.<br />

Kennel, C. F., J. P. Edmiston, and T. Hada (1985), A quarter century of collionless shock<br />

research, in Collisionless Shocks in the Heliosphere: A Tutorial Review, Geophys.<br />

Monogr. Ser., vol. 34, edited by Stone and Tsurutani, pp. 1–36.<br />

Kirk, J. G. and R. O. Dendy (2001), Shock acceleration of cosmic rays—a critical review,<br />

J. Phys. G: Nucl. Part. Phys., 27, 1589-1595, doi: 10.1088/0954-3899/27/7/316.<br />

Lee, M. A., V. D. Shapiro and R. Z. Sagdeev (1996), Pickup ion energization by shock<br />

surfing, J. Geophys. Res. , 101, 4777.<br />

Leroy, M. M., D. Winske, C. C. Goodrich, C. S. Wu, and K. Papadopoulos (1982), The<br />

structure of perpendicular bow shocks, J. Geophys. Res., 87 (A7), 5081–5094,<br />

doi:10.1029/JA087iA07p05081.<br />

Leroy, M. M. (1983), Structure of perpendicular shocks in collisionless plasma, Phys.<br />

Fluids, 26, 2742, doi:10.1063/1.864468.


179<br />

Lembege, B., J. Giacalone, M. Scholer, T. Hada, M. Hoshino, V. Krasnoselskikh, H.<br />

Kucharek, P. Savoini and T. Terasawa (2004), Selected problems in collisionless-<br />

shock physics, Space Sci. Rev., vol.110, 161-226.<br />

Lever, E. L., K. B. Quest and V. D. Shapiro (2001), Shock surfing vs. shock drift<br />

acceleration, Geophys. Res. Lett., 28, 1367.<br />

Li, H., C. Wang, and J. D. Richardson (2008), Properties of the termination shock<br />

observed by voyager 2, Geophys. Res. Lett., 35, L19,107, doi:10.1029/2008GL034869.<br />

Liewer, P. C., B. E. Goldstein, and N. Omidi (1993), Hybrid simulations of the effects of<br />

interstellar pickup hydrogen on the solar wind termination shock, J. Geophys. Res., 98,<br />

15,200.<br />

Lipatov, A. S. and G. P. Zank (1999), Pickup ion acceleration at low- β p perpendicular<br />

shocks, Phy. Rev. Letts., 82, 3609.<br />

Lucek, S. G. and A. R. Bell (2000), Non-linear amplification of a magnetic field driven<br />

by cosmic ray streaming, Mon. Not. R. Astron. Soc., 314, 65–74.<br />

Marshall, W. (1955), The structure of magnetohydrodynamic shock waves, Proc. R. Soc.,<br />

London, Ser. A, 233,367.<br />

McComas, D. J. , F. Allegrini, P. Bochsler, M. Bzowski, M. Collier, H. Fahr, H. Fichtner,<br />

P. Frisch, H. Funsten, S. Fuselier, G. Gloeckler, M. Gruntman, V. Izmodenov, Paul


180<br />

Knappenberger, M. Lee, S. Livi, D. Mitchell, E. Möbius, T. Moore, D. Reisenfeld, E.<br />

Roelof, N. Schwadron, M. Wieser, M. Witte, P. Wurz, and G. Zank (2004), The<br />

Interstellar Boundary Explorer (IBEX), Physics of the Outer Helioshell, CP719.<br />

McComas, D. J., F. Allegrini, P. Bochsler, M. Bzowski, E. R. Christian, G. B. Crew, R.<br />

DeMajistre, H. Fahr, H. Fichtner, P. C. Frisch, H. O. Funsten, S. A. Fuselier, G.<br />

Gloeckler, M. Gruntman, J. Heerikhuisen, V. Izmodenov, P. Janzen, P. Knappenberger,<br />

S. Krimigis, D. Reisenfeld, E.Roelof, L. Saul, N. A. Schwadron, P. W. Valek, R.<br />

Vanderspek, P. Wurz, and G. P. Zank (2009), Global Observations of the instellar<br />

interaction from the Interstellar Boundary Explorer (IBEX), Science,<br />

doi:10.1126/science.1180906.<br />

McKean, M. E., N. Omidi, and D. Krauss-Varban (1995), Wave and ion evolution<br />

downstream of quasi-perpendicular bow shocks, J. Geophys. Res., 100(A3), 3427–<br />

3437.<br />

Möbius, E., P. Bochsler, M. Bzowski, G. B. Crew, H. O. Funsten, S. A. Fuselier, A.<br />

Ghielmetti, D. Heirtzler, V. V. Izmodenov, M. Kubiak, H. Kucharek, M. A. Lee, T.<br />

Leonard, D. J. McComas, L. Petersen, L. Saul, J. A. Scheer, N. Schwadron, M. Witte,<br />

and P. Wurz (2009), Direct observations of interstellar H, He, and O by the Interstellar<br />

Boundary Explorer, Science, doi:10.1126/science.1180906.


181<br />

Müller, H.-R., G. P. Zank, and A. S. Lipatov (2000), Self-consistent hybrid simulations<br />

of the interaction of the heliosphere with the local interstellar medium, J. Geophys.<br />

Res., 105, 27,419.<br />

Opher, M., E. Stone, P. C. Liewer, T. I. Gombosi (2006), Global Asymmetry of the<br />

Heliosphere, in Physics of the inner heliosheath, AIP Conference Proceedings, edited<br />

by J. Heerikhuisen, V. Florinski, G. P. Zank and N. V. pogorelov, vol. 858, pp. 45-50.<br />

Paschmann, G., N. Sckopke, S. J. Bame, and J. T. Gosling (1982), Observations of<br />

gyrating ions in the foot of the nearly perpendicular bow shock, Geophys. Res. Lett., 9<br />

(8), 881–884.<br />

Pesses, M. E. , D. Eichler and J. R. Jokippi (1981), Cosmic ray drift, shock wave<br />

acceleration, and the anomalous component of cosmic rays, Astrophy. J., 246, p. L85-<br />

L88, doi: 10.1086/183559.<br />

Phillips, P. E., and A. E. Robson (1972), Influence of reflected ions on the magnetic<br />

structure of a collionless shock front, Phys. Rev. Lett., 29, 154,<br />

doi:10.1103/PhysRevLett.29.154.<br />

Pogorelov, N. V., J. Heerikhuisen and G. P. Zank (2008), Probing heliospheric<br />

asymmetries with an MHD-kinetic model, Astrophy. J., 675, L41-44,<br />

doi:10.1086/529547.


182<br />

Prested, C., N. A. Schwadron, J. Passuite, B. Randol, B. Stuart, G. Crew, J Heerkhuisen,<br />

N. Pogorelov, G. Zank, M. Opher, F. Allegrini and D. J. McComas, M. Reno, E.<br />

Roelof, S. Fuselier, H. Funsten, E. Moebius and L. Saul (2008), Implications of solar<br />

wind suprathermal tails for IBEX ENA images of the heliosheath, J. Geophys. Res.,<br />

113, A06102, doi:10.1029/2007JA012758.<br />

Quest, K. B. (1985), Simulation of high-Mach-number collisionless perpendicular shock<br />

in astrophysical plasmas, Phys. Rev. Lett., 54 (16), 1872–1874,<br />

doi:10.1103/PhysRevLett.54.1872.<br />

Richardson, J. D. (2008), Plasma temperature distributions in the heliosheath, Geophys.<br />

Res. Lett., 35, L23, 104.<br />

Richardson, J. D., J. C. Kasper, C. Wang, J. W. Belcher, and A. J. Lazarus (2008), Cool<br />

heliosheath plasma and deceleration of the upstream solar wind at the termination<br />

shock, Nature, 454, 63–66, doi:10.1038/nature07024.<br />

Scholer, M., H. Kucharek, and J. Giacalone (2000), Cross-field diffusion of charged<br />

particles and the problem of ion injection and acceleration at quasi-perpendicular<br />

shocks, J. Geophys. Res., 105(A8), 18,285–18,293.<br />

Sckopke, N., G. Paschmann, S. J. Bame, and J. T. Gosling (1983), Evolution of ion<br />

distribution across the nearly perpendicular bow shock: Specularly and non-specularly<br />

reflected-gyrating ions, J. Geophys. Res., 88 (A8), 6121–6136,<br />

doi:10.1029/JA088iA08p06121.


183<br />

Scudder, J. D., A. Mangeney, C. Lacombe, C. C. Harvey, T. L. Aggson, R. R. Anderson,<br />

J. T. Gosling, G. Paschmann, and C. T. Russell (1986a), The resolved layer of a<br />

collionless, high β , supercritical, quasi-perpendicular shock wave 1. Rankine–<br />

Hugoniot geometry, currents, and stationarity, J. Geophys. Res., 91 (A10), 11,019–<br />

11,052.<br />

Scudder, J. D., A. Mangeney, C. Lacombe, C. C. Harvey and T. L. Aggson (1986b), The<br />

resolved layer of a collisionless, high β , supercritical, quasi-perpendicular shock wave<br />

2. dissipative fluid electrodynamics, J. Geophys. Res., 91 (A10), 11,053–11,073.<br />

Schwadron, N. A. (1998), A model for pickup ion transport in the heliosphere in the limit<br />

of uniform hemispheric distribution. J. Geophys. Res., 103, 20,643–20,649.<br />

Schwadron, N. A, and D. J. McComas (2006), Particle Acceleration at a Blunt<br />

Termination Shock, Physics of the Inner Heliosheath.<br />

Schwadron, N. A., G. Crew, R. Vanderspek, F. Allegrini, M. Bzowski, R. DeMagistre, G.<br />

Dunn, H. Funsten, S. A. Fuselier, K. Goodrich, M. Gruntman, J. Hanley, J.<br />

Heerikuisen, D. Heirtlzer, P. Janzen, H. Kucharek, G. Loeffler, K. Mashburn, K.<br />

Maynard, D. J. McComas, E. Moebius, C. Prested, B. Randol, D. Reisenfeld, M. Reno,<br />

E. Roelof and P. Wu (2009), The Interstellar Boundary Explorer science operations<br />

center, Space Sci.e Rev., doi:10.1007/s11214-009-9513-x.


184<br />

Schwadron, N. A., M. Bzowski, G. B. Crew, M. Gruntman, H. Fahr, H. Fichtner, P. C.<br />

Frisch, H. O. Funsten, S. Fuselier, J. Heerikhuisen, V. Izmodenov, H. Kucharek, M.<br />

Lee, G. Livadiotis, D. J. McComas, E. Moebius, T. Moore, J. Mukherjee, N. V.<br />

Pogorelov, C. Prested, D. Reisenfeld, E. Roelof, and G. P. Zank (2009), Comparison of<br />

Insterstellar Boundary Explorer observations with 3D global heliospheric models,<br />

Science, doi:10.1126/science.1180986.<br />

Sonett, C. P., and I. J. Abrams (1963), The distant geomagnetic field, 3, Disorder and<br />

shocks in the magnetopause, J. Geophys. Res., 68, 1233.<br />

Stone, E. C. (2001), News from the edge of interstellar Space, Science, 299 (5527), 55-56,<br />

doi:10.1126/science.1060090.<br />

Stone, E. C., A. C. Cummings, F. B. McDonald, B. C. Heikkila, N. Lal, and W.<br />

R.Webber (2005), Voyager 1 explores the termination shock region and the<br />

heliosheath beyond, Science, 309 (5743), 2017–2020, doi:10.1126/science.1117684.<br />

van Heemert, A. (1957), Cyclic Permutations with Sequences and Related Problems, J.<br />

reine angew. Math., 198, 56 – 72.<br />

Vasyliunas, V. M. and G. L. Siscoe (1976), On the flux and energy spectrum of<br />

interstellar ions in the helioshell, J. Geophys. Res., 81, 1247.


185<br />

Whang, Y. C., and L. F. Burlaga (2000), Anticipated Voyager crossing of the termination<br />

shock, Geophys. Res. Lett., 27(11), 1607–1610, doi:<br />

10.1029/2000GL003748.<br />

Winske, D. (1984), Microtheory of collisionless shock current layers, in Collisionless<br />

Shocks in the Heliosphere: Reviews of Current Research, Geophys. Monogr. Ser., 35,<br />

195–205.<br />

Winske, D, and M. M. Leroy (1984), Hybrid simulation techniques applied to the Earth’s<br />

bow shock, in Computer simulation of space plasmas, edited by H. Matsumoto and T.<br />

sato, p. 255 – 278, Terra Scientific Publishing Company, Tokyo.<br />

Winske, D., and K. B. Quest (1986), Electromagnetic Ion Beam Instabilities: Comparison<br />

of One- and Two-Dimensional Simulations, J. Geophys. Res., 91(A8), 8789–8797.<br />

Winske, D. and K. B. Quest (1988), Magnetic field and density fluctuations at<br />

perpendicular supercritical collisionless shocks, J. Geophys. Res., 93 (A9), 9681–9693.<br />

Winske, D., and N. Omidi (1993), Hybrid codes: Methods and applications, in Computer<br />

Space Plasma Physics: Simulation Techniques and Software, edited by H. Matsumoto<br />

and Y. Omura, 103–160, Terra Scientific Publishers, Tokyo.<br />

Winske, D., L. Yin, N. Omidi, H. Karimabadi, and K. B. Quest (2003), Hybrid simulation<br />

codes: Past, present and future-A tutorial, in Space Plasma Simulation, edited by J.<br />

Buechener, C. T. Dum, and M. Scholer, 140–169, Springer, Germany.


186<br />

Winske, D., S. P. Gary, P. Wu and N. A. Schwadron (2009), Pickup ion energization at<br />

the termination shock, J. Geophys. Res., in preparation.<br />

Woods, L. C. (1969), Critical Alfvén Mach numbers for transverse field MHD shocks,<br />

Phys. Plasmas, 11, 25–34, doi:10.1088/0032-1028/11/1/004.<br />

Wu, C. S., D. Winske, Y. M. Zhou, S. T. Tsai, P. Rodriguez, M. Tanaka, K.<br />

Papadopoulos, K. Akimoto, C. S. Lin, M. M. Leroy and C. C. Goodrich (1984),<br />

Microinstabilities associated with a high Mach number, perpendicular bow shock,<br />

Space Sci. Rev., 37, 63–109.<br />

Wu. P., D. Winske, S. P. Gary, N. A. Schwadron and M. A. Lee (2009), Ion heating and<br />

energy dissipation at the heliospheric termination shock, J. Geophys. Res., 114,<br />

A08103, doi:10.1029/2009JA014240.<br />

Wu. P., D. Winske, S. P. Gary, and N. A. Schwadron, Hybrid simulations of<br />

perpendicular termination shocks: consequences of variation in pickup ion<br />

distributions, J. Geophys. Res., in preparation.<br />

Yang, Z. W., Q. M. Lu, B. Lembege and S. Wang (2009), Shock front nonstationarity<br />

and ion acceleration in supercritical perpendicular shocks, J. Geophys. Res., 114,<br />

A03111.<br />

Vasyliunas, V. M. and G. L. Siscoe (1976), On the flux and energy spectrum of<br />

interstellar ions in the heliosphere, J. Geophys. Res., 81, 1247.


187<br />

Vogl, D. F. , D. Langmay, N. V. Erkaev, H. K. Biernat, C. J. Farrugia and S. Mühlbachler<br />

(2003), The anisotrpic jump equations for oblique fast shocks in a Kappa distributed<br />

medium, Adv. Space Res., 32 (4), 519-523.<br />

Zank, G. P. (1999), Interaction of the Solar wind with the local interstellar medium: A<br />

theoretical perspective, Space Sci. Rev., vol.89, 413–688.<br />

Zank, G. P., H. L. Pauls, I. H. Cairns, and G. M. Webb (1996), Interstellar pickup ions<br />

and quasi-perpendicular shocks: Implications for the termination shock and<br />

interplanetary shocks, J. Geophys. Res., 101, 457–477.


Curriculum Vitae<br />

EDUC<strong>AT</strong><strong>ION</strong><br />

Pin Wu<br />

Ph.D. Candidate, Department of Astronomy, Boston University<br />

Affiliate, ISR-1, Los Alamos National Laboratory<br />

Department of Astronomy, 725 Commonwealth Ave, Boston, MA 02215<br />

pwu@bu.edu | (857)753-7725 (cell) | http://people.bu.edu/pwu<br />

B.S., Modern Physics, University of Science and Technology of China (USTC), 2001<br />

M.A., Astronomy, Boston University, 2003–2005<br />

Ph.D., Astronomy, Boston University, 2003–2009<br />

Dissertation: Ion Kinetics at the Heliospheric Termination Shock<br />

Advisor: Nathan A. Schwadron, Boston University<br />

Mentors: S. Peter Gary and Dan Winske, Los Alamos National Laboratory<br />

JOURNAL PUBLIC<strong>AT</strong><strong>ION</strong>S<br />

As the First Author (Chronologically)<br />

1. Wu, P., T. A. Fritz, B. Larvaud, and E. Lucek (2006), Substorm associated<br />

magnetotail energetic electrons pitch angle evolutions and flow reversals: Cluster<br />

observation, Geophys. Res. Lett., 33, L17101, doi:10.1029/2006GL026595.<br />

2. Wu, P., N. A. Schwadron, G. L. Siscoe, and P. Riley (2008), Initial condition<br />

influence on coronal mass ejection propagation, J. Geophys. Res., 113, A00B05,<br />

doi:10.1029/2008JA013082.<br />

3. Wu, P., D. Winske, S. P. Gary, N. A. Schwadron and M. A. Lee (2009), Energy<br />

dissipation and ion heating at the heliospheric termination shock, J. Geophys. Res.,<br />

114, A08103, doi:10.1029/2009JA014240.<br />

4. Wu, P., D. Winske, S. P. Gary, and N. A. Schwadron (2009), Hybrid simulations of<br />

perpendicular termination shocks: consequences of variations in pickup ion<br />

distributions, in preparation.<br />

As a Contributing Author<br />

5. Winske, D., S. P. Gary, P. Wu and N. A. Schwadron (2009), Pickup ion energization<br />

at the termination shock, in preparation.


189<br />

6. Lee, M. A., H. J. Fahr, H. Kucharek, E. Moebius, C. Prested, N. Schwadron and P.<br />

Wu (2009), Physical processes in the outer heliosphere, Space Science Reviews, doi:<br />

10.1007/s11214-009-9522-9.<br />

7. Schwadron, N. A., G. Crew, R. Vanderspek, F. Allegrini, M. Bzowski, R. DeMagistre,<br />

G. Dunn, H. Funsten, S. A. Fuselier, K. Goodrich, M. Gruntman, J. Hanley, J.<br />

Heerikuisen, D. Heirtlzer, P. Janzen, H. Kucharek, G. Loeffler, K. Mashburn, K.<br />

Maynard, D. J. McComas, E. Moebius, C. Prested, B. Randol, D. Reisenfeld, M.<br />

Reno, E. Roelof and P. Wu (2009), The Interstellar Boundary Explorer Science<br />

Operations Center, Space Science Reviews, doi:10.1007/s11214-009-9513-x.<br />

INVITED TALKS (Chronologically)<br />

1. Substorm associated magnetotail energetic electrons pitch angle evolutions and flow<br />

reversals: Cluster observation, Los Alamos National Laboratory, ISR-1<br />

magnetospheric physics group Seminar, July 6, 2005, Los Alamos, New Mexico.<br />

2. Hybrid simulation of the heliospheric termination shock, Los Alamos National<br />

Laboratory ISR-1 solar & heliospheric physics group Seminar, September 26, 2007,<br />

Los Alamos, New Mexico.<br />

3. Energy dissipation and ion heating at the heliospheric termination Shock, New<br />

England Space Consortium, MIT, January 26, 2009, Boston, Masschusetts.<br />

4. Energy dissipation and ion heating at the heliospheric termination Shock, Los Alamos<br />

National Laboratory ISR-1 solar & heliospheric physics group Seminar, April 22,<br />

2009, Los Alamos, New Mexico.<br />

5. Energy dissipation and ion heating at the heliospheric termination Shock, 8 th Annual<br />

International Astrophysics Conference, May 1–7, 2009, Big Island, Hawaii<br />

APPOINTMENTS<br />

1. Teaching Fellow, Boston University, 2003–2004<br />

2. Graduate Research Assistant, Boston University, 2004–2009<br />

3. Visitor, sponsors/mentors: Reiner Friedel and Geoff Reeves, ISR-1, Los Alamos<br />

National Laboratory, July 5, 2005–July 15, 2005<br />

4. Staff Research Assistant (Intern), sponsors/mentors: S. Peter Gary and Dan Winske,<br />

ISR-1, Los Alamos National Laboratory, July 2, 2007–September 28, 2007<br />

5. Visitor, sponsors/mentors: S. Peter Gary and Dan Winske, ISR-1, Los Alamos<br />

National Laboratory, April 20, 2009 - May 15, 2009<br />

6. Visitor, sponsors/mentors: S. Peter Gary and Dan Winske, ISR-1, Los Alamos<br />

National Laboratory, September 21, 2009 – October 2, 2009


190<br />

CONFERENCE/WORKSHOP EXPERIENCES<br />

1. Center for Integrated Space Weather Modeling (CISM) Summer School (2003:<br />

attended)<br />

2. Geospace Environment Modeling (GEM) workshop (2004: gave a poster; 2005: gave<br />

a student tutorial talk and a poster)<br />

3. Solar Heliospheric & INterplanetary Environment (SHINE) workshop (2008: gave a<br />

poster, chaired an invited talk section; 2009: gave a student tutorial and a poster)<br />

4. AGU Fall Meeting (2004, 2005, 2007, 2008: gave a poster; 2005: gave a talk)<br />

5. IBEX Science Working Team meeting (2007: gave a talk at Orbital Science<br />

Corporation; 2008: gave a talk at BU)<br />

6. New England Space Consortium (2005–now: attended every season)<br />

7. Solar Physics Summer School (2009: attended)<br />

TITLES/AWARDS<br />

Student ambassador, Boston Ballet, 2007–2009<br />

Outstanding student scholarship, USTC, 2000<br />

2 nd prize, dance, College student art festival, Anhui TV, 1999<br />

Outstanding student scholarship, USTC, 1998<br />

DiAo outstanding student scholarship, USTC, 1997<br />

AFFILI<strong>AT</strong><strong>ION</strong>S<br />

Los Alamos National Laboratory, NASA COSMICOPIA, Boston Ballet<br />

PROFESS<strong>ION</strong>AL MEMBERSHIP<br />

Student Member, American Geophysics Union (AGU), 2004–present<br />

Student Member, American Physics Society (APS), 2005–2006<br />

SKILLS U<br />

Operating System: Linux, MAC, Windows<br />

Programming: IDL, FORTRAN, OpenDX, CismDX, C, HTML, LaTex<br />

Language: English (fluent), Mandarin (native), Taiwanese (native)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!