<<

UNIVERSITY OF CALIFORNIA

Los Angeles

Charged Particle Energization and Transport

in the Magnetotail during Substorms

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Physics

by

Qingjiang Pan

2015

ABSTRACT OF THE DISSERTATION

Charged Particle Energization and Transport

in the Magnetotail during Substorms

by

Qingjiang Pan

Doctor of Philosophy in Physics

University of California, Los Angeles, 2015

Professor Maha Ashour-Abdalla, Chair

This dissertation addresses the problem of energization of particles (both and ions)

to tens and hundreds of keV and the associated transport process in the magnetotail during

substorms. Particles energized in the magnetotail are further accelerated to even higher energies

(hundreds of keV to MeV) in the radiation belts, causing space weather hazards to human activities

in space and on ground. We develop an analytical model to quantitatively estimate flux changes

caused by betatron and Fermi acceleration when particles are transported along narrow high-speed

flow channels from the magnetotail to the inner . The model shows that energetic

particle flux can be significantly enhanced by a modest compression of the magnetic field and/or

ii

shrinking of the distance between the magnetic mirror points. We use coordinated spacecraft

measurements, global magnetohydrodynamic (MHD) simulations driven by measured upstream

solar wind conditions, and large-scale kinetic (LSK) simulations to quantify local

acceleration in the near-Earth reconnection region and nonlocal acceleration during plasma earthward transport. Compared to the analytical model, application of the LSK simulations is much less restrictive because trajectories of millions of test particles are calculated in the realistically determined global MHD fields and the results are statistical. The simulation results validated by the observations show that electrons following a power law distribution at high energies are generated earthward of the reconnection site, and that the majority of the energetic electrons observed in the inner magnetosphere are caused by adiabatic acceleration in association with magnetic dipolarizations and fast flows during earthward transport. We extend the global

MHD+LSK simulations to examine ion energization and compare it with electron energization.

The simulations demonstrate that ions in the magnetotail are first nonadiabatically accelerated in the weak field region close to the reconnection site, and then adiabatically accelerated in the high- speed flow channels as they catch up with and ride on the earthward propagating dipolarization structures. The nonlocal adiabatic acceleration mechanism for ions is very similar to that for electrons. However, the motion of energetic electrons is adiabatic except in very limited regions near the reconnection site while the motion of energetic ions is marginally adiabatic in the dipolarization regions. The simulations also show that the earthward transport of both species is controlled by the high-speed flows via the dominant ExB drift in the magnetotail. To understand how the power law electrons are initially produced in the magnetotail, we use an implicit particle- in-cell (PIC) code to model the processes in the near-Earth reconnection region. We find that the power law electrons are produced not in the reconnection diffusion region, but in the immediate

iii downstream of the reconnection outflow in the course of dipolarization formation and intensification. Our study illustrates that during substorms, particles are accelerated via a multi- step process, including local acceleration in the reconnection region and nonlocal acceleration during the earthward transport, and the multi-step acceleration occurs on multiple spatial scales ranging from a few kilometers (the scale of electron diffusion region) to more than ten Earth radii

(the transport scale).

iv

The dissertation of Qingjiang Pan is approved.

George Morales

Christopher T. Russell

Raymond J. Walker

Maha Ashour-Abdalla, Committee Chair

University of California, Los Angeles

2015

v

To my mother

vi

TABLE OF CONTENTS

ABSTRACT ii

DEDICATION vi

LIST OF FIGURES xi

LIST OF SYMBOLS xiii

ACKNOWLEDGEMENTS xiv

VITA xvii

1. Background and Purpose of this Study 1

1.1. Introduction ...... 1

1.2. Acceleration by Magnetic Reconnection ...... 7

1.3. Acceleration during Plasma Earthward Transport ...... 13

1.4. Purpose of this Study ...... 24

1.5. Structure of the Dissertation ...... 26

2. Theory of Adiabatic Acceleration of Charged Particles 28

2.1. Introduction ...... 28

2.2. Characteristics of Adiabatic Particle Orbits ...... 29

vii

2.3. An Analytical Model of Adiabatic Acceleration ...... 33

2.3.1. Motivation ...... 33

2.3.2. The Adiabatic Acceleration Model ...... 34

2.3.3. Comparisons with Observations ...... 38

2.3.4. Discussions ...... 49

3. Modeling Electron Energization and Transport in the Magnetotail during

a Substorm 52

3.1. Introduction ...... 52

3.2. Observations of the March 11, 2008 Substorm Event ...... 52

3.2.1. Geotail Observations of the Solar Wind ...... 52

3.2.2. THEMIS Observations in the Magnetotail ...... 54

3.3. Simulation Methodology ...... 60

3.4. MHD and Electron LSK Simulations of the March 11, 2008 Substorm Event ...... 62

3.4.1. MHD Simulation Results ...... 62

3.4.2. LSK Simulation Results and Comparisons with Observations ...... 67

3.5. Discussions ...... 75

3.6. Conclusions ...... 79

4. Modeling Ion Energization and Transport Associated with Magnetic

Dipolarizations during a Substorm 81

viii

4.1. Introduction ...... 81

4.2. Observations of the February 07, 2009 Substorm Event ...... 81

4.3. MHD Simulation of the February 07, 2009 Substorm Event ...... 88

4.4. Ion LSK Simulation of the February 07, 2009 Substorm Event ...... 92

4.4.1. LSK Simulation Set-up ...... 92

4.4.2. LSK Simulation Results and Comparisons with Observations ...... 93

4.5. Conclusions and Discussions ...... 100

5. A Comparison Study of Ion and Electron Energization and Transport

Mechanisms during a Substorm 103

5.1. Introduction ...... 103

5.2. Comparisons of the Electron and Ion LSK Simulation Set-up for the February

07, 2009 Substorm Event ...... 104

5.3. Simulation Results ...... 107

5.3.1. Comparisons of Simulation Results with Observations ...... 107

5.3.2. Comparisons of Ion and Electron Acceleration Mechanisms ...... 111

5.4. Conclusions and Discussions ...... 121

6. Particle-in-cell (PIC) Simulation of Electron Acceleration by Magnetic

Reconnection 125

6.1. Introduction ...... 125

ix

6.2. Simulation Methodology ...... 126

6.3. Simulation Results and Comparisons with Observations ...... 130

6.3.1. Reconnection Structure ...... 130

6.3.2. Electron Acceleration ...... 134

6.4. Conclusions ...... 147

7. Conclusions and Problems for the Future 149

7.1. Conclusions ...... 149

7.2. Unsolved Problems and Future Work ...... 159

APPENDIX 1: Particle Sources for LSK Simulations 165

BIBLIOGRAPHY 169

x

LIST OF FIGURES

1.1 A schematic of the magnetospheric convection driven by magnetic reconnection ...... 2

2.1 Energy flux and power law index at THEMIS P2 in the March 11, 2008 event ...... 41

2.2 Energy flux and power law index at THEMIS P4 in the March 11, 2008 event ...... 42

2.3 Betatron and Fermi acceleration in the March 11, 2008 event...... 44

2.4 Energy flux and power law index at THEMIS P1 in the February 27, 2009 event ...... 46

2.5 Energy flux and power law index at THEMIS P4 in the February 27, 2009 event ...... 47

2.6 Betatron and Fermi acceleration in the February 27, 2009 event ...... 48

3.1 Geotail observations of the solar wind on March 11, 2008 ...... 53

3.2 THEMIS locations and the observed magnetic fields in the March 11, 2008 event ...... 55

3.3 THEMIS P2 observations in the March 11, 2008 event ...... 57

3.4 THEMIS P4 observations in the March 11, 2008 event ...... 59

3.5 Snapshots of the MHD and LSK simulations for the March 11, 2008 event ...... 64

3.6 Two types of electron source distributions for the LSK simulation ...... 68

3.7 Comparison of the energy flux for THEMIS P2 in the March 11, 2008 event ...... 71

3.8 Comparison of the energy flux for THEMIS P4 in the March 11, 2008 event ...... 72

3.9 Comparison of the differential flux as a function of energy for THEMIS P2 ...... 73

3.10 Comparison of the differential flux as a function of energy for THEMIS P4 ...... 74

3.11 Electron acceleration mechanism during transport in the March 11, 2008 event ...... 76

3.12 Electron transport in the March 11, 2008 event ...... 78

xi

4.1 Solar wind measured by WIND and positions in the February 07, 2009 event ...83

4.2 THEMIS P2 observations in the February 07, 2009 event ...... 85

4.3 THEMIS P3 observations in the February 07, 2009 event ...... 87

4.4 Snapshots of the MHD and ion LSK simulations for the February 07, 2009 event ...... 90

4.5 Comparison of observations and simulations for P3 for the February 07, 2009 event ...... 94

4.6 Characteristics of a representative test ion ...... 98

5.1 Comparisons of observations with simulation results for THEMIS P3 ...... 110

5.2 A Snapshot of the MHD and LSK simulations for the February 07, 2009 event ...... 113

5.3 Characteristics of a representative ion ...... 115

5.4 Characteristics of a representative electron ...... 117

5.5 Kappa in the February 07, 2009 event ...... 121

6.1 Reconnection rate normalized by the upstream Alfvén velocity and magnetic field ...... 132

6.2 Snapshots of the reconnection at t=3.05 sec and t=5.25 sec ...... 133

6.3 Electron heating and production of energetic electrons at t=3.05 sec and t=5.25 sec .....136

6.4 Dipolarization and reconnection outflow at the center of the current sheet (Z=0) ...... 138

6.5 Normalized electron distribution functions at the DF and in the EDR ...... 141

6.6 Electron velocity distributions at the DF and in the EDR at t=5.25 sec...... 144

6.7 Densities of electrons and ions from the current sheet and background ...... 146

7.1 Distribution functions produced in the model of particle multistep energization on

multiple scales in the magnetotail during substorms ...... 157

xii

LIST OF SYMBOLS t time u guiding-center parallel velocity r position uE guiding-center EB drift velocity v particle velocity uB guiding-center gradient drift velocity V fluid velocity uc guiding-center curvature drift f distribution function velocity q electrical charge uac guiding-center acceleration drift m mass velocity

E electric field vA Alfvén speed E electric field strength or  resistivity kinetic energy  magnetic field compressional factor

B magnetic field  contraction factor of bounce distance B magnetic field strength between mirror points c speed of light in vacuum  solid angle  particle gyro radius  pitch angle w particle gyration velocity  gyro phase

W kinetic energy  kappa parameter

c particle gyro frequency D Debye length s unit length along field line  pe plasma frequency M magnetic moment di ion inertial length R guiding-center position

de electron inertial length uGC guiding-center velocity

xiii

ACKNOWLEDGEMENTS

I wish to first thank my thesis advisor, Professor Maha Ashour-Abdalla. Her guidance, advice and enthusiasm have been invaluable to me every step of my graduate study and research.

She has supported my graduate study and research by providing the best research tools, by creating good opportunities for academic activities, and most of all, by pointing to me a correct direction of research. Every time I came up with unrealistic and “ambitious” ideas, she reminded me of current research status in space physics. This is a very important reason that I could conduct good scientific projects and graduate in time. I thank her also because she has offered me a fair amount of freedom in details of my work and has encouraged me to explore new area. I have been grateful to her and hopefully will “become more mature in the future” (in her words).

I am grateful to Professor Raymond J. Walker, who has also advised me every step of my research. His patience and kindness have been influential to me. He has helped me revised every manuscript, poster, presentation slide, and of course this dissertation. He has taught me how to write in an elaborating fashion. He and Maha constantly suggested me writing what progresses I have made instead of criticizing others’ work. I would like to thank Professor George Morales, who taught me plasma physics. His clear and thorough teaching is apparently invaluable to my research. He has also been kind enough to give me advice on multiple occasions and to serve on my thesis committee. I would like to thank Professor Christopher Russell and Professor Christoph

Niemann for serving on my thesis committee and offering advice on my research. Professor

Giovanni Lapenta was very generous to let me use his state-of-the-art implicit PIC code and has helped me tremendously in finishing Chapter 6 of the thesis. I appreciate Professor Richard Sydora for his advice and hospitality during my visit at University of Alberta.

xiv

I learned from scientists in the space plasma simulation group. Dr. Dave Schriver developed the electron LSK code, which has been an indispensable tool in my research. He has also helped me understand particle motion in the magnetotail. Dr. Mostafa El-Alaoui has helped me on the MHD simulations and provided excellent MHD fields that are used in this dissertation.

Dr. Robert Richard was kind enough to read some of my manuscripts and provided good comments.

I am grateful to Haoming Liang, who has been a good friend and my companion of research. He has helped me deal with numerous details.

I would like to thank Greg Kallemeyn in helping me edit part of the thesis and some of my manuscripts.

Most of all, I wish to thank my parents. Their unconditional love and support make me feel fortunate. My love, gratitude and debt to them are beyond measure.

The research in this dissertation was support by a Magnetospheric Multiscale Mission

Interdisciplinary Scientist grant (NASA grant NNX08AO48G at UCLA), Geospace grant (NASA grant NNX12AD13G) and NASA grant NNX10AQ47G. We acknowledge V. Angelopoulos and the THEMIS Science Support team for the use of data and software (TDAS) from the THEMIS

Mission, and specifically, C.W. Carlson and J.P. McFadden for the use of ESA data, D. Larson and R.P. Lin for the use of SST data, D.L. Turner for calibration of SST data, K.H. Glassmeier, U.

Auster, and W. Baumjohann for the use of FGM data, J.W. Bonnell and F.S. Mozer for the use of

EFI data, and A. Roux and O. LeContel for the use of SCM data. Part of the computations were carried out with the support of the NASA Advanced Supercomputing Facility at the Ames

Research Center, and the Gordon supercomputer at San Diego. The Gordon supercomputer is part of the Extreme Science and Engineering Discovery Environment (XSEDE) project, which is supported by National Science Foundation grant OCI-1053575. We would like to acknowledge

xv

high-performance computing support from Yellowstone (ark:/85065/d7wd3xhc) provided by

NCAR's Computational and Information Systems Laboratory, sponsored by the National Science

Foundation. This work used computational and storage services associated with the Hoffman2

Shared Cluster provided by UCLA Institute for Digital Research and Education's Research

Technology Group.

Chapter 2, Chapter 3, Chapter 4, Chapter 5, and Chapter 6 are respectively reproduced with

significant modifications from manuscripts Pan et al. [2012], Pan et al. [2014a], Pan et al. [2014b],

Pan et al. [2015a], and Pan et al. [2015b]. I would like to thank my co-authors and the publishers

for this matter. Future use of the content of these copyrighted materials shall not infringe the

copyright of their respective owners.

xvi

VITA

Education 07/2010 B.S. (with honors), in Space Physics, Peking University 03/2012 M.S. in Physics, UCLA

Selected Awards 2007-2010 “Zeng Xianzi” Scholarship 2007-2010 Xin Changcheng Scholarship 2007-2009 Chancellor Fund for Undergraduate Student Research 2007-2008 Scholarship for Excellent Academic Performance

Research Experience 09/2007-06/2010 Undergraduate Student Research Assistant Peking/Beijing University, Beijing, China Advisors: Prof. Zuyin Pu and Prof. Lun Xie Analyzed ultra-low frequency waves data obtained by conjugate NASA THEMIS multi-spacecraft and IMAGE ground-based .

10/2012-12/2012 Visiting Scholar University of Alberta, Edmonton, Canada Collaborator: Prof. Richard D. Sydora Simulated and analyzed magnetic reconnection and plasma waves with an implicit particle-in-cell (PIC) code.

09/2010-present Graduate Student Researcher University of California, Los Angeles, USA PhD thesis topic: charged particle energization and transport in the magnetotail during substorms. Advisors: Prof. Maha Ashour-Abdalla and Prof. Raymond J. Walker Responsibilities include:  Modeling particle energization and transport in the terrestrial space with a parallelized magnetohydrodynamics (MHD) code and large- scale kinetic (LSK) codes.  Developing a code for particle sources of the LSK simulations.  Modeling particle energization by processes near reconnection sites with a parallelized implicit PIC code.  Developing theoretical plasma physics models to quantitatively predict effects of particle energization.  Calibrating, analyzing and visualizing multi-spacecraft time series data from NASA and ESA missions, including THEMIS, CLUSTER, Geotail; and using the satellite data to test and validate (or invalidate) simulation and analytical models.

xvii

Publications Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker, and Melvin L. Goldstein (2012), Adiabatic Acceleration of Suprathermal Electrons Associated with Dipolarization Fronts, JGR-Space Physics, 117, A12224, doi:10.1029/2012JA018156.

Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker (2014), Electron Energization and Transport in the Magnetotail during Substorms, JGR-Space Physics, 119, doi:10.1002/2013JA019508.

Pan Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, Mostafa El-Alaoui (2014), Ion Energization and Transport Associated with Magnetic Dipolarizations, GRL-Space Physics, 41, doi:10.1002/2014GL061209.

Melvyn L. Goldstein, Maha Ashour-Abdalla, Adolfo F. Viñas, John Dorelli, Deirdre Wendel, Alex Klimas, Kyoung-Joo Hwang, Mostafa El-Alaoui, Raymond J. Walker, Qingjiang Pan, Haoming Liang(2015), Mission Oriented Support and Theory (MOST) for MMS—the Goddard Space Flight Center/UCLA Interdisciplinary Science Program, Space Science Reviews, DOI 10.1007/s11214- 014-0127-6 (book chapter).

Pan Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, Mostafa El-Alaoui (2015a), A Comparison of Electron and Ion Energization and Transport Mechanisms in the Magnetotail during Substorms, submitted to JGR-Space Physics.

Pan Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, Giovanni Lapenta (2015b), Production of Power Law Electrons: Magnetic Field Diffusion or Dipolarization, in preparation for publication.

Talks Pan Qingjiang, Energization of Electrons Associated with Dipolarization Fronts, space physics seminar, University of Alberta, Edmonton, Canada, November 2012.

Pan Qingjiang, Ion Energization and Transport in the Magnetotail during Substorms, the NASA Magnetospheric Multiscale Mission (MMS) meeting, University of Iowa, Iowa city, USA, March 2014.

Pan Qingjiang, Particle Energization and Transport in the Magnetotail during Substorms, space physics seminar, UCLA, Los Angeles, USA, April 2014.

Posters Pan Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, What Breaks Magnetic Field Lines- A Revisit of Reconnection Theory, the 10th International School/Symposium for Space Simulations (ISSS-10), Banff, Alberta, Canada, July 2011.

Ashour-Abdalla, Maha., Mostafa El-Alaoui, David Schriver, Pan Qingjiang, Robert Richard, Meng Zhou, and Raymond J. Walker, Electron Acceleration Associated with Earthward

xviii

Propagating Dipolarization Fronts, AGU Fall meeting, San Francisco, California, USA, December 2011.

Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker, David Schriver, and Robert L. Richard, Adiabatic Acceleration of Suprathermal Electrons Associated with Dipolarization Fronts, Geospace Environment Modeling (GEM) meeting, Snowmass, Colorado, USA, June 2012.

Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker, David Schriver, and Robert L. Richard, On the Importance of Magnetic Reconnection and Adiabatic Acceleration for Suprathermal Electrons Associated with Dipolarization Fronts, AGU Fall meeting, San Francisco, California, USA, December 2012.

Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker, Electron Energization and Transport in the Magnetotail during a Substorm, Geospace Environment Modeling (GEM) meeting, Snowmass, Colorado, USA, June 2013.

Pan Qingjiang, Maha Ashour-Abdalla, Mostafa El-Alaoui, Raymond J. Walker, Electron Energization and Transport in the Magnetotail during a Substorm, AGU Fall meeting, San Francisco, California, USA, December 2013.

Pan, Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, Mostafa El-Alaoui, Ion Energization and Transport in the Magnetotail during Substorms, Geospace Environment Modeling (GEM) meeting, Portsmouth, Virginia, USA, June 2014.

Pan Qingjiang, Maha Ashour-Abdalla, Raymond J. Walker, Mostafa El-Alaoui, A Comparison Study of Electron and Ion Energization and Transport Mechanisms in the Magnetotail during Substorms, AGU Fall meeting, San Francisco, California, USA, December 2014.

xix

CHAPTER 1

Background and Purpose of this Study

1.1. Introduction

The problem of how particles in the magnetosphere are energized to tens of keV to MeV

has been under intensive investigations since the beginning of the space age [Kivelson and Russell,

1995, and references therein]. These magnetospheric electrons and ions originate from the solar wind and the ionosphere with typical initial energies of a few eV. The energization occurs through a multistep process occurring in multiple regions: (1) the bow shock formed in front of the magnetopause, through which the super-Alfvénic solar wind is slowed down to adjust to the magnetospheric obstacle; (2) the dynamic magnetotail and its connection to aurora; and (3) the radiation belts located in the inner magnetosphere where particles are trapped. This dissertation addresses the problem of energization of particles (both electrons and ions) to tens and hundreds of keV [e.g. Baker et al., 1981] and the associated transport process in the magnetotail during substorms.

In a two-page paper, Dungey [1961] proposed for the first time an open magnetospheric convection model driven by magnetic reconnection at the magnetopause and in the magnetotail.

This model has been a major framework to study the dynamics of the Earth’s magnetosphere.

Figure 1 (top panel) shows a schematic view of the open magnetospheric convection model

[Hughes, 1995]. For simplicity, we assume that the interplanetary field is predominantly southward.

Then the magnetic field driven by the solar wind flowing against the front of the magnetosphere will be approximately antiparallel to the geomagnetic field on the other side of the magnetopause.

1

1’ bow shock 2 3 4 magnetosheath magnetopause

5

6 1 7 9 8 jet solar reconnection 7’ wind 6’

5’

3’ 2’ 4’ z 1’ x

y

Vi By>0 Ve By<0

δe δi

By<0 By>0 yellow: IDR green: EDR Le Li

Figure 1.1. A schematic of the magnetospheric convection driven by magnetic reconnection. (Top)

Flow of plasma within the magnetosphere. The numbered field lines show the succession of

2 configuration of the geomagnetic field line (1) after reconnection with an interplanetary magnetic field line (1’) at the magnetopause. Field lines 6 and 6’ reconnect at an X-line in the tail, after which the field line returns to the dayside at lower latitudes, adapted from Hughes [1995]. (Bottom)

A schematic of the night side reconnection region. In the upstream inflow region, electrons and ions are magnetized and convected with the magnetic field toward the diffusion region. The ion inflow is diverted in the ion diffusion region (IDR), where ions are demagnetized and electrons are magnetized. After the ion inflow diversion, the magnetized electrons continue to flow toward the center, until they become demagnetized in the electron diffusion region (EDR). The separation of electron and ion flows in the IDR generates the quadrupole Hall magnetic fields in the Y- direction, adapted from Drake and Shay [2006].

Reconnection occurs between field line 1 and 1’. The newly reconnected field line is dragged by the solar wind successively via the numbered locations 2, 3, and 4, entering the magnetopause on the night side (location 5). The field line is reconnected at location 6, becoming a closed one. It then returns to the Earth through locations 7, 8 and eventually to the dayside through location 9, completing a global magnetospheric convection cycle. The reconnection on the magnetopause is asymmetric, while the reconnection on the night side tends to be symmetric about the current.

During geomagnetic quiet times, the average location of the night side reconnection is at

XRGSM~110 E [McPherron, 1991]. The geocentric solar magnetospheric (GSM) coordinate system is used in this dissertation. In this system, the X-direction points to the , the Z-direction points to northern magnetic pole, and Y-direction completes the orthogonal system according to the right-hand rule, pointing to the west or dusk side.

3

Unlike the steady large-scale magnetospheric convection, magnetospheric substorms are

dynamic processes, in which magnetic energy is explosively released and transferred to particle

kinetic energy [e.g. Akasofu, 1964; McPherron, 1972; McPherron et al., 1973; Russell and

McPherron, 1973; Baker et al., 1981; McPherron, 1991; Lui, 1996; Baker et al., 1996;

Angelopoulos et al., 2008, and references therein]. Consider the magnetosphere as a system. The solar wind drives the system through dayside reconnection and the magnetosphere responds internally in a passive fashion through convection throttled by the distant tail reconnection. During the substorm growth phase, the external solar wind driving exceeds the capacity of internal

convection, resulting in accumulation of magnetic flux in the magnetotail. The accumulated

nightside magnetic flux increases magnetic pressure and energy, compressing and stretching the

magnetotail. In the subsequent expansion phase, the stretched configuration becomes unstable to

tearing. A new neutral line, called the near-Earth X-line, typically forms at

30RXE GSM  20 R E , sometimes as close as XRGSM~10 E [McPherron, 1991], where the

stretched field lines reconnect. The near-Earth reconnection changes field line topology and

generates Alfvénic outflow jets. As the jets propagate to (and away from) the Earth, the magnetic

pressure and energy stored in the stretched tail is released via an abrupt reconfiguration of the

magnetic field to a dipole-like shape. The near-Earth reconnection, the earthward flow jets, the

global dipolarization of the magnetic field, and the associated complex processes closer to the

Earth (e.g. disruption of the cross-tail electric current) give rise to the auroral activity associated with substorms [Akasofu, 1964]. The abrupt reconfiguration of the tail magnetic field following an intensive external driving and instability triggering is analogous to the sawtooth oscillation in a tokamak plasma [Biskamp, 2005]. Eventually, substorm expansion exhausts excess flux and energy in the tail, the X-line retreats to distant tail and the magnetosphere is stabilized during the

4

substorm recovery phase. This qualitative framework of substorms is called the near-Earth neutral

line (NENL) model (see the review by Baker et al. [1996]). The NENL model seems to be

generally accepted by a majority of the magnetospheric community as the basic model of the

processes in the tail during substorms. The data and simulations in this dissertation are consistent

with this model. However, the details of the model have been under intensive debate mainly

because satellite observations are only conducted at a few points. There are also substorm models

that do not emphasize the role of the near-Earth magnetic reconnection. Interested readers are

encouraged to read the comprehensive review on substorms by McPherron [1991] and a more

recent review by Sergeev et al. [2012]. Note the overview presented in Figure 1.1 for

magnetospheric convection is also applicable to the substorm expansion phase, with three caveats.

First, the night side reconnection shown should be understood as the near-Earth reconnection X-

line. Second, in the beginning of the expansion phase, due to reconnection in the distant tail

( XRGSM~110 E ) and the newly formed near-Earth X-line, plasmoids with a transverse scale of

tens of RE are generated and ejected tailward, which are not shown in the figure. Third, the figure does not show variations in the Y-direction; during substorm expansion phase, the azimuthal variations are expected to be substantially different from that during steady magnetospheric convection (details are in section 1.3).

In the global substorm model, the critical near-Earth reconnection occurs on a microscopic scale. Figure 1.1 (bottom panel) shows schematically the reconnection region. In the inflow region, ions and electrons are magnetized and convected with magnetic field toward the diffusion region.

The ion inflow is diverted and ejected in the ion diffusion region (IDR), where ions are demagnetized whereas electrons are magnetized. After the ion inflow diversion, the electrons continue to flow with the magnetic field toward the center, and they are demagnetized in the

5

electron diffusion region (EDR). The separated electron and ion flows in the IDR constitute the

electric Hall currents, which generate quadrupole Hall magnetic fields By near the separatrices

[Sonnerup, 1979]. The IDR width i is several times the ion inertial length di ; its length Li is

tens of times di . The EDR width e is a few times the electron inertial length de and the EDR

length Le is a few to tens of times the di , shorter than but comparable to Li [Daughton et al.,

2006]. Note that close to each of the separatrices, there are counter-streaming fast electron beams consisting of an incoming beam toward the X-line in the inflow region and an outgoing beam away from the X-line in the outflow region, which also contribute to the Hall electric currents [Hoshino et al., 2001; Pritchett, 2001]. The beams are not depicted in the diagram.

The near-Earth magnetic reconnection plays an instrumental role in the magnetic-to-kinetic energy transfer during substorms. On one hand, close to the X-line, magnetic energy is directly

converted to kinetic energy via localized magnetic diffusion, resulting in plasma heating and

production of energetic electrons. On the other hand, the topological change of the tail field lines by reconnection enables subsequent particle energization when the reconnection outflow jets

propagate to the Earth. Accordingly, particle acceleration mechanisms in the magnetotail proposed

by previous studies can be categorized in two classes: those operating near the reconnection

diffusion region and those occurring during plasma earthward transport (see sections 1.2 and 1.3 for detailed references). The former class is termed local acceleration, and the latter class is termed

nonlocal acceleration. We stress that there is no clear distinction between local acceleration and

nonlocal acceleration in the Earth’s magnetotail. In addition, some of the studies emphasizing local

acceleration considered energization in the immediate reconnection outflow region, while some of

the studies on nonlocal acceleration also discussed particle energization resulting from

6

reconnection acceleration. Nevertheless, we feel it is an approximate framework to organize the

large and growing volume of literature on particle energization in the magnetotail. Therefore, we

review local and nonlocal acceleration separately in the following two sections.

1.2. Acceleration by Magnetic Reconnection

The suggestion that magnetic reconnection could be a significant cause of particle

acceleration in cosmic plasmas was first made by Giovanelli [1947]. In the Earth’s magnetotail,

magnetic reconnection is believed to be particularly important in accelerating particles, because

the magnetotail-like field reversal across the current sheet provides favorable conditions for

reconnection, through which magnetic energy can be rapidly transferred to particle kinetic energy.

Using measurements by the (IMP7) satellite, Sarris et al., [1976] reported particle

bursts with proton energy E p  0.29MeV and electron energy EMeVe  0.22 in the plasma sheet

at X ~ 35RE , and suggested that one of the possible acceleration mechanisms was the annihilation of magnetic field lines at the region of the neutral line. Using

measurements at X ~ (20 30)RE , Terasawa and Nishida [1976] reported energetic electron

bursts of EMeVe  0.3 ~ 1.0 in association with the southward turning of Z-component of the

local magnetic field. They suggested that these electrons are accelerated at the neutral line and

trapped in the magnetic flux rope tailwards of the neutral line. Similar electron bursts of

EkeVe  200 in association with southward turning of Bz and tailward plasma jets have also been

observed at X ~ 30RE with the IMP8 spacecraft [Baker and Stone, 1976, 1977]. Möebius et al.

[1983] reported measurements of energetic protons of E p  20 500keV and energetic electrons

7

of EkeVe  75 from ISSE-1 and ISSE-2 and suggested that the energetic particles were accelerated near an earthward moving X-line in the event analyzed. Energetic particle bursts in these early measurements have been related to magnetic reconnection. However, no direct evidence of magnetic reconnection were reported in association with the energetic particles.

Recent observations have made significant progresses in directly connecting energetic

particles with reconnection characteristics. Øieroset et al. [2002] reported an event in which the

WIND spacecraft at X ~ 60RE detected energetic electrons (up to 300 keV) in the IDR with Hall

reconnection characteristics. They found no enhancements of energetic ion fluxes in this event.

Hoshino et al. [2001] identified an event of earthward crossing of a reconnection X-line by the

Geotail spacecraft at X ~ 24RE , and found that a power law distribution of suprathermal

electrons (EkeVe  20 ) that extended from the Maxwellian distribution of hot thermal electrons

(2-3 keV) was generated by the magnetic reconnection. Using data obtained from Cluster

spacecraft at X ~ 16RE and YR ~ 8 E , Imada et al. [2007] identified a reconnection event and

found that electrons were accelerated to about 127 keV via a two-step acceleration process, namely

initial acceleration at the X-line and subsequent acceleration in the immediate downstream reconnection outflow region. Note that the 127 keV energy was the highest energy channel

presented in the study (same for the studies with Cluster data discussed below). The energy channel

was determined by the RAPID instrument onboard Cluster [Wilken et al., 2001]. The two-step

acceleration scenario derived from an event study by Imada et al. [2007] is consistent with their

earlier statistical study of the average profiles of energetic and thermal electrons in the magnetotail

reconnection regions by using Geotail data, which showed that due to the Earth’s dipole field

energetic electron flux is much stronger in the earthward side of X-line than in the tail side [Imada

8 et al., 2005]. Their later statistical study with Geotail data found that electrons are efficiently accelerated to E 38 keV in a thin current sheet during fast reconnection events [Imada et al.,

2011]. Chen et al. [2008] analyzed Cluster data during a reconnection event and found that electrons are accelerated in multiple magnetic islands (flux ropes). Using data from Cluster

spacecraft at X ~ 17RE , Retinò et al. [2008] reported the spacecraft crossed a flux rope in the

IDR. They found that the flux of electrons is largest within the flux rope where they are mainly directed perpendicular to the magnetic field. At the magnetic separatrices, the fluxes are smaller, but the energy spectra are harder and electrons are mainly field aligned. Similar events were reported by Wang et al. [2010] and Huang et al. [2012a]. During these events, energetic electron flux is larger in a downstream magnetic island than that in the reconnection current sheet. Even though all the aforementioned Cluster observations were limited to below 127 keV, they did show that the flux increases at 127 keV were weak, suggesting electron acceleration upper limit was close to that energy. To summarize, the overwhelming evidence in recent observations show electrons are accelerated to tens of keV to about a hundred keV by processes close to the X-line, while there are very few corresponding observations of energetic ions.

A number of analytical models of particle energization by reconnection have been developed. The simplest analytical models are stationary [Bulanov and Sasorov, 1976; Burkhart et al., 1990; Vekstein and Priest, 1995]. These models assume simple magnetic field

BeBzL00()zz BxL () xx e with characteristic lengths Lx and Lz in the X- and Z-directions,

and a constant and uniform electric field along the Y-direction E  E0e y . The basic idea is to solve particle trajectories in the prescribed stationary fields (differential equations), and to estimate particle asymptotic energy gain from the electric field when they are thrown out of the

reconnection region by the finite Lorentz force vByz . A more complicated class of models

9

consider the effects related to the formation of the X-line in the course of substorm expansion,

which can be described as the growth of an ion tearing mode [Galeev et al., 1978, 1979]. The

growth of the ion tearing mode generates an inductive electric field that can accelerate particles in

the vicinity of the X-line. With appropriate assumptions of the length of the X-line and the

evolution of the dynamic fields, it is estimated that the inductive electric field can accelerate both

electrons and protons to MeV energies in the Earth’s magnetotail [Zelenyi et al., 1984, 1990a].

Surprisingly, models like these are the only analytical ones for considering particle acceleration

by magnetic reconnection. Various mechanisms of particle acceleration near the diffusion region

are derived from numerical simulations and/or suggested by observational data. These models

include:

a. acceleration by a reconnection electric field in the vicinity of the X-line. Essentially,

all the aforementioned analytical models deal with this mechanism, which was later

confirmed by kinetic simulations [Ricci et al., 2003; Pritchett, 2006a, 2006b].

b. acceleration by an electric field parallel to the magnetic field; the parallel electric field

is supported by electron density cavities and/or pressure anisotropy [Drake et al., 2003,

2005; Egedal et al., 2005, 2012, 2013]. This mechanism operates favorably near the

reconnection separatrices, where the fast electron beams are subjected to streaming

instabilities, leading to nonlinear turbulence and generating electron density cavities

[Goldman et al., 2008].

c. acceleration due to particle Speiser motion in the diffusion region. Speiser motion was

originally proposed for particle acceleration in the tail current sheet with Bx reversal

across current sheet, a weak Bz and a constant Ey [Speiser 1965, 1967]. However,

because in the classical two-dimensional reconnection picture, the geometry in the

10

diffusion region satisfies the conditions for Speiser motion (recall the fields in the

aforementioned stationary models). It was suggested that electrons and ions can be

accelerated to their respective Alfvén velocities in the IDR [Shay et al., 2001; Hoshino

et al., 2001].

d. acceleration due to gradient and curvature drift along the electric field ( Ey ) direction

in the immediate downstream pileup region [Hoshino et al., 2001; Imada et al., 2007].

This mechanism is similar to the adiabatic acceleration during plasma earthward

transport farther away the reconnection site (see section 1.3). The difference is the

distance between the acceleration site and the X-line, and therefore the nature of

particle motion is different. Particle motion tends to be nonadiabatic close to the X-line,

and becomes more adiabatic farther away.

e. wave-particle interaction [Cattell et al., 1994; Okada et al., 1994; Shinohara et al.,

1998; Cairns and McMillan, 2005]. This mechanism is mainly speculated from

observations of waves with frequency close to the lower-hybrid frequency in the

reconnection region. However, it was suggested that hybrid waves scatter electrons to

larger pitch angles so that they can gain more energy in the equatorial plane via

mechanism (d) [Hoshino et al., 2001].

f. acceleration by a reconnection electric field when particles are trapped in the O-point

type field geometry (magnetic islands) [Vasyliunas, 1980; Mattaeus et al., 1984;

Goldstein et al., 1986].

g. acceleration due to contraction of magnetic islands [Drake et al., 2006], coalescence of

magnetic islands [Pritchett, 2008; Oka et al., 2010], and particle stochastic motion

11

across multiple magnetic islands [Drake et al., 2006; Hoshino, 2012]. These

mechanisms are derived from simulations.

The observational evidence for mechanisms (f) and (g) are that there are high-

energy particles, in particular high-energy electrons, in multiple magnetic islands (flux

ropes in 3D) near the reconnection region [Chen et al., 2008; Retinò et al., 2008; Wang

et al., 2010; Huang et al., 2012a], although it less definitive that these electrons are

accelerated by the magnetic islands.

h. surfing/surfatron acceleration [Sagdeev and Shapiro, 1973]. This mechanism describes

particle acceleration by the reconnection electric field ( Ey ) when they are trapped in

the current sheet. The trapping is due to the polarization electric field ( Ez ) developed

in the boundary layer between the current sheet and the lobe for thin current sheet

reconnection [Hoshino, 2005]. The polarization electric field ( Ez ) is induced in

association with the Hall reconnection [Birn et al., 2001; Nagai et al., 2001].

Each of these models favors a particular kind of scenario, and obviously has been and will

continue evolving as satellite measurements improve and computer capacity increases. To a large

extent, all the aforementioned models (except for (e)) are derived from two-dimensional (2D)

reconnection studies. Three-dimensional (3D) reconnection has been shown to be substantially

different from 2D reconnection [Daughton et al., 2011]. For example, with a finite guide field, the

3D evolution is dominated by the formation and interaction of helical flux ropes (corresponding

to magnetic islands in 2D), resulting from secondary instabilities within the electron layers. New

flux ropes spontaneously appear within these layers, leading to a turbulent evolution. How particles

are energized in realistic 3D dynamics remains poorly understood. In addition to the reduction of

dimensionality, simulations from which these mechanisms are derived are often compromised by

12

small simulation domains and unrealistic proton-to-electron mass ratio. With small simulation

domains, it is difficult to develop a large picture of particle energization. An artificial mass ratio,

along with other unrealistic physical parameters such as the magnitude of the Alfvén speed and

particle thermal speed in terms of the speed of light, pose difficult questions on interpreting the

final particle energies and quantifying particle energization. Complicated by the use of unrealistic

physical parameters, the simulation results referenced above were usually not vigorously

compared with observations in the same units. Relativistic electrons with energies close to the rest

mass energy are easily produced in these simulations [e.g. Hoshino et al., 2001; Drake et al., 2005;

Pritchett, 2006a, 2006b], while the aforementioned observations in or very close to the diffusion region showed that electrons are accelerated only to tens of keV to about a hundred keV energy range in the reconnection diffusion region [Øieroset et al., 2002; Imada et al., 2005, 2007; Chen

et al., 2008; Retinò et al., 2008; Wang et al., 2010; Imada et al., 2011; Huang et al., 2012a].

1.3. Acceleration during Plasma Earthward Transport

While magnetic reconnection in the magnetotail is actively examined as a major candidate

for particle energization during substorms, studies of particle energization beyond the reconnection

region focus on dipolarizations and substorm injections.

Dipolarization refers to a reconfiguration of the magnetosphere from a stretched tail-like

to a dipole-like configuration. The leading edges of dipolarizations, known as dipolarization fronts

(DFs), are characterized by rapid increases in the north-south component of the magnetic field

()Bz and are frequently observed in the magnetotail during substorms [Russell and McPherron,

1973; Sergeev et al., 1996; Ohtani et al., 2004; Runov et al., 2009; Schmid et al., 2011; Fu et al.,

13

2012a]. In addition to the characteristic increase of Bz , common features of a DF passing a

spacecraft include a dip of Bz ahead of the increase of Bz , a drop in the plasma density, a decrease

in plasma beta, an increase in plasma temperature, a decrease in plasma pressure and an increase

in magnetic pressure, and a decrease in plasma entropy [Hwang et al., 2011]. The hot and tenuous plasmas in the flux tubes following DFs are plasma bubbles characterized by low plasma entropy

[Wolf et al., 2009 and references therein]. From an MHD perspective, DFs are tangential discontinuities [e.g. Sergeev et al., 1996; Fu et al., 2012b], which are related to the nonlinear slow

mode [Kivelson and Russell, 1995]. However, DFs have their kinetic structure, resulting from

diamagnetic drift of electrons and ions in the presence of a steep magnetic field gradient. Naturally,

the thickness of DFs is comparable with the ion inertial length or the ion gyro radius, whichever

is larger [e.g. Runov et al., 2009; Sergeev et al., 2009]. From a multi-fluid perspective, physics related to the Hall term in the generalized Ohm’s law governs the structure of DFs [Biskamp, 2005].

The Hall electric current near DFs has been confirmed by observations [Zhou et al., 2009; Zhang et al., 2011; Fu et al., 2012b].

Dipolarizations are often accompanied by high-speed earthward (and tailward) flows

[Baumjohann et al., 1990; Angelopoulos et al., 1992]. These flows (~ 1 min) known as bursty bulk

flows (BBFs) are often embedded in intervals of enhanced plasma flows (~ 10 min) and are

believed to be the primary mechanism for the earthward transport of mass, energy and magnetic

flux in the magnetotail [Angelopoulos et al., 1994]. Previous studies have suggested that the width

of these flow channels is 23 RE [Angelopoulos et al., 1996; Sergeev et al., 1996; Nakamura et al., 2004]. Dipolarizations and flows are associated with large transverse electric fields [e.g.

Sergeev et al., 2009; Fu et al., 2012b] and strong wave activity whose frequency ranges from the

lower-hybrid frequency to the electron gyro frequency [Zhou et al., 2009].

14

DFs and associated flows observed in the near-Earth plasma sheet at XRGSM~10 E , are often interpreted as magnetic flux pileup associated with flow braking [Hesse and Birn, 1991;

Shiokawa et al., 1997; Baumjohann et al., 1999; Runov et al., 2012] and current disruption [Lui et al., 1988]. Transient DFs in the mid-tail are interpreted as signatures of nightside flux transfer events (NFTE) resulting from sporadic and spatially localized reconnection [Sergeev et al. 1992;

Fu et al., 2013]. DFs have been found in both quasi-local and global MHD simulations [e.g.

Wiltberger et al., 2000; Birn et al., 2004a, 2011; El-Alaoui et al., 2012, 2013], hybrid simulations

[Krauss-Varban and Karimabadi, 2003] and PIC kinetic simulations [Sitnov et al., 2009; Sitnov and Swisdak, 2011]. Dipolarizations in global MHD simulations are intensified as the reconnection jets brake in the inner magnetosphere, whereas the dipolarizations in particle-in-cell (PIC) simulations are more pulse-like and transient, and are intensified near the reconnection region.

Other mechanisms such as the ballooning interchange instability [Hurricane et al., 1996; Pritchett and Coroniti, 2010] may also be important for dipolarization formation and for determining their azimuthal extent.

Large increases in high-energy fluxes of electrons associated with dipolarizations and high- speed flows have been observed by the Cluster spacecraft and Time History of Events and

Macroscale Interactions during Substorms (THEMIS) spacecraft [e.g. Runov et al., 2009; Asano et al., 2010; Vaivads et al., 2011; Runov et al., 2011]. For most of these dipolarization events, as fronts pass by an observing spacecraft, the electron energy fluxes shift to higher energy. This is observed as a simultaneous increase of the high-energy electron fluxes (tens of keV to hundreds of keV) and decrease of the low-energy electron fluxes (a few keV or less) [e.g. Deng et al., 2010;

Hwang et al., 2011]. For some DF events, the energy fluxes of high-energy electrons increase as the fronts arrive, while in other events the high-energy electron energy fluxes decrease just before

15

the fronts arrive and then increase. For some events the high-energy electron pitch angle

distributions peak near 90o, while others peak at smaller pitch angles (<45o). Both types of

distributions can appear in the inner magnetosphere and the mid-tail plasma sheet, depending on

the particular event [e.g. Deng et al., 2010; Fu et al., 2011]. However, flux increases associated

with DFs are not limited to electrons. Ion fluxes that demonstrate similar increases associated with

DFs were observed by the THEMIS spacecraft for the February 27, 2009 event [Runov et al., 2009;

Deng et al., 2010]. (Note that the THEMIS ion observations do not allow us to differentiate

between ion species.) Observations of high-energy flux increases associated with DFs for both

electrons and ions were also presented in a multi-event study, in which the high-energy ion flux

increases on a time scale of tens of seconds, whereas the electron flux increases within a few

seconds [Runov et al., 2011]. While it remains unclear whether these similarities and differences

between electron fluxes and ion fluxes are universal across events, that fluxes of both species

increase upon the arrival of DFs and high-speed flows indicates that certain energization and

transport mechanisms operate for both species in these events.

As a consequence of particle energization and transport in the magnetotail, injections of

energetic particles are frequently observed in the inner magnetosphere. Early observations of

injections at geosynchronous orbit during substorms led to the idea of an “injection boundary”, an

azimuthal boundary moving earthward fromX ~ (9 12)RE to the geosynchronous orbit under

an enhanced cross-tail electric field [McIlwain, 1974; Mauk and McIlwain, 1974]. The injection

boundary covers the region where fluxes become simultaneously enhanced at different energies,

called dispersionless injection. The dispersionless injection boundary/region was later found to be

azimuthally narrow [Belian et al., 1978; Reeves et al., 1991; Gabrielse et al., 2014]. Energetic

particles observed azimuthally distant from the injection boundary get there by energy-dependent

16

drifts and hence appear dispersed in energy, with electrons drifting eastward [Arnoldy and Chan,

1969; Pfitzer and Winckler, 1969] and ions drifting westward [Bogott and Mozer, 1973]. When

the injected particles become trapped and complete full drift orbits, drift echoes are observed as

injections of multiple times with progressively increased dispersion [Lanzerotti et al., 1967]. Note

that to avoid the effect of energy-dependent drifts, dispersionless injections are caused either by simultaneous acceleration of particles across energies in the injection region or by energy-

independent transport of accelerated particles from other regions. The injection boundary idea was

subsequently further developed. Birn et al. [1997a] demonstrated statistically that a spatially-

dependent pattern exists for dispersionless injections, which can be explained as two injection

boundaries—one for electrons, one for ions—that are offset from each other. Typically, ion

dispersionless injections are inclined towards dusk (~3 hours before midnight) and electron

dispersionless injections towards dawn (~2 hours after midnight). Simultaneous injections of both

species are observed near midnight. This pattern was confirmed by Thomsen et al. [2001], who

used two geosynchronous satellites to show that the pattern occurs not only statistically, but also

for individual injection events. Energetic particle injections and magnetic field dipolarizations are

found to occur concurrently [Sauvaud and Winckler, 1980]. Sharp inner fronts of substorm

injections (injection fronts) have been observed to coincide with DFs in the inner magnetosphere

[Sergeev et al., 1998].

Notably, the observations of substorm injections are mainly in the inner magnetosphere,

within the geosynchronous orbit in particular. In contrast, observations of high-energy fluxes of

electrons associated with dipolarizations and fast flows are in the tail as well as in the inner

magnetosphere. Baker et al. [1979] made an important connection between observations of

energetic particle bursts in the tail and observations of substorm injections. They conjectured that

17

in events with no drift echoes under extremely strong solar wind driving, Ey in the plasma sheet

is strong; the impulsive bursts of energetic protons ( EMeV 0.3 ) observed in the plasma sheet at

XR~18 E are caused by magnetic reconnection; these energetic protons are convected to the

inner magnetosphere rapidly without significant gradient drift, so they are observed as

dispersionless injections at the geosynchronous orbit. In contrast, in drift echo events under

relatively weaker solar wind driving, the transverse electric field Ey is relatively weaker, the

earthward transport is slower, and protons may often gradient drift substantially during their transport, resulting in dispersed substorm injections for lower energy protons ( E 200 keV ).

Various adiabatic and nonadiabatic acceleration mechanisms have been proposed to

account for energetic particles in the magnetotail. The adiabatic acceleration mechanisms include

betatron acceleration as particles are transported from a weak magnetic field region in the tail to a

stronger field region in the inner magnetosphere [e.g. Tverskoy, 1969; Kivelson et al., 1973;

Ashour-Abdalla et al., 2011; Pan et al., 2012], and Fermi acceleration resulting from contraction

of field line length between mirror points during convection [e.g. Tverskoy, 1969; Sharber and

Heikkila, 1972; Pan et al., 2012; Ashour-Abdalla et al., 2013]. Particles, especially ions can gain

or lose energy via nonadiabatic motion in the tail current sheet. The tail current sheet is

characterized by a magnetic field reversal across the current sheet, a weak magnetic field normal

to the current sheet and a dawn-dusk electric field. Particles in the current sheet undertake Speiser

motion consisting of bounce motion within the current sheet and gyration about the normal

magnetic field, until they are eventually ejected out of the current sheet along the magnetic field

line. During the trapped motion in the current sheet, ions (electrons) drift in the Y-direction

(negative Y-direction), so they gain energy from the dawn-dusk electric field before they are

ejected out [Speiser, 1965, 1967; Lyons and Speiser, 1982; Lyons, 1984]. In the magnetotail, ions

18

can also gain or lose energy when they encounter the current sheet multiple times, through resonant

orbits [Burkhart and Chen, 1991; Büchner, 1991; Ashour-Abdalla et al., 1993, 1995; Zelenyi et

al., 2007] and quasi-adiabatic orbits [Zelenyi et al., 1990b]. These particle energization

mechanisms were reviewed by Birn et al. [2012]. Regarding energization related to dipolarizations, wave-particle scattering has been suggested to be important for electron acceleration and heating

because lower-hybrid waves, electron cyclotron harmonics and whistler waves have been observed

in association with dipolarizations, [e.g. Deng et al., 2010; Khotyaintsev et al., 2011; Huang et al.,

2012b], although it is not clearly demonstrated that wave-particle interaction can result in electron

energy gain. Ions can be trapped or quasi-trapped by the electric potentials caused by moving

dipolarizations [Artemyev et al., 2012; Ukhorskiy et al., 2013]. The quasi-trapped ions can gain

energy by encountering and reflecting from DFs multiple times. This process is similar to ion

acceleration by perpendicular shocks.

Particle tracing calculations have been used to investigate the particle energization in the magnetotail and the resultant substorm injections in the inner magnetosphere. Using 3D fields

resembling dipolarization events, including a transient electric field surge, and a magnetic field

whose time scale of variation is comparable with ion gyro period, Delcourt and Sauvaud [1994]

calculated nonadiabatic ion trajectories and showed that ions are accelerated to high energies by

the rapidly changing magnetic field and the transient electric field. Li et al. [1998] and Zaharia et

al. [2000] calculated particle guiding-center orbits in the equatorial plane, using assumed localized

earthward propagating electromagnetic pulses in a fixed background magnetic field. They found

that particles transported from the tail to the inner magnetosphere undertake betatron acceleration,

resulting in dispersionless injections at the geosynchronous orbit. Zaharia et al. [2004] extended

their earlier model by including the background field evolving from a nondipolar (stretched) to a

19

dipole shape as the pulse propagates to the Earth. The resultant energization was increased and

more realistic due to the “dipolarization” of the background magnetic field. Birn et al. [1997b]

examined ion (proton) acceleration by calculating ion orbits in the dynamic electric and magnetic

fields of a 3D MHD simulation. The MHD fields included the generic characteristics of neutral

line formation and dipolarization. The energetic proton flux changes obtained from the test proton

orbits agreed well with observations that demonstrate rapid ion flux increases at energies above

about 20 keV during substorm injections. The protons are accelerated by the transverse electric

field associated with the dipolarization. The acceleration mechanism is equivalent to the betatron

effect. In a complementary study, Birn et al. [1998] examined electron acceleration using the

electron guiding-center approximation in the equatorial plane and the same 3D MHD fields. Their

simulations were able to reproduce observed energy-dependent characteristics of substorm

electron injections, and showed that the dipolarization region earthward of the reconnection site is

more significant than reconnection in accelerating electrons. They later extended the electron

orbits to include nonequatorial drifts [Birn et al., 2004b]. A more recent study by Birn et al. [2013]

confirmed that the major acceleration mechanisms responsible for the energetic electrons injected

into the inner magnetosphere are betatron acceleration and Fermi acceleration associated with the

dipolarization. Ions are accelerated in a similar fashion, despite that ion orbits are nonadiabatic.

Since 1990s, the UCLA space plasma simulation group has been developing the particle

tracing technique for ions [Ashour-Abdalla et al., 1990] and electrons [Schriver et al., 1998] to

study three types of problems. The first type concerns the formation of magnetospheric structures

such as the central plasma sheet, the plasma bulk flows, the plasma sheet boundary layer (PSBL),

and the “beamlets” within the PSBL [Ashour-Abdalla et al., 1992a, 1993, 1995, 1996, 1999a,

1999b, 2000, 2005; Peroomian and Ashour-Abdalla, 1995, 1996]. The second type of problem

20

addressed is the origin of the ion-to-electron temperature ratio [Schriver et al., 1998]. Below we

review the literature on the third type problem, which is energization of particles to tens and

hundreds of keV during substorms.

Regarding ion energization, Ashour-Abdalla et al. [1992b, 1992c, 1994] calculated ion

trajectories in a 2D Tsyganenko [1989] empirical magnetic field model and a uniform dawn-dusk electric field. They showed that the different regions of the plasma sheet could be characterized by

varying values of the adiabaticity parameter,  , which is defined as the square root of the ratio of

the curvature radius of the magnetic field line ( Rcurv ) to the ion gyro radius (  ), namely

 Rcurv [Büchner and Zelenyi, 1989]. Both Rcurv and  are determined locally at the ion

position. In the region near the X-line,   1 and the ions can be described as quasiadiabatic or

nonadiabatic. In the other extreme, close to the Earth or away from the current sheet,   1 and ion motion is adiabatic. Between these two regions, in the wall region,  ~1 and the ions are chaotic [Büchner and Zelenyi, 1989]. In the wall region with a dawn-dusk electric field, ions gain

energy continuously as they are demagnetized and move rapidly along the dawn-dusk direction.

Ion acceleration in the wall region was further demonstrated in a substorm event study in which ion trajectories were calculated in a more sophisticated electromagnetic field derived from a global

MHD simulation driven by realistic upstream solar wind conditions [Ashour-Abdalla et al., 2009].

In that study, Ashour-Abdalla et al. [2009] also showed that the resultant high-energy ions were observed by the THEMIS spacecraft just after substorm onset. The technique of tracing a large number of particle trajectories in a given electromagnetic field and then calculating collective quantities such as the distribution function and energy flux is called large-scale kinetic (LSK) simulations [Ashour-Abdalla et al., 1993, 2005]. Zhou et al. [2011] used the global MHD+LSK method to examine ion injections during a substorm event. They found that the injected energetic

21

ions observed at X ~ 7RE were accelerated in two regions. One region was around the near-

Earth X-line (X ~ 20RE ), where particles were mostly accelerated nonadiabatically by strong

electric fields. The other consisted of several localized regions with  ~1 3 betweenX ~ 7RE

and X ~ 18RE , where particles were accelerated in nonadiabatic motion by the potential electric

field. This latter mechanism is the aforementioned nonadiabatic acceleration in the wall region.

Pan et al. [2014b] applied the MHD+LSK method to examine ion energization associated with

magnetic dipolarizations and found that ions that originated near the reconnection site initially

gained energy nonadiabatically, and then gained energy adiabatically as the ions caught up with

and then rode on the earthward propagating dipolarizations. Pan et al. [2014b] also demonstrated

that high-speed flows in narrow channels controlled the earthward transport of ions in the outer

magnetosphere due to the dominance of the EB drift compared to the gradient and curvature

drifts.

The MHD+LSK simulation scheme has also been applied to study electron energization in

the magnetotail during substorms. Ashour-Abdalla et al. [2011] simulated a substorm event that

occurred on February 15, 2008. They found that only low-energy electrons (e.g. 6–12 keV) were

present near the reconnection site and most high-energy electrons (e.g. 41–95 keV) were generated

in association with dipolarizations and fast flows. The electron fluxes from the LSK simulation,

which used Maxwellian distributions as electron sources, were consistent with the measured fluxes

from THEMIS P4 at X ~ 9.8RE . They suggested that reconnection produced low-energy

electrons and the electromagnetic fields associated with the DF accelerated electrons to high

energy via nonlocal betatron process. However, in the February 15, 2008 event, there were no

observations in the tail that could be used to quantify high-energy electrons near the reconnection

22

site. Furthermore, using Maxwellian sources in the LSK simulation may have underestimated the

flux of high-energy electrons resulting from reconnection. Therefore it was not clear how much of

the acceleration occurred near the reconnection site and how much of it occurred as the DF

propagated toward the Earth in that particular substorm. To continue this effort, Pan et al. [2014a]

studied a substorm event that occurred on March 11, 2008 and demonstrated that adding a high-

energy power law tail with E~10 keV to the LSK source distribution was necessary to reproduce

energetic electrons observed by THEMIS P2 at X ~ 14.6RE , suggesting that acceleration near

the reconnection region was important in generating these suprathermal electrons. Meanwhile, by

comparing THEMIS P2 measurements in the tail and P3/P4 measurements at X ~ 10RE , Pan et

al. [2014a] also showed that nonlocal acceleration during plasma earthward transport was

responsible for the majority of the high-energy electrons observed in the inner magnetosphere. In

light of the convincing evidence for nonlocal electron acceleration in the magnetotail during substorms, Liang et al. [2014] examined the nonlocal acceleration mechanism in two very different substorm events that occurred on February 15, 2008 and August 15, 2001. They found that in the

February 15, 2008 event, the high-speed flows in narrow channels produced by azimuthally localized reconnection swept the electrons and adiabatically accelerated them, leading to pancake-

like electron distributions ( f ()vfv  () ) in the inner magnetosphere. In contrast, in the August

15, 2001 event, an X-line extending across the tail was formed and the flows were slow. The electrons were nonadiabatically accelerated in the weak field region close to the X-line, resulting

in cigar-like electron distributions ( f ()vfv  ( ) ).

In the event studies of particle energization from the reconnection region to the inner

magnetosphere during substorms, electron and ion acceleration were addressed separately. Pan et

al. [2015a] used THEMIS measurements and MHD+LSK simulations to investigate energization

23

and transport mechanisms for both species during a substorm event. The LSK simulation results

showed that thermal ions and electrons (a few keV) observed at the dipolarizations originated from

a relatively wide region of the tail near the reconnection site and were convected to the inner

magnetosphere. Higher-energy particles (tens of keV up to ~100 keV) were produced far away

from the reconnection site by the perpendicular electric fields associated with the dipolarizations

and accompanying high-speed flows. Electrons that originated from the reconnection site were

adiabatically accelerated during earthward transport, and surprisingly ions undertook adiabatic acceleration in a manner similar to that of electrons. However, electron motion was adiabatic in

the magnetotail except in very limited regions close to the X-line, while ion motion was marginally

adiabatic in the dipolarization regions. It was demonstrated that high-speed flows in narrow

channels controlled the earthward particle transport of both electrons and ions in the magnetotail

due to the dominant EB drift in this event.

We point out that the aforementioned particle tracing calculations differ in the electromagnetic fields in which particle trajectories are calculated. The fields are either from

relatively simple analytical description [Delcourt and Sauvaud, 1994; Li et al., 1998; Zaharia et al., 2000, 2004], from empirical models [Ashour-Abdalla et al., 1992b, 1992c, 1994], from generic

MHD fields [Birn et al., 1997b, 1998, 2004, 2013], or from event-dependent global MHD fields

[Ashour-Abdalla et al., 2009; Zhou et al., 2011; Ashour-Abdalla et al., 2011; Pan et al., 2014a,

2014b; Liang et al., 2014; Pan et al., 2015a].

1.4. Purpose of this Study

24

The question addressed in this dissertation is where and how charged particles are

accelerated to tens and hundreds of keV in the magnetotail during substorms. The purpose of our

present study is to integrate simultaneous THEMIS spacecraft measurements in the tail and in the

inner magnetosphere, theoretical models, global MHD+LSK simulations, and implicit PIC

simulations, to quantify both electron and ion energization by various processes and incorporate

the energization processes into a global and complete picture. The computer simulations and

spacecraft measurements are complementary to each other for understanding particle energization

and transport on multiple scales. Specifically, the coordinated observations by identical THEMIS

spacecraft in the magnetotail and in the inner magnetosphere allow us to quantify energization by

processes near the X-line and energization during earthward plasma transport. They provide

information for the LSK particle sources, and constraint MHD+LSK and PIC simulations. The

MHD+LSK simulations provide a global picture of the energization and transport processes that

are responsible for the high-energy particle fluxes observed at a few points by THEMIS spacecraft.

The implicit PIC simulations resolve kinetic effects and allow us to model processes near the X-

line in a relatively large simulation domain with mass ratio and physical parameters that are close

to the realistic values [Brackbill and Forslund, 1982; Vu and Brackbill, 1992; Lapenta et al., 2006;

Markidis et al., 2010 and references therein; Lapenta, 2012]. The large simulation domain allows

us to quantify nonlocal acceleration in the reconnection outflows and clarify the relationship

between local acceleration in the diffusion region and nonlocal acceleration during transport. The

PIC simulations provide explanations to the high-energy fluxes that are observed by THEMIS

spacecraft in the tail. The high-energy fluxes in the magnetotail are generated by processes near

the reconnection, which cannot be properly handled by the MHD+LSK simulations. Throughout

our study, we provide comparisons of simulation results with observations if possible. “In the end,

25

the proof of the value of computer modeling will come from the comparison of its results with

experiments or observations.” (John Dawson, 1985)

Particle energization and transport in the magnetotail is an interesting and challenging

physics question because it involves processes on various scales from a few kilometers (the EDR

scale) to tens of RE (the transport scale). Moreover, developing a quantitative and global picture

of electron and ion energization is also practically desirable because tens to hundreds of keV

electrons injected from the tail are thought to be the source of electrons further accelerated to even

higher energies in the radiation belts [e.g. Green and Kivelson, 2004 and references therein; Horne

et al., 2005]. To model very energetic electrons (hundreds of keV to MeV) trapped in the radiation

belts correctly, we must understand and be able to model the particle sources injected from the

magnetotail. In addition, the tens to hundreds of keV injected ions are one major component of the sources of the ring current [e.g. Thomsen et al., 1998; Nose et al., 2001]. The ring current is a major element that determines the inner magnetospheric dynamics during magnetic storms. The sources of the very energetic particles and development of the ring current are critical issues of space weather because the very energetic particles and the ring current can influence the performance and reliability of space-borne and ground-based technological systems, e.g. causing disruption to satellite operations, telecommunications, navigation, and electric power distribution grids, leading to a variety of socioeconomic losses [Bothmer and Daglis, 2007 and reference therein].

1.5. Structure of the Dissertation

26

Following this introduction, in Chapter 2 we first briefly review the characteristics of

adiabatic particle orbits. The “classical” derivation of equation of guiding-center motion was presented by Northrop [1963] and Banõs [1967]. We comment on the implications of the theory with regard to numerical modeling of particle energization in the magnetotail. Following this review, we present a simple model of adiabatic acceleration to estimate flux changes when particles are transported from the magnetotail to the inner magnetosphere. The model and comparisons of its predictions with observations were published by Pan et al. [2012].

The following three chapters present studies of substorm events, focusing on electron

energization (Chapter 3), ion energization (Chapter 4), and a comparison between them (Chapter

5). The central goal is to use the coordinated THEMIS spacecraft measurements and the global

numerical MHD+LSK simulations to quantify charged particle energization on the meso to global

scale and identify the major acceleration mechanisms. These three chapters are reproduced with

modifications from Pan et al. [2014a], Pan et al. [2014b] and Pan et al. [2015a], respectively.

Chapter 6 deals with electron acceleration on the micro to meso scale in the critical near-

Earth reconnection region. As will be pointed out in Chapter 3, high-energy electrons that follow

a power law distribution near the reconnection X-line are indispensable for achieving consistency

between the global MHD+LSK simulation results and the THEMIS measurements. The problem

of how these power law distributed electrons are generated is examined by using the state-of-the-

art implicit PIC code. The simulation results can also be found in the manuscript by Pan et al.

[2015b].

In Chapter 7, we summarize our findings, discuss caveats of the present study and unsolved

problems, and point out possible major progresses in the future.

27

CHAPTER 2

Theory of Adiabatic Acceleration of Charged Particles

2.1. Introduction

In this chapter, we first briefly review orbit characteristics for particles undertaking

adiabatic guiding-center drifts. The “classical” derivation of guiding-center equation of motion

was presented by Northrop [1963] and Banõs [1967] and will not be repeated here. Instead, the

results will be quoted. The guiding-center theory, as part of the gyrokinetic theory, has received

considerable efforts and progresses in the plasma physics research related to fusion. The classical

derivation is based on physical intuition rather than mathematical rigorousness. The readers are

encouraged to read literature in fusion research, e.g. Brizard and Hahm [2007] for a comprehensive

review of gyrokinetic theory, in which the authors described how to overcome mathematical

vagueness in the procedure of averaging the gyro phase and how to preserve symmetries (e.g.

conservation of energy and phase space) of the guiding-center Hamiltonian system in modern

derivation. Note that the classical results are corrected to the lowest order [Northrop, 1963]. We

comment on implications of the theory with regard to numerical modeling of particle energization

in the magnetotail.

Following this review, we present a simple model to estimate flux changes when particles

are adiabatically transported from the magnetotail to the inner magnetosphere. The model and

comparisons of the model predictions with spacecraft measurements were published by Pan et al.

[2012].

28

2.2. Characteristics of Adiabatic Particle Orbits

“The understanding of the individual orbits is not the same as complete understanding of

plasma physics. The science, or the art, of treating cooperative phenomena must be added to the

thorough discussion of the motion of the individual particles. But it is a basic step in the development of plasma physics to lay the foundation by discussing individual particle motions.

The result is, as always when meaningful progress is made in physics, that a seemingly involved

subject is beginning to show signs of regularity and therefore simplicity. ” (Edward Teller, 1963)

The origins of the adiabatic theory of charged particle motion can be traced back to Alfvén

in the 1940s [Alfvén, 1950]. The fundamental assumption of adiabatic theory is that a separation

of scales of particle motion exists, e.g. between gyration and guiding-center drift with respect to

ambient fields. The derivation of the guiding-center equation of motion relies on a small parameter

for the Taylor expansion of the Lorentz-force equation about the guiding-center position. The

small parameter used by Northrop [1963] and Banõs [1967] is    L , where   mv0 c qB0 is

1 the particle gyro radius with typical velocity v0 in the magnetic field B0 , and L  ln B is the

characteristic spatial scale of the magnetic field variation. Because   mv00 c qB L m q , mq is a proxy for the small parameter in the following discussion where physical quantities are not normalized by their characteristic ones [Northrop, 1963]. Expanding the Lorentz-force equation of motion about the guiding-center position and averaging out the gyro phase, we have the equation

for the guiding-center drift velocity uGC for non-relativistic particles

29

dR u  GC dt Mc cBE  (2.1) q e1 2 e1u  () B mc ee112   uE uu uueEE1 u u uu EE qt s  t  s where Rrρ is the guiding-center position, eB1  B is a unit vector in the direction of the

dR magnetic field, u e is the guiding-center parallel velocity, M is the magnetic moment (the dt 1

cEe  first adiabatic invariant), u  1 is the EB drift velocity, and  e is the projection E B s 1 of the gradient operator in the magnetic field direction. The magnetic moment is defined as

2 q mw2 M c , where  and w are particle gyro frequency and gyro velocity, respectively. 22cB c

The right-hand-side quantities of equation (2.1) are evaluated at the guiding-center position. The drift velocity in the perpendicular direction includes the EB drift term on the order of (1) or

() , depending on the order of perpendicular electric field, the gradient drift term on the order of () , and the “acceleration” drift on the order of () (terms in the square bracket). The

mu2 c ee contribution by u 11 (the first term in the square bracket) is termed curvature drift c qB s

[e.g. Chen, 2006].

The perpendicular drift velocity depends on electromagnetic fields and guiding-center parallel velocity. To follow particle guiding-center motion in given fields, we need update the parallel velocity, which is given by

mdu M B m 2 Eu uueEE   1 () (2.2) qdt q s q t s

30

mu2 mu2 The equation for the change of particle kinetic energy WMBE is given by 22

1(,)dW dRR M B t ER(,)t   ( 2 ) (2.3) qdt dt q t The change of the first adiabatic invariant to the lowest order is

1 dM  () 2 (2.4) qdt

Note that the right hand side of equation (2.4) is to the order of () 2 rather than () . It is because M q on the left hand side contains a small parameter mq~()  . The same situation occurs in equation (2.3).

In the derivation of these equations, it is required that E ~()  and that

tu~~  suE  acting on e1 and uE are at most (1) . It should be understood that the orderings of these quantities are not necessarily expressed in term of the same small parameter, instead there should be one small parameter for each of them. However, if we adopt the parameter for the Taylor expansion to be the maximum of these small parameters, i.e. take the strongest constraint, the theory is still valid. Let’s examine some of the constraints relevant to particle motion in the magnetotail.

1 B First, it is obvious that  B,T  1 is required, since an appreciable change in the c Bt magnetic field on the time scale of gyration shall invalidate the guiding-center picture. This constraint is important for ions in intensifying dipolarizations because ion gyro period (~seconds) is comparable to dipolarization intensification time scale.

31

Second, the time scale of particle parallel traveling a radius of curvature of the magnetic field line should be much larger than the gyration time scale, namely

1 uuuue1 1 e1  221 , where Rc  is the radius of curvature of the ccccsR w s

magnetic field line, and  Rc is an adiabaticity parameter [Büchner and Zelenyi, 1989]. For a 90 pitch angle particle, u  0 , therefore a small kappa puts no constraint. For a small pitch angle particle on the equatorial plane, its parallel velocity is large, its kappa value is small, hence its parallel motion imposes a strong constraint on the adiabaticity of particle motion.

Third, the time scale of perpendicular EB drift across a characteristic length should be

uE  ln B cEb larger than gyration time scale, namely  ~12 B  . In the presence of ccB high-speed flows, the perpendicular convection electric field is large, so the perpendicular drift imposes a strong constraint.

The adiabatic theory has important implications on test particle simulations (the method is used in the following three chapters). First, the theory sets forth a series of criteria for particle

1 B u cEb motion to be adiabatic. They are  B,T  1,  ~12  ,  ~12  B  and c Bt w B c

1/2 1/4   5 (empirical). Since  B,T  m ,   m ,    m ,   m for given parallel and perpendicular energies, ion motion is more likely to become nonadiabatic than electron’s. This requires that we solve Lorentz-force equation for ions in the magnetotail, but we may solve the guiding-center equation for electrons. Second, if the adiabaticity criteria are satisfied, the guiding- center drift velocity may be estimated as

32

cMcmucEe e2 e e uuuu~~11 B 1  1. The EB drift does not depend on GC E B c B Bq qBs energy or charge, whereas the gradient and curvature drifts depend on both of them. For electrons and ions with the same parallel and perpendicular energies, their guiding-center velocities are the same, but they gradient and curvature drift in opposite directions. In the magnetotail, localized high-speed flows carry strong dawn-dusk convection electric fields. If the EB drift is dominant, electrons and ions are transported along the flow channels in the tail. On the other hand, if gradient and curvature drifts dominate, the major feature of transport is that electrons drift eastward and ions drift westward in the magnetosphere. The relative magnitude of the EB drift compared to the gradient and curvature drifts varies with particle kinetic energy. We will show the relative magnitude of the drifts is of great importance in determining how particle gain energy in the

1 dW M B magnetotail. Third, consider the adiabatic energy gain equation uE . The first qdtGC q t term on the right-hand-side (RHS) is charge-dependent, so in order for electrons and ions to gain energy via the first term, they need to drift in opposite directions, namely ions need to drift parallel to the dawn-dusk electric field in the magnetotail, whereas electrons need to do the opposite. This is made possible by the charge-dependent gradient and curvature drifts. The second term is charge- independent (the charges on both sides are cancelled), so an intensification of magnetic field (e.g. dipolarization) will accelerate both electrons and ions. These applications to the Earth’s magnetosphere are discussed in the following three chapters.

2.3. An Analytical Model of Adiabatic Acceleration

2.3.1. Motivation

33

Large increases in the energy fluxes of high-energy electrons (tens of keV to hundreds of keV) are frequently reported in association with DFs. It is unclear from observations how much of the acceleration occurs as the DFs propagates toward the Earth and how much occurs near the reconnection site. Recent studies using data from the four Cluster satellites indicate that electrons can be accelerated by multi-step processes near the site of magnetic reconnection as well as far away from the diffusion region in the same events [Asano et al., 2010; Vaivads et al., 2011].

In this section, we explore the collective effect on electron fluxes when electrons are adiabatically transported from the magnetotail to the inner magnetosphere. Based upon the analysis of adiabatic acceleration, we attempt to qualitatively clarify the roles of magnetic reconnection and adiabatic acceleration in producing suprathermal electrons associated with DFs in the inner

magnetosphere (XGSM~-10 RE ). In section 2.2.2 we discuss the adiabatic acceleration model. In section 2.2.3 we present comparisons of the model predictions with data from the THEMIS spacecraft during two geo-active intervals (March 11, 2008 and February 27, 2009). In section

2.2.4 we discuss these observations in terms of what we would expect for acceleration as the DFs propagate earthward and use those inferences to consider what may be happening nearer the reconnection site.

2.3.2. The Adiabatic Acceleration Model

In the Earth’s magnetosphere, the time scales of changes in the magnetic and electric fields are much larger than time periods of the gyration of a particle about a magnetic field line, the bounce motion between mirror points, or the azimuthal drift of a particle about the Earth. There are three adiabatic invariants corresponding to these three types of motion, respectively [Northrop,

34

1963; Roederer, 1970; Schulz and Lanzerotti, 1974; Green and Kivelson, 2004]. The first invariant

(magnetic moment) is given by

P 2 M   (2.5) 2mB where P is the relativistic momentum perpendicular to the magnetic field. The increase in P due to a slowing varying B is termed betatron acceleration. The second adiabatic invariant involves the parallel momentum along the bounce motion between mirror points

JPds (2.6)   where P and ds are the relativistic momentum parallel to the magnetic field and the distance

along a field line, respectively. An increase in P due to a decrease of the distance between mirror points is usually referred to as Fermi acceleration.

The azimuthally integrated differential flux jE(,,,) r t and distribution function

f (,,,)Et r are functions of kinetic energy E , pitch angle  , position r and time t . The relation of differential flux and distribution function is [Schulz and Lanzerotti, 1974; Lyons and Williams,

1986]

j(,,,)Et r fE(,,,) r t (2.7) P2 where P is the relativistic momentum.

From Liouville’s theorem, which states that the particle distribution function in phase space is constant along the particle characteristic trajectory without collisions or diffusion caused by wave-particle interaction, namely

jE11111(,,,) rr t jE 2 (, 2 222 ,,) t 22constant (2.8) PP12

35 where the subscripts 1 and 2 indicate that the quantities are evaluated at different positions and times.

We assume that particles undergo slow magnetic field compression given by

B 2   (2.9) B1 where  is the magnetic field compression factor. Assume the first adiabatic invariant is conserved

2 P,2 2   (2.10) P,1 and PP  sin , we have

22 P22sin  22  (2.11) P11sin  Assume the second adiabatic invariant is conserved, the effect of Fermi acceleration on particles due to contraction of mirror distance is

P ,2   (2.12) P,1 where PP  cos .  is the ratio of the distances between mirror points for a particular particle on different mirror bounces, which we call the contraction factor.  should vary for particles with different pitch angles, energies and positions; however, in this model it is assumed to be uniform in order to obtain analytical estimates (see the discussion section below). Hence,

P cos 22  (2.13) P11cos With betatron and Fermi acceleration operating simultaneously, a particle trajectory in phase space is given by equations (2.11) and (2.13),

2222 2 PP21(cos  1 sin)  1 2 (2.14) 2 sin 1 sin 2  22 2  cos11 sin 

36

Combining equations (2.8) and (2.14), we have

22 2 jE2(,,,)( 2 222rr t cos 1 sin)(,,,) 11 jE 1 111 t (2.15) where the relationship of the pitch angles and energies indicated by the subscripts follow from

22 22 equation (2.14) after noting that PEmcEc(20 )/, where m0 is the particle rest mass.

To quantify the spectra of particles, we assume that the differential flux of energetic particles varies as a power law in energy [Øieroset et al., 2002; Imada et al., 2007]; that is

j(,,,)EtCWtErr ( ,,,)n (2.16) where W is kinetic energy range with EW . The exponent n is the power law index, which quantifies the slope of differential flux as a function of energy. If we insert the power law into

22 22 equation (2.15) and use non-relativistic approximation PEmcEcmE(200 )/2  , we obtain:

jE20222(,,,) r t

22 2 E0 ( cos111 sin )jt (22 2 ,  111 ,r , ) cos11 sin n 22 2 E0 ( cos110111 sin )CW ( ,  ,r , t )22 2 (2.17) cos11 sin 22 2nn 1 ( cos1101110 sin )CW ( ,  ,r , t ) E 22 2n 1 (cos1110111 sin)jE (,,, r t) Thus,

jE20222(,,,) r t 22 2n 1 (cos11 sin) (2.18) jE10111(,,,) r t

The differential flux at energy E0 with EW00 is larger than its source flux at E0 by a factor of

2 2 2 n1 ( cos 1   sin 1 ) . An important implication of this is that adiabatic enhancement of the differential flux depends strongly on the power law index of the source flux. For example, if the power law index of suprathermal electrons associated with DFs is in the range -4 to -6 and the

37 compression factor is 1.5 when transported from the outer magnetosphere to the inner magnetosphere, the differential flux can increase by a factor of about 7 to 17. However, if the power law index is, say, between 0 and -1 in low energy range, the differential flux will increase only by a factor of about 1 to 2 under the same compression. Moreover, adiabatic acceleration does not change the power law index because

jE20222(,,,) r t 22 2nn 1  (cos1101110 sin)CW (,,,) r t E (2.19) n  CW(,,,)02220 r t E where

22 2n 1 CW(,,,)(cos0222rr t 1 sin) 1 CW (,,,) 0111 t (2.20)

2.3.3. Comparisons with Observations

We have selected two DF events observed by the THEMIS satellites. The first event was on March 11, 2008. The second event was on February 27, 2009 and has been studied extensively

[Runov et al., 2009; Deng et al., 2010; Ge et al., 2011]. Our approach is to use differential flux data at a spacecraft in the tail (usually, THEMIS P1 or P2) to calculate the resultant adiabatic accelerated flux at a spacecraft closer to Earth (usually P4), and then compare the result with the data at P4. The method assumes that the electrons at the distant spacecraft are the same set of electrons that are observed closer to Earth. With that assumption, we can determine if the electron flux observed nearer to the Earth is consistent with adiabatic acceleration of the flux observed at the more distant satellite. Note that electrons also drift mainly in the Y-direction due to magnetic field curvature and gradient. However, the large convective electric field drives electrons earthward from P1 (or P2) to P4 within 1~2 minutes. The width of fast flow channels in

magnetosphere is typically a few RE [e.g. Sergeev et al., 1996; Nakamura et al., 2004], we estimate

38 electron drift in the Y-direction to be about one RE , which is smaller than the flow channel width, hence electrons tend to stay in the same fast flow channel as they are transported earthward.

To more easily estimate the effects of betatron and Fermi acceleration we selected quasi- perpendicular electrons to evaluate betatron acceleration and quasi-parallel electrons for Fermi

  acceleration. For betatron acceleration, we selected electrons such that 1 ~ 90  2 ~ 90 so that

2 2 P2  P1 and EE21  . Then the equation relating the differential fluxes of quasi-perpendicular electrons at different locations becomes

 jE21(,90,,)rr 2211 t jE (,90,,) 11 t (2.21)

  2 2 2 2 For Fermi acceleration, the criteria were 1 ~ 0  2 ~ 0 so that P2   P1 E21  E , and

22 jE21221111( ,0,rr , t ) jEt ( ,0, , ) (2.22)

A. Event #1 March 11, 2008

THEMIS P2 at (XYZ , , )GSM = (-14.7, 5.4, -1.8) RE and P4 at (XYZ , , )GSM = (-10.4, 5.3,

-1.6) RE probed similar structures and they were on the same flow channel in a MHD simulation

[Pan et al., 2014a]. The first three panels in Figure 2.1 are data from P2 showing in the top panel

128 Hz resolution magnetic field observations from the Flux-Gate (FGM) [Auster et al., 2008] in GSM coordinates. The differential energy fluxes observed by P2 are presented in the second panel. The lowest three energy channels of the fluxes plotted were measured by the

Electrostatic Analyzer (ESA) instrument [McFadden et al., 2008], and other channels in the energy range from 26 keV to 113 keV were obtained by the Solid State Telescope (SST) instrument

[Angelopoulos, 2008]. Median values giving the approximate energy of each of the energy channels are summarized in the legend on the right. The power law index of electron differential flux from SST data (26keV to 113keV) was calculated by using a least squares fit to the differential

39 flux as a function of energy on a log-log scale. Red and green lines plotted along with the power law index are plus or minus one standard deviation. They provide upper and lower limits to the power law index, indicating the error of the fit. Note that in the theoretical section we discussed the changes in power law distributed differential fluxes. The differential fluxes are different from the differential energy fluxes in the second panel. The DF of interest observed by P2 was

characterized by a sudden intensification of Bz at 06:22:58 UT after modest bumps and dips. The front was followed by large transient fluctuations. Starting from 06:20:00 UT, concomitant with a bump/dip in the magnetic field, the energy fluxes gradually decreased to their minima just before the front arrived. At the end of this decrease, the energy fluxes were about half of those at 06:20:00

UT. This decrease was followed by a significant but gradual increase (by a factor of 5~6 in about

2 minutes) associated with the fronts. The power law index decreased and reached a minimum of

-3.5 as the front passed by. This change in power law index means that flux changes at different energies were different. After the transient fluctuations in the magnetic field, the power law index returned to -2.5 after 06:26 although the energy fluxes stayed at higher levels.

Figure 2.2 shows data from P4; the format is the same as Figure 2.1. The DF arrived at P4 at 06:23:53. The energy fluxes increased by almost an order of magnitude and did so more abruptly than at P2 as the front passed by. After the DF passed (~06:28) the energy fluxes at P4 returned to the level before the front’s arrival. The power law index was roughly anti-correlated with the energy fluxes. It decreased to -4.5 at the front and increased to -2.5 as the energy fluxes decreased to the pre-front level. The large magnitude of the power law index at the front indicates the differential flux is steeper than that of the preexisting plasma sheet electrons. The flux of suprathermal electrons at P4 is larger than that at P2 by a factor of 6 to 7 in the fronts. The power law index during the fluctuation period at P4 was smaller than that at P2 by about 0.8.

40

P2 20 Bx 10

0 By [nT] B (gsm) -10 Bz -20 15keV(ESA) 107 20keV 26keV 106 31keV(SST) 41keV 105 52keV 65keV [eV/s/cm^2/sr/eV] 104 93keV Differential energy flux -1.5 index -2.0

-2.5 lower limit -3.0

Powerlaw index -3.5 upper limit -4.0 X(Re) -14.8 -14.7 -14.7 Y(Re) 5.4 5.3 5.3 Z(Re) -1.7 -1.8 -1.8 hhmm 0620 0625 0630 2008 Mar 11

Figure 2.1. Energy flux and power law index at THEMIS P2 in the March 11, 2008 event. From top to bottom showed are magnetic field, electron differential energy flux and suprathermal electron power law index. The first three channels of energy flux are measured by the Electrostatic

Analyzer (ESA), and the other channels by the Solid State Telescope (SST).

41

P4 30 Bx 20

10 By [nT] 0 B (gsm) -10 Bz -20 15keV(ESA) 107 20keV 26keV 106 31keV(SST) 41keV 105 52keV 65keV [eV/s/cm^2/sr/eV] 104 93keV Differential energy flux -1.5 -2.0 index -2.5 -3.0 lower limit -3.5

Powerlaw index -4.0 upper limit -4.5 X(Re) -10.4 -10.4 -10.4 Y(Re) 5.4 5.3 5.3 Z(Re) -1.6 -1.6 -1.6 hhmm 0620 0625 0630 2008 Mar 11

Figure 2.2. Energy flux and power law index at THEMIS P4 in the March 11, 2008 event. Same format as in Figure 2.1.

Figure 2.3 (top panel) shows the differential fluxes of electrons. The blue, green lines show the fluxes observed by P2 and P4, respectively. The red line is the predicted flux at P4 assuming betatron acceleration using the flux at P2 as the source. The data points of electron fluxes shown were picked from the positions when the magnetic field reached its peak during the dipolarization.

Those times are 06:22:59 UT at P2 and 06:23:55 UT at P4. Quasi-perpendicular electrons were selected for the betatron acceleration calculation. The compression ratio of the total magnetic field was  1.3 due to either a global change of magnetic field or a local compression inside a growing

42 pileup region [Fu et al., 2011]. Pitch angle ranges of P2, 1 and P4, 2 were selected to be

 1270  110 . The calculated flux of electrons in the energy of 3 keV-200 keV at P4 agrees with the data at P4. The flux at P4 significantly increases (by a factor of 6 to 7) in the high energy range although the compression of magnetic field from P2 to P4 is just by a factor of 1.3. Because the magnitude of the power law index of the electron flux is relatively large, a slight shift in the x- axis on the log-log plot of flux as function of energy generates a large flux increase.

The observed flux of electrons below ~3 keV at P4 is half an order smaller than the calculated one. Several possible factors could lead to this discrepancy. First, whistler-mode waves with electric wave field ~5 mV/m were detected by P4 from 06:23:55.6 UT to 06:23:56.4 UT (not shown); those waves might scatter electrons at large pitch angles to small pitch angles within seconds due to pitch-angle scattering near the resonant energy, which was a few keV in this case

[Khotyaintsev et al., 2011]; this is consistent with the data that, in this time period, electrons below

~3 keV are predominantly in the parallel direction. Second, Bx and By were about -10 nT for P2 and P4 at the DF, which indicates that the satellites were not in the center of plasma sheet, leading to the variation of electron density.

Figure 2.3 (bottom panel) shows the differential fluxes of electrons assuming Fermi acceleration. Again the blue, green lines give observations from P2 and P4, The red line is the predicted flux at P4 assuming Fermi acceleration using the flux at P2 as the source. Quasi-parallel

 electrons were used (12020).   1.5 from P2 to P4 gave the best fit. For suprathermal electrons, the calculated flux fits the observations very well and the discrepancy at low energy is modest.

43

106

P2 observation P4 observation betatron operating 104

102

100 Differential flux (#/s/cm^2/ev/sr) 10-2

10-4 102 103 104 105 Energy (eV)

106

P2 observation P4 observation Fermi operating 104

102

100 Differential flux (#/s/cm^2/ev/sr) 10-2

10-4 102 103 104 105 Energy (eV)

Figure 2.3. Betatron and Fermi acceleration in the March 11, 2008 event. The top (bottom) panel shows differential flux of quasi-perpendicular (quasi-parallel) electrons. The blue and green lines are observations from P2 and P4, respectively, and the red line is the predicted flux at P4 using the flux at P2 as the source assuming betatron acceleration (top) and Fermi acceleration (bottom).

44

B. Event #2 February 27, 2009

We applied our theory to the well-studied February 27, 2009 event. The figures of this event are in the same format as those for the March 11, 2008 event. THEMIS P1, P2, P3 and P4 observed similar DF signatures and were in the same flow channel in a MHD simulation (see

Figure 12-d in Ge et al. [2011]). As seen in Figures 2.4-2.6, significant flux increments of high energy electrons are associated with DFs and are consistent with predicted fluxes from adiabatic acceleration. The electron fluxes shown in Figure 2.6 were picked from 07:51:29 UT at P1 and

07:54:15 UT at P4. The compression ratio of the total magnetic field from P1 to P4 was   1.6 and contraction factor was   2.0 . There are several differences in this event. First, the energy fluxes of suprathermal electrons increased immediately when the fronts arrived at both spacecraft

(Figures 2.4-2.5). Unlike the event on March 11, 2008, there was no decrease in the energy fluxes before the increase. Second, the calculated differential fluxes from betatron and Fermi acceleration are consistent with the observed ones (Figures 2.6); the discrepancy between the observations and theory for quasi-perpendicular electrons at low energy is much smaller than on March 11, 2008.

This might be because P1 and P4 were both in the central plasma sheet. More importantly, the spectrum of suprathermal electrons associated with the DF captured by P1 (Figure 3.4) was very

similar to those observed closer to the Earth by P2 at -15 RE (not shown) and P4 at  10RE (Figure

3.5). P1 was at  20RE , which was close to the location of the reconnection site in the MHD simulation [Ge et al., 2011], suggesting these suprathermal electrons were produced by magnetic reconnection and convected with the reconnection jets to the location of P1. As the reconnection jets propagated earthward from P1 to P4, the high-energy electron fluxes were further enhanced.

45

In Figure 2.6 the predicted betatron and Fermi acceleration fluxes at high energy respectively increased by factors of 5 and 10 from P1 to P4.

P1 30 Bx 20

10 By [nT] B (gsm) 0 Bz -10 15keV(ESA) 6 10 20keV 105 26keV 4 31keV(SST) 10 41keV 103 52keV 102 65keV [eV/s/cm^2/sr/eV] 1 93keV Differential energy flux 10 -3 index -4

-5 lower limit

-6

Powerlaw index upper limit -7 X(Re) -20.1 -20.1 -20.1 -20.0 Y(Re) -0.6 -0.6 -0.6 -0.6 Z(Re) -1.5 -1.5 -1.5 -1.5 hhmm 0750 0752 0754 0756 2009 Feb 27

Figure 2.4. Energy flux and power law index at THEMIS P1 in the February 27, 2009 event.

Same format as in Figure 2.1.

46

P4 40 Bx 20

0 By [nT] B (gsm) -20 Bz -40 15keV(ESA) 6 10 20keV 105 26keV 4 31keV(SST) 10 41keV 103 52keV 102 65keV [eV/s/cm^2/sr/eV] 1 93keV Differential energy flux 10 -3.5 -4.0 index -4.5 -5.0 lower limit -5.5

Powerlaw index -6.0 upper limit -6.5 X(Re) -11.1 -11.1 -11.1 -11.1 Y(Re) -1.7 -1.8 -1.8 -1.8 Z(Re) -2.4 -2.4 -2.4 -2.4 hhmm 0750 0752 0754 0756 2009 Feb 27

Figure 2.5. Energy flux and power law index at THEMIS P4 in the February 27, 2009 event.

Same format as in Figure 2.1.

47

106

P1 observation P4 observation betatron operating 104

102

100 Differential flux (#/s/cm^2/ev/sr) 10-2

10-4 102 103 104 105 Energy (eV)

106

P1 observation P4 observation Fermi operating 104

102

100 Differential flux (#/s/cm^2/ev/sr) 10-2

10-4 102 103 104 105 Energy (eV)

Figure 2.6. Betatron and Fermi acceleration in the February 27, 2009 event. Same format as in

Figure 2.3.

48

2.3.4. Discussions

Particle transport and nonlocal acceleration associated with DFs is complicated in part because of structures embedded within the fronts. There is coupling between processes on multiple spatial and temporal scales [e.g. Sergeev et al., 2009; Deng et al., 2010]. Analysis is further complicated because we have observations at only a limited number of points. Despite this complexity we believe that THEMIS observed suprathermal electrons that originated near the outer satellites (P1 or P2) and subsequently were detected in the inner tail region near P4. This interpretation seems reasonable because of similarities in the observed DFs and because the virtual satellites were in the same flow channel in the MHD simulations. Our analysis of the data, in combination with theory, reveals that adiabatic enhancement of flux strongly depends on the slope of the source differential flux. In particular, for suprathermal electrons large increases in flux can be induced by modest magnetic field compressions and modest contractions of distance between mirror points.

What cause magnetic compressions during transport? Consider the magnetic field evolution

dcB  BVBV()()     (  B ) (2.23) dt 4 This equation describes the change in the magnetic field in the frame moving with the MHD flow.

The RHS includes terms of plasma compression, flow gradients along a field line, and diffusion.

Intuitively, compression (V 0 ) increases the magnetic field whereas diffusion decreases it.

By neglecting the diffusion term because it is insignificant in the region of flux pileup [e.g. Pan et al., 2014b], the frozen-in condition is satisfied. We then have VEB  and

ddtBBVBV()()    . As the plasma propagates toward the inner magnetosphere, it is slowed and compressed, i.e. V  0 , therefore B  0 and  1.

49

Adiabatic acceleration theory predicts invariance of the power law index during adiabatic transport. In the observations, the power law indices were roughly anti-correlated with the energy fluxes. The power law indices at the DFs changed within 0.8 from P2 (or P1) to P4. Thus, the suprathermal electrons associated with dipolarizations characterized by these power law indices can be traced back to the source regions of DFs, which are most likely near the reconnection region.

We suggest that a combination of local processes near the magnetic reconnection site and nonlocal adiabatic acceleration during earthward transport is responsible for the high flux of suprathermal electrons associated with DFs in the inner magnetosphere.

We simplified the effect of Fermi acceleration by setting the contraction factor  to be uniform for electrons with different energies and pitch angles, which is determined by contraction of mirror point distance during earthward propagation. However, a more detailed analysis would need to take into account of the non-uniformity of the parameters that describe Fermi acceleration because electron trajectories in non-dipole and dynamic magnetic fields are expected to be complex, which makes it difficult to assign a simple contraction factor to account for the effect of

Fermi acceleration in the variation of the electron flux. Moreover, immediately downstream of reconnection outflow, electrons are expected to behave nonadiabatically because of the small radius of curvature of magnetic field lines and large magnetic gradient [e.g. Lyons, 1984; Büchner and Zelenyi, 1989; Schriver et at., 1998; Hoshino et al., 2001; Imada et al., 2007]. The process in this region cannot be addressed by the analytical model because it is not clear exactly how far P1

(or P2) is from the diffusion region and how to quantify the nonadiabatic effects. In the next chapter, we will refine our calculation by using a large-scale kinetic (LSK) simulation, in which we follow a large number of electron trajectories in the time-dependent electric and magnetic fields derived from a global MHD simulation [Ashour-Abdalla et al., 2005]. That study will enable us to include

50 the effects due to nonadiabatic motion in the outflow region immediately downstream from the reconnection site and demonstrate nonlocal acceleration for an ensemble of particles. A statistical study using THEMIS data should provide clues that reveal the relationship between the structures in DFs and reconnection, and thereby help us understand production of high-energy particles in these processes. Our present analysis suggests, however, that such detailed simulation and statistical studies will not fundamentally change the physics described here that is both local processes near reconnection site and nonlocal adiabatic enhancement in the downstream region contribute to the observed characteristics of suprathermal electrons in the inner magnetosphere.

51

CHAPTER 3

Modeling Electron Energization and Transport in the Magnetotail

during a Substorm

3.1. Introduction

Having gained insights about nonlocal acceleration from the simple theoretical model

presented in Chapter 2, we now apply the more powerful global MHD+LSK calculations to the

March 11, 2008 event. The goals are to determine the major electron energization processes in the

magnetotail and incorporate their contributions into a global scenario under realistic

magnetospheric conditions, and especially to quantify the local energization near the reconnection

sites and nonlocal energization during earthward transport. In section 3.2, we present the solar

wind observations from the Geotail spacecraft, which was upstream from the bow shock and the

THEMIS observations in the magnetotail. The simulation methodology is discussed in section 3.3.

The MHD+LSK results are presented in section 3.4. Electron energization mechanisms and

transport features are discussed in section 3.5. We summarize in section 3.6.

3.2. Observations of the March 11, 2008 Substorm Event

3.2.1. Geotail Observations of the Solar Wind

Figure 3.1 shows the solar wind conditions on March 11, 2008 observed by the Geotail

spacecraft at (X ,YZ , )GSE (15, 2, -1) R E (upstream of the sub-solar point of the bow shock) in the geocentric solar ecliptic (GSE) coordinate system. The GSE system has its X-axis pointing from

52

the Earth towards the Sun and its Y-axis is chosen to be in the ecliptic plane pointing towards dusk.

Its Z-axis is parallel to the ecliptic pole. The pertinent interval is indicated by the two black vertical

dashed lines (05:40-06:51 UT). The earthward component of the solar wind velocity Vx was about

680km/s (fast solar wind) for the interval starting at least from 02:00 UT to 08:00 UT. The IMF

Z-component Bz was northward and fluctuated around 2nT from 02:00 UT to 05:40 UT and

turned southward at about 05:40 UT (first black vertical dashed line) and then stayed around -2nT

until 06:50 UT (second black dashed line), when it turned northward again. A sizable substorm

onset occurred around 05:51 UT where the AE index substantially increased (blue vertical dashed

line in the last panel), and it continued to increase to a peak of about 700nT at 06:40UT.

Geotail Observation of Solar Wind 6 4 2 0 Bx (nT) -2 -4 -6 6 4 2 0 By (nT) -2 -4 -6 6 4 2 0 Bz (nT) -2 -4 -6 200 Vz 0 -200 Vy Vi -400 (km/s) -600 Vx -800 5 4 3 P

(nP) 2 1 0

AL 500

0 AU (nT) Index Geomagnetic -500 AE

hhmm 0200 0400 0600 0800 2008 Mar 11

53

Figure 3.1. Geotail observations of the solar wind on March 11, 2008. From top to bottom, the

first three panels show X-, Y-, and Z-components of the magnetic field in GSE coordinates, the

fourth panel shows solar wind velocity, and the fifth panel shows solar wind dynamic pressure.

The last panel gives the geomagnetic indices from the World Data Center (WDC) for

Geomagnetism.

3.2.2. THEMIS Observations in the Magnetotail

During the substorm interval, THEMIS P2, P3 and P4 were in the nightside magnetosphere.

THEMIS P1 was in the magnetosheath, and P5 was in the radiation belt region. Because we are

interested in the transient dipolarization structures in the magnetotail, we focus on the data

obtained by P2, P3, and P4. There were at least three dipolarization events, as indicated by the

vertical dashed lines in Figure 3.2. They occurred during the time intervals 05:50-06:10 UT, 06:20-

06:40 UT and 06:40-07:20 UT. The first dipolarization occurred during the beginning of the

expansion phase, while the second and third dipolarizations occurred during the expansion phase.

It is interesting to notice that there were strong magnetic field oscillations before the onset of the

expansion phase. In this chapter, we study the second dipolarization event, because the Bx values are small from 06:20 UT to 06:30 UT, indicating that the spacecraft were close to the plasma sheet.

54

-5 10

0 P5 5

P3 5 P2 0 P4 P5 P3

Z_gsm (Re) P4 Y_gsm (Re) P2 10 -5

15 -10 5 0 -5 -10 -15 -20 5 0 -5 -10 -15 -20 X_gsm (Re) X_gsm (Re)

THEMIS Observation of Magnetic Field 40 Bz 20

0 By P2

B_GSM (nT) -20 Bx -40 40 Bz 20

0 By P3

B_GSM (nT) -20 Bx -40 40 Bz 20

0 By P4

B_GSM (nT) -20 Bx -40 hhmm 0500 0600 0700 0800 2008 Mar 11

Figure 3.2. THEMIS locations (top) and the observed magnetic fields (bottom) in the March 11,

2008 event.

55

During the interval from 06:20 UT to 06:30 UT, THEMIS spacecraft P2 was at

(X ,YZ , )GSM  (-14.7, 5.4, -1.8) RE , P3 was at (X ,YZ , )GSM  (-10.7, 4.5, -1.6) RE , and P4 was

at (X ,YZ , )GSM  (-10.7, 4.5, -1.6) RE . Figure 3.3 shows THEMIS P2 observations from 06:20

UT to 06:30 UT. From top to bottom showed are 128 Hz resolution magnetic field observations

from the Fluxgate Magnetometer (FGM) [Auster et al., 2008] in GSM coordinates, ion bulk

velocity calculated from measurements by the Electrostatic Analyzer (ESA) instrument

[McFadden et al., 2008] and Solid State Telescope (SST) instrument [Angelopoulos, 2008], the

differential energy flux of high-energy electrons (>26keV) from SST, the differential energy flux

of low-energy electrons (<26keV) from ESA, the electric wave fields from the Electric Field

Instrument (EFI) [Le Contel et al., 2008], and the magnetic wave fields from the Search Coil

Magnetometer (SCM) [Bonnell et al., 2008]. The black and red dashed lines in the last two panels

are the electron gyro frequency fce and the lower hybrid frequency fcef ci , respectively. The DF

of interest observed by P2 was characterized by the sudden intensification of Bz at 06:22:58 UT

(vertical dashed line), following a series of modest bumps and dips. The front was followed by a region of large transient fluctuations. The earthward flow was large and increasing. The ion bulk flow velocity was 300 km/s and gradually increased to 600 km/s. Starting from 06:22 UT, concomitant with a bump and dip in the magnetic field, the high-energy electron flux gradually decreased to its minimum just before the front arrived. This decrease was followed by a gradual

increase as the front passed by P2. The flux reached its maximum value at 06:26 UT when the

dipolarized magnetic field became relatively steady. The thermal electron energy flux behaved similarly. Electromagnetic waves between the lower hybrid frequency and the electron gyro frequency became active as the front passed by.

56

THEMIS P2 Observation 20 bz 10 0 by FGM B (nT) -10 bx -20 600 vz 400 200 vy

Vi (km/s) 0

ESA+SST vx

-200 ) -1 8 10 eV 107 -1 5 6 sr 10 10 -2 5 (eV) SST

10 cm DEflux 104 -1 103

10000 108 (eV s 107 6 1000 10

(ev) 5 ESA 10 DEflux 104 100 103 1.00 1000

100 0.10 EFI E (Hz) 10 (mV/m) 1 0.01 0.100 1000

100 0.010 (nT) SCM B (Hz) 10

1 Sun Apr 13 16:00:51 2014 0.001 X_gsm(Re) -14.8 -14.7 -14.7 Y_gsm(Re) 5.4 5.3 5.3 Z_gsm(Re) -1.7 -1.8 -1.8 hhmm 0620 0625 0630 2008 Mar 11

Figure 3.3. THEMIS P2 observations in the March 11, 2008 event. Shown from top to bottom are the magnetic field observations in GSM coordinates, the ion bulk velocity, the differential energy flux of high-energy electrons (>26keV), the differential energy flux of low-energy electrons

(<26keV), the electric wave fields, and the magnetic wave fields. The black and red dashed lines in the last two panels are respectively the lower hybrid frequency and the electron gyro frequency.

57

THEMIS P4 observations are shown in Figure 3.4 in the same format as Figure 3.3.

Because P3 and P4 were close to each other and their data are similar, we show only P4 data. As

the front arrived at 06:23:53 UT (vertical dashed line), P4 detected the transient DF structure and

subsequent large fluctuations, which are similar to the P2 observations. Compared with the P2 observations, the earthward flow slowed down to 200 km/s. Unlike the flux observed by P2, the

energy flux measured by P4 remained steady before the front arrived, and then suddenly shifted to

lower-energy for about 20 seconds and then shifted back to higher energy. At this time, the high-

energy electron flux was well above the level before the front arrival. The wave activity began as the front arrived and persisted during the subsequent period of magnetic field fluctuations.

Due to the ambiguity between spatial and temporal changes in the measurements, there are two possibilities for the increase in the high-energy electron fluxes: (1) hot and tenuous plasmas

are transported into the regions surrounding the satellites, while the local cold and dense plasmas

are excluded by electromagnetic fields; (2) the cold and dense plasmas are locally energized by

electromagnetic fields into hot and tenuous ones. The validity of either interpretation depends on

particle inertia. Ions move with the magnetic field if their gyro radius i is much smaller than the

spatial scale of the magnetic field gradient B / B ; otherwise they are loosely connected with

magnetic field lines. The scale of the DF structure is on the order of the ion inertial length and

comparable with the ion gyro radius [e.g. Runov et al., 2009; Sergeev et al., 2009]. Hence, it is

likely that ions near the front are not tightly bound to the magnetic field lines. In contrast, electrons

have a smaller inertia than ions, and the scale of the magnetic field gradient is much larger than

their gyro radius e , so they tend to follow the magnetic field lines they initially gyrate about. An

equivalent argument can be made in terms of plasma fluids: the validity of either interpretation

depends on whether the fluids (electron or ion) are frozen-in with the magnetic field. Near the DF,

58

ions with larger gyro radii are demagnetized, while electrons are convected with the magnetic field.

This discrepancy between ion and electron fluid motions results in an appreciable Hall current

given by the generalized Ohm’s law [Biskamp, 2005]. The Hall current near DFs has been

confirmed by observations [Zhou et al., 2009; Zhang et al., 2011; Fu et al., 2012b]. Thus we expect that the hot and tenuous electrons are transported to the satellite locations while interaction of preexisting ions with DFs is important.

THEMIS P4 Observation 30 bz 20 10 by

FGM 0 B (nT) -10 bx -20 600 vz 400 200 vy 0 Vi (km/s)

ESA+SST -200 vx -400 ) -1 8 10 eV 107 -1 5 6 sr 10 10 -2 5 (eV) SST

10 cm DEflux 104 -1 103

10000 108 (eV s 107 6 1000 10

(ev) 5 ESA 10 DEflux 104 100 103 1.00 1000

100 0.10 EFI E (Hz) 10 (mV/m) 1 0.01 0.100 1000

100 0.010 (nT) SCM B (Hz) 10

1 Sun Apr 13 16:19:10 2014 0.001 X_gsm(Re) -10.4 -10.4 -10.4 Y_gsm(Re) 5.4 5.3 5.3 Z_gsm(Re) -1.6 -1.6 -1.6 hhmm 0620 0625 0630 2008 Mar 11

59

Figure 3.4. THEMIS P4 observations in the March 11, 2008 event. The same format as in Figure

3.3 is used.

3.3. Simulation Methodology

In order to develop a quantitative and global scenario of particle energization and transport

in the magnetotail during substorms, we employ simulations to complement the satellite data. The

vastness of the space, the longtime duration of events, and the multi-scale nature of the

magnetospheric plasmas, prevent making a global kinetic simulation. In order to obtain realistic

global electromagnetic fields while capturing kinetic features related to individual particles, we

use a scheme that combines a global MHD model and a LSK simulation. The MHD model [Raeder

et al., 1998, 2001; El-Alaoui et al., 2001, 2009] driven by the measured solar wind conditions

provides a global realistic description of the magnetospheric electromagnetic fields. In LSK

simulation, large numbers of particle trajectories are followed in the MHD electromagnetic fields

to obtain a global picture of the particle behavior [Ashour-Abdalla et al., 1993]. The particle orbits

are limited by the MHD field temporal resolution. High frequency waves are omitted from the

MHD approximations, so their effects on the particle trajectories are not included.

It might be argued that LSK as a test particle method is not self-consistent because: (1)

calculations of particle trajectories do not provide feedback to the calculations of moments and

fields, and (2) interactions between test particles are not included. However, a feedback process in

the scheme of MHD+LSK simulations would be redundant because the electromagnetic fields

assume a MHD approximation, which provides a self-consistent description with certain

limitations. Incorporating interactions between test particles is also unnecessary because these

interactions have been included via the force of the mean electromagnetic fields. To better

60

understand this, let us consider the testing of a large number of particles in the electromagnetic

fields derived from a fully kinetic simulation. First, it is unnecessary to provide feedback from the

test particles to the calculation of the fields because the fields are self-consistent. Second, the test

particle orbits are exact because the interactions between particles have been incorporated through

the field calculations.

The remaining issue concerning the MHD+LSK simulations is the problem of

normalization. For the test particle method, this involves converting the individual particle

information recorded by virtual detectors such as recording time, position, particle energy, and

pitch angles into collective physical quantities such as the differential flux and distribution function,

which can be directly compared with observations. This has been discussed previously [Ashour-

Abdalla et al., 1993; Kress et al., 2007; Richard et al., 2009]. The directional differential flux

JE(,,,  rD , tD ) can be calculated as

ND ()nddtvA 1 JE(,,,r , t ) C MHD LL (3.1) DD  i1 NLDDD dAn v, dt dEd

where E is particle energy,  is pitch angle,  is azimuthal angle, rD is detector location, tD is

detection time, N L ( ND ) is the number of launched (detected) test particles, ()nv MHD is the

product of MHD number density and velocity, dA L ( dA D ) is particle launch (detector) area, nv,D

is particle velocity direction at detection, dtL ( dtD ) is launch (detection) time interval, and

dddsin  is the solid angle differential. C is a constant on the order of unity. Similarly,

the directional differential energy flux JEeDD ( , , ,r , t ) is

ND ()nddtvA E JE(,,,r , t ) C MHD LL (3.2) eDD i1 NLDDD dAn v, dt dEd

61

Because in the magnetotail, the weak magnetic field may limit the accuracy of the pitch angle

information, we average the flux over the solid angle. The averaged differential flux JE ( ,rD , tD )

and the averaged differential energy flux JEeDD ( ,r , t ) are

C ND ()nddtvA1 JE(,r , t ) MHD LL (3.3) DD  4 i1 NLDDD dAn v, dt dE

CEND ()nddtvA JE(,r , t ) MHD LL (3.4) eDD  4 i1 NLDDD dAn v, dt dE

3.4. MHD and Electron LSK Simulations of the March 11, 2008 Substorm Event

3.4.1. MHD Simulation Results

We modeled the electromagnetic fields using a MHD simulation of the magnetosphere

driven by solar wind conditions from the Geotail spacecraft shown in Figure 3.1. The MHD

simulation results provide the global three-dimensional time-varying electric and magnetic fields

for the Earth's magnetosphere. The GSE coordinate system was used in the MHD simulation. The

MHD simulation data (and LSK simulation data shown later) were processed and will be presented

in a GSE-maximum pressure surface system. In this system, the Z-axis value of maximum pressure

surface is used to replace the virtual spacecraft Z-axis location in the simulation. For instance,

THEMIS P2 location at 06:20 UT is (X ,YZ , )GSE (-14.8, 4.0, -3.0) R E , the virtual detector of P2

in simulation is set as (,,)X YZGSE (-14.8, 4.0, Z mp ) R E , where Zmp is the corresponding

maximum pressure surface location. This system is used because (1) electromagnetic fields and

plasma moments vary significantly within a few RE in the magnetotail along Z-direction of GSE

coordinate and MHD simulation might not accurately reproduce the variation along Z-axis [Raeder

et al., 2001], and (2) the maximum pressure surface is a good approximation to the center of the

62

plasma sheet on the night side [Ashour-Abdalla et al., 2002]. The system is applied when

spacecraft measurements are performed close to the center plasma sheet.

Figure 3.5 (left column) shows nine snapshots from the MHD simulation results at

06:21:40 UT, 06:22:00 UT, 06:22:20 UT, 06:22:40 UT, 06:23:00 UT, 06:32:00 UT, 06:23:40 UT,

06:24:00 UT, and 06:24:20 UT. The color-coded variable is the north-south component of the

magnetic field Bz on the maximum pressure surface. DFs moving earthward can be seen as

increases in the Bz component (red region). The DF of interest was formed at 06:21:40 UT at

X GSE~-14R E and on the dawn side of P2. Bz was ~5 nT, which is relatively weak. The DF

propagated along a flow channel. The earthward flow speed was 400-600 km/s. The flow channel

was limited to a small region in the Y-direction, about 2-3 RE wide. The earthward flow originated

from a region of (,)X YRGSE ~(-21,2) E , where a flow reversal occurred. This flow reversal region is where the near-Earth reconnection occurred in the global MHD simulation. Considering the locations of the DF formation and the flow reversal, it appears that the DF strength did not grow to an appreciable level near the reconnection region, although the flow is generated by magnetic

reconnection. The front was intensified during its earthward propagation. Bz was ~10 nT at

06:23:00 UT, and it was ~14 nT at 06:23:40 when the front encountered P3 and P4 at

X GSE~ -10.4R E . The front merged with the preexisting strong magnetic field at P3 and P4

locations. The MHD results indicate that THEMIS P2 was on the dusk edge of the DF when the

front passed it. We point out that in general dipolarizations in our global MHD simulations are

formed and intensified in the region far away from the reconnection X-line, as a consequence of

flow braking and plasma compression.

63

Bz and flow 062140UT DEflux 41-95keV 062120-062140UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062200UT DEflux 41-95keV 062140-062200UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062220UT DEflux 41-95keV 062200-062220UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

64

Bz and flow 062240UT DEflux 41-95keV 062220-062240UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062300UT DEflux 41-95keV 062240-062300UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062320UT DEflux 41-95keV 062300-062320UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

65

Bz and flow 062340UT DEflux 41-95keV 062320-062340UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062400UT DEflux 41-95keV 062340-062400UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Bz and flow 062420UT DEflux 41-95keV 062400-062420UT 1.0 20 107

-4 -4 )

6 -1 0.8 10 -2 15 -2 eV -1 5 0 0 10 sr 0.6 10 -2

2 2 4 cm P2 P2 10 -1 0.4 P3 P3 Y(RE) Y(RE) 4 5 Bz(nT) 4 103 60.2 P4 6 P4 0 2 80.0 8 10 DEflux(eV s 10 1.0 1.2 1.4-5 10 1.6 1.8 102.01 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300 km/s X(RE) X(RE)

Figure 3.5. Snapshots of the MHD and LSK simulations for the March 11, 2008 event. Left column: magnetic field Z-component and flow vectors (black arrows) on the maximum pressure surface from the MHD simulation. The locations of THEMIS P2, P3, and P4 are also shown. Right

66

column: differential energy flux of electrons in the range of 41keV to 95keV from the LSK

simulation and flow vectors from the MHD simulation on the maximum pressure surface for the

same time intervals as the left column.

3.4.2. LSK Simulation Results and Comparisons with Observations

We present the LSK simulation results in two subsections. In subsection 3.4.2.1, we

compare the differential energy fluxes observed by THEMIS and simulated by LSK and discuss

the implications concerning energization near the reconnection site. In subsection 3.4.2.2, we

present the DF from the MHD simulation and electron energization from the LSK simulation side-

by-side and discuss the electron energization during earthward transport.

3.4.2.1. Electron Energization due to Reconnection

In principle, to quantify the acceleration due to processes near the reconnection site, we

can electron trajectories backwards in time starting from THEMIS P2 location, which is

closer to the reconnection site in the simulation, and thereby examine the electron flux near the

reconnection site. Conversely, if the electron flux observed by THEMIS P2 is reproduced by LSK

via forward pushing of the prescribed electron source launched near the reconnection site, then the

prescribed source probably represents a reasonable flux due to reconnection. We have adopted the

forward pushing method and performed a LSK simulation with two types of source distributions.

The first type of source distribution includes only thermal electrons obeying a Maxwellian

distribution whose one dimensional thermal energy is 1 keV (hereafter, it is referred as 1 keV

Maxwellian distribution). The second type of source distribution is a combination of the 1 keV

Maxwellian distribution and a power law distribution at high energies. The power law distribution

67

has a lower-energy boundary Emin  9 keV and a power law index n  4.5 (Figure 3.6 and also see

Appendix 1 for details). The combined distribution is close to the electron distribution measured

by THEMIS P2. Previous observations near the reconnection region suggested high-energy

electrons follow a power law distribution [Øieroset et al., 2002; Imada et al., 2007]. Theoretical

study suggested that power law distributions can be generated by localized electric fields in

collisionless plasmas [Morales and Lee, 1974]. The combination of a Maxwellian distribution of

thermal electrons and a power law distribution of high-energy tail is a special approximation of

the generalized Lorentzian (Kappa) distributions [Vasyliunas, 1968; Summers and Thorne, 1991].

This combination was adopted because it helps handle high-energy electrons.

102

Maxwell distribution

100 Powerlaw distribution

10-2 f(E) 10-4

10-6

10-8 0.01 0.10 1.00 10.00 100.00 1000.00 E(keV)

Figure 3.6. Two types of electron source distributions for the LSK simulation. The first type

includes only thermal electrons of 1 keV Maxwellian distribution (black line); the second type

includes both thermal electrons and high-energy electrons that obey a power law distribution (blue

line).

68

In the LSK simulation, 37500 electrons following Maxwellian distribution and 4000 electrons obeying power law distribution were launched for every 20 seconds from 06:13 UT to

06:29 UT. The electrons were initially launched uniformly from a planar region given by

-18REGSEEXR -16 , 15REGSEEYR, ZGSE  2.5 . This region is on the MHD maximum pressure surface. By inspection of MHD simulation, this region is also the origin of the earthward flow near the reconnection site. The electron trajectories were followed by using a combination of full particle and guiding-center calculations [Schriver et al., 2011]. We used the adiabaticity parameter  that relates to electron trajectory (not the  in Kappa distribution) to determine whether full particle dynamics was necessary. The  parameter is defined as the square root of

the local magnetic field radius of curvature divided by the local gyro radius [Büchner and Zelenyi,

1989]. If  was decreased to below 10, we switched from the guiding-center approximation to full

particle dynamics; if  was increased to above 15, we performed the opposite switch. The data

about the electrons were collected with virtual detectors that were placed throughout the simulation

domain. We calculated fluxes and energy fluxes by using these detector records and equations (3.3)

and (3.4) above. Notice that flux contribution by different sources can be conveniently added up.

Figure 3.7 shows the differential energy flux for THEMIS P2. The second panel contains

P2 observed energy flux. The energy ranges of 5 channels are covered by both the ESA and SST

instruments. The third panel gives the energy flux from the LSK simulation with only thermal electrons as the source, and the fourth panel shows the results using both the thermal electron source and the high-energy electron source. It is clear that the electron flux calculated by using only a Maxwellian source disagrees with the THEMIS P2 observations at high energies. In contrast, the energy flux calculated by using the source including both thermal Maxwellian and high-energy power law electrons is consistent with the observations. This consistency is manifested in three

69

aspects: the trend in time, the flux levels, and that fluxes of different energies move together. It is

not surprising that including high-energy electrons significantly improves comparison between the

simulation and observations if we consider: (1) P2 is not far from the source region where the electrons were launched; (2) the electron flux observed by P2 has a high-energy tail; and (3) no magnetic structures such as DFs formed between the electron source region and the P2 location, therefore substantial acceleration is unlikely to occur between the source location and P2. These comparisons show that the simulation including high-energy electrons with a power law distribution is the correct one. More importantly, it suggests that electrons were accelerated to up to 95 keV near the reconnection site. The power law distribution reasonably quantifies the production of high-energy electrons by processes near the reconnection site.

Figure 3.8 shows the energy flux for THEMIS P4 in the same format as in Figure 3.7. Two characteristics are noticeable. First, unlike the observed and simulated fluxes at THEMIS P2, the energy flux at P4 is different for different energy channels. Below 12 keV, the flux fluctuates and generally does not increase, but for channels above 12 keV, the flux significantly increases at

06:24:00 UT in observations and at 06:23:00 UT in the simulation. Second, adding a power law distribution of high-energy electrons to the source in the LSK simulation improves the comparison between the observed and simulated fluxes, especially for the 25-41 keV channel. However, this improvement is less significant than that seen in the P2 data. This is expected because the flux of source high-energy electrons is much smaller than that observed at P4 when the front passed by.

For example, the 41-95 keV flux is on the level of 1056 ~10 eV/(cm 2 s sr  eV) at P2 while on the

level of 1067 ~10 eV/(cm 2 s sr  eV) at P4. The majority of the high-energy electrons at P4 are not

from convection of the source. Instead, they are from acceleration of lower-energy electrons by

the dipolarization during earthward transport. This will be discussed in the next subsection.

70

Magnetic field and DEflux (P2) 20 bz 10 0 by (nT)

B(GSM) -10 bx -20 ) 9 -1 10 2-6keV 8 eV 10

-1 6-12keV 7 sr 10 -2 12-25keV 6

cm 10 25-41keV -1

Observed 5 10 41-95keV 104 (eV s 9 10 108 107 106 (MXL ele) Simulated 105 104 9 10 108 107 106

Simulated 105 4 (MXL-PWL ele) 10 X_gsm(Re) -14.8 -14.7 -14.7 Y_gsm(Re) 5.4 5.3 5.3 Z_gsm(Re) -1.7 -1.8 -1.8 hhmm 0620 0625 0630 2008 Mar 11

Figure 3.7. Comparison of the energy flux for THEMIS P2 in the March 11, 2008 event. From

top to bottom, the first panel shows the measured magnetic field. The second panel shows THEMIS

P2 observations. Energy ranges of 5 channels are covered by both the ESA and SST instruments.

The third panel gives the simulated differential energy flux using only Maxwellian thermal

electrons as the source, and the fourth panel gives the result including both Maxwellian thermal

electrons and high-energy electrons. Because in the MHD simulation, the flow associated with DF

arrives at P2 about 30 seconds earlier than observation, the simulated flux is also about 30 seconds

ahead of the corresponding observed flux, as indicated by the vertical lines.

71

Magnetic field and DEflux (P4) 30 bz 20 10 by

(nT) 0 B(GSM) -10 bx -20 ) 9 -1 10 2-6keV 8 eV 10

-1 6-12keV 7 sr 10 -2 12-25keV 6

cm 10 25-41keV -1

Observed 5 10 41-95keV 104 (eV s 9 10 108 107 106 (MXL ele) Simulated 105 104 9 10 108 107 106

Simulated 105 4 (MXL-PWL ele) 10 X_gsm(Re) -10.4 -10.4 -10.4 Y_gsm(Re) 5.4 5.3 5.3 Z_gsm(Re) -1.6 -1.6 -1.6 hhmm 0620 0625 0630 2008 Mar 11

Figure 3.8. Comparison of the energy flux for THEMIS P4 in the March 11, 2008 event. The same format is used as in Figure 3.7. The simulated flux is about one minute ahead of the corresponding observed flux, as indicated by the vertical lines.

Before moving on to the next subsection, we show the comparison of the differential fluxes as functions of energy between the observations and the simulation. The differential fluxes are calculated by using equation (3.3). Figure 3.9 shows a comparison of the differential fluxes from

P2 observations and the LSK simulation at 06:24:00 UT. Because the traces of the energy flux from the THEMIS P2 data and the LSK simulation are similar in Figure 3.7, we selected the same times in observation and simulation. The differential flux as a function of energy from the LSK

72

simulation matches well with the THEMIS data. Figure 3.10 shows a comparison of the differential

flux for P4 at 06:25:00 UT. As shown in Figure 3.8, the flux change in LSK simulation occurs

about one minute before the change in the THEMIS P4 observations, so the simulated flux from

about one minute ahead was selected for comparison. Figure 3.10 shows that the simulated flux is consistent with the corresponding THEMIS P4 data. These remarkable agreements validate the

MHD+LSK simulations and demonstrate the ability to use LSK simulations to reproduce the differential fluxes in a wide range of energies.

105

P2 Obs(062400UT) P2 Sim(062340-062400UT)

) 4

-1 10 P2 Sim(062400-062420UT) eV -1 sr -1

s 3

-2 10

102

101 Differential flux (cm

100 102 103 104 105 Energy (eV)

Figure 3.9. Comparison of the differential flux as a function of energy for THEMIS P2. The black line shows the measured flux by P2 at 06:24:00 UT, the blue line shows the simulated flux at

06:23:40-06:24:00 UT, and the red line shows the simulated flux at 06:24:00-06:24:20 UT.

73

105

P4 Obs(062500UT) P4 Sim(062320-062340UT)

) 4

-1 10 P4 Sim(062340-062400UT) eV -1 sr -1

s 3

-2 10

102

101 Differential flux (cm

100 102 103 104 105 Energy (eV)

Figure 3.10. Comparison of the differential flux as a function of energy for THEMIS P4. The

black line shows the measured flux by P4 at 06:25:00 UT, the blue line shows the simulated flux

at 06:23:20-06:23:40 UT, and the red line shows the simulated flux at 0623:40-06:24:00 UT.

3.4.2.2. Electron Energization during Transport

In Figure 3.5 (right column), the differential energy flux data for the energy range 41 keV

to 95 keV are plotted on the maximum pressure surface at 20 second intervals from 06:21:40 UT

to 06:24:20 UT. Each plot corresponds to the snapshot of the MHD simulation shown next to it in

the left column. Two key points can be drawn from Figure 3.5.

First, it is evident that the high-energy electron flux at P3 and P4 increased by almost an

order of magnitude compared with the flux near P2 as the front arrived. This increase occurred far

away from the reconnection sites. The high-energy electrons at P3 and P4 were from acceleration

during earthward transport, rather than convection of the high-energy source.

74

Second, the high-energy electrons at P3 and P4 were generated where and when the

dipolarization formed and intensified, as indicated by Bz from the MHD simulation shown in the

left column for Figure 3.5. Before the dipolarization, the flux close to the source region was

510eV/(cmssreV)52 for the 41-95 keV energy range. As the dipolarization formed at about

06:21:40 UT at X GSE~-14R E and continued to intensify till 06:22:40 UT at X GSE~-12R E , electrons were energized in about one minute. The flux of 41-95 keV electrons increased to

310eV/(cmssreV)62 during that interval. After that, the flux of high-energy electrons

propagated to P3 and P4, and were accumulated in the region near X GSE~-11R E as the MHD flow

was braked and diverted. At X GSE~-11R E , the electron flux flowed toward the dawn side, resulting

from magnetic field gradient and curvature drifts.

3.5. Discussions

To identify the energization mechanism during earthward transport, we followed the flow

and inferred the trajectory of the DF on the maximum pressure surface. Then we calculated the

differential flux along this trajectory. The results are summarized in Figure 3.11. The blue line is

the simulated differential flux at 06:21:40-06:22:00 UT at the position of

(X ,YX ) (P4 5 RYR E , P4  2.5 E ) . (,)X PP44Y is virtual P4 location in the simulations. This

represents the electron source before the formation of the front. The green line is the simulated

flux at a relative position of (X ,YX ) (P4 2 RYR E , P4  0.5 E ) at 06:22:20-06:22:40 UT, which is

where and when the DF in the MHD simulation formed and intensified. The red line is the

simulated flux at the position of P4 at 06:23:20-06:23:40 UT, when the front propagated to P4.

The black line is the flux measured by P4 at 06:25:00 UT. It is clear that the acceleration of the

75

test particles (from blue line to green line) occurred where the DF formed and intensified.

Furthermore, two key points are noticeable. First, the energization is nearly uniform for all

electrons with energy above 1 keV. Second, the magnetic field increased from ~5nT to ~12nT, but the flux of high-energy electrons increased by almost an order of magnitude. The acceleration is

rather effective in increasing the high-energy flux. These two points support that the acceleration

is adiabatic, as showed in the previous chapter.

105 P4 observation source distribution energized distribution

) 4

-1 10 P4 simulation eV -1 sr -1

s 3

-2 10

102

101 Differential flux (cm

100 102 103 104 105 Energy (eV)

Figure 3.11. Electron acceleration mechanism during transport in the March 11, 2008 event. The

blue line is the simulated flux at position of (X ,YX ) (P4 5 RYR E , P4 2.5 E ) at 06:21:40-06:22:00

UT, the green line is simulated flux at a relative position of (X ,YX ) (P4 2 RYR E , P4 0.5 E ) at

06:22:20-06:22:40 UT, the red line is the simulated flux at P4 at 06:23:40-06:24:00 UT, and the

black line is the flux observed by P4 at 06:25:00 UT.

76

Now let’s discuss electron transport. Figure 3.12 shows the Z-component of the magnetic

field, low-energy (2-6 keV) electron flux and high-energy (41-95 keV) electron flux in a broader region. Electrons closely followed the flow channels, both in earthward and tailward directions.

Low- and high-energy electrons followed a similar transport pattern determined by the flows. This similarity across energies demonstrates that the energy-independent E×B drift played a central

role in electron transport in the magnetotail. Close to the Earth, low-energy electrons were

accumulated and penetrated the region of strong magnetic field while high-energy electrons circled

around it. The high-energy electrons dawnward drifted more significant than low-energy electrons.

This is because close the Earth, energy-dependent gradient and curvature drifts are the main

component of electron guiding-center motion. We point out that the transport features are

statistical. Specifically, many test electrons were launched uniformly in a region covering the flow

channel, hence the transport was averaged across the flow channel. In addition, the flux on

maximum pressure surface is bounce averaged. That the flux closely follow the flows in the

magnetotail does not to require that E×B drift is much larger than curvature/gradient drift all the

time for test electrons.

77

1.0 Bz and flow 062340UT -10 20

-5 15

0 10 P3 P2 Y(RE) 5 5 Bz(nT) 0.8 P4

10 0

15 -5 -5 -10 -15 -20 -25 -30

DEflux 2-6keV 062320-062340UT -10 0.6 108 )

-5 -1 eV -1

106 sr 0 -2 cm

P3 P2 -1 Y(RE) 5 P4 104

10 0.4 DEflux(eV s 102 15 -5 -10 -15 -20 -25 -30

DEflux 41-95keV 062320-062340UT -10 107

6

10 )

-5 -1 eV

0.2 -1 105 sr

0 -2

4 cm

P3 P2 10 -1 Y(RE) 5 P4 103

10 102 DEflux(eV s

150.0 101 -51.0 -101.2 -15 1.4 -20 1.6-25 1.8-30 2.0 300 km/s X(RE)

Figure 3.12. Electron transport in the March 11, 2008 event. From top to bottom, the first panel is

the Z-component of the magnetic field and flow vectors (black arrows) on the maximum pressure

surface from the MHD simulation at 06:23:40UT. The locations of THEMIS P2, P3, and P4 are

78

also shown. The second panel is differential energy flux of electrons in the range of 2 keV to 6

keV accumulated during 06:23:20-06:23:40 UT from the LSK simulation and flow vectors from

the MHD simulation on the maximum pressure surface. The third panel is for 41 keV to 95 keV electrons, same format is used as in the second panel.

Finally, let's look at the simulation results in the context of the THEMIS wave observations.

First, electromagnetic whistler mode waves with electric wave fields ~5 mV/m were detected by

P4 from 06:23:55.6 UT to 06:23:56.4 UT (not shown here), and their frequencies are about

0.7 ~ 0.9 fce . Previous studies showed whistler waves associated with DFs can scatter electrons

near the resonant energy [e.g. Khotyaintsev et al., 2011]. However, the resonant energy is a few

keV for this event, which is well below the energy range we are considering (tens of keV). Second,

electrostatic waves with electric fields ~2 mv/m and frequencies ~ 1.2 fce were also detected from

06:23:23 UT to 06:23:24 UT and from 06:23:25 UT to 06:23:26 UT by P3. Similar waves detected at DFs by THEMIS have been reported [Zhou et al., 2009]. Considering the short presence and small amplitude, they are unlikely to energize electrons significantly. Third, as can be seen in the last two panels in Figure 3.3 and Figure 3.4, there are broad band waves with frequencies close to

the lower hybrid frequency fcef ci associated with the DFs. According to the resonance

condition, they are expected to affect ion orbits. These waves are not incorporated in the MHD

simulation, hence not in the LSK simulation. The good agreements between the MHD+LSK

simulations and observations suggest that including the effect of these waves is unlikely to change

the electron energization and transport picture in this event.

3.6. Conclusions

79

Using analyses of coordinated multi-point observations, a realistic global MHD simulation,

and an electron LSK simulation, we have developed a global and quantitative picture of electron

energization and transport in the magnetotail during a substorm event.

The global scenario of electron energization is as follows: the electrons are initially

accelerated near the reconnection site, and subsequently, they are further accelerated adiabatically

by convection electric fields far away the reconnection site. For the March 11, 2008 event, the

electron energy flux for the energy range 41-95keV is 510eV/(cmssreV)52 due to

acceleration near the reconnection site; it is further increased to 310eV/(cmssreV)62 due to

adiabatic acceleration of lower-energy electrons. The adiabatic enhancement occurs when and

where the DF forms and intensifies. This two-step acceleration process has been previously

suggested by observational and analytical studies [e.g. Asano et al., 2010; Vaivads et al., 2011;

Pan et al., 2012]. Our simulations provide quantitative evidence supporting this picture.

The global transport of electrons occurs as follows: in the magnetotail, the electron fluxes follow along the flow channel both in earthward and tailward directions, with small diffusion in configuration space. This is because the energy-independent E×B drift driven by strong

convection electric fields is statistically dominant in the flow channels. The gradient and curvature

drifts become dominant in the inner magnetosphere (within X ~ 11RE in the March 11, 2008

event). As a consequence, the high-energy electrons drift towards the dawn side.

80

CHAPTER 4

Modeling Ion Energization and Transport Associated with Magnetic

Dipolarizations during a Substorm

4.1. Introduction

In this chapter, we extend the global MHD+LSK simulation scheme to examine ion

energization in the magnetotail during a substorm event that occurred on February 07, 2009. The

main task is to develop a global scenario of ion energization and transport associated with magnetic

dipolarizations under realistic magnetospheric conditions. Since in the next chapter the February

07, 2009 event is also used for a comparison of energization and transport mechanisms between

electrons and ions, the other task in this chapter is to present observations and MHD simulation

results. This chapter is organized as follows: the observations are presented in Section 4.2; the

MHD simulation results are presented in Section 4.3; the ion LSK simulation results are discussed

in Section 4.4; the simulation results are summarized in Section 4.5.

4.2. Observations of the February 07, 2009 Substorm Event

During the time of interest (03:00-05:00 UT) the WIND spacecraft was located at

(X ,YZ , )GSM (202, 73, 38) R E . Figure 4.1 shows its measurements of the solar wind. The IMF

was southward and weak staring from 00:10 UT (the black vertical line). The solar wind was slow;

Vx was about 320 km/s. Geomagnetic indices show a weak substorm onset at ~03:40 UT. The peak

AE index was ~ 120 nT. The solar wind observed by the WIND spacecraft needs about 55 minutes

to propagate toX  25RE , the upstream boundary of the global MHD simulation. During the

81

substorm of interest (03:00-05:00 UT), the five THEMIS spacecraft were all on the dawn side of

the magnetotail. At time 04:00 UT, THEMIS P4, P5, and P3 were located close to each other at

X 8.4RE , X  8.5RE and X  9.4RE down tail. P2 was located at X  18.6RE , and the

P1 was located atX  30.6RE . The geosynchronous spacecraft GOES-10, GOES-11, and GOES-

12 were respectively near 0, 19, and 23 h magnetic local time. They were on the dusk side of the

THEMIS fleet. Lyons et al. [2012] investigated the relation between aurora streamers observed by

THEMIS all sky imagers and dipolarization and flow activities in the tail in this substorm event.

Aurora streamers are longitudinal finger-like structures. They determined the substorm onset to be at 03:47:12 UT (the blue vertical line in Figure 4.1) near the longitude of GOES-10. Subsequently,

the aurora activity expanded westward to GOES-12 location and eastward to P4, P5, and P3 locations. They found that GOES-10 did not observe dipolarizations when the substorm onset there,

instead significant dipolarization was observed by P3, P4, and P5 at ~ 04:06 UT in association

with aurora streamers. Oka et al. [2011] showed that after the dipolarization seen by P3, P4, P5 at

~ 04:06 UT, the pressure in the inner magnetosphere increased. This pressure increase propagated

tailward and was detected by P2 at X ~ 18.6RE at ~ 04:14 UT. Meanwhile, the tail seen by P3,

P4 and P5 became more stretched. At ~ 04:18 UT, P1 at X   30.6RE detected a reversal of flow

from tailward to earthward, suggesting the X-line retreated to P1 location at that time. They

concluded that the pressure increase after dipolarizations eventually caused the X-line to move

tailward.

82

WIND Observation of Solar Wind 6 4 2 0 Bx (nT) -2 -4 -6 6 4 2 0 By (nT) -2 -4 -6 6 4 2 0 Bz (nT) -2 -4 -6 100 Vz 0 -100 Vy Vi -200 (km/s) -300 Vx -400 4 3

P 2

(nPa) 1 0 150 AL 100

50 AU (nT) Index 0

Geomagnetic -50 AE -100 hhmm 0000 0200 0400 0600 2009 Feb 07

-10

-5 P2 P3 P1 P5 0 G10 G12 P4

Y_gsm (Re) 5 G11

10 0 -10 -20 -30 X_gsm (Re)

Figure 4.1. Solar wind measured by WIND and satellite positions in the February 07, 2009 event.

(Top) Solar wind data. From top to bottom showed are the IMF, the solar wind velocity, the proton

83

number density, the dynamic pressure, and geomagnetic indices. The black vertical line is 00:10

UT and the blue vertical line is 03:47:12 UT, adapted from El-Alaoui et al. [2013]. (Bottom)

Projection of spacecraft trajectories onto the equatorial plane between 01:00 UT and 07:00 UT

(marked points are at 04:06 UT).

Since we are interested in particle energization in the earthward direction of the

reconnection site, we present the detail observations from P2 at X ~ 18.6RE in Figure 4.2. The vertical line indicates an earthward propagating reconnection jet. The measured electric field spikes were about 20 mV/m. The measured electric field was larger than the convection electric

field ( VBi ), indicating P2 was close to the IDR. The measured electron density is smaller than

the ion density, probably due to overkill of photoelectron effect. The photoelectrons were removed

before calculating the electron density. The energy below which the electrons were regarded as

photoelectrons is represented by the black line in electron energy flux (the last panel). Although

the substorm event was weak, the earthward flow reached 500-600km/s. The local Alfvén speed

3 with BnT~10 and ncmi ~0.1 was vkmsA ~ 689 / . Hence the outflow velocity was close to the

local Alfvén speed. The electron temperature was TkeVe ~ 2 and the ion temperature was

TkeVi ~3 4 . The temperature ratio was TTie / ~ 1.5 2 . The temperature of both species in the

reconnection outflow region were larger than that in the lobe region, suggesting that the plasmas

were heated in the reconnection outflow. The total pressure was almost constant, even though the

magnetic pressure and plasma pressure varied significantly. The high-energy electron distribution

5 in the outflow region is fitted to a power law distribution fe ()EE with E 11 keV . The high-

6 energy ion distribution is fitted to fi ()EE with E  25 keV .

84

THEMIS P2 observation 20 Bz 10 0 By (nT) -10 B(GSM) Bx -20 20 Ez 10

0 Ey

(mV/m) -10 E(GSM) Ex -20 10 5 0 -5 (mV/m) -10 Evixb(GSM) 1.00 Ne 0.10 Ni Density (cm^-3) 0.01 600 400 200

(km/s) 0 Vi(GSM) -200 8 Ti/Te 6 Te T 4 (keV) 2 Ti 0 1.000 Pb 0.100 Pe Pi

(nPa) 0.010

Pressure Pt 0.001 106 107 105 106 104 105

3 4

ion 10 10

Eflux 2 3

10 ) 1 10 10 102 -1 6 8 105 107 eV 10 10 -1 4 6 10 105 sr 3 -2 10 104

Eflux 2 10 10 3 cm electron 10 -1 101 102 X_gsm(Re) -18.6 -18.6 -18.6 -18.6 Y_gsm(Re) -2.5 -2.6 -2.7 -2.8 (eV s Z_gsm(Re) -4.2 -4.2 -4.2 -4.2 hhmm 0350 0400 0410 0420 2009 Feb 07

Figure 4.2. THEMIS P2 observations in the February 07, 2009 event. From top to bottom showed are the magnetic field in GSM coordinates, the measured electric field, the convection electric

field ( VBi ), the density, the ion bulk velocity, the temperature, the pressure, the ion energy

flux, and the electron energy flux. The vertical line indicates an earthward propagating

reconnection jet.

85

Figure 4.3 shows the observations of the dipolarization of interest by THEMIS P3 in the

inner magnetosphere. THEMIS P4 and P5 measurements are similar to those from P3. The electric

field associated with the dipolarization was mainly in the X- and Y- direction, due to large Vx and

Vy . The measured electron density was higher than the ion density, probably resulting from

secondary electron effect in the ESA instrument. The ion temperature was TkeVi ~ 6 8 and the

electron temperature was Te ~ 3 4 keV , larger than those prior the arrival of the front. The tenuous

plasmas carried by the dipolarization to the inner magnetosphere were also hotter than the plasmas in the reconnection outflow at P2, suggesting that substantial heating occurred during earthward transport. The magnetic pressure change was balanced by the plasma pressure change, resulting in

an almost constant total pressure across the DF. High-energy ( 25keV ) ion and electron energy

fluxes measured by the SST instrument increased as the front arrived. However, the electron flux

gradually shifted to higher-energy beginning from 03:59 UT, ~6 min prior the arrival of the DF.

Recall the aforementioned analyses by Lyons et al. [2012]. This gradual increase was likely caused by the substorm activity that occurred on the dusk side of P3 location prior to the observation of the dipolarization. We found that the electron flux spectra before the dipolarization were dispersed, i.e. higher-energy flux increased earlier than lower-energy flux, suggesting that the flux increase

resulted from energy-dependent dawnward gradient and curvature drifts. In contrast, the change

in ion fluxes prior to the dipolarization was much weaker, suggesting after injection from the tail,

ions drifted (duskward) farther away from P3. We will examine the high-energy particles in detail

when we discuss LSK simulation results.

86

THEMIS P3 observation 30 Bz 20 10 By

(nT) 0

B(GSM) -10 Bx -20 15 Ez 10 5

Ey 0 (mV/m) E(GSM) -5 Ex -10 10 5 0 -5 (mV/m) -10 Evixb(GSM) 1.0 Ne

Ni Density (cm^-3) 0.1

100 0

(km/s) -100 Vi(GSM) -200 8 Ti/Te 6 Te T 4 (keV) 2 Ti 0 1.000 Pb 0.100 Pe Pi

(nPa) 0.010

Pressure Pt 0.001 106 107 105 106 104 105

3 4

ion 10 10

Eflux 2 3

10 ) 1 10 10 102 -1 6 8 105 107 eV 10 10 -1 4 6 10 105 sr 3 -2 10 104

Eflux 2 10 10 3 cm electron 10 -1 101 102 X_gsm(Re) -9.2 -9.3 -9.3 -9.4 Y_gsm(Re) -1.7 -1.9 -2.1 -2.2 (eV s Z_gsm(Re) -3.4 -3.4 -3.4 -3.4 hhmm 0350 0400 0410 0420 2009 Feb 07

Figure 4.3. THEMIS P3 observations in the February 07, 2009 event. The same format is used as in Figure 4.2. The vertical line indicates the arrival of the DF.

87

4.3. MHD Simulation of the February 07, 2009 Substorm Event

The global MHD simulation used the WIND observation of the solar wind as input at the

upstream boundary at X  25RE . The simulation results were validated and described by El-

Alaoui et al. [2013]. They found that the MHD simulation reproduced the dipolarizations and high-

speed flows observed by P3, P4 and P5 in the magnetosphere reasonably well. Detailed analysis

of the simulation revealed a global picture of the magnetotail dynamics for this substorm event.

Specifically, the substorm related reconnection occurred at X ~20 RE and extended only

partway across the magnetotail; the localized reconnection generated high-speed outflow

propagating earthward (and tailward) in narrow channels; multiple magnetic dipolarizations

formed and intensified at the earthward ends of the flow channels; together with flows, the

dipolarizations propagated toward a region with a strong magnetic field in the inner magnetosphere

at X ~7 RE that acted like a “wall”, where the dipolarized magnetic fluxes merged with the

strong magnetic field and piled up, while the flows were slowed and subsequently diverted,

forming large vortexes.

Figure 4.4 shows the snapshots of the Z-component of the magnetic field ( Bz ) and the Y-

component of the electric field ( Ey ) every 40 sec from 04:04:00 UT to 04:09:20 UT. At 04:04:00

UT, a small dipolarization structure intensified at (,)~(11,2)X YR E (white arrow). It

propagated earthward and merged with a strong dipole magnetic field at X ~ ( 7,2) RE at

04:06:40 UT. The dipolarization was driven by a high-speed flow in a narrow channel. The flow

originated in a region at X ~20 RE and YR~0 7 E , where the reconnection occurred. The flow

curved first toward the dawn side and then toward the dusk side, and finally slowed and diverted

as it encountered the strong-field wall. Subsequently, a larger magnetic dipolarization intensified

88

near 04:06:40 UT at (X ,YR ) ~ ( 12,0) E (black arrow), and gradually increased to 30 nT at

04:10:00 UT as it propagated along a path dawnward of the earlier dipolarization. This larger

dipolarization was also driven by a similar curved flow. The characteristic flow speed was 300

km/s. The characteristic width of the flow channels was about 23 RE . A large Ey with a peak

value of ~5-6 mV/m coincided with these two dipolarizations. The electric field in an MHD

simulation is given by EVBj   , where V is the MHD flow velocity,  is the resistivity,

and j is the current [Raeder et al., 1998]. We checked that the large electric field associated with

the dipolarization structures was dominated by the VB term, which represents the convection

electric field carried by flows. This point can also be confirmed by a simple estimate: a 20 nT

dipolarized magnetic field in a 300 km/s flow channel induces a 6 mV/m electric field. Similar as

those in the global MHD simulation of the March 11, 2008 event presented in Chapter 3, the

simulated dipolarizations in the February 07, 2009 event also formed and intensified far away from

the reconnection region, resulting from flow braking and plasma compression.

89

Bz and flow Eflux(>25keV) and flow Ey and flow -10 1.0 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 0 0 0 2 10 5 Y(Re) Y(Re) Y(Re) 10 Bz(nT) Ey(mV/m) 040400 UT 5 0 5 104 5 0 Eflux(keV/s/cm

10 -10 10 103 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 0.8 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 0 0 0 2 10 5 Y(Re) Y(Re) 10 Y(Re) Bz(nT) Ey(mV/m) 040440 UT 5 0 5 104 5 0 Eflux(keV/s/cm

10 -10 10 103 10 -2 -5 0.6-10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 0 0 0 2 10 5 Y(Re) Y(Re) Y(Re) 10 Bz(nT) Ey(mV/m) 040520 UT 5 0 5 104 5 0 Eflux(keV/s/cm 10 0.4 -10 10 103 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 0 0 0 2 10 5 Y(Re) Y(Re) Y(Re) 10 Bz(nT) Ey(mV/m) 040600 UT 5 0 5 104 5 0 0.2 Eflux(keV/s/cm 10 -10 10 103 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =19.6keV

test 0 0 0 2 10 5 Y(Re) Y(Re) 10 Y(Re) Bz(nT) Ey(mV/m) 5 0.0 0 5 104 5 0 1.0 1.2 1.4 1.6 Eflux(keV/s/cm 1.8 2.0 3 040640 UT W 10 -10 10 10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300km/s X(Re) X(Re) X(Re)

90

Bz and flow Eflux(>25keV) and flow Ey and flow -10 1.0 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =19.3keV

test 0 0 0 2 10 5 Y(Re) Y(Re) Y(Re) 10 Bz(nT) Ey(mV/m) 5 0 5 104 5 0 Eflux(keV/s/cm

3 040720 UT W 10 -10 10 10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 0.8 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =23.3keV

test 0 0 0 2 10 5 Y(Re) Y(Re) Y(Re) 10 Bz(nT) Ey(mV/m) 5 0 5 104 5 0 Eflux(keV/s/cm

3 040800 UT W 10 -10 10 10 10 -2 -5 0.6-10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =32.1keV

test 0 0 0 2 10 5 Y(Re) Y(Re) 10 Y(Re) Bz(nT) Ey(mV/m) 5 0 5 104 5 0 Eflux(keV/s/cm

3 040840 UT W 10 0.4 -10 10 10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =36.1keV

test 0 0 0 2 10 5 Y(Re) Y(Re) 10 Y(Re) Bz(nT) Ey(mV/m) 5 0 5 104 5 0 0.2 Eflux(keV/s/cm 3 040920 UT W 10 -10 10 10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) X(Re) -10 -10 -10 6 30 107 -5 -5 -5 4 /sr) 20 106 2 =37.9keV

test 0 0 0 2 10 5 Y(Re) Y(Re) 10 Y(Re) Bz(nT) Ey(mV/m) 5 0.0 0 5 104 5 0 1.0 1.2 1.4 1.6 Eflux(keV/s/cm 1.8 2.0 3 041000 UT W 10 -10 10 10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300km/s X(Re) X(Re) X(Re)

91

Figure 4.4. Snapshots of the MHD and ion LSK simulations for the February 07, 2009 event. Left

column: The Z-component of the magnetic field from the MHD simulation. Middle column:

Integrated energy flux with energy above 25 keV from the ion LSK simulation. Right column: The

Y-component of the electric field from the MHD simulation. Flow vectors (black arrows) from the

MHD simulation are plotted over the magnetic field, energy flux and electric field. All of the

quantities are plotted on the maximum pressure surface. Each row of plots corresponds to the same

time, which is labeled next to the Y-axis on the left. The representative ion trajectory, shown in

Figure 4.6, is superimposed on these quantities (white lines). The solid circle on each line

represents the ion location at the corresponding time. The energy of the test ion is also labeled next

to the Y-axis.

4.4. Ion LSK Simulations of February 07, 2009 Substorm Event

4.4.1. LSK Simulation Set-up

Applying the electric and magnetic fields derived from the MHD simulation, we performed

an ion LSK simulation. The ion LSK code solves the Lorentz-force equation of motion by using a

fourth-order Runge-Kutta method. For the LSK simulation ion source in the magnetotail, we

adopted a distribution that combines a Maxwellian distribution representing the thermal ions and

a power law distribution representing the high-energy tail. Two parameters, the temperature

(4Ti  keV ) of the Maxwellian distribution and the power law index (n=6) were derived from ion

distribution observed by THEMIS P2 in the reconnection outflow at X 18.6RE during this

event. For the simulation, 80,000 ions obeying an isotropic Maxwellian distribution and 80,000

ions obeying an isotropic power law distribution were launched separately every 20 sec from 03:50

UT to 04:20 UT. The ions were launched uniformly in space from a surface given by

92

-19 REEXR-17 , 5 REEYR10 , and Z  Zmp , where Zmp is the maximum pressure surface location. The maximum pressure surface is a good approximation to the center of the plasma sheet on the night side [Ashour-Abdalla et al., 2002]. This launch location was selected because by inspecting the flows and magnetic field from the MHD simulation, we found that the high-speed flows originated there. The ion data were collected using virtual detectors placed throughout the simulation domain. Energy fluxes were calculated by using the detector data and

equation (3.4).

4.4.2. LSK Simulation Results and Comparisons with Observations

4.4.2.1. Ion Energy Fluxes

In order to validate the LSK simulation, we compare our simulation results with THEMIS

observations. Because THEMIS P3, P4 and P5 were within 1.5 RE to each other they observed

similar high-energy flux increases, only the comparison for P3 is shown in Figure 4.5. The key

observational features upon the arrival of the dipolarization are the increase of Bz , the earthward

flow, and the dramatic increase of high-energy ion fluxes in contrast to the modest change of low-

energy fluxes. The dipolarization of interest accompanied by the earthward flow is reproduced by

the global MHD simulation. Compared to observations, the simulated dipolarization is weaker,

and the magnetic field and flow speed have fewer variations on the shorter time-scale. The

simulation reproduces the observed dramatic enhancement of the high-energy ion fluxes. Similar

to the observations, the simulated low-energy flux changes little.

93

30 Bx 20 10

By B

(nT) 0 -10 Bz -20 200 Vx 100

Vy V 0

(km/s) -100 Vz )

-1 -200 107 25-37keV

eV 6

-1 10 5 37-48keV

sr 10 48-65keV -2 104 3 65-77keV cm (SST) 10 77-116keV -1 Ion eflux 102 101 116-175keV 7

(eV s 10 1.5-2.7keV 106 5 2.7-4.6keV 10 104 4.6-8.0keV 3

(ESA) 10 8.0-13.8keV Ion eflux 102 13.8-25keV 101 30 Bx 20 10 By B

(nT) 0 -10 Bz -20 200 Vx 100 Vy V 0

(km/s) -100 Vz -200 107 25-37keV 106 5 37-48keV 10 48-65keV 104 3 65-77keV (SST) 10 77-116keV Ion eflux 102 101 116-175keV 107 1.5-2.7keV 106 5 2.7-4.6keV 10 104 4.6-8.0keV 3

(ESA) 10 8.0-13.8keV Ion eflux 102 13.8-25keV 101 X_gsm(Re) -9.2 -9.3 -9.3 -9.4 Y_gsm(Re) -1.7 -1.9 -2.1 -2.2 Z_gsm(Re) -3.4 -3.4 -3.4 -3.4 hhmm 0350 0400 0410 0420 2009 Feb 07

Figure 4.5. Comparison of observations and simulations for P3 in the February 07, 2009 event.

From top to bottom, the first four panels are observed magnetic field, plasma velocity, high-energy ion flux from the Solid State Telescope (SST), low-energy ion flux from the Electrostatic Analyzer

(ESA). The next four panels are the corresponding simulated quantities at the virtual P3 position

94

in the simulation domain. The spike in the observed flux in between 04:03 UT and 04:04 UT is

because of the attenuator set-up in the SST detector. The vertical lines indicate the arrival of the

dipolarization of interest. The ~2 minute difference between the observed DF and the simulated

DF is reasonable, see Raeder et al. [1998, 2001], El-Alaoui et al. [2001, 2009].

To examine the energization and transport process, the integrated energy flux with energy

greater than 25 keV derived from the LSK simulation is plotted in Figure 4.4 (middle column).

 The integrated energy flux is defined as JE(,,)r tdE, which gives the total energy flux above 25keV e

25 keV and has the unit of keV s-1 cm -2  sr -1 . Three features are remarkable for the total energy

flux. First, the flux intensity pattern was similar to that of Bz . Enhancements of the high-energy

flux coincided with the two dipolarizations described in section 4.3 (as indicated by the arrows in

each snapshot). Quantitatively, near the reconnection site, the flux was on the order of

~106-1-2-1 keV s cm  sr , it increased to ~107-1-2-1 keV s cm sr as the two dipolarizations

intensified during their propagation to the inner magnetosphere. This illustrates that dipolarizations

are powerful accelerators acting far distant from the reconnection site. Second, in the magnetotail

( X ~9 RE ), the high-energy flux closely followed the flow channels, which is a combined

effect of acceleration and convection. Specifically, the reconnection outflow carried a large

convection electric field ( Eyy~(VB )as described above) which pushed the ions in the flow direction via the EB drift. When pushed along the flows to the dipolarization regions, the ions

were accelerated almost adiabatically (described below) to higher energy and contributed to the high-energy flux in the flow channels. Third, that the flux (statistical quantities) closely followed the flow channels is evidence that the energy-independent EB drift dominated along the flow

95

channels. The flows slowed at the “wall” ( X ~7 RE ), where the ion gradient and curvature drifts

became significant so the ions moved toward the dusk side.

4.4.2.2. Ion Trajectories

In this section, we examine ion trajectories to identify the acceleration mechanism. We

focus on the ions that contribute to the high-energy flux enhancements associated with the

dipolarizations showed in Figure 4.4. The trajectory of a representative ion and its characteristics

are presented in Figure 4.6. This ion was launched near the reconnection site at 04:06:20 UT at

 X 18.3 RE and YR 0.27 E . It had an initial energy of 13.1 keV and a pitch angle near 90 .

This initial energy is in the high-energy tail of the plasma sheet distribution. We choose a high- energy tail ion because dipolarizations act as adiabatic accelerators; for ions to be accelerated to high energy, their initial energy needs to be sufficiently large. As the ion was pushed toward the

Earth (until X ~9 RE ), its energy increased to ~40 keV. The ion was accelerated through a two- stage process. The first-stage acceleration occurred from 04:06:20 UT to 04:07:26 UT in the region

of X ~ 13 RE (before the first blue vertical dashed line). Due to the weakness of the magnetic

field (frequently below 10 nT), the first adiabatic invariant was violated during this interval. Thus,

the ion was nonadiabatically accelerated and the energy gain was ~10 keV. The second-stage

acceleration occurred from 04:07:26 UT to 04:10:00 UT (between the two blue vertical dashed

lines), when the ion drifted into a region with a magnetic field greater than 20 nT. Its orbit became

stable and had a regular bounce motion. The first adiabatic invariant was conserved and the ion

was adiabatically accelerated. The scenario of the two-stage process is manifested by the  parameter. According to Büchner and Zelenyi [1989], in magnetotail-like field reversals, the first adiabatic invariant is conserved if   1 because the frequencies of the gyration about magnetic

96

field lines and the bounce motion about the current sheet are well separated. As these two

frequencies approach each other,  1, the interference of gyration and bounce motion breaks

the first adiabatic invariant. Note that the radius of curvature of the magnetic field line is determined by the magnetic field from the MHD simulation, and the energy dependence of the

RqBRcurv curv 1/4 kappa is   W , therefore, this two-stage acceleration process works for i 2mW

a broad energy range of ions having similar trajectories. Such a two-stage acceleration process also

has been reported in studies of electrons in the magnetotail [e.g. Ashour-Abdalla et al., 2013].

Let’s examine the physics of the acceleration mechanism. The electric field was perpendicular to the magnetic field throughout the motion. The ion exchanged energy with the perpendicular electric field. It gained energy during one-half gyration and lost energy during the other half, as implied by the oscillations in the energy and the vector product of the electric field and the ion velocity. Because of this energy exchange, the periodic oscillation in the first adiabatic invariant was significantly reduced by subtracting the energy related to the EB term. The energy

11 associated with the EB drift is calculated as WmumV 22~~1 keV, which is about E×B 22E MHD

one tenth of the ion total energy. The net effect of the energy exchange was that the ion energy

gain exceeded its energy loss as it moved into regions with a stronger magnetic field. This process

is known as betatron acceleration [Northrop, 1963; Baños, 1967; Birn et al., 2013]. As discussed

in Chapter 2, the gain of kinetic energy W can be expressed as:

1 dW M B uE (4.1) qdtGC q t

where uGC is the guiding center velocity, M is the first adiabatic invariant and q is the electric

charge. The energy gain results from two effects. The first is due to the guiding center motion

97

along the electric field direction. The second is caused by the temporal variation of the magnetic field.

5 Rx -10 0 -5 Ry -10 R(Re) -15 Rz -5 -20 40 30 04:07:26 20 0 04:10:00 W(keV)

10 Y(Re) 0 10 8 6 5 4 Kappa 2 0 3.0 10 2.5 subtract 2.0 ExB -5 -10 -15 -20 -25

(keV/nT) 1.5 X(Re) T 1.0

/B no subtract 0.5 10 per 0.0 W 40 Bx 20 By 0 Bz 5 B(nT) -20 Bt -40 8 6 Epar 0 4 Z(Re) 04:10:00 2 Eper E(mv/m) 0 -2 15 -5 04:07:26 10 5 0 -5 -10

E*V(keV/s) -10 -15 -5 -10 -15 -20 -25 hhmm 0408 0410 2009 Feb 07 X(Re) Figure 4.6. Characteristics of a representative test ion. Left column: from top to bottom, the variables are the position vector, the total energy, kappa, the first adiabatic invariant (the green

(black) line is the perpendicular energy over the total magnetic field before (after) subtraction of

energy associated with the EB drift), the magnetic field, the electric field, and the vector product

of the electric field and the ion velocity. The interval between the two vertical dashed blue lines is

04:07:26-04:10:00 UT. Right column: the trajectory of the test ion. The two panels show projection of the orbit onto the X-Y and X-Z planes. The blue solid circles correspond to the initial and final points of the 04:07:26-04:10:00 UT interval.

98

To examine how exactly the ion are accelerated by the electric field associated with the

dipolarization, the trajectory is superimposed on the magnetic field, the ion energy flux and the

electric field shown in Figure 4.4. At 04:06:20 UT, the aforementioned second dipolarization had

already formed near X ~12 RE . The ion launched near the reconnection site was quickly pushed

toward the curved dipolarization path. From 04:06:20 UT to 04:70:20 UT, the ion gained its energy

nonadiabatically because of the weak magnetic field. At 04:07:20 UT, the magnetic field at the

particle location is greater than 20 nT. The ion motion became adiabatic at that time. As the ion

continued to move earthward, at about 04:08:40 UT, it caught up with the dipolarization structure.

The ion was able to catch up with the dipolarization structure because the EB drift along the

flow channel was fast due to the large Ey . This fast drift was in contrast to the slow propagation

of the dipolarization structure. The dipolarization was slow because of plasma compression. The

adiabatic energy gain during this catching-up process from ~04:07:20 UT to ~04:08:40 UT was due to EB drift toward a stronger magnetic field. As noted by Birn et al. [2013],

M uE u B , namely the energy gained by a particle as it moves into a stronger magnetic B q E

field due to EB drift, can be expressed as a gradient drift in an electric field direction. After

catching up with the DF the ion moved with the dipolarization. The ion energy further increased

as the dipolarization continued to intensify. This energy gain was due to the temporal variation of

the magnetic field strength ( Bt 0 ). Finally, as the dipolarization merged with the strong

magnetic field region, the ion gradient drifted duskward. Its energy changed little.

In the previous section, we presented the integrated energy flux of high-energy ions side by side with the magnetic field and electric field while in this section we examined the characteristics of an ion trajectory. Because this ion trajectory is in the region of intense high-

99 energy flux associated with the dipolarization, it represents typical ions that are energized by the dipolarization. Conversely, the high-energy flux increase gives a statistical measure of the energy gained by ions via the dipolarization. Therefore, the acceleration scenario developed by examining the single ion trajectory complements and is consistent with that from inspecting the high-energy flux in the previous section.

4.5. Conclusions and Discussions

Using data from the realistic MHD and ion LSK simulations, we presented a global scenario of ion energization and transport. We found:

1. Most of the high-energy ion flux enhancements are due to nonlocal energization by the

dipolarizations, which are driven by high-speed flows in narrow channels.

2. Ions originating from the reconnection site undergo a two-stage energization process. Not

far from the reconnection site, where the magnetic field is weak, the ions are

nonadiabatically accelerated. Subsequently, they adiabatically gain energy as they catch up

with and ride on the earthward propagating dipolarizations.

3. In the magnetotail, the high-speed flows control ion transport via the EB drift, whereas

close to the Earth, ions gradient and curvature drift toward the dusk side.

Let’s compare these results with previous studies. First, previous analytical studies suggested that ions upstream of DFs can be captured by the fronts and accelerated via trapping

(resonant) and quasi-trapping interaction with the fronts [Artemyev et al., 2012; Ukhorskiy et al.,

2013]. Our simulations, as well as the simulations by Birn et al. [2013], show that most of the high-energy ions are generated in the dipolarized regions, i.e. behind the fronts. They are

100

accelerated during their earthward drift to stronger magnetic field. Second, Birn et al. [2013]

performed test particle simulations in electric and magnetic fields from a generic MHD simulation.

The MHD simulation reproduces the magnetotail reconnection, localized flow bursts and

dipolarizations during substorms. The test particle simulations reproduce a rapid rise of energetic

particle fluxes upon the arrival of dipolarizations. Ions are nonadiabatically accelerated by the

electric fields associated with dipolarizations. Our study complements and extends their results. In

our global MHD simulation driven by realistic upstream solar wind conditions, the flow speed and

width, the magnetic and electric fields, and the dynamics of the system are realistically determined

and event-dependent. In our LSK simulation, ions originating in the flows are first nonadiabatically

accelerated and then adiabatically accelerated as they catch up with and ride on the dipolarizations.

Furthermore, the ions closely follow the flow channels in the tail because of the dominant EB drift; therefore ion transport in the magnetotail is directly controlled by the flows. We have estimated drift velocities for the representative ion and found that in presence of high-speed flows, the EB drift velocity is much larger than the curvature and gradient drift velocities. It appears

that the ions simulated by Birn et al. [2013] drift across the flow channel much faster (see Figure

3 therein). They did not discuss ion collective transport in detail. The differences in the results and

emphases of the results perhaps stem from differences in the methodology. Specifically, the

electric and magnetic fields, which govern the adiabaticity of particle motion and magnitude of

particle drifts, are appreciably different in the two MHD simulations. In addition, our LSK

simulation traces many particles forward in time from a source region near the reconnection site

and collects them with virtual detectors. This is more efficient for tracking particle transport, but

less efficient for finding all the sources of particles than the backward tracing method used by Birn

et al. [2013].

101

The difference between the ion earthward motion and dipolarization propagation has significant implications. First, without this difference, ions would need to ride exactly on dipolarizations in order to gain energy through dipolarization intensification. The difference allows

more ions to gain energy by catching up with the dipolarizations from behind and gain energy as

they drift toward stronger magnetic field regions. The difference is made possible by the large

electric fields carried by the dynamic high-speed flows and plasma compression at the

dipolarization regions. Second, ions can gain more energy by catching up with the dipolarizations

than by riding on them because the catching-up ions originate from the reconnection site in the

magnetotail, where the magnetic field is weaker than in the dipolarizations intensification regions

(usually several RE away from the reconnection site in MHD simulations [El-Alaoui et al., 2013]).

Third, this difference and the dynamic nature of flows require us to launch particles continuously

and uniformly in time so that the fluxes and distribution functions become statistically meaningful.

These measures have been taken in our MHD+LSK simulations.

102

CHAPTER 5

A Comparison Study of Ion and Electron Energization and

Transport Mechanisms during a Substorm

5.1. Introduction

Compared with studies by Delcourt and Sauvaud [1994], Birn et al. [1997b, 1998], Li et al. [1998], Zaharia et al. [2000, 2004] and Birn et al. [2004b, 2013], one major improvement of studies by Ashour-Abdalla et al. [2011], Pan et al. [2014a, 2014b], and Liang et al. [2014] is that we used fields from global MHD modeling in which upstream boundaries are set by event- dependent solar wind measurements. However, there has been no event-study that compares electron and ion acceleration in the magnetotail for the same substorm event, which is of great interest considering the observed differences and similarities in particle flux characteristics (see section 1.3). This chapter aims to apply the global MHD+LSK simulation scheme to the February

07, 2009 substorm. The emphases are the differences and similarities between ions and electrons in terms of the observations, the simulation approaches, and the simulation results. This chapter is organized as follows: the simulation methodology is presented in Section 5.2; the LSK simulation results are presented in Section 5.3; the simulation results are summarized and their implications are discussed in Section 5.4. Note that since the MHD results for the February 07, 2009 event were described by El-Alaoui et al. [2013] and in Chapter 4, we will not repeat the detailed discussions.

In Chapter 4, we have presented the ion LSK simulation results for this event; some of those ion

LSK results are included in this chapter to enable a detailed comparison between the ion and electron results.

103

5.2. Comparisons of the Electron and Ion LSK Simulation Set-up for the February 07, 2009

Substorm Event

We combine a global MHD model, an ion LSK simulation, and an electron LSK simulation

to model magnetospheric dynamics and particle energization for the February 9, 2009 substorm

event. Table 1 provides a summary of the LSK simulations for this event. The electromagnetic

fields used for test particles are from the global MHD simulation [Raeder et al., 1998, 2001; El-

Alaoui et al., 2001; El-Alaoui et al., 2009]. Solar wind obtained from the WIND spacecraft at

(X ,YZ , )GSM (202, 73, 38) R E was used to drive the MHD simulation. The MHD simulation

results were validated and described by El-Alaoui et al. [2013] and in Chapter 4. The LSK code

for ions (protons) solves the Lorentz-force equation of motion by using a fourth-order Runge-Kutta

method. The particle mover in the electron LSK code switches between guiding-center and

Lorentz-force equation. The switch is determined by the adibaticity parameter of electron motion

 Rcurv , where Rcurv is the curvature radius of the magnetic field, and  is the electron gyro

radius [Büchner and Zelenyi, 1989]. Both Rcurv and  are determined locally at the position of the

electron. The thresholds of switch (   7 and   5 ) are based on the study of particle orbit characteristic in the tail-like field reversal [Büchner and Zelenyi, 1989]. Note that if 57 , no switch occurs and electrons are pushed forward in time by using the same method as in their prior step. A detailed description of the electron LSK code is given by Schriver et al. [2011].

For the particle sources in the LSK simulations, we use a distribution that combines a thermal Maxwellian distribution with a high-energy power law tail [Pan et al., 2014a]. The combination of a Maxwellian distribution for thermal electrons and a power law distribution for the high-energy tail is a special approximation of a broad class of generalized Lorentzian (Kappa)

104

distributions [Vasyliunas, 1968; Summers and Thorne, 1991]. This combination was adopted

because it helps handle high-energy electrons. The parameters for the distribution are estimated by

using THEMIS P2 data atX  18.6RE , which was earthward of the magnetic reconnection site

according to the analysis of the observational data by Oka et al. [2011]. Note that the temperature

ratio TTie/ 2 was derived from reconnection outflow plasma measured by P2 near the center of

plasma sheet during this event. To determine launch times of particle sources, we estimate particle

convection time from the launch location to the region of interest. In this event, the convection

time from the reconnection region to the inner magnetosphere (~10 RE ) is ~5 minutes with a flow speed of ~200km/s. The dipolarization of interest was observed at ~04:06:00 UT, therefore we launched particles starting at 03:50:00 UT to ensure sufficient time for convection. The number of source electrons are based on the simulation described in Chapter 3. Note that 3,024 power law source electrons per 20 seconds are sufficient to determine the effect of high-energy electron source on producing the high-energy electron fluxes in the regions of interest. Virtual detectors are placed throughout the simulation domain. In this study, we use the information recorded by the maximum pressure surface detector, which approximates the location of the center of the plasma sheet in the magnetotail [Ashour-Abdalla et al., 2002]. In the LSK simulations, we need to convert information about individual particles recorded by the virtual detectors (e.g. recording time, position, energy, and pitch angle) into collective physical quantities (e.g. energy fluxes and distribution functions). This so-called normalization process has been discussed in Chapter 3.

105

Table 5.1. Set-up of the ion and electron LSK simulations for the February 07, 2009 substorm

Proton Electron

Electromagnetic field A global MHD simulation [El-Alaoui et al., 2013];

Solar wind monitor: WIND spacecraft

Particle mover Lorentz-force equation Switch to guiding-center

equation if   7 ;

Switch to Lorentz-force

equation if   5

[Schriver et al., 2011]

Particle sources An isotropic Maxwellian An isotropic Maxwellian

(Ti  4 keV ) and an isotropic (Te  2 keV ) and an isotropic

power law tail power law tail

( fE( ) E6 , E 25 keV ) ( fE( ) E5 , E 11 keV )

[Pan et al., 2014a, 2014b]

Particle launch location -19 REE XR -17 ; 5 REEYR 10 ; Z  Zmp , Zmp is the

Maximum pressure surface [Ashour-Abdalla et al., 2002]

Launch times Every 20 seconds from 03:50:00 Every 20 seconds from

to 04:20:00 03:50:00 to 04:13:00

Number of source 80,000 Maxwellian protons; 19,440 Maxwellian electrons; particles per 20 seconds 80,000 power law protons 3,024 power law electrons

Diagnostics Detectors on the maximum pressure surface, and planes at fixed X,

Y, Z locations.

106

Normalization CEND ()nddtvA JE(,r , t ) MHD LL eDD  4 i1 NLDDD dAn v, dt dE

[Ashour-Abdalla et al., 1993; Richard et al., 2009]

5.3. Simulation Results

5.3.1. Comparisons of Simulation Results and Observations

In Figure 5.1 we compare our simulation results with THEMIS observations. Because

THEMIS P3, P4 and P5 were within 1.5 RE to each other, they observed similar high-energy flux increases, so only the comparisons for P3 are shown. The Geocentric Solar Magnetospheric (GSM) coordinate system is used. The first four panels summarize the observations. The key observational

features of the dipolarization of interest are the increase of Bz associated with the earthward flow, and the dramatic increase of high-energy ion fluxes (e.g. 37keV-48keV and 77keV-116keV) in contrast to the modest decrease of the low-energy fluxes (e.g. 4.6keV-8.0keV) upon the arrival of the dipolarization. Like the low-energy ion flux, the low-energy electron flux (e.g. 1.94keV-

3.36keV) change is modest. However, the high-energy electron fluxes demonstrate an “anomaly”: the 17.5keV-23.0keV electron flux gradually increases for ~5 min before the dipolarization front arrival, and then increases further after the dipolarization front passes, although the latter increase is less dramatic than the high-energy ion flux increase. The 58keV-73keV electron flux shows enhancement after the dipolarization; there is no high temporal resolution data before 04:05:10

UT. The increase prior to the dipolarization front in the electron flux is likely due to the substorm activity that occurred west of THEMIS P3 location prior to the observation of the dipolarization.

Lyons et al. [2012] showed a series of aurora activities associated with substorm onset at 03:47:12

UT and subsequent substorm expansion west of all the THEMIS spacecraft. THEMIS P3, P4 and

107

P5 did not enter the active region until ~04:02:06 UT, after which they observed the dipolarization

and earthward flows. Two other observed features (not shown) also support our interpretation of

the “anomaly”: (1) the electron flux spectra before the dipolarization are dispersed, namely higher-

energy flux increases earlier than lower-energy flux, suggesting that the flux increase resulted from

energy-dependent gradient and curvature drifts; (2) a weaker change is observed for ion fluxes

prior to the dipolarization, suggesting that after injection from the tail, ions drifted (westward)

farther away from THEMIS P3.

The next four panels of Figure 5.1 show the corresponding simulated quantities at the virtual P3 position in the simulation domain. The dipolarization accompanied by the earthward flow is reproduced by the global MHD simulation. Compared to observations, the simulated dipolarization is weaker, and the magnetic field and flow speed have fewer variations on the shorter time-scale. Two features about the ion fluxes are notable. First, the low-energy (4.6keV-8.0keV) flux changes modestly beginning from 04:00:00 UT, which is consistent with the observed flux.

The simulation does not reproduce the dip right after the dipolarization. This is probably because

compared to observations, the Bz increase is less in the MHD simulation so that the sweeping

effect of dipolarization front is less effective. However, the low-energy flux provides a benchmark for the LSK simulations, suggesting that particles are launched sufficiently early to allow convection to the inner magnetosphere. Second, the simulation reproduces the observed dramatic enhancement of the high-energy ion fluxes as the dipolarization arrives. Similar to that seen in the observational data, the increase of 36keV-46keV flux in the electron LSK simulation from

04:02:00 UT to 04:06:20 UT is due to electron acceleration associated with an earlier

dipolarization on the west of the virtual spacecraft location (see Figure 4.4 in Chapter 4). After

04:06:00 UT, the 36keV-46keV and 58-73keV fluxes increase upon the arrival of the

108

dipolarization front of interest. This increase is less dramatic than that seen in the simulated high-

energy ion fluxes, which is consistent with the observations. We point out that the remarkable difference between the low-energy fluxes and the high-energy fluxes is captured by the simulations,

suggesting that the simulations have correctly preserved the major underling physical process. This process is described in the next section.

109

30 Bx 20 10

By B

(nT) 0 -10 Bz -20 200 Vx 100

Vy V 0

(km/s) -100 Vz )

-1 -200 107 eV 6 4.6-8.0keV

-1 10

sr 5

-2 10 37-48keV 104 cm

-1 3 Ion eflux 10 77-116keV 102 8

(eV s 10 7 1.94-3.36keV 10 6 10 17.5-23.0keV 105 4 Ele eflux 10 58-73keV 103 30 Bx 20 10 By B

(nT) 0 -10 Bz -20 200 Vx 100 Vy V 0

(km/s) -100 Vz )

-1 -200 107 eV 6 4.6-8.0keV

-1 10

sr 5

-2 10 37-48keV 104 cm

-1 3 Ion eflux 10 77-116keV 102 8

(eV s 10 7 1.94-3.36keV 10 6 10 36-46keV 105 4 Ele eflux 10 58-73keV 103 X_gsm(Re) -9.3 -9.3 -9.3 Y_gsm(Re) -1.9 -2.0 -2.1 Z_gsm(Re) -3.4 -3.4 -3.4 hhmm 0400 0405 0410 2009 Feb 07

110

Figure 5.1. Comparisons of observations with simulation results for THEMIS P3. From top to

bottom, the first four panels show the observed magnetic field, plasma velocity, ion differential

energy fluxes, and electron differential energy fluxes from THEMIS P3. For ions, the 4.6keV-

8.0keV energy flux is measured by the Electrostatic Analyzer (ESA) [McFadden et al., 2008], whereas the 37keV-48keV and 77keV-116keV energy fluxes are measured by the Solid State

Telescope (SST) instrument [Angelopoulos, 2008]. The spike in the observed 37keV-48keV and

77keV-116keV ion fluxes in between 04:03 UT and 04:04 UT is due to the attenuator set-up in the

SST detector. For electrons, the 1.94keV-3.36keV and 17.5keV-23.0keV fluxes are obtained by the ESA instrument, and the 58keV-73keV flux is obtained by the SST instrument. There is no burst mode data for the 58keV-73keV electron flux before 04:05:10 UT. The next four panels show the corresponding simulated quantities at the virtual P3 position in the simulation domain. The

vertical lines indicate the arrival of the dipolarization of interest. The ~2 minute difference between

the observed dipolarization and the simulated one is reasonable, see Raeder et al. [1998, 2001],

El-Alaoui et al. [2001, 2009].

5.3.2. Comparisons of Ion and Electron Acceleration Mechanisms

In order to identify the physical processes underlying the observed fluxes, we examine

particle energy fluxes and characteristics of representative particle trajectories. The energy fluxes

give a statistical measure of the particle energy changes, whereas the single particle trajectories

enable us to determine the acceleration mechanisms.

Figure 5.2 shows a snapshot of the MHD and LSK simulations at 04:09:00 UT, in which

the magnetic and electric fields are plotted along with the particle energy fluxes. The first row

111 shows Bz and Ey from the MHD simulation. The second (third) row displays the ion (electron) high-energy and low-energy fluxes from the LSK simulations.

Regarding the energy fluxes, two critical remarks are necessary. First, for each species, the low-energy particles are spread cross the magnetotail and are convected from the reconnection site to the inner magnetosphere along the flow channels, whereas the majority of high-energy particles are present in association with the dipolarization. The high-energy flux concentration at the dipolarization location is a consequence of nonlocal acceleration, rather than convection, because the earthward convection driven by the energy-independent EB× drift would result in the same pattern for the high- and low- energy fluxes, which is not the case.

Second, the convection-dominated low-energy flux pattern is the same for both species, and both high-energy electrons and high-energy ions are associated with the dipolarization, suggesting that the underlying nonlocal energization and transport mechanisms apply to both

species. A notable difference is that in the region inside of X ~10 RE , the high-energy ion fluxes flow toward the dusk side, whereas high-energy electron fluxes flow to the dawn side. This difference is less obvious for the low-energy fluxes. This is because in the inner magnetosphere the energy- and charge-dependent gradient and curvature drifts comprise the major component of particle guiding-center drift (see below).

112

Bz1.0 and flow 040900UT Ey and flow 040900UT -10 -10 6 30 -5 -5 4 20 P3 P3

0 0 2

Y(Re) 10 Bz(nT) Ey(mV/m) 0.8 5 0 5 0

10 -10 10 -2 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 300km/s X(Re) X(Re) Ion eflux(37keV-48keV) Wtest=32.9keV Ion eflux(4.6keV-8.0keV) 6 7 -10 0.6 10 -10 10

105 106 -5 -5 /sr/keV) /sr/keV)

4 2 5 2 P3 10 P3 10

0 0

Y(Re) 103 104

5 5 102 103

0.4 Eflux(keV/s/cm Eflux(keV/s/cm

10 101 10 102 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(Re) X(Re) Ele eflux(36keV-46keV) Wtest=35.8keV Ele eflux(1.94keV-3.36keV) -10 107 -10 108

0.2 106 107 -5 -5 /sr/keV) /sr/keV)

5 2 6 2 P3 10 P3 10

0 0

Y(Re) 104 105

5 5 103 104 Eflux(keV/s/cm Eflux(keV/s/cm

10 0.0 102 10 103 -5 1.0-10 -15 -201.2 -25 1.4 -5 1.6-10 -15 -201.8 -25 2.0 X(Re) X(Re)

113

Figure 5.2. A snapshot of the MHD and LSK simulations for the February 07, 2009 event. First

row: Z-component of the magnetic field ( Bz ) and Y-component of the electric field ( Ey ) from the

MHD simulation. Black arrows represent flow vectors. Second row: energy fluxes of high-energy

(37-48keV) and low-energy (4.6-8.0keV) ions from the ion LSK simulation. Third row: energy fluxes of high-energy (36-46keV) and low-energy (1.94-3.36keV) electrons from the electron LSK simulation. All the quantities are plotted on the maximum pressure surface. The representative ion

(electron) trajectory, shown in Figure 5.3 (Figure 5.4), is superimposed on the fields and high- energy ion (electron) flux diagrams. The white line shows the ion trajectory and the black line shows the electron trajectory. The solid circles represent particle locations at 04:09:00 UT. The

energies of the test particles (Wtest ) are labeled at the top of the high-energy flux plots. The location

of the virtual spacecraft P3 is represented by the black square.

To clearly identify acceleration mechanisms, we examine individual particle trajectories.

Figure 5.3 is a reproduction of Figure 4.3 in Chapter 4, summarizing the characteristics of a

representative ion of a large pitch-angle from the ion LSK simulation. The trajectory of this ion is

also superimposed on the fields and the 37keV-48keV ion flux in Figure 5.2 above (white line).

This ion was launched near the reconnection site at 04:06:20 UT at (X ,YR ) ( 18.3,0.27) E . The

ion first undergoes nonadiabatic acceleration from 04:06:20 UT to 04:07:26 UT (before the first

vertical line) in the weak field region with X  ~ 13RE , where the gyro radius is comparable to

the characteristic scale of the magnetic field variation. It then gains energy adiabatically from

04:07:06 UT to 04:10:00 UT (between the vertical lines) as it drifts earthward toward a stronger

magnetic field region. The adiabatic acceleration mechanism is characterized by the relatively

114 large adiabaticity parameter (  ~4) and the stable first adiabatic invariant ( M ~1keV / nT ).

During the adiabatic acceleration process, the ion exchanges energy with the perpendicular electric field. It gains energy during one-half gyration and loses energy during the other half, as implied by the oscillations in the energy and the vector product of the electric field and the ion velocity.

The net effect of the energy exchange is that the ion energy gain exceeds its energy loss as it moves into regions with a stronger magnetic field. This process is known as betatron acceleration

[Northrop, 1963; Baños, 1967; Birn et al., 2013]. A more detailed description of the trajectory and characteristics of this ion was given in the previous chapter (section 4.4.2.2).

5 Rx -10 0 -5 Ry -10 R(Re) -15 Rz -5 -20 40 30 04:07:26 20 0 04:10:00 W(keV)

10 Y(Re) 0 10 8 6 5 4 Kappa 2 0 3.0 10 2.5 subtract 2.0 ExB -5 -10 -15 -20 -25

(keV/nT) 1.5 X(Re) T 1.0

/B no subtract 0.5 10 per 0.0 W 40 Bx 20 By 0 Bz 5 B(nT) -20 Bt -40 8 6 Epar 0 4 Z(Re) 04:10:00 2 Eper E(mv/m) 0 -2 15 -5 04:07:26 10 5 0 -5 -10

E*V(keV/s) -10 -15 -5 -10 -15 -20 -25 hhmm 0408 0410 2009 Feb 07 X(Re) Figure 5.3. Characteristics of a representative ion. Left column: from top to bottom, the variables are the position vector, the total energy, kappa, the first adiabatic invariant (the green (black) line is the perpendicular energy over the total magnetic field before (after) subtraction of the energy

115 associated with the EB drift), the magnetic field, the electric field, and the vector product of the electric field and the ion velocity. The time interval between the two vertical dashed blue lines is

04:07:26-04:10:00 UT. Right column: projections of the ion trajectory onto the X-Y and X-Z planes. The blue solid circles correspond to the initial and final points of the 04:07:26-04:10:00

UT interval.

Figure 5.4 shows the characteristics of a typical electron from the electron LSK simulation.

Its trajectory is also plotted in Figure 5.2 (black line). In a very limited region at

(X ,YR ) ( 20,10) E before 04:04:20 UT (the first vertical line), the electron motion is nonadiabatic. Its energy increases from a few keV to ~30 keV from 04:04:20 UT to 04:09:00 UT

(between the vertical lines) as its motion quickly becomes adiabatic (  5). The first adiabatic invariant is M ~1.1keV / nT . The electron gains energy from the perpendicular convection electric field. The parallel electric field is negligible. Note that in the electron LSK code, the electron with   5 is pushed according to the guiding-center drift equation and the first invariant does not change. The increase of electron energy is numerically realized by conserving the first invariant. Because the electron undertakes adiabatic motion, the guiding-center drift velocities were conveniently calculated and extracted from the simulation. From 04:04:20 UT to 04:09:00

UT, the X-component of the EB drift velocity is about 200km/s. Because the MHD flow is curved, a substantial Y-component (~100 km/s) is also present. The gradient drift is about 100-

200 km/s for a short period from 04:04:20 UT to 04:06:00 UT. The curvature drift velocity for this quasi-perpendicular electron shows some spikes with peaks of 200 km/s when the electron encounters the current sheet during its bounce motion. It is clear that the EB drift velocity is consistently much larger than the gradient and curvature drift velocities in the tail. Therefore, the

116

electron closely follows the flow channel, as demonstrated by the trajectory in Figure 5.2. The gradient drift later on dominates the electron guiding-center motion in the inner magnetosphere

for this quasi-perpendicular electron. After ~04:11:00 UT, the gradient drift velocity is about 100

km/s insideX  8 RE while the EB and curvature drift velocities are negligible. We have

checked that the curvature drift is comparable to the gradient drift in the inner magnetosphere for

electrons with relatively small pitch angles.

105 Rx -10 0 -5 Ry -10 R(Re) Rz -20-15 40 Wper -5 30 20 Wpar 10 W(keV) W 04:09:00 0 T 30 0 20 y(Re) 10 Kappa 0 4 5 04:04:20

T 3 /B 2 per

W 1 (keV/nT) 0 10 Bx 3040 20 By -5 -10 -15 -20 -25 10 Bz x(Re)

B(nT) 0 B -20-10 T 10 6 4 Epar 2 0 Eper E(mV/m) -2 5 400 Vx 200 0 Vy -200 Vz 0 Veb(km/s) -400 400 Vx z(Re) 04:09:00 200 0 Vy -200 Vz -5 04:04:20

Vgb(km/s) -400 400 Vx 200 0 Vy -200 Vz -10 Vc(km/s) -400 -5 -10 -15 -20 -25 hhmm 0405 0410 0415 2009 Feb 07 x(Re) Figure 5.4. Characteristics of a representative electron. Left column: from top to bottom, the variables are the position vector, the perpendicular, parallel and total energies, kappa, the first adiabatic invariant, the magnetic field, the electric field, the EB drift velocity, the gradient drift

velocity, and the curvature drift velocity. The interval between the two vertical dashed blue lines

117

is 04:04:20-04:09:00 UT. Right column: projections of the electron trajectory onto the X-Y and

X-Z planes. The blue solid circles correspond to the initial and final points of the 04:04:20-

04:09:00 UT interval.

In Chapter 4, we estimated the ion drift velocities and found that the EB drift is the main

component in the high-speed flow channel in the outer magnetosphere, whereas the gradient and

curvature drifts dominate in the inner magnetosphere. Because the ion LSK simulation solves the

Lorentz-force equation of motion rather than the guiding-center equation, it was inconvenient to directly calculate the drift velocities for ions. Here given electron guiding-center velocity calculation, we present a straightforward argument on the relative importance of the ion guiding-

center drifts. As discussed in Chapter 2, the guiding-center velocity (uGC ) for a particle in the

magnetotail can be approximated by:

cEe Mc e mu2 c e e uuuu~~u+ u+11   B 11  (5.1) GC E B c B qB q B s

where u is the guiding-center parallel velocity,uE , uB ,uc are respectively the EB , gradient,

mw2 and curvature drift velocities, q is the electric charge and M  is the first adiabatic invariant, 2B

B e  and s are respectively a unit vector and a unit length along the magnetic field [Northrop, 1 B

1963]. Given the magnetic field, the gradient and curvature drifts are determined by perpendicular

energy and parallel energies respectively. They both depend on charge. Therefore, with the similar

energy for the quasi-perpendicular test particles, the representative ion has gradient and curvature

drift velocities close to those of the representative electron, but in the opposite direction. On the

118

other hand, the EB drift is independent of the particle data. Therefore for the representative ion,

the relative importance of its EB drift compared to its gradient and curvature drifts is the same as that of the representative electron, namely the EB drift dominates in the tail while the

gradient and curvature drifts take over the guiding-center motion in the inner magnetosphere.

The change in the particle kinetic energy (W ) during adiabatic motion can be expressed

as:

1 dW M B uE (5.2) qdtGC q t

The energy gain results from two effects. The first is due to the guiding-center motion along the

electric field direction and the second is due to the temporal variation of the magnetic field

[Northrop, 1963]. The second term is independent of charge. However, the first term does depend

on charge. If both electrons and ions gain energy via a process related to the first term, one expects

them to drift in opposite directions. This is demonstrated by the trajectories superimposed on the

Ey in Figure 5.2. The electron crosses the flow channel from the dusk side to the dawn side, while

the ion does the opposite. A subtle point is that with Ey ~4 6 mV/m, the particle energy changes

by 25.6 keV-38.4 keV if it drifts 1 RE across the flow channel, so to gain tens of keV energy the

required cross-tail drift distance is much smaller in a strong localized electric field than in a weak

large-scale cross-tail electric field. In summary, except for drifting in opposite directions, ions and

electrons with similar energies and pitch angles are not only transported earthward in a similar

way, but also accelerated by the same nonlocal process during transport. This explains the similar

patterns across species seen in Figure 5.2.

119

Because the nonlocal acceleration mechanism depends critically on the adiabaticity of

particle motion, we calculate the kappa parameter on the global scale. The results are shown in

Figure 5.5. The middle (right) column shows kappa for ions/protons (electrons) obtained by fixing

the first adiabatic invariant ( M  1 keV/nT ) for the given magnetic field from the MHD simulation.

This generic first adiabatic invariant value is close to that of the representative ion and electron. It

corresponds to W ~ 30 keV particles in the dipolarized region where Btotal ~ 30 nT . It is very

interesting to note the pattern of kappa. For ions, inside of X ~ 20RE , the dipolarized regions

have larger kappa values than the non-dipolarized regions. This is because: (1) magnetic field in dipolarized regions is large so the gyro radius is small; and (2) the magnetic field is more dipole- like so the radius of field line curvature is large. Were there no dipolarizations, the ion motion

would be nonadiabatic outside of X ~8 RE . Kappa is about 3-4 near the dipolarization region,

confirming that the kappa value derived from the representative ion trajectory is typical. In contrast,

the electrons are adiabatic except in very limited regions close to the reconnection site for this

event. Note that during each cycle of particle bounce motion, kappa reaches its minimum at the

center of the current sheet, thus Figure 5.5 shows the lowest value. In this event ion acceleration

in the second stage becomes adiabatic because of the relative large kappa in the dipolarization

regions, whereas electron motion is adiabatic almost everywhere and not just in the dipolarization

structures.

120

0.8

0.6 Bz and flow 040900UT Kappa [proton: 1keV/nT] Kappa [electon: 1keV/nT] -10 -10 6 -10 6 30 0.4 5 5 -5 -5 -5 20 4 4

0 0 3 0 3 0.2 10 Y(RE) Kappa Kappa Bz(nT) 2 2 5 0 5 5 1 1 0.0 10 1.0 1.2 -10 10 1.4 1.60 10 1.8 2.00 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 -5 -10 -15 -20 -25 X(RE) X(RE) X(RE) 300km/s

Figure 5.5. Kappa in the February 07, 2009 event. Left column: Z-component of the magnetic field from the MHD simulation at 04:09:00 UT for the February 7, 2009 event. Middle column:

Kappa for M  1 keV/nT protons in the magnetic field. Right column: Kappa for M  1 keV/nT electrons in the magnetic field. The black arrows represent flow vectors from the MHD simulation.

All the quantities are presented on the maximum pressure surface.

5.4. Conclusions and Discussions

By using THEMIS observations, a global MHD simulation, and ion and electron LSK simulations for the February 07, 2009 event, we attempted to develop a global scenario for particle energization in the magnetotail. Specifically, the distributions obtained by THEMIS P2 at

X ~ 18.6RE (earthward of the reconnection site) were set as the particle sources in the LSK

simulations, therefore the resultant energization by processes near the reconnection site was

quantified. We found that magnetic reconnection produces thermal particles (a few keV) and high-

energy particles that obey a power law distribution (more than 10 keV). These particles are

accelerated far away from the reconnection site (nonlocally) to tens of keV up to a hundred keV

by perpendicular electric fields associated with earthward propagating dipolarizations and fast

flows in narrow channels. The picture of nonlocal particle energization applies to both electrons

121

and ions. The nonlocal acceleration is adiabatic, which is supported by calculations of the

adiabaticity parameter and the first adiabatic invariant for typical high-energy test particles. The

adiabaticity parameter pattern reveals that electron motion is adiabatic except in very limited

regions near the reconnection site, while ion motion is marginally adiabatic in the dipolarized

regions, where the nonlocal acceleration is most effective. As the ion motion becomes adiabatic,

the nonlocal acceleration process for ions is the same as that for electrons. The charge- and energy-

independent EB drift dominates over the charge- and energy-dependent gradient and curvature

drifts in the flow channels in the magnetotail, therefore particles closely follow the flow channels.

On the other hand, the gradient and curvature drifts are the major component of the guiding-center

motion in the inner magnetosphere (within X ~ 8 RE for the February 07, 2009 event), hence

ions drift toward the dusk side and electrons drift toward the dawn side. The higher the particle energy, the more significant the azimuthal drift.

Our study shows that injections of particles during substorms are highly dependent on dipolarizations and flow channels. First, the local time of the injections is determined by the location of flow channels because acceleration by dipolarizations occurs in narrow flow channels, and because in the magnetotail both low-energy and high-energy particles are transported along

flow channels. Second, dispersionless injections occur in the flow channels because the EB drift is dominant. Outside of the injection regions, particle spectra become dispersed as a consequence of the energy-dependent gradient and curvature drifts. Third, dipolarizations are powerful accelerators due to the effectiveness of adiabatic acceleration, so the magnitude of injections depends on dipolarization strength and flow speed. In large substorms, the dipolarizations are expected to be stronger and the flow speeds are expected to be greater, so the convection electric fields are larger, and the nonlocal particle energization is therefore stronger. These findings are

122

consistent with the observed characteristics of substorm injections. Specifically, Gabrielse et al.

[2014] performed a statistical study using THEMIS data that ranged from within geosynchronous

orbit to X ~ 30RE , and found a clear correlation between injections and azimuthally localized dipolarizations, fast flows and impulsive dawn-dusk electric fields. They also found that with increased geomagnetic activity, the dipolarizations are stronger, the electric fields are larger, the injection occurrence rates are higher, the energy spectra are harder (more high-energy particles), and the aforementioned correlation is better. Injections in fast flow channels are found to be dispersionless.

It is notable that in the MHD simulation, physical processes beyond the MHD temporal and spatial scales are missing, therefore the LSK simulations cannot reproduce the shorter-scale spatial and temporal variations seen in the observed fluxes. Perhaps more importantly, using the global MHD+LSK simulation scheme, we cannot properly model the physical processes near the reconnection region. In the present study, particle sources similar to THEMIS P2 measurement

(closer to the simulated reconnection site) are used to represent the energized distributions near the reconnection site, and then the nonlocal energization from the sources to the inner magnetosphere is examined. It is not clear from the observations how far P2 was from the

reconnection site. The region from the assumed X point/line continuing a few RE away along the

reconnection outflow is important for particle energization [e.g. Speiser, 1965; Hoshino et al.,

2001; Imada et al., 2007]. Nevertheless, the combined Maxwellian and power law sources in the

LSK simulations quantify the overall particle energization due to magnetic reconnection and

processes operating within possibly a few RE to the diffusion region. The question as to which

process(es) the source particles in the LSK simulations should be attributed can be resolved by

using PIC simulations, in which the region from the diffusion region to a few RE away is covered

123 by a self-consistent particle simulation and particle energization can be thoroughly examined. Such a study is the subject of the next chapter.

124

CHAPTER 6

Particle-in-cell (PIC) Simulation of Electron Acceleration by

Magnetic Reconnection

6.1. Introduction

As demonstrated in Chapters 3-5, with the MHD+LSK simulations, particle energization

by magnetic reconnection processes can be better quantified by using distributions measured by

spacecraft closer to the assumed or simulated reconnection site as the sources than using empirical

Maxwellian distributions. In particular, in the study of the March 11, 2008 event described in

Chapter 3, we discovered that adding high-energy electrons that follow a power law distribution

near the reconnection X-line in the LSK simulation is crucial in achieving consistency between

simulated fluxes and those obtained by THEMIS P2 atX ~ 14.7RE . This suggests that the

physical processes occurring near the reconnection site are important for understanding particle

energization in the magnetotail as a whole. However, it was not clear how far the spacecraft was

from the reconnection site during that event, because the measurements showed no signatures of

X-line crossing, e.g. simultaneous reversals of Bz and Vx , or signatures related to Hall

reconnection physics, e.g. quadrupole Hall magnetic field [Øieroset et al., 2001]. As described in

Chapter 1, the region from the assumed X-line to a few RE away in the reconnection outflow is

important for particle energization. The question as to which processes the source particles in the

LSK simulations should be attributed to was not answered by the global MHD+LSK simulations,

which cannot properly model the physical processes close to the X-line because electron physics

and Hall physics rather than macroscopic MHD physics are the major players.

125

In this chapter, the problem of how these power law distributed electrons are generated

near the reconnection site is investigated by using an implicit particle-in-cell (PIC) code. By using

a relatively large simulation domain and following the dipolarization propagation, we separate and

quantify electron energization by the reconnection electric field and by dipolarization. To our

knowledge, there is no study that directly quantifies them both.

6.2. Simulation Methodology

PIC simulations solve the Maxwell-Vlasov equation system self-consistently. A PIC

simulation typically contains three components: (1) interpolating particles to the grid points and

integrating density and current on the grid points with the particle data; (2) calculating electric and

magnetic fields on the grid points by solving Maxwell’s equations given the density and current;

(3) interpolating the fields solved on the grid points to particle locations and pushing particles according to the Lorentz-force equation. Note the Lorentz-force equation for the particle equation

of motion is derived from the Vlasov equation with appropriate interpolation of particles to grid

points and interpolation of fields to particle locations [Birdsall and Langdon, 2004; Lapenta, 2012].

The equations of particle motion are coupled to Maxwell’s equations. They can be decoupled by

using an explicit scheme. That is using particle position and velocity from a prior step to calculate

the fields, and then using the newly calculated fields to update particle data. The procedure is

basically cycling the aforementioned three components in the order of (1)(2)(3)(1).

However the explicit scheme is subject to three severe numerical stability constraints [Hockney and Eastwood, 1988; Birdsall and Longdon, 2004]. First, the explicit differentiation of the

Maxwell’s equations requires that a Courant–Friedrichs–Lewy (CFL) condition must be satisfied on the speed of light, ct x. Second, the explicit discretization of the equations of motion

126

introduces a constraint related to the fastest electron response time,  pet 1 . Third, the

interpolation between grid and particles causes a loss of information and an aliasing instability

called finite grid instability that produces an additional stability constraint x D , where the

proportionality constant  is of order one. This requires the grid spacing to be of the order of the

Debye length D or smaller. To overcome the stability constraints in the explicit numerical

scheme, a fully implicit PIC method has been developed [Langdon et al., 1983; Lapenta, 2012 and

references therein]. In the fully implicit method, the field equations and Lorentz-force equations

are coupled: the source terms (density and current) in the field equations contain particle data, and

the particle equations of motion contain field information. Because the number of particles and the

number of grid points in a PIC simulation are usually very large, it is very expensive to solve the

coupled equation system.

The aforementioned difficulties in the fully implicit PIC method are significantly reduced

by using an implicit moment method [Brackbill and Forslund, 1982; Vu and Brackbill, 1992;

Lapenta et al., 2006; Markidis et al., 2010 and references therein; Lapenta, 2012]. The cycle of

the implicit moment method is as follows: (1) predicting particle positions with present and

(unknown) future fields, and approximating the charge and current density (moments of particles)

using Taylor expansion with the predicted particle positions; (2) solving the fields with the

approximated charge and current density; (3) pushing particles using the newly computed fields.

Compared with the explicit scheme, the major change in the implicit moment method is in step (1):

the charge and current density values are extrapolated using a Taylor expansion and are expressed

in terms of the present and future fields. The implicit moment scheme is linearly unconditionally

stable and the stability constraints of the explicit scheme do not apply to it [Brackbill and Forslund,

1982]. Moreover, it has been shown that the implicit method gives an accurate estimate of the

127

evolution of the system when the particle displacement per time step is smaller than the grid

spacing, vte x, where ve is electron thermal velocity [Brackbill and Forslund, 1985]. The

inequality can be satisfied with large time steps even when the grid spacing is large compared to

the Debye length. It removes the need to resolve small time scales (plasma frequencies) without

eliminating them. The unresolved scales are kept in an approximate way allowing the coupling

with slower scales that are fully resolved by the large time step [Brackbill and Forslund, 1985].

The state-of-the-art iPIC3D code implements the implicit moment method and is used in the

present study [Markidis et al., 2010 and references therein; Lapenta, 2012].

To simulate magnetic reconnection, we assume a 2D Harris equilibrium [Harris, 1962] as

2 the initial condition, with Be00(xz , ) B tanh( z / ) x and nxz(,) n0 sec( h z / ) nb , where

  0.5di is the current sheet width, B0 is the lobe magnetic field, n0 is the density at the center

of the current sheet, and nnb  0.1 0 is the uniform background density representing the level of the

lobe density. The densities are for both electrons and ions. There is an uncertainty in setting the

lobe density relative to the plasma sheet density because in general spacecraft cannot accurately measure the very low energy (a few to tens of eV) electrons due to the photoelectron effect and the relatively high low-energy limits of instruments. In some other PIC simulations, the

background density is set as nnb  0.2 0 [e.g. Hoshino et al., 2001; Pritchett, 2001]. The

temperature is spatially uniform for both species, and the temperature ratio is TTie / 5 . This temperature ratio is larger than that used for the particle sources in the LSK simulations of the

February 07, 2009 event. This is because the LSK source particle temperature is for the reconnection outflow plasmas, whereas the temperature in the present study is for plasmas before

reconnection occurs. The mass ratio is mmie / 400 , which is larger than those used in typical

128

explicit PIC simulations studying electron acceleration by magnetic reconnection [e.g. Hoshino et

al., 2001; Drake et al., 2005; Pritchett, 2006a, 2006b] and dipolarization formation [e.g. Sitnov et

al. 2009; Sitnov and Swisdak, 2011]. The time step size is  pit  0.05 , corresponding to

 pet 1 and cet 0.277 . The time step in the implicit PIC simulation is determined by the physics of interest. Our study requires us to resolve the electron gyro frequency, which is satisfied

by cet 0.277 . The plasma frequency to electron gyro frequency ratio is  pe/7.2 ce  . Using

3 the observed lobe magnetic field of B0 ~20nT and plasma sheet density of ncm0 ~0.2 observed

by THEMIS P2 in the February 07, 2009 event, the ratio is  pe/~7.17 ce . Explicit PIC

simulations typically set this value lower, e.g.  pe/~2 ce [e.g. Drake et al., 2005; Pritchett,

2006a, 2006b] because they require pet ~0.1. The grid size is x zd  0.05 i , corresponding

to x zd1 e . Because of this, the electron diffusion region (EDR) on the scale of several de

is resolved. Electron thermal velocity in terms of the speed of light is vce  0.04 , and electron

2 thermal energy is kTee 0.0016 m c . Explicit PIC simulations require x ~ D  dvee c d e, so

the electron thermal speed is usually set much higher, e.g. vce ~0.2 [e.g. Drake et al., 2005;

Pritchett, 2006a, 2006b]. The simulation domain is (LLx ,zi ) (60,30) d. The system size is

significantly larger than that used in the study of dipolarization formation [e.g. Sitnov et al., 2009;

Sitnov and Swisdak, 2011]. The relatively large domain allows us to examine energization by the dipolarizations before they reach the boundaries, therefore we can differentiate electron acceleration by reconnection and acceleration by dipolarizations. The coordinate system is similar to GSM. Boundaries for particles and fields are open in the X- and Z- directions [Divin et al.,

2007], and they are periodic in the Y-direction.

129

The physical parameters are set to be close to the observed ones, except for the artificial

mass ratio. With the artificial mass ratio, a nontrivial and open question is how to interpret the

simulated quantities in terms of physical quantities. In particular, we are interested in interpreting

the energetics of electrons. The electron acceleration process is strongly coupled with both ion and

electron dynamics [Hoshino et al., 2001]. We feel that with mmie / 400 , it is reasonable to

assume that the electron mass is realistic, whereas the ion mass is smaller than its realistic value.

2 With this assumption, kTee0.0016 m c 0.82 keV , close to the electron temperature

(TkeVe ~ 0.5 1.0 ) observed before reconnection during the February 07, 2009 event. The

reconnection electric field (reconnection rate) is expected to be larger than its realistic value by

about a factor of 2.14 because the reconnection rate normalized by the upstream Alfvén velocity

1/2 ( vmA  ) is approximately independent of mass ratio [Shay et al., 2001]. The energy gain by electrons is therefore larger by about a factor of 2.14 at most, since direct acceleration by the reconnection electric field is only one of a few possible acceleration mechanisms. To compare with observations, we present our results in physical units by setting the lobe magnetic field

3 B0  20nT and plasma sheet density ncm0  0.2 . The lobe magnetic field B0 and the plasma

sheet density n0 are used for normalization, e.g. the Alfvén velocity is calculated as

B0 vA 0.006928c ~2078 km/s. 4 nm0 i

6.3. Simulation Results and Comparisons with Observations

6.3.1 Reconnection Structure

130

The purpose of the present study is to investigate electron energization in the EDR and at the DF. Before proceeding to analyze electron energization, let’s compare the reconnection structure with previous results as a validation of our simulation. Figure 6.1 shows the magnetic reconnection rate. The reconnection rate is normalized by the upstream Alfvén velocity and the reconnection magnetic field [Shay et al., 1999]. The fast reconnection starts at t~2 sec, when the reconnection rate dramatically increases. The reconnection rate reaches its peak at t~2.9 sec. The peak value is about 0.13, which is comparable with previously reported Hall collisionless reconnection rates [Shay et al., 1999]. Subsequently, the reconnection rate gradually drops to below 0.1. Previous studies suggested two possible mechanisms that can cause the rate drop. The first one is that during the fast reconnection, an electron current layer is formed and elongated at the center of the current sheet, throttling the reconnection outflow [Daughton et al., 2006]. The second mechanism is that in the decay stage of the reconnection there are not enough inflow electrons to support the reconnection electric field, so the rate decreases accordingly [Wan and

Lapenta, 2008]. Our simulation shows elongation of the current layer when the rate drops (see

Figure 6.2(h) below). However, to determine the major underlying mechanism that is responsible for the rate drop requires much more detailed analysis, which is beyond the scope of the present study. Interested readers should read the study by Wan and Lapenta [2008], in which they compared different set-up of simulations and the resultant reconnection rates. The electron layer formed in our simulation is unstable to the tearing instability. A secondary island is generated at t~5.4 sec, and the reconnection rate increases accordingly. Formation of magnetic islands due to the unstable electron current layer and its possible effect on reconnection rate have been reported in explicit PIC simulations [Daughton et al., 2006].

131

0.14

0.12

0.1

/c) 0.08 A,up v 0 0.06 (B γ

0.04

0.02

0 0 1 2 3 4 5 6 7 t(sec)

Figure 6.1. Reconnection rate normalized by the upstream Alfvén velocity and magnetic field.

The vertical line corresponds to t=5.40 sec, after which the effect of the magnetic island sets in.

Figure 6.2 shows two snapshots of the fast reconnection at t=3.05 sec and t=5.05 sec. We

select these two times because t =3.05 sec is the time when the DF and the EDR are separated (see

below), and t=5.25 sec is the time just before the magnetic island is produced. During the fast

reconnection, DFs are generated and propagate outwards. The DFs are confined to a region near

the equatorial plane (Z=0). In the magnetotail, the earthward DF is stronger than the tailward DF

due to the background magnetic field. In our simulations, because the initial magnetic field is set

to be uniform in the X-direction, the system is anti-symmetric about the X-line in the X-direction.

Hereafter we focus on the DF that propagates in the positive X-direction. The strength of the DF

132

a) e)

b) f)

c) g)

d) h)

Figure 6.2. Snapshots of the reconnection at t=3.05 sec (a-d) and t=5.25 sec (e-h). (a) and (e):

magnetic field Z-component Bz ; (b) and (f): electric field Y-component E y ; (c) and (g): electron

density; (d) and (h): electron velocity X-component. The black lines on top of the color plots are

contours of magnetic field potential Ay , representing projection of magnetic field lines onto the

XZ plane. The coordinates are relative to the center of the system.

133

is Bz ~ 15nT at X~0.3 Re at t=3.05 sec. It intensifies to Bz ~ 25nT at X~1.03 Re at t=5.25 sec.

The associated convection electric field at the DF is about 30 mV/m at t=3.05 sec and about 60

mV/m at t=5.25 sec. Both values are much larger than the typical observed value (~10-20 mV/m),

including those in the February 07, 2009 event (see Figure 4.2). This is mainly because the Alfvén

velocity in the simulation with lighter ions is larger by a factor of 2.14, therefore the simulated

outflow velocity and convection electric field are larger by a factor of 2.14 than the observed values.

In addition, the lobe (background) density is set to be one tenth of the current sheet density, which

is half of the value used in other studies [e.g. Hoshino et al., 2001; Pritchett, 2001]. The lower

lobe density likely increases the outflow velocity in our simulation by a factor of 2 . The dipolarization pulse sweeps the electrons, resulting in a low density plasma region behind it. The ion density (not shown) demonstrates the same feature as the electron density. As the fast reconnection proceeds, a thin electron current layer is formed and elongated near the X-line (panels

(d) and (h)). The length of the current layer at t=3.05 sec is ~0.16 Re at t=3.05 sec, and is increased to ~0.8 Re at t=5.25 sec. Close to each of the separatrices, there are counter-streaming fast electron beams consisting of an incoming beam toward the X-line in the inflow region and an outgoing beam away from the X-line in the outflow region. Fast electron beams [Hoshino et al., 2001;

Pritchett, 2001] and formation of an electron current layer [Daughton et al., 2006; Fujimoto, 2006] have been respectively demonstrated in explicit PIC simulations. Both features were also presented in the study of reconnection using an implicit PIC code called CELESTE3D (similar to iPIC3D) by Wan and Lapenta [2008].

6.3.2. Electron Acceleration

134

Electron heating and acceleration to suprathermal energies are examined in this section. Figure 6.3

shows the electron temperature and energetic electrons at t=3.05 sec and t=5.25 sec, corresponding

to the snapshots shown in Figure 6.2. At t=3.05 sec, the parallel temperature is Te, ~1 2 keV along the separatrices. From t=3.05 sec to t=5.25 sec, the parallel temperature along the

separatrices is further increased to TkeVe, ~3 4 . Parallel acceleration by electric fields in the

separatrix regions is a well-known aspect of kinetic reconnection [Lapenta et al., 2014]. The

parallel electric field is supported by kinetic effects, including electron density cavities and

pressure anisotropy [Drake et al., 2003, 2005; Egedal et al., 2005, 2012]. As an indicator of the

strength of the parallel electric field, the potential drop along a separatrix for the X-point to the

Boundary system boundary is Eds 22.3 kV ( ds is a length along the separatrix). The potential drop X-point

is much larger than the corresponding parallel temperature. The strong parallel electric field produces fast electron beams [Hoshino et al., 2001; Pritchett, 2001; Lapenta et al., 2010] (also see

Figure 6.2(h) above), which are unstable to streaming instabilities, leading to nonlinear turbulence and electron heating [Goldman et al., 2008]. In contrast to the parallel temperature, the perpendicular temperature is larger in the outflow region behind the DF, indicating that perpendicular heating occurs mainly in the outflow region, although substantial heating also occurs in a thin layer near the X-line. This thin layer is within the electron current layer described above

(Figure 6.2(h)). The maximum of the temperature in the outflow region at t=5.25 sec is

Te, ~4 keV . The thermal electron distribution is isotropic in the perpendicular plane (the other

perpendicular temperature is not shown). From t=3.05 sec to t=5.25 sec, the region of electron heating expands as the reconnection jets propagate outwards. To show production of energetic

electrons, the averaged differential energy flux JEe ( ) at 25.5 keV-51 keV is calculated by using

135

a) e)

b) f)

c) g)

d) h)

Figure 6.3. Electron heating and production of energetic electrons at t=3.05 sec (a-d) and t=5.25 sec (e-h). (a) and (e): electron temperature parallel to the magnetic field; (b) and (f): electron temperature perpendicular to the magnetic field; (c) and (g): averaged differential energy flux

JEe () (including the current sheet electrons and the background electrons) in the energy range of

25.5 keV-51 keV; (d) and (h): differential energy flux of the current sheet electrons in the energy

range of 25.5 keV-51 keV.

136

12E the formula JE() EfE (), where E is the kinetic energy and f ()E is the e 4 m distribution function. Similar to the pattern of perpendicular temperature, the high-energy flux concentrates in the outflow region, representing an extension of the heated distribution to suprathermal energies as a result of continuous production of energetic electrons by reconnection exhaust. However, the high-energy flux peaks at the DF, indicating the DF is a powerful accelerator for energetic electrons. Panels (d) and (h) contain the energy flux contributed only by the current sheet energetic electrons. The current sheet electrons do not enter the diffusion region; they are swept away by the DF. However, they contribute significantly to the high-energy flux at the DF (the reasons are discussed below).

To separately identify energization by reconnection and by dipolarization, we need to identify the

EDR and DF locations. The EDR is defined as the region where the electron flow speed perpendicular to the magnetic field deviates from the perpendicular E×B drift velocity. The DF

location ( X DF ) is defined as the location of the maximum Bz . Figure 6.4 shows two snapshots of

the X-components of the EB drift velocity, the electron velocity, the ion velocity, and Bz at the

center of the current sheet (Z=0). In the center of the current sheet, the X-components of velocities

are perpendicular to the magnetic field, whose X- and Y-components vanish due to anti-symmetry

about Z=0 plane. Therefore the X-components of velocities can help us identify the EDR. At the

beginning of fast reconnection, the dipolarization pulse is embedded in the EDR. As the fast

reconnection proceeds, the DF propagates outward and intensifies. At t=3.05 sec, the DF with

Bz ~15nT at X ~ 0.3RE is well separated from the EDR, whose outer edge is X ~ 0.16RE . At

t=5.25 sec, it propagates to X~1.03 RE and its peak value is Bz ~ 25nT . During the fast

reconnection the EDR is elongated. However, the elongation is not as fast as the DF propagation.

137

Therefore, after t=3.05 sec, the DF is further separated from the EDR. This separation provides an

opportunity to identify acceleration of electrons by the reconnection electric field in the EDR and

acceleration by the DF during its propagation.

v and B (t = 3.05 sec) v and B (t = 5.25 sec) 4 4 x 10 x z x 10 x z 6 30 6 30 v v ExB,x ExB,x v v 4 e,x 20 4 e,x 20 v v i,x i,x B B 2 z 10 2 z 10

0 0 (nT) 0 0 (nT) z z (km/s) (km/s) x x B B v v −2 −10 −2 −10

−4 −20 −4 −20

−6 −30 −6 −30 −2.4 −1.8 −1.2 −0.6 0 0.6 1.2 1.8 2.4 −2.4 −1.8 −1.2 −0.6 0 0.6 1.2 1.8 2.4 x(Re) x(Re)

Figure 6.4. Dipolarization and reconnection outflow at the center of the current sheet (Z=0) at t=3.05 sec (left) and at t=5.25 sec (right). The green line is the X-component of the EB drift velocity, the blue line is the X-component of electron velocity, the black line is the X-component of the ion velocity, and the red line is the Z-component of the magnetic field. The vertical line represents the DF location.

In the following discussions, we will quantify electron acceleration by calculating electron distributions from the simulation. We intend to fit each electron distribution to a Maxwellian distribution or a combination of a Maxwellian distribution with a power law distribution,

depending on whether a substantial high-energy tail is present. The Maxwellian distribution fM with temperature T is defined as:

138

E 1 EE   fM exp   with E [0, ) (6.1) Eth22 EE th  th 

where ETth  2 is the one-dimensional thermal energy. The normalized power law distribution

f p can be defined as [Pan et al., 2014a, also please see Appendix 1 of the thesis]:

1/2n EE1  f p     with EEE[,min max ] (6.2) EabEth  th

where Emin is the lower-energy bound, Emax is the higher-energy bound,  n 1.5 ,

aE min Eth and bE max Eth . We combine the power law distribution with the Maxwellian

distribution, namely

 E  xfM  E [0, Emin ] E Eth f   (6.3) Eth E yf E [ E , E ]  p  min max  Eth

where the coefficients x and y are the weights of the Maxwellian distribution and the power law

distribution, respectively. By requiring the combined distribution and its first-order derivative to be continuous at the lower-energy bound, we have

EEmin  min xfMp yf  (6.4) EEth  th

E n  min (6.5) 2Eth

Finally, the combined distribution satisfies the normalization condition

139

Emin EE xfEth d y1 (6.6) 0 M  EEth th

Solving for the coefficients x and y , the combined distribution is completely determined.

We calculate the electron distributions at the DF by following the DF propagation. Figure

6.5 shows the distributions (data points) at t=0 sec, t=3.05 sec and t=6.61 sec. The electrons are

selected in the region of X DFidXX DF and  4dZee 4 d. Note ddie 20 ~ 0.08 R E. The distributions at t=0 sec and t=3.05 sec are fitted to Maxwellian distributions (the black and the blue

lines). The fitting is done by inspection of trials of different temperatures. The criteria are that the

results need to be approximately evenly spread on both sides of the fitted lines and that the results

near the maximum of the distributions are weighted more heavily because they are expected to

have least uncertainty. The electron temperature increases from initial value TkeVe ~ 0.82 to

TkeVe ~ 2.66 at t=3.05 sec. After the separation of the DF from the EDR at t=3.05 sec, electron

temperature increases a little. However, a high-energy tail is generated from t=3.05 sec to t=6.61

sec. The high-energy tail follows a power law distribution, which is a straight line in the plot with

a log-log scale. Therefore, the distribution at the DF at t=6.61 is fitted to a combined distribution

(the red line). The fitted temperature is TkeVe ~ 2.86 , the lower-energy bound is EkeVmin ~ 14.7 ,

and the power law index is n ~5.15. The power law electrons are about 2.1% of the total electron

population. The high-energy tail generated at the DF is consistent with the feature reflected in

Figure 6.3(g), in which the 25.5 keV-51 keV energy flux peaks at the DF, reaffirming that DFs are

powerful in accelerating electrons to suprathermal energies.

For a comparison, the distributions in a fixed box with  2dXii 2 d and  4dZee 4 d

are also shown in Figure 3. This box approximates the EDR at t=3.05 sec, whose length is Ldei ~ 4

140

Electron distribution (at DF) 1 10

0 10

−1 10

−2 10 )

−1 −3 10 f(keV

−4 10 t=0 sec T =0.82 keV e −5 10 t=3.05 sec T =2.66 keV e −6 10 t=6.61 sec T =2.86 keV e

−7 n=5.15; E>14.7 keV 10 −2 −1 0 1 2 3 10 10 10 10 10 10 E(keV)

Electron distribution (−0.16Re < x < 0.16Re) 1 10

0 10

−1 10

−2 10 )

−1 −3 10 f(keV

−4 10 t=0 sec −5 T =0.82 keV 10 e t=3.05 sec

−6 T =2.04 keV 10 e t=6.61 sec T =4.29 keV −7 e 10 −2 −1 0 1 2 3 10 10 10 10 10 10 E(keV)

141

Figure 6.5. Normalized electron distribution functions at the DF (top) and in the EDR (bottom).

The data points are from the simulation. The solid lines are Maxwellian distributions, except for

the red line in the top panel, which is a combination of a Maxwellian distribution and a power law

distribution. The horizontal arrow points to the lower-energy bound that delineates the power law

distribution from the Maxwellian distribution.

and whose width is ee ~ 8d . At t=3.05 sec, the temperature in the EDR is TkeVe ~ 2.04 , which

is smaller than that at the DF TkeVe ~ 2.66 , suggesting that the increase in temperature at the DF

is not due to temporal effects associated with the non-uniform reconnection rate. More importantly,

we have checked that throughout the simulation, the electron distributions in the EDR are close to

Maxwellian, without substantial high-energy tails. The electrons are heated to TkeVe ~ 4.29 at

t=6.61 sec, resulting from magnetic island formation and contraction after t~5.25 sec.

Three notes are necessary before moving on to compare the power law distribution with

observations. First, we found that the power law tail is gradually enhanced as the DF intensifies and propagates away from the EDR beginning from t=3.05 sec. The tail at t=6.61 sec is the strongest before the DF reaches the boundary of the simulation domain. Second, we do not focus on electron heating by the magnetic island near the initial X-line because extensive studies have reported that magnetic islands can be formed in the reconnection region and accelerate electrons

[e.g. Mattaeus et al., 1984; Drake et al., 2005, 2006; Hoshino, 2012]. Third, we have assumed that electrons are convected with the DF as they are magnetized outside of the EDR, so the electrons

heated in the magnetic island are not expected to enter the distribution at the DF.

142

The power law distribution at the DF qualitatively agrees with observations. The

parameters derived from the electron distribution observed by THEMIS P2 at X  18.6RE in the

February 07, 2009 substorm event are TkeVe ~ 2 , EkeVmin ~11 and n ~5.5 [Pan et al., 2015a].

In this event, P2 was in the earthward reconnection outflow region, and these parameters were calculated with a DF passing the spacecraft. The parameters of electron distribution in the

reconnection outflow at X ~ 20RE inferred from a global MHD+LSK simulation during the

March 11, 2008 substorm event are TkeVe ~ 2 , Emin ~9 keV , and n ~4.5 [Pan et al., 2014a].

The simulated distribution at the DF is also comparable with that observed by Cluster spacecraft

at X ~ 16.4RE in the reconnection downstream region with a DF passing [Imada et al., 2007].

Moreover, the Cluster observations demonstrated that the electron spectrum is harder at the DF than that close to the X-line, in agreement with the simulation. The simulated distributions are not verifiable with observations from WIND spacecraft in the distant tail-reconnection region reported by Øieroset et al. [2002], because no signatures of dipolarization were observed there [Øieroset et al., 2001].

To identify the heating and acceleration mechanisms, we calculate electron velocity distributions. The results are shown in Figure 6.6. As in Figure 6.5, the distributions at the DF are

calculated from electrons in X DFidXX DF and  4dZee 4 d , and the electrons in the

EDR are from  2dXii 2 d and  4dZee 4 d. At the DF, electrons are energized predominantly in the perpendicular direction (panel (a)). The distribution is isotropic in the perpendicular plane (panel (b)). Note we have also checked the distribution at the DF as a function

of perpendicular energy (f (E ) ) and as a function of parallel energy ( f ()E ), and found that the acceleration is in the perpendicular direction. In the EDR, electrons are accelerated in the X-

143

direction and ejected outward, forming a thin current layer (panel (c)). Electrons are also clearly

accelerated and heated by the reconnection electric field in the negative Y-direction (panel (d)).

The acute readers might find it puzzling that the electron distribution in the EDR at t=5.25 sec can

be fitted to a Maxwellian distribution as a function of energy, even though the velocity distributions

clearly demonstrate non-Maxwellian features. However, the key observation is that the velocity

distribution f (,vvx y ) is close to a half-Maxwellian distribution, namely

fv(,xy v 0)~(, f Mxy v v 0) and fv(,xy v 0)~0, where fM (,vvxy ) represents the Maxwellian

distribution.

(a) (c)

(b) (d)

Figure 6.6. Electron velocity distributions at the DF (a-b) and in the EDR (c-d) at t=5.25 sec.

144

When analyzing 25.5 keV-51 keV electrons shown in Figure 6.3 (panels (c-d) and (g-h)),

we found that the current sheet electrons contribute significantly to the energy flux at the DF, even

though they do not enter the diffusion region. We now discus the reason. We now discus the reason.

The densities of electrons and ions from the current sheet and background at Z=0 at t=5.25 sec are

shown in Figure 6.7. As expected, the background density peaks of both species are behind the DF

whereas the current sheet density peaks are ahead of the DF because of the sweeping and pileup

effect. However, two features are subtle and important. First, the shape of the electron background

density is similar to that of the DF, indicating the background electrons are convected with the DF and may be adiabatically accelerated as the DF intensifies. Second, the current sheet particles penetrate the DF. The electrons that penetrate into the slowly varying DF (on the ion time scale) are expected to be accelerated very efficiently because their first adiabatic invariant is expected to be conserved. The limited penetration is due to kinetic effect (finite-Larmor-radius effect and/or inertia effect), as suggested by the different penetration depths for electrons and ions. These pieces of evidence, together with the information that the electrons are accelerated predominated in the perpendicular direction, suggest that the acceleration mechanism operating at the DF is betatron acceleration. The betatron effect in the present study seems to be different from that discussed in

Chapter 2 (also see Pan et al. [2012]). There both the initial distributions and accelerated distributions were power law distributions, and the acceleration occurred in the region from the tail to the inner magnetosphere. The result of betatron acceleration was uniformly shifting the distribution to a higher-energy, without significantly changing the shape of distributions. In the present PIC simulation, Maxwellian initial distributions are applied. Observations have shown that the plasma sheet prior to reconnection has significant high-energy electrons [e.g. Deng et al., 2010].

Future studies of electron acceleration by magnetic reconnection should add high-energy electrons

145

to the initial distributions, in a same approach as for the source particles in the LSK simulations.

It is likely that the electron distributions as functions of energy will no longer be Maxwellian in

the EDR and that the difference between distributions in the EDR and at the DF is quantitative rather than qualitative. Another possible reason causing the difference in betatron effect is that some of the electrons immediately downstream of the outflow may encounter nonadiabatic orbits, as demonstrated by Hoshino et al. [2001]. Nonadiabatic and irregular orbits can prevent electrons from convecting with the magnetic field, leading to mixing of different populations during DF formation. Without electron orbit information, it is difficult to pinpoint the significance of the nonadiabatic effect. Nevertheless, with the evidence that electrons are predominantly accelerated in the perpendicular direction, that the background electrons are convected with the magnetic field,

and that the current sheet electrons penetrate into the DF, we feel betatron acceleration is a major

contributor to the energetic electrons in the outflow and at the DF in particular.

n and B (t = 5.25 sec) z 0.6 30 n e,c n i,c 0.5 n 25 e,b n i,b B 0.4 z 20

) 0.3 15 −3 (nT) z B n (cm 0.2 10

0.1 5

0 0

−0.1 −5 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 x(Re)

146

Figure 6.7. Densities of electrons and ions from the current sheet ( ne,c and ni,c ) and background

( ne,b and ni,b ) at the center of the current sheet (Z=0) at t=5.25 sec.

6.4. Conclusions

In this chapter, we used the implicit PIC code to simulate 2D magnetic reconnection with

physical parameters that are close to the observed values in the magnetotail. The simulation results

were validated against satellite measurements and previous explicit PIC simulations. Specifically,

the DF strength (~15-25 nT within X~1.03 Re) is comparable with observations. The associated

electric field (~30-60 mV/m) is much larger than the observed value (~10-20 mV/m) because of

lighter ion mass is applied in the simulation. We found that electrons are heated to TkeVe, ~2 4 in the parallel direction along the separatrix, similar to previous results. Electrons are heated to

TkeVe, ~2 4 in the perpendicular direction in the reconnection exhaust, with the temperature

maximum located behind the front. The energetic electrons (25.5 keV-51 keV) are generated in

the region where electrons are heated in the perpendicular direction. However, the peak flux of

energetic electrons occurs at the DF. The electron distribution function as a function of energy in

the EDR is Maxwellian. In contrast, in addition to a thermal Maxwellian component with

TkeVe ~ 2.86 , the electron distribution at the DF (~1.7 Re away from the X-line) at t=6.61 sec has

a substantial high-energy tail, with E14.7 keV and n ~5.15. These characteristic values are

comparable to those observed by THEMIS P2 (X   18.6RE ) in the reconnection outflow region with a DF passing in the February 07, 2009 substorm event, and those in the outflow region

(~20X  RE ) inferred from the global MHD+LSK simulation of the March 11, 2008 substorm

147 event. The simulated power law distribution is also comparable with that from the Cluster observations reported by Imada et al. [2007]. Furthermore, the simulation reproduced the Cluster observed feature that electron spectrum at the DF in the downstream region is harder than that in the diffusion region. The power law tail in our simulation is caused by acceleration associated with the DF rather than by acceleration in the EDR because it is generated after the DF separates from the EDR. To our knowledge, there has been no simulation study which directly quantifies both the acceleration in the EDR and the acceleration associated with the DF. The energization mechanism operating in the EDR is the reconnection electric field acceleration, and that in the outflow and at

DF is likely betatron acceleration.

148

CHAPTER 7

Conclusions and Problems for the Future

7.1. Conclusions

We have studied in this dissertation the problem of energization of particles (both electrons

and ions) to tens and hundreds of keV and the associated transport process in the magnetotail

during substorms.

We developed a simple analytical model to quantify nonlocal adiabatic acceleration effects

[Pan et al., 2012]. In this model, given the particle differential flux at the presumed source in the tail, the flux in the inner magnetosphere is calculated by assuming conservation of the first and second adiabatic invariants. The model is characterized by two parameters, the magnetic field compression factor accounting for the enhancement in the perpendicular flux due to betatron acceleration and the contraction factor accounting for the enhancement in the parallel flux due to

Fermi acceleration. The model shows that the adiabatic enhancement of flux strongly depends on the slope of the source flux. For example, at high energies, a large increase in perpendicular

(parallel) flux can be induced by a small compression (contraction) factor given a steep source

spectrum. When compared with THEMIS data, the model worked surprisingly well in predicting

flux enhancement of high-energy electrons from the tail to the inner magnetosphere, despite the

fact that betatron and Fermi acceleration were characterized by just two parameters. In retrospect,

the main reason for this was that high-speed flows were present in the events analyzed, which

confined the electrons to narrow flow channels and carried electrons of all energies earthward via

the charge- and energy-independent EB drift. As reviewed in Chapters 1-2, the theory of

149 adiabatic particle motion has been developed since 1940s [Alfvén, 1950; Northrop, 1963; Banõs,

1967] and has been extensively applied to the magnetosphere [e.g. Tverskoy, 1969; Sharber and

Heikkila, 1972; Kivelson et al., 1973; Lyons, 1984; Artemyev et al., 2011; Ashour-Abdalla et al.,

2011; Birn et al., 2012 and references therein]. In particular, it was demonstrated that by applying conservation of the first and second adiabatic invariants in an analytical model of the magnetic field, particle distributions at different locations can be mapped from the distribution at a given position [Tverskoy, 1969; Lyons, 1984; Artemyev et al., 2011; Birn et al., 2012]. The success of our two-parameter model, and the factor that differentiate it from previous models is that our model applies to cases in which the initial and final distributions (fluxes) are in the same high-speed flow channel.

The analytical model is rather restrictive because the particles are assumed to be uniformly accelerated. A more powerful approach is to combine a LSK simulation with a global MHD simulation driven by measured upstream solar wind. The approach is much less restrictive because in the LSK simulation trajectories of millions of particles are calculated in the realistically determined MHD fields and the results are statistical. The MHD+LSK method was applied to model electron acceleration and transport in the March 11, 2008 substorm event [Pan et al., 2014a].

We found that adding a high-energy tail of electrons obeying a power law distribution in the LSK electron source was crucial in obtaining consistency between the simulated energetic electron flux

and that measured by THEMIS P2 at X ~ 14.7RE , suggesting that the power law distributed electrons were produced from the source region, which was close to reconnection in the MHD simulation. Meanwhile, the LSK simulation reproduced the flux enhancement from P2 in the tail to P3/P4 in the inner magnetosphere. This flux increase was by about an order of magnitude for the 41-95 keV energy channel. The enhancement occurred in the dipolarization region where the

150

magnetic field was compressed by about a factor of 2. The enhancement was uniform for energies

greater than ~ 1 keV, suggesting that the nonlocal acceleration process was adiabatic. The adiabatic

acceleration mechanism was further confirmed by characteristics of electrons trajectories, which

showed that for typical tens of keV electrons the adiabaticity parameter ( ) was much greater than

unity (  1) in the flow channels and the first adiabatic invariant was approximately conserved.

This study provided convincing quantitative evidence of local acceleration and nonlocal

acceleration in a substorm event. In addition, electron transport in the magnetotail was found to be

determined by high-speed flows generated by magnetic reconnection. The EB drift was

statistically dominant even for tens of keV electrons. Electrons gradient and curvature drifted towards the dawn side in the inner magnetosphere. Ashour-Abdalla et al. [2011] applied the

MHD+LSK simulations to study electron acceleration during a substorm event that occurred on

February 15, 2008. They found that nonlocal betatron acceleration was the major mechanism that

produced energetic electrons (approaching 100 keV) observed by THEMIS P4 in the inner magnetosphere in that substorm. In the February 15, 2008 event, there was no observation in the

tail that could be used to quantify the presence of energetic electrons closer to the reconnection

region. Ashour-Abdalla et al. [2011] used a Maxwellian source in the LSK simulation, which may

have underestimated the flux of energetic electrons resulting from reconnection. Our study

extended their work by utilizing simultaneous measurements in the tail and in the inner

magnetosphere and adding power law electrons to the source in the LSK simulation. As a result,

both local acceleration and nonlocal acceleration were quantified in a substorm event.

Having used the MHD+LSK simulations and THEMIS spacecraft measurements to study

electron energization in the magnetotail, we applied them to examine ion energization associated

with magnetic dipolarization during the weak February 07, 2009 substorm event [Pan et al.,

151

2014b]. We found that the major high-energy ion flux enhancements observed in the inner magnetosphere were due to nonlocal acceleration by the dipolarizations and high-speed flows. Ions originating from the reconnection site underwent a two-stage energization process. Not far from the reconnection region, where the magnetic field was weak, the ions were nonadiabatically accelerated. Subsequently, they adiabatically gained energy as they caught up with and rode on the earthward propagating dipolarization structures. They could catch up with the dipolarizations because of plasma compression at the dipolarization regions. For ion transport, we found that in

the magnetotail, the high-speed flows controlled ion transport via the EB drift, whereas close to

the Earth, ions gradient and curvature drifted towards the dusk side. The ion acceleration scenario

is similar to the electron acceleration scenario presented by Pan et al. [2014a], but it is significantly different from the scenario of nonadiabatic acceleration in the wall region with  ~1 [e.g. Ashour-

Abdalla et al. 1992b, 1992c, 2009; Zhou et al., 2011], even though in both scenarios, ions gain

energy from the perpendicular electric field in the Y-direction. Dipolarizations and high-speed

flows in narrow channels are critical for ion acceleration in the present study, while they are not

required for the acceleration in the wall region. Therefore, it will be interesting to examine ion

energization in events with different characteristics of flows and dipolarizations (see the discussion

in the next section). The acceleration of ions when they drift across high-speed flow channels was

also discussed by Birn et al. [2013]. However, compared to that study, the ion motion in our study

is much more adiabatic. Moreover, in our LSK simulation, ions originating in the flows closely

follow the flow channels in the tail while it appears that the ions simulated by Birn et al. [2013]

drift across the flow channels much faster (see Figure 3 therein). These differences probably stem

from the differences in the electric and magnetic fields. In our global MHD simulation driven by

realistic upstream solar wind conditions, the flow speed and width, the magnetic and electric fields,

152

and the dynamics of the system are realistically determined and event-dependent, while their MHD simulation has generic characteristics of neutral line formation and dipolarizations in the magnetotail (see more discussion in the next section).

We extended the study of the February 07, 2009 event to electron energization and transport, and compared the global energization and transport mechanisms between electrons and ions [Pan et al., 2015a]. We found that thermal ions and electrons (a few keV) observed at the dipolarizations originated from a relatively wide region of the tail near the reconnection site and were convected to the inner magnetosphere. High-energy particles (tens of keV up to ~100 keV) were produced by the perpendicular electric fields associated with the dipolarizations and accompanying high- speed flows. The particle trajectories showed that electrons that originated from the reconnection site were adiabatically accelerated during earthward transport, and, surprisingly, ions were accelerated in a manner similar to that of electrons. However, the high-energy electron motion was adiabatic except in very limited regions near the reconnection while high-energy ion motion was marginally adiabatic in the dipolarization regions. Different from the aforementioned study by Birn et al. [2013], we showed that both electrons and ions closely follow the flow channels in the tail due to the dominant EB drift. To our knowledge, this was the first study that compared global electron and ion energization in the same substorm event.

To understand the power law electrons observed by THEMIS P2 in the tail and inferred from the MHD+LSK simulations, we applied the state-of-the-art implicit PIC code [Brackbill and

Forslund, 1982; Vu and Brackbill, 1992; Lapenta et al., 2006; Markidis et al., 2010 and references therein; Lapenta, 2012] to examine electron acceleration in the reconnection region [Pan et al.,

2015b]. We were able to extend previous simulations of magnetic reconnection by setting a relatively large simulation domain and using realistic physical parameters. The large simulation

153

domain allowed us to examine electron acceleration by the reconnection electric field and

acceleration associated with dipolarization pulses in the outflow region [e.g. Sitnov et al., 2009;

Sitnov and Swisdak, 2011]. Previous studies of electron acceleration using explicit PIC

simulations typically chose the proton-to-electron mass ratio as mmie / ~ 100 , the plasma

frequency to electron gyro frequency ratio as  pe/~2 ce , and the electron thermal speed in terms

of the speed of light as vce / ~ 0.2 [e.g. Hoshino et al., 2001; Drake et al., 2005; Pritchett, 2006a,

2006b]. In our simulations, these parameters were improved to mmie / 400 ,  pe/7.2 ce  ,

vce / 0.04 . Note that the realistic values are mmie / 1836 ,  pe/~7.17 ce , vce / ~ 0.04. With

the larger mass ratio and other realistic parameters, the problem of interpreting electron energies

was significantly reduced, hence direct comparisons of simulation results with observations in the

same physical units were made possible. We found that in the reconnection region, electrons with

uniform initial temperature (TkeVe ~ 0.82 ) are heated to Te, ~2 4 keV in the parallel direction

along the separatrices by strong parallel electric fields. Electrons are heated to TkeVe, ~2 4 in

the perpendicular direction by the reconnection electric field in the EDR and by betatron

acceleration in the outflow region. The energy flux of 25.5-51 keV electrons peaks at the DF due

to betatron acceleration of electrons in the reconnection exhaust and penetration of current sheet

electrons into the DF. The electron distribution function in the EDR as a function of energy is

Maxwellian whereas at the DF it has a high-energy tail ( n ~4 6, EkeVmin ~10 20 ) in addition

to a thermal component (TkeVe ~ 2 4 ). These characteristic values are comparable to those

observed by THEMIS P2 (X  18.6RE ) in the reconnection outflow region with a DF passing in

the February 07, 2009 substorm event [Pan et al., 2015a], and those in the outflow region

(~20X  RE ) inferred from the global MHD+LSK simulation of the March 11, 2008 substorm

154 event [Pan et al., 2014a]. The simulated power law distribution is also comparable with that from the Cluster observations reported by Imada et al. [2007]. As in the PIC simulation, Imada et al.

[2007] reported that Cluster observed a harder electron spectrum at the DF than in the diffusion region. The high-energy tail results from acceleration associated with the dipolarization pulse

(~1 2RE away from the X-line) rather than acceleration in the EDR because it is generated after the DF separates from the EDR. To our knowledge, there has been no simulation study which simultaneously quantifies the acceleration in the EDR and the acceleration associated with the DF.

To summarize, we have developed a complete model of particle multi-step energization on multiple scales in the magnetotail during substorms, including acceleration localized near the reconnection site and acceleration during plasma earthward transport. Figure 7.1 shows a schematic of the normalized particle distribution functions produced in this model:

(a) In the reconnection region, particles (electrons in particular) are heated to a few keV in

the parallel direction along the separatrices by reconnection parallel electric fields.

Electrons are heated to a few keV in the perpendicular direction by the reconnection

electric field in the EDR and by betatron acceleration in the outflow region. The

electron distribution function in the EDR as a function of energy is Maxwellian

(~24TkeVe  ) whereas at the DF it has a high-energy tail ( n ~4 6 ,

EkeVmin ~10 20 ) in addition to a thermal component (TkeVe ~ 2 4 ) [Pan et al.,

21015b]. The DF and the high-energy tail are observed by a spacecraft close to the X-

line, e.g. THEMIS P2.

(b) The keV thermal particles and tens of keV power law distributed high-energy particles

produced by processes near the reconnection region are further nonlocally accelerated

155

to tens of keV up to about a hundred keV by perpendicular electric fields associated

with the dipolarizations and reconnection jets as they propagate towards the Earth in

narrow channels [Ashour-Abdalla et al., 2011; Pan et al., 2014a, 2014b, 2015a; Liang

et al., 2014; Pan et al., 2015a]. The resultant energetic electrons are observed by a

spacecraft in the inner magnetosphere, e.g. THEMIS P4. The jet plasmas are hotter than

preexisting plasma sheet plasmas and their distributions are preferably increased at a

few tens of keV (see Figure 6 by Deng et al. [2010] and Figure 8 and Figure 9 by Birn

et al. [2014]), so a spacecraft would observe an increase of power law index (absolute

value) for energetic particles (e.g. EkeV 25 ) during a DF pass [Pan et al., 2012]. The

nonlocal acceleration is adiabatic for both electrons and ions, with a caveat that the

high-energy electron motion is adiabatic in a much wider region in the tail than that of

the ions [Pan et al., 2014b, 2015a]. The power law index for energetic particles under

adiabatic acceleration from the tail to the inner magnetosphere remains relatively

unchanged [Pan et al., 2012].

156

P2 before DF (a) (b) EDR P2 at DF P2 at DF P4 at DF f(E) f(E)

Emin

E E0 E

Figure 7.1. Distribution functions produced in the model of particle multistep energization on

multiple scales in the magnetotail during substorms. (a) Particle (electron) acceleration near the

reconnection region. The electron distribution in the EDR is Maxwellian, whereas the electron

distribution at the DF in the immediate downstream of the outflow is hotter and has a high-energy

tail. (b) Particle acceleration during transport. Close to the X-line, the electron distribution in the

preexisting plasma sheet has a high-energy tail. The distribution at the DF is hotter and preferably

enhanced at a few tens of keV. The particles at the DF are adiabatically accelerated as the DF

propagates to the inner magnetosphere. The power law index above EkeV0 ~25 (absolute value)

increases as the DF passes a spacecraft, whereas at the DF it does not change significantly from the tail to the inner magnetosphere. Part (a) is derived from the THEMIS P2 measurements and the PIC simulations in Chapter 6, and part (b) is derived from the analytical model, the THEMIS measurements, and the MHD+LSK simulations presented in Chapters 2-5.

157

The acceleration mechanism by early-stage dipolarizations ( ~ 1 2RE to the X-line) in our

PIC simulations is similar to that by late-stage dipolarizations (a few to more than 10RE to the X- line) in our global MHD+LSK simulations. However, the dipolarizations in the PIC simulation are different from those in the MHD simulations. The former dipolarizations are intensified to

Bz ~25nT within ~ 2RE from the X-line, and the dipolarizations are more pulse-like (a few to

tens of seconds long). Pulse-like dipolarizations are also produced in explicit PIC simulations

[Sitnov et al., 2009; Sitnov and Swisdak, 2011]. The dipolarizations in the MHD simulations,

however, are formed in the region close to the Earth. They are intensified as the plasmas carried

by the flows are compressed against the strong near-Earth magnetic field. They are more

sustainable (a few to tens of minutes long) and less pulse-like. In addition to the MHD simulations

in the two events presented in this dissertation, other global MHD simulations also reproduce

dipolarizations resulting from plasma compression and flow braking [Birn et al., 2004, 2011; El-

Alaoui et al., 2012, 2013]. In observations, multiple dipolarizations separated by tens of seconds

to a few minutes have been observed near the reconnection region [Fu et al., 2013] and in the inner

magnetosphere [Zhou et al., 2009]. In fact multiple dipolarization pulses proceeding a sustained

dipolarization were presented for the March 11, 2008 event in Chapter 3, see Figure 3.3-3.4. In

that event, the interval of the dipolarization pulses at P4 (X ~ 10.4RE ) was shorter that that at P2

(X ~ 14.7RE ), indicating the dipolarization pulse region was compressed as the flow speed was

slowed. Combining these pieces of evidence, it seems that dipolarization pulses are generated near

the reconnection region and pile up in the inner magnetosphere as they propagate towards the Earth.

The PIC simulations demonstrated the production of dipolarization pulses, whereas the MHD

simulations showed the plasma compression and the magnetic field pileup. This interpretation

seems to be consistent with a recent 2D multi-scale study of reconnection and dipolarizations by

158

Ashour-Abdalla et al. [2015], which showed that a chain of dipolarization pulses are generated by

the unsteady reconnection at X ~ 32RE . The dipolarization pulses propagate earthward, and

merge atX ~ (15 20)RE . The dipolarizations are generated more frequently in the simulation

(one dipolarization in every two seconds) than in the aforementioned observations described in

Chapter 3, Zhou et al. [2009] and Fu et al. [2013]. Despite these indicative evidence, the possible connection between these two kinds of dipolarizations remains to be determined by future large- scale fully kinetic simulations. Therefore, connecting particle energization near the reconnection region demonstrated by the PIC simulations (part (a) of Figure 7.1) and nonlocal energization demonstrated by the global MHD+LSK simulations (part (b) of Figure 7.1) is not as straightforward as it appears.

7.2. Unsolved Problems and Future Work

The results that we have presented in this dissertation suggest directions of future work.

Having studied particle energization in a few substorm events, one can ask a natural question: how

do particle energization and transport depend on particular substorm events? Is the model of

particle multistep energization on multiple scales in the magnetotail still valid in large substorms?

This is an interesting question because of three reasons. First, the substorm events studied in this

dissertation are weak and moderate, reflecting the solar minimum effect during the THEMIS

mission. As we pointed out in Chapter 4 and commented above, a significant difference between

our study of the February 07, 2009 event and the study by Birn et al. [2013], who used idealized boundary conditions for their MHD simulation, is that the MHD fields are appreciably different.

The different MHD fields cause significant differences in ion trajectories. In our study, the ions

159

follow the flow channels closely, whereas it appears that the ions simulated by Birn et al. [2013]

drift across the flow channels much faster. Moreover, the motion of ions of tens of keV energy is

marginally adiabatic in the dipolarization regions in our study, whereas it is much less adiabatic in

their study. The nonlocal adiabatic acceleration picture depends on the adiabaticity of particle

motion, which depends on the electromagnetic fields. Second, Liang et al. [2014] examined the

nonlocal acceleration mechanism in two substorm events that occurred on February 15, 2008 and

August 15, 2001. In these two events, the substorm magnitudes were approximately the same---

AE index peaks were ~300 nT and ~200 nT respectively (moderate substorms)---but the magnetotail configuration and electron acceleration were very different. During the February 15,

2008 event, the dayside reconnection occurred nearer the subsolar point, and the high-speed flows in narrow channels produced by azimuthally localized tail reconnection swept the electron sources.

The electrons were adiabatically accelerated, and the electron distribution was pancake-like

( f ()vfv  () ) in the inner magnetosphere. In contrast, during the August 15, 2001 event, the

dayside reconnection occurred on the flanks of magnetopause. An X-line extending across the tail

was formed and the earthward flows were slow. The electrons were nonadiabatically accelerated

in the weak field region close to the X-line, resulting in cigar-like electron distributions

( f ()vfv  ( )). Third, in Chapter 5 we suggested that characteristics of injections of energetic

particles during substorms are determined by dipolarizations and flow channels through three

factors: the locality of flow channels, the relative importance of EB drift compared to

gradient/curvature drifts, and the adiabatic acceleration efficiency. In larger substorms, the

dipolarizations are stronger and flow speeds are greater [e.g. Gabrielse et al., 2014]. Therefore, the convection electric fields are larger and the adiabatic acceleration is stronger. However, the dependence of dipolarization and flow characteristics (e.g. the width of the flow channels) on

160

geomagnetic activity level is not well established. Considering these three reasons, we feel that it

is important to examine our model in many more events. For example, by comparing weak

substorm events with strong substorm events, we can better understand the robustness of the nonlocal adiabatic acceleration mechanism (especially for ions), and the consequences on the

substorm injections.

Future studies should further investigate the differences and similarities between electron

and ion acceleration in the magnetotail. In this dissertation, we have presented a comparison study

of electron energization with ion energization for the February 07, 2009 substorm event. In this

substorm, ion and electron fluxes demonstrated similar features, e.g. the simultaneous increase of

high-energy fluxes and decrease of low-energy fluxes for both species upon the arrival of the

dipolarization. The ion energization was stronger than the electron’s for this event. However, there

are events in which electron energization is stronger, e.g. the February 15, 2008 event studied by

Zhou et al. [2009] and Ashour-Abdalla et al. [2011]. In both studies, ion fluxes were not shown

because no appreciable ion energization was associated with the dipolarization in the event. A six-

case study by Runov et al. [2011] indicated that both electron and ion high-energy fluxes increase

upon a DF arrival; ion high-energy flux increase is slower and less dramatic than that for electrons

(see Figure 5 therein). However, the limited number of events compromised this conclusion.

Previous statistical studies of many events investigated high-energy electron fluxes associated

dipolarizations by using data from Cluster [Fu et al., 2012c] and THEMIS [Wu et al., 2013], but did not investigate ion fluxes. Therefore, extending the statistical studies to ion fluxes and comparing them with electron fluxes are needed to quantify the similarities and differences between electron energization and ion energization. To better understand the differences and similarities across species, MHD+LSK simulations should be employed to study events in different

161

categories, e.g. an event with stronger electron acceleration, an event with stronger ion acceleration,

and an event with approximately equal acceleration for electrons and ions.

Another intriguing problem concerns the pitch angle distributions. Recent studies of

electron distributions showed various features that were not expected. Using Cluster data, Fu et al.

[2012c] performed a statistical study and found that the distributions are pancake-like (maximum

at 90 pitch angle) in growing flux pileup regions (FPRs, i.e. dipolarization regions), characterized

  by the Vx peaks lagging behind the Bz peaks. They are cigar-like (maximum at 0 and 180 ) in

decaying FPRs in which the Bz peaks lag behind the Vx peaks instead. This statistical study is

consistent with their earlier event study [Fu et al., 2011]. Their interpretation is that in growing

FPRs, the magnetic fields are compressed by higher-speed plasmas behind the DFs, so electrons

are betatron accelerated in the perpendicular direction. In the decaying FPRs, the expanding

magnetic flux tubes cool electrons in the perpendicular direction. The relationship between spatial

locations of FPRs and pitch angle distributions was not determined in their studies, but was

considered in a statistical study using THEMIS data by Wu et al. [2013]. They demonstrated that

the distributions are more pancake-like in the midtail (X   15RE ) than in the near-Earth region

(15REEXR 10 ). They interpreted this result as a consequence of stronger compression in

the midtail because of higher flow velocities there than in the near-Earth region. This observational feature and its interpretation were supported by a MHD plus test particle simulation [Birn et al.,

2014]. The simulation reproduced the observed spatial dependence of pitch angle distributions.

The simulation also reproduced the triple-peak structure (maximum at 0 , 90 and 180 ) in

electron distributions observed by THEMIS [Wu et al., 2013; Runov et al., 2013], although the

cause of the structure was not clearly explained. Future studies using MHD+LSK simulations may

162

examine the pitch angle distributions in simulations and compare them with observations in

substorm events.

The study of particle energization using PIC simulations should be further improved in the

future. First, as pointed out in Chapter 6, the artificial mass ratio poses a nontrivial and open

question on interpreting the energetics of electrons because the electron acceleration process is

strongly coupled with both ion and electron dynamics [Hoshino et al., 2001]. In our simulations,

the mass ratio was mmie / 400 , and the electron energies were accurate within about a factor of

2. So the problem was reduced when compared to typical explicit PIC simulations [e.g. Hoshino

et al., 2001; Drake et al., 2005; Pritchett, 2006a, 2006b]. For an implicit PIC simulation of a

D/2 system of space-time dimension D , the computing time scales with mass ratio as mmie/  . It

is feasible to increase the mass ratio to its realistic value mmie / 1836 by increasing the

computing time by a factor of about 8, and the total computing time will increase to about half of

a million CPU hours. With the realistic mass ratio, the problem of interpreting electron energetics

given by the simulations will be completely eliminated. Second, the initial set-up in the PIC

simulations needs to be further adjusted according to in-situ measurements. An obvious

improvement would be to add high-energy particles to the initial distributions. In particular,

observations have shown that the preexisting plasma sheet has a significant population of high-

energy electrons [e.g. Deng et al., 2010]. It is likely that the electron distributions as functions of

energy will no longer be Maxwellian in the EDR if high-energy electrons are added to the initial

distributions. If so, the difference between distributions in the EDR and at the DF is quantitative

rather than qualitative. Another improvement would be to use different temperatures for the plasma

sheet and the lobe, e.g. a cold population for the lobe and a warm population for the plasma sheet

[Hoshino et al., 2001]. However, to truly resolve the issue of unrealistic initial distributions, we

163

need to completely embrace measured distributions in the lobe and in the plasma sheet. The

ongoing Magnetospheric Multiscale (MMS) mission comprising four identically instrumented

spacecraft will study magnetic reconnection and particle acceleration in the magnetotail [Burch et

al., 2015]. Designed for measuring fields and particles on the electron scale, MMS will provide

unprecedented high quality data to constrain PIC simulations. Third, ion acceleration by magnetic reconnection needs to be examined and compared with electron acceleration. Recent observations

have overwhelmingly shown that electrons are accelerated to tens of keV to about a hundred keV

by processes close to the X-line [Øieroset et al., 2002; Imada et al., 2005, 2007; Chen et al., 2008;

Retinò et al., 2008; Wang et al., 2010; Imada et al., 2011; Huang et al., 2012a], while there are

very few corresponding observations of energetic ions. Previous PIC simulations showed that ion

distributions are non-Maxwellian near the X-line [e.g. Hoshino et al., 1998], and ions in the

outflow region are heated with effective thermal speeds on the order of Alfvén velocity based on

the reconnecting magnetic field [Drake et al., 2009], but no significant energetic ions are present.

With the high-quality MMS data, combined with the implicit PIC code and well-constrained

simulation set-up, significant progresses can be made in understanding energy partition and production of high-energy particles in the near-Earth reconnection region.

164

APPENDIX 1

Particle Sources for LSK Simulations

For the particle sources in the LSK simulations, we used a Maxwellian distribution to

represent the thermal particles and a power law distribution to represent the high-energy tail.

Here is a full description of the combined distribution.

First, the general form of the power law distribution, f p , as a function of energy,

EEE[,min max ] is

n E()v fcp ()v   (A1) Eth

where c is the normalization coefficient, Eth is a quantity with energy unit, n is the power law

index and Emin ( Emax ) are lower-energy (upper-energy) limits. Subscript p indicates a power

law distribution and will only appear below when necessary. The normalization condition is

vEEmax max / th 2EdE 4()fvdvvv2  1 4() f 1 (A2) vEEmin min / th mm 1 where Em v2 is used. By substituting (A1) into (A2), we have 2 n 3/2 a EEdE211  m 41cc  (A3) b     Ethmm42  E th a b

where  n 3/2, aE min / Eth , and bE max / Eth . Hence, the normalized distribution takes the form,

nn3/2 Em11   E fc()v     (A4) EEabEth42  th  th

3/2 nn 3 v 3 vth mE 111   E ffv() ()v th        (A5) vEabEabEth22 th   th 2  th

165

This is an isotropic, three-dimensional velocity vector distribution rather than velocity magnitude

(scalar) distribution. The velocity magnitude distribution is

1n vv2 v 1  E ff()42 ()2  (A6) vvvabEth th th  th Transforming the argument to energy, we have 11/2nn Evvvth11 th  E  E ff() ()2     (A7) EvvabvEabEth22 th th  th Note that the selection of the arguments ( v or E ) of the distribution function matters when referring to the power law index. The differential flux as a function of energy is

1n EEvv11 E jf() () f ()  (A8) EEvvabEth th th2 th  th

Second, the distribution function requires that the connection between the Maxwellian distribution and the power law distribution be continuous and smooth. This requirement gives the

exact relationship between the power law distribution lower-energy boundary, Emin and the power law index. The Maxwellian distribution is

2 v2   2 vv2 2vth feM ()  (A9) vvth  th Transforming the argument to energy, we have

2 vv22 E    22 Evvvvth211 th 22vvth v th E 2 E th ffMM() ()  e  e  e (A10) Evvth2222 th vv th v th E th Thus 1 f () e /2 (A11) M 2

where   EE/ th . Here Eth is one dimensional thermal energy of the Maxwellian distribution,

and we assign this meaning to the Eth in power law distribution, (A1). Now we calculate the

slope of Maxwellian distribution as a function of energy by taking the natural log of both sides,

166

1ln  lnf ( ) ln (A12) M 22

Setting   lnfM ( ) and   ln , we have:

111  ed e  ln (A13) 22d 2222

By requiring the same slope for the Maxwellian and power law distributions at Emin , the

relationship between the power law index and lower-energy boundary is

11  1Emin nn  (A14) 22 2 2 2E EEmin  min th

Third, we calculate the weights of the Maxwellian and power law distributions. The

accumulated flux of the Maxwellian distribution, I M is

2 vvvvvvvv/ th 2  IfderfMM() () ( ) exp() (A15) 0 2 vvvvvth th th2vth  th2 th

EEEEE/ th 2 I ()fderfe () ()  /2 (A16) MM0 EEEth th th 2  We set the weight of the Maxwellian distribution to be x , and the weight of the power law distribution to be y . Neglecting the contribution by the Maxwellian distribution tail, which is

1( I EEmin /)th and very small, we have a set of equations due to normalization and the

requirement that the combined distribution be continuous,

I(/)EExymin th   1 (A17)

yfpthMth(/) Emin E xf (/) E min E (A18)

1/2n 11EEmin min Equation (A18) yields yxEE exp(min 2th ) . Substitute it ab  Eth2 E th

into (A17) to obtain the weights. For example, for electrons in the March 11, 2008 event, we

167 have n  4.5 , Eth 1 keV , Emin  9 keV , Emax  450 keV , IEMth(min / E ) 0.97 , yx 0.04 , and x 1/1.01.

Both a Kappa distribution and the combined distribution derived here have two free parameters. For a Kappa distribution, they are the effective temperature and the Kappa value, 

(e.g. equation (5) in Summers and Thorne [1991]). For the combined distribution, they are the

Maxwellian temperature and the power law index (or lower-energy boundary). The combined

distribution resembles one Kappa distribution with    in the domain of EE min and another

Kappa distribution with   2 in the domain of EE min , because Maxwellian distribution corresponds the Kappa distribution with    and Kappa distribution approximates to power law distribution at high energy.

168

BIBLIOGRAPHY

Akasofu, S. I. (1964), The development of the auroral substorm, Planetary and Space Science, 12, 273-282.

Alfvén, H (1950), Cosmical Electrodynamics, Clarendon Press, Oxford.

Angelopoulos, V. (2008), The THEMIS mission, Space Sci. Rev., 141, 5 -34.

Angelopoulos, V., et al. (2008), Tail reconnection triggering substorm onset, Science, 321, 931- 935, doi: 10.1126/science.1160495

Angelopoulos, V., W. Baumjohann, C. F. Kennel, F. V. Coroniti, M. G. Kivelson, R. Pellat, R. J. Walker, H. Lühr, and G. Paschmann (1992), Bursty bulk flows in the inner central plasma sheet, J. Geophys. Res., 97(A4), 4027-4039.

Angelopoulos, V., C. Kennel, F. Coroniti, R. Pellat, M. Kivelson, R. Walker, C. Russell, W. Baumjohann, W. Feldman, and J. Gosling (1994), Statistical characteristics of bursty bulk flow events, J. Geophys. Res., 99, 21,257, doi:10.1029/94JA01263.

Angelopoulos, V., et al. (1996), Multipoint analysis of a bursty bulk flow event on April 11, 1985, J. Geophys. Res., 101(A3), 4967-4989.

Arnoldy R.L., and K.W. Chan (1969), Particle substorms observed at the geostationary orbit, J. Geophys. Res., 74, 5019.

Artemyev, A. V., V. N. Lutsenko, and A. A. Petrukovich, (2012), Ion resonance acceleration by dipolarization fronts: Analytic theory and spacecraft observation, Ann. Geophys., 30, 317-324.

Artemyev, A. V., L. M. Zelenyi, A. A. Petrukovich, and R. Nakamura (2011), Hot electrons as tracers of large-scale structure of magnetotail current sheets, Geophys. Res. Lett., 38(14).

Asano, Y., et al. (2010), Electron acceleration signatures in the magnetotail associated with substorms, J. Geophys. Res., 115, A05215, doi:10.1029/2009JA014587.

Ashour-Abdalla, M., J. P. Berchem, J. Büchner, and L. M. Zelenyi (1990), Chaotic scattering and acceleration of ions in the Earth's magnetotail, Geophys. Res. Lett., 17(13), 2317-2320.

Ashour-Abdalla, M., J. P. Berchem, J. Büchner, and L. M. Zelenyi (1993), Shaping of the magnetotail from the mantle: Global and local structuring, J. Geophys. Res., 98, 5651-5676.

Ashour-Abdalla, M., J. M. Bosqued, M. El-Alaoui, V. Peroomian, L. M. Zelenyi, R. J. Walker, and J. Wright (2005), A stochastic sea: The source of plasma sheet boundary layer ion structures observed by Cluster, J. Geophys. Res., 110, A12221, doi:10.1029/2005JA011183.

169

Ashour-Abdalla, M., J. M. Bosqued, M. El-Alaoui, V. Peroomian, M. Zhou, R. Richard, R. Walker, A. Runov, and V. Angelopoulos (2009), A simulation study of particle energization observed by THEMIS spacecraft during a substorm, J. Geophys. Res., 114, A09204, doi:10.1029/2009JA014126.

Ashour-Abdalla, M., M. El-Alaoui, F. V. Coroniti, R. J. Walker, and V. Peroomian (2002), A new convection state at substorm onset: Results from an MHD study, Geophys. Res. Lett., 29(20), 1965, doi:10.1029/2002GL015787.

Ashour-Abdalla, M., M. El-Alaoui, M. L. Goldstein, M. Zhou, D. Schriver, R. Richard, R. Walker, M. G. Kivelson, and K. Hwang (2011), Observations and simulations of non‐local acceleration of electrons in magnetotail magnetic reconnection events, Nat. Phys., 7, 360–365, doi:10.1038/nphys1903.

Ashour-Abdalla, M., M. El-Alaoui, V. Peroomian, R. J. Walker, J. Raeder, L. A. Frank, and W. R. Paterson (1999b), Source distributions of substorm ions observed in the near-Earth magnetotail, Geophys. Res. Lett., 26(7), 955-958, doi:10.1029/1999GL900112.

Ashour-Abdalla, M., M. El-Alaoui, V. Peroomian, R. J. Walker, J. Raeder, L. A. Frank, and W. R. Paterson (2000), The origin of the near-Earth plasma population during a substorm on November 24, 1996, J. Geophys. Res., 105(A2), 2589–2605, doi:10.1029/1999JA900389.

Ashour-Abdalla, M., M. El-Alaoui, V. Peroomian, R. J. Walker, L. M. Zelenyi, L. A. Frank, and W. R. Paterson (1999a), Localized reconnection and substorm onset on December 22, 1996, Geophys. Res. Lett., 26, 3545-3548, doi:10.1029/1999GL003630.

Ashour-Abdalla, M., L. A. Frank, W. R. Paterson, V. Peroomian, and L. M. Zelenyi (1996), Proton velocity distributions in the magnetotail: Theory and observations, J. Geophys. Res., 101(A2), 2587–2598, doi:10.1029/95JA02539.

Ashour-Abdalla, M., G. Lapenta, R. J. Walker, M. El-Alaoui1, and H. Liang (2015), Multiscale Study of Electron Energization during Unsteady Reconnection Events, J. Geophys. Res., doi: 10.1002/2014JA020316.

Ashour-Abdalla, M., D. Schriver, M. E. Alaoui, R. Richard, R. Walker, M. L. Goldstein, E. Donovan, and M. Zhou (2013), Direct auroral precipitation from the magnetotail during substorms, Geophys. Res. Lett., 40, 3787–3792, doi:10.1002/grl.50635.

Ashour-Abdalla, M., L. M. Zelenyi, J. M. Bosqued, and R. A. Kovrazhkin (1992a), Precipitation of fast ion beams from the plasma sheet boundary layer, Geophys. Res. Lett., 19(6), 617– 620.

Ashour-Abdalla, M., L. M. Zelenyi, J. Bosqued, V. Peroomian, Z. Wang, D. Schriver, and R. L. Richard (1992b), The formation of the wall region: Consequences in the near Earth magnetotail, Geophys. Res. Lett., 19, 1739, doi:10.1029/92GL01810.

170

Ashour-Abdalla, M., L. M. Zelenyi, J. M. Bosqued, V. Peroomian, Z. Wang, D. Schriver, and R. Richard (1992c), Effects of near-Earth stochastic acceleration and reflections of magnetotail ions on the formation of auroral arcs, in Proceedings of the International Conference on Substorms (ICS-1), Eur. Space Agency Spec. Publ., 335, 545, Eur. Space Agency, Paris.

Ashour-Abdalla, M., L. M. Zelenyi, V. Peroomian, and R. L. Richard (1994), Consequences of magnetotail ion dynamics, J. Geophys. Res., 99(A8), 14891-14916, doi:10.1029/94JA00141.

Ashour-Abdalla, M., L. M. Zelenyi, V. Peroomian, R. L. Richard, and J. M. Bosqued (1995), The mosaic structure of plasma bulk flows in the Earth's magnetotail, J. Geophys. Res., 100(A10), 19191–19209, doi:10.1029/95JA00902.

Auster, U., et al. (2008), The THEMIS fluxgate magnetometer, Space Sci. Rev., 141, 235–264, doi:10.1007/s11214-008-9365-9.

Baker, D. N., R.D. Belian, P.R. Higbie, and E.W. Hones Jr. (1979), High-energy magnetospheric protons and their dependence on geomagnetic and interplanetary conditions, J. Geophys. Res., 84, 7138.

Baker, D. N., E. W. Hones, P. R. Higbie, R. D. Belian, and P. Stauning (1981), Global properties of the magnetosphere during a substorm growth phase: A case study, J. Geophys. Res., 86, (A11), 8941-8956.

Baker, D. N., and E. C. Stone (1976), Energetic electron anisotropies in the magnetotail: Identification of open and closed field lines, Geophys. Res. Lett., 3(9), 557-560.

Baker, D.N., and E.C. Stone (1977), Observations of energetic electrons energy ( E 200 keV ) in the earth’s magnetotail: plasma sheet and fireball observations, J. Geophys. Res., 82, 1532- 1546, doi:10.1029/JA082i010p01532.

Baker, D. N., T. I. Pulkkinen, V. Angelopoulos, W. Baumjohann, and R. L. McPherron (1996), Neutral line model of substorms: Past results and present view, J. Geophys. Res., 101, A6: 12975-13010.

Baños, A. (1967), The guiding centre approximation in lowest order, Journal of Plasma Physics, 1(03), 305-316.

Baumjohann, W., G. Paschmann, and H. Lühr (1990), Characteristics of high-speed ion flows in the plasma sheet. J. Geophys. Res., 95(A4), 3801-3809.

Baumjohann, W., M. Hesse, S. Kokubun, T. Mukai, T. Nagai, and A. A. Petrukovich (1999), Substorm dipolarization and recovery, J. Geophys. Res., 104, 24,995– 25,000.

Birdsall C., and A. Langdon (2004), Plasma Physics Via Computer Simulation, McGraw-Hill, New York.

171

Birn, J., et al. (2001), Geospace Environmental Modeling (GEM) magnetic reconnection challenge, J. Geophys. Res., 106( A3), 3715-3719.

Birn, J., A. V. Artemyev, D. N. Baker, M. Echim, M. Hoshino, and L. M. Zelenyi (2012), Particle acceleration in the magnetotail and aurora, Space Sci. Rev., 173(1-4), 49-102.

Birn, J., M. Hesse, R. Nakamura, and S. Zaharia (2013), Particle acceleration in dipolarization events, J. Geophys. Res., 118, 1960–1971, doi:10.1002/jgra.50132.

Birn, J., R. Nakamura, E. V. Panov, and M. Hesse (2011), Bursty bulk flows and dipolarization in MHD simulations of magnetotail reconnection, J. Geophys. Res., 116, A01210, doi:10.1029/2010JA016083.

Birn, J., J. Raeder, Y. L. Wang, R. A. Wolf, and M. Hesse (2004a), On the propagation of bubbles in the geomagnetic tail, Annales Geophysicae, 22(5), 1773-1786.

Birn, J., A. Runov, and M. Hesse (2014), Energetic electrons in dipolarization events: Spatial properties and anisotropy, J. Geophys. Res., 119, 3604–3616, doi:10.1002/2013JA019738.

Birn, J., M. F. Thomsen, J. E. Borovsky, G. D. Reeves, D. J. McComas, and R. Belian (1997a), Characteristic plasma properties during dispersionless substorm injections at geosynchronous orbit, J. Geophys. Res., 102(A2), 2309.

Birn, J., M. F. Thomsen, J. E. Borovsky, G. D. Reeves, D. J. M. R. D. Belian, and M. Hesse (1997b), Substorm ion injections: Geosynchronous observations and test particle orbits in three-dimensional dynamic MHD fields, J. Geophys. Res., 102, 2325.

Birn, J., M. F. Thomsen, J. E. Borovsky, G. D. Reeves, D. J. McComas, R. D. Belian, and M. Hesse (1998), Substorm electron injections: Geosynchronous observations and test particle simulations, J. Geophys. Res., 103, 9235.

Birn, J., M. F. Thomsen, and M. Hesse (2004b), Electron acceleration in the dynamic magnetotail: Test particle orbits in three-dimensional MHD simulation fields, Phys. Plasmas, 11, 1825, doi:10.1063/1.1704641.

Biskamp, D. (2005), Magnetic Reconnection in Plasmas, Cambridge University Press, New York.

Belian, R. D., D. N. Baker, P. R. Higbie, and E. W. Hones Jr. (1978), High-resolution energetic particle measurements at 6.6 RE, 2. High-energy proton drift echoes, J. Geophys. Res., 83(A10), 4857–4862, doi:10.1029/JA083iA10p04857.

Bogott, F.H., and F.S. Mozer (1973), Nightside energetic particle decreases at the synchronous orbit, J. Geophys. Res., 78, 8119.

Bonnell, J. W., et al. (2008), The electric field instrument (EFI) for THEMIS. Space Sci. Rev., 141, 303-341.

172

Bothmer V., and I.A. Daglis, Space Weather-Physics and Effects, Springer, 2007.

Brackbill J. U., and D. W. Forslund (1982), An implicit method for electromagnetic plasma simulation in two dimensions, Journal of Computational Physics, 46, 271-308.

Brackbill J.U., and D.W. Forslund (1985), Simulation of low-frequency, electromagnetic phenomena in plasmas, Multiple Time Scales, edited by J.U. Brackbill, B.I. Cohen, Academic Press, Orlando, pp. 271-310.

Brizard, A. J., and T. S. Hahm (2007), Foundations of nonlinear gyrokinetic theory, Reviews of modern physics, 79(2), 421.

Büchner, J. (1991), Correlation-modulated chaotic scattering in the Earth’s magnetosphere, Geophys. Res. Lett., 18, 1595-1598, doi:10.1029/91GL01905.

Büchner, J., and L. M. Zelenyi (1989), Regular and chaotic charged particle motion in magnetotaillike field reversals: 1. Basic theory of trapped motion, J. Geophys. Res., 94(A9),11821–11842, doi:10.1029/JA094iA09p11821.

Bulanov, S.V., and P.V. Sasorov (1976), Energy spectrum of particles accelerated in the neighborhood of a line of zero magnetic field, Sov. Astron., 19, 464–468.

Burch, J. L., T. E. Moore, R. B. Torbert, B. L. Giles (2015), Magnetospheric multiscale overview and science objectives, Space. Sci. Rev., doi:10.1007/s11214-015-0164-9.

Burkhart G.R., and J. Chen (1991), Differential memory in the Earth’s magnetotail, Geophys. Res. Lett., 96, 14033 doi:10.1029/91JA01137.

Burkhart G.R., J.F. Drake, J. Chen (1990), Magnetic reconnection in collisionless plasmas- prescribed fields, J. Geophys. Res., 95, 18833–18848, doi:10.1029/JA095iA11p18833.

Cairns, I. H., and B. F. McMillan (2005), Electron acceleration by lower hybrid waves in magnetic reconnection regions, Physics of Plasmas, 12, 102110.

Cattell, C., F. Mozer, K. Tsuruda, H. Hayakawa, M. Nakamura, T. Okada, and T. Yamamoto (1994), Geotail observations of spiky electric fields and low-frequency waves in the plasma sheet and plasma sheet boundary, Geophys. Res. Lett., 21(25), 2987-2990.

Chen, F. F. (2006), Introduction to plasma physics and controlled fusion, 2nd edition, Springer.

Chen, L.-J., et al. (2008), Observation of energetic electrons within magnetic islands, Nature Phys., 4, 19-23, doi:10.1038/nphys777.

173

Daughton, W., V. Roytershteyn, H. Karimabadi, L. Yin, B. J. Albright, B. Bergen, and K. J. Bowers (2011), Role of electron physics in the development of turbulent magnetic reconnection in collisionless plasmas, Nat. Phys., 7(7), 539-542.

Daughton, W., J. Scudder, and H. Karimabadi (2006), Fully kinetic simulations of undriven magnetic reconnection with open boundary conditions. Physics of Plasmas, 13(7), 072101.

Delcourt, D., and J. Sauvaud (1994), Plasma sheet ion energization during dipolarization events, J. Geophys. Res., 99, 97, doi:10.1029/93JA01895.

Deng, X., M. Ashour-Abdalla, M. Zhou, R. Walker, M. El‐Alaoui, V. Angelopoulos, R. E. Ergun, and D. Schriver (2010), Wave and particle characteristics of earthward electron injections associated with dipolarization fronts, J. Geophys. Res., 115, A09225, doi:10.1029/2009JA015107.

Divin, A. V., M. I. Sitnov, M. Swisdak, and J. F. Drake (2007), Reconnection onset in the magnetotail: Particle simulations with open boundary conditions, Geophys. Res. Lett., 34, L09109, doi:10.1029/2007GL029292.

Drake, J. F., and M. A. Shay (2006), The fundamentals of collisionless reconnection. In Reconnection of Magnetic Fields: Magnetohydrodynamics and Collisionless Theory and Observations, edited by J. Birn and E. Priest, Cambridge University Press.

Drake, J. F., M. A. Shay, W. Thongthai and M. Swisdak (2005), Production of energetic electrons during magnetic reconnection. Phys. Rev. Lett. 94, 095001.

Drake, J. F., M. Swisdak, H. Che, and M. A. Shay (2006), Electron acceleration from contracting magnetic islands during reconnection, Nature, 443, 553–556, doi:10.1038/nature05116.

Drake, J. F., M. Swisdak, T. D. Phan, P. A. Cassak, M. A. Shay, S. T. Lepri, R. P. Lin, E. Quataert, and T. H. Zurbuchen (2009), Ion heating resulting from pickup in magnetic reconnection exhausts, J. Geophys. Res., 114 (A5).

Drake, J. F., M. Swisdak, M. A. Shay, B. N. Rogers, A. Zeiler, and C. Cattell (2003), Formation of electron holes and particle energization during magnetic reconnection, Science, 299, pp. 873.

Dungey, J. W. (1961), Interplanetary Magnetic Field and the Auroral Zones, Phys. Rev. Lett., 6, pp. 47-48

Egedal, J., W. Daughton, and A. Le (2012), Large-scale electron acceleration by parallel electric fields during magnetic reconnection, Nat. Phys., 8, 321–324.

Egedal, J., A. Le, and W. Daughton (2013), A review of pressure anisotropy caused by electron trapping in collisionless plasma, and its implications for magnetic reconnection, Physics of Plasmas, 20(6), 061201.

174

Egedal, J., M. Øieroset, W. Fox, and R. P. Lin (2005), In situ discovery of an electrostatic potential, trapping electrons and mediating fast reconnection in the earth's magnetotail, Phys. Rev. Lett., 94, 025006.

El-Alaoui, M. (2001), Current disruption during November 24, 1996, substorm. J. Geophys. Res., 106, 6229-6245, doi:10.1029/1999JA000260.

El-Alaoui, M., M. Ashour-Abdalla, R. J. Walker, V. Peroomian, R. L. Richard, V. Angelopoulos, and A. Runov (2009), Substorm evolution as revealed by THEMIS satellites and a global MHD simulation, J. Geophys. Res., 114, A8.

El-Alaoui, M., R. L. Richard, M. Ashour-Abdalla, M. L. Goldstein, and R. J. Walker (2013), Dipolarization and turbulence in the plasma sheet during a substorm: THEMIS observations and global MHD simulations, J. Geophys. Res. Space Physics, 118, doi:10.1002/2013JA019322.

El-Alaoui, M., R. L. Richard, M. Ashour-Abdalla, R. J. Walker, and M. L. Goldstein (2012), Turbulence in a global magnetohydrodynamic simulation of the Earth's magnetosphere during northward and southward interplanetary magnetic field, Nonlinear Process. Geophys., 19(2), 165-175.

Fu, H. S., et al. (2013), Dipolarization fronts as a consequence of transient reconnection: In situ evidence, Geophys. Res. Lett., 40, 6023-6027, doi:10.1002/2013GL058620.

Fu, H. S., Y. V. Khotyaintsev, M. André, and A. Vaivads (2011), Fermi and betatron acceleration of suprathermal electrons behind dipolarization fronts, Geophys. Res. Lett., 38, L16104, doi:10.1029/2011GL048528.

Fu, H. S., Y. V. Khotyaintsev, A. Vaivads, M. André, and S. Y. Huang (2012a), Occurrence rate of earthward-propagating dipolarization fronts, Geophys. Res. Lett., 39, L10101, doi: 10.1029/2012GL051784.

Fu, H. S., Y. V. Khotyaintsev, A. Vaivads, M. André, and S. Y. Huang (2012b), Electric structure of dipolarization front at sub-proton scale, Geophys. Res. Lett., 39, L06105, doi:10.1029/2012GL051274.

Fu, H. S., Y. V. Khotyaintsev, A. Vaivads, M. André, V. A. Sergeev, S. Y. Huang, E. A. Kronberg, and P. W. Daly (2012c), Pitch angle distribution of suprathermal electrons behind dipolarization fronts: A statistical overview, J. Geophys. Res., 117, A12221, doi:10.1029/2012JA018141.

Fujimoto, K. (2006), Time evolution of the electron diffusion region and the reconnection rate in fully kinetic and large system, Physics of Plasmas, 13(7).

Galeev A.A. (1979), Reconnection in the magnetotail, Space Sci. Rev., 23, 411–425, doi:10.1007/BF00172248.

175

Galeev A.A., F.V. Coroniti, M. Ashour-Abdalla (1978), Explosive tearing mode reconnection in the magnetospheric tail, Geophys. Res. Lett., 5, 707–710, doi:10.1029/GL005i008p00707.

Gabrielse, C., V. Angelopoulos, A. Runov, and D. L. Turner (2012), The effects of transient, localized electric fields on equatorial electron acceleration and transport toward the inner magnetosphere, J. Geophys. Res., 117, A10213, doi:10.1029/2012JA017873.

Gabrielse, C., V. Angelopoulos, A. Runov, and D. L. Turner (2014), Statistical characteristics of particle injections throughout the equatorial magnetotail, J. Geophys. Res. Space Physics, 119, 2512–2535, doi:10.1002/2013JA019638.

Ge, Y. S., J. Raeder, V. Angelopoulos, M. L. Gilson, and A. Runov (2011), Interaction of dipolarization fronts within multiple bursty bulk flows in global MHD simulations of a substorm on 27 February 2009, J. Geophys. Res., 116, A00I23, doi:10.1029/2010JA015758.

Giovanelli, R. G. (1947), Magnetic and electric phenomena in the sun’s atmosphere associated with sunspots, MNRAS, 107(4), 338–355.

Goldman, M. V., D. L. Newman, D. L., and P. Pritchett (2008), Vlasov simulations of electron holes driven by particle distributions from PIC reconnection simulations with a guide field. Geophys. Res. Lett., 35(22), doi: 10.1029/2008GL035608.

Goldstein, M. L., W. H. Matthaeus, and J. J. Ambrosiano (1986), Acceleration of charged particles in magnetic reconnection: Solar flares, the magnetosphere, and solar wind, Geophys. Res. Lett., 13(3), 205-208.

Green, J. C., and M. G. Kivelson (2004), Relativistic electrons in the outer radiation belt: Differentiating between acceleration mechanisms, J. Geophys. Res., 109, A03213, doi: 10.1029/2003JA010153.

Harris, E. G. (1962). On a plasma sheath separating regions of oppositely directed magnetic field, Nuovo Cimento, 23(1), 115-121.

Hesse, M., and J. Birn (1991), On reconnection and its relation to the substorm current wedge, J. Geophys. Res., 96, 19417– 19426.

Hockney R., and J. Eastwood (1988), Computer Simulation using Particles, Taylor & Francis, Bristol.

Horne, R. B., R. M. Thorne, S. A. Glauert, J. M. Albert, N. P. Meredith, and R. R. Anderson (2005), Timescale for radiation belt electron acceleration by whistler mode chorus waves, J. Geophys. Res., 110, A03225, doi:10.1029/2004JA010811.

Hoshino M (2005), Electron surfing acceleration in magnetic reconnection, J. Geophys. Res., 110(A9), A10215. doi:10.1029/2005JA011229.

176

Hoshino, M. (2012), Stochastic Particle Acceleration in Multiple Magnetic Islands during Reconnection, Phys. Rev. Lett., 108, 135003, doi:10.1103/PhysRevLett.108.135003.

Hoshino, M., T. Mukai, T. Terasawa and I. Shinohara (2001), Suprathermal electron acceleration in magnetic reconnection. J. Geophys. Res. 106, 25979–25978.

Hoshino, M., T. Mukai, T. Yamamoto, and S. Kokubun (1998), Ion dynamics in magnetic reconnection: Comparison between numerical simulation and Geotail observations, J. Geophys. Res..103 (A3), 4509-4530.

Huang, S. Y., et al. (2012a), Electron acceleration in the reconnection diffusion region: Cluster observations, Geophys. Res. Lett., 39, L11103, doi:10.1029/2012GL051946.

Huang, S. Y., et al. (2012b), Kinetic structure and wave properties associated with sharp dipolarization front observed by Cluster, Ann. Geophys., Vol. 30. No. 1, pp. 97-107 Copernicus GmbH.

Hughes W. J. (1995), The magnetopause, magnetotail and magnetic reconnection. In Introduction to Space Physics, edited by M. G. Kivelson and C.T. Russell, Cambridge Univ. Press, New York.

Hurricane, O. A., R. Pellat, and F. V. Coroniti (1996), Instability of the Lembége-Pellat equilibrium under ideal magnetohydrodynamics, Phys. Plasmas, 3, 2472.

Hwang, K-J., M. L. Goldstein, E. Lee, and J. S. Pickett (2011), Cluster observations of multiple dipolarization fronts, J. Geophys. Res., 116, A00I32, doi:10.1029/2010JA015742.

Imada S., M. Hirai, M. Hoshino, T. Mukai (2011), Favorable conditions for energetic electron acceleration during magnetic reconnection in the Earth’s magnetotail, J. Geophys. Res., 116(A15), A08217, doi:10.1029/2011JA016576.

Imada S., M. Hoshino, T. Mukai (2005), Average profiles of energetic and thermal electrons in the magnetotail reconnection regions, Geophys. Res. Lett., 320, L09101. doi:10.1029/2005GL022594.

Imada, S., R. Nakamura, P. W. Daly, M. Hoshino, W. Baumjohann, S. Muhlbachler, A. Balogh, and H. Reme (2007), Energetic electron acceleration in the downstream reconnection outflow region, J. Geophys. Res., 112, A03202, doi:10.1029/2006JA011847.

Kavanagh, L. D., J. W. Freeman, and A. J. Chen (1968), Plasma flow in the magnetosphere, J. Geophys. Res., 73(17), 5511-5519.

Khotyaintsev, Y. V., C. M. Cully, A. Vaivads, M. Andre, and C. J. Owen (2011), Plasma jet braking: energy dissipation and non-adiabatic electrons, Phys. Rev. Lett., 106, 165001.

177

Kivelson, M. G., T. A. Farley, and M. P. Aubry (1973), Satellite studies of magnetospheric substorms on August 15, 1968: 5. Energetic electrons, spatial boundaries and wave-particle interactions at OGO-5, J. Geophys. Res., 78, 3079.

Kivelson, M. G. and C. T. Russell (1995), Introduction to Space Physics, Cambridge University Press, New York.

Krauss-Varban, D., and H. Karimabadi (2003), Timing and localization of reconnection signatures - Is there a substorm model problem? Geophys. Res. Lett., 30(6).

Kress, B. T., M. K. Hudson, M. D. Looper, J. Albert, J. G. Lyon, and C. C. Goodrich (2007), Global MHD test particle simulations of >10 MeV radiation belt electrons during storm sudden commencement, J. Geophys. Res., 112, A09215, doi:10.1029/2006JA012218.

Kowal, G., A. Lazarian, E. T. Vishniac, and K. Otmianowska-Mazur (2009), Numerical tests of fast reconnection in weakly stochastic magnetic fields. Astrophys. J. 700, 63–85.

Langdon A., B. Cohen and A. Friedman, Direct implicit large time-step particle simulation of plasmas, Journal of Computational Physics, 51, 107-138.

Lanzerotti, L. J., C. S. Roberts, and W. L. Brown (1967), Temporal Variations in the Electron Flux at Synchronous Altitudes, J. Geophys. Res., 72(23), 5893–5902, 1350 doi:10.1029/JZ072i023p05893.

Lapenta, G. (2012), Particle simulations of space weather, Journal of Computational Physics, 231(3), 795-821.

Lapenta, G., J. U. Brackbill, and P. Ricci (2006), Kinetic approach to microscopic-macroscopic coupling in space and laboratory plasmas, Physics of Plasmas, 13(5), 055904.

Lapenta, G., S. Markidis, A. Divin, M. Goldman, and D. Newman (2010), Scales of guide field reconnection at the hydrogen mass ratio, Physics of Plasmas, 17(8), 082106, doi: 10.1063/1.3467503.

Lapenta, G., S., Markidis, A. Divin, D. Newman, and M. Goldman (2014), Separatrices: the crux of reconnection, J. of Plasma Phys., 1-39, doi: 10.1017/S0022377814000944.

Le Contel, O., et al. (2008), First results of the THEMIS search coil magnetometers. Space Sci. Rev., 141, 509-534.

Li, X., D. N. Baker, M. Temerin, G. D. Reeves, and R. D. Belian (1998), Simulation of dispersionless injections and drift echoes of energetic electrons associated with substorms, Geophys. Res. Lett., 25, 3763–3766, doi:10.1029/1998GL900001.

Lyons, L. R. (1984), Electron Energization in the Geomagnetic Tail Current Sheet, J. Geophys. Res., 89(A7), 5479–5487, doi:10.1029/JA089iA07p05479.

178

Lyons, L. R., and T. W. Speiser (1982), Evidence for current sheet acceleration in the geomagnetic tail, J. Geophys. Res., 87(A4), 2276-2286.

Lyons, L. R., and D. J. Williams (1986), Quantitative Aspects of Magnetospheric Physics. Geophysical Journal of the Royal Astronomical Society, 86, 221-222.

Lui, A. T. Y. (1996), Current disruption in the Earth's magnetosphere: Observations and models, J. Geophys. Res., 101, A6, 13067-13088.

Lui, A. T. Y., R. E. Lopez, S. M. Krimigis, R. W. McEntire, and L. J. Zanetti (1988), A case study of magnetotail current sheet disruption and diversion, Geophys. Res. Lett., 15, 721-724.

Markidis, S., G. Lapenta, and Rzwan-udding (2010), Multi-scale simulations of plasma with ipic3d, Mathematics and Computers in Simulation, 80 (7) 1509-1519, doi:10.1016/j.matcom.2009.08.038.

Matthaeus, W. H., J. J. Ambrosiano, and M. L. Goldstein (1984), Particle acceleration by turbulent magnetohydrodynamic reconnection, Phys. Rev. lett., 53(15), 1449.

Mauk, B. H., and C. E. McIlwain (1974), Correlation of Kp with the substorm injected plasma boundary, J. Geophys. Res., 79, 3193.

McFadden, J. P., et al. (2008), The THEMIS ESA plasma instrument and in-flight calibration, Space Sci. Rev., 141, 277– 302.

McIlwain, C. (1974), Substorm injection boundaries, in Magnetospheric Physics, edited by B. M. McCormac, p. 143, D. Reidel, Hingham, Mass.

McPherron, R. L. (1972), Substorm related changes in the geomagnetic tail: The growth phase, Planetary and Space Science, 20(9), 1521-1539.

McPherron, R.L. (1991), Physical processes producing magnetospheric substorms and magnetic storms, in Geomagnetism, ed. J.A. Jacobs (Academic press, London), Vol. 4, p.593.

McPherron, Robert L., C. T. Russell, and M. P. Aubry (1973), Satellite studies of magnetospheric substorms on August 15, 1968: 9. Phenomenological model for substorms, J. Geophys. Res., 78 (16), 3131-3149.

Möebius, E., M. Scholer, D. Hovestadt, G. Paschmann, G. Gloeckler (1983), Energetic particles in the vicinity of a possible neutral line in the plasma sheet, J. Geophys. Res., 88, 7742–7752 doi:10.1029/JA088iA10p07742.

Morales, G. J., and Y. C. Lee (1974), Effect of localized electric fields on the evolution of the velocity distribution function, Phys. Rev. Lett., 33(26), 1534-1537.

179

Nagai T., I. Shinohara, M. Fujimoto, M. Hoshino, Y. Saito, S. Machida, T. Mukai (2001), Geotail observations of the Hall current system: Evidence of magnetic reconnection in the magnetotail, J. Geophys. Res., 106, 25929–25950, doi:10.1029/2001JA9000380.

Nakamura, R., et al. (2004), Spatial scale of high-speed flows in the plasma sheet observed by Cluster, Geophys. Res. Lett., 31, L09804, doi:10.1029/2004GL019558.

Northrop, T. G. (1963), The Adiabatic Motion of Charged Particles, John Wiley, New York.

Nose, M., S. Ohtani, K. Takahashi, A.T.Y. Lui, R.W. McEntire, D. J. Williams, S.P. Christon, and K. Yumoto, (2001), Ion composition of the near-Earth plasma sheet in storm and quiet intervals: Geotail/EPIC measurements, J. Geophys. Res., 106, 8391.

Ohtani, S., M. A. Shay, and T. Mukai (2004), Temporal structure of the fast convective flow in the plasma sheet: Comparison between observations and two-fluid simulations, J. Geophys. Res., 109, 3–8, doi:10.1029/2003JA010002.

Oka M., T.D. Phan, S. Krucker,M. Fujimoto, I. Shinohara (2010), Electron acceleration by multi- island coalescence, Astrophys. J, 714, 915–926, doi:10.1088/0004-637X/714/1/915.

Okada, T., K. Tsuruda, H. Hayakawa, M. Nakamura, T. Mukai, T. Yamamoto, and H. Matsumoto (1994), GEOTAIL observations of electrostatic waves in the lower hyrid frequency range in the plasma sheet boundary layer, Geophys. Res. Lett., 21(25), 2931-2934.

Øieroset, M., R. P. Lin, T. D. Phan, D. E. Larson, and S. D. Bale (2002), Evidence for electron acceleration up to 300 keV in the magnetic reconnection diffusion region in the earth's magnetotail, Phys. Rev. Lett., 89, 195001.

Øieroset, M., T. D. Phan, M. Fujimoto, R. P. Lin, and R. P. Lepping (2001), In situ detection of collisionless reconnection in the Earth's magnetotail, Nature, 412(6845), 414-417.

Pan, Q., M. Ashour-Abdalla, M. El-Alaoui, R. J. Walker, and M. L. Goldstein (2012), Adiabatic acceleration of suprathermal electrons associated with dipolarization fronts, J. Geophys. Res., 117, A12224, doi:10.1029/2012JA018156.

Pan, Q., M. Ashour-Abdalla, R. J. Walker, and M. El-Alaoui (2014a), Electron energization and transport in the magnetotail during substorms, J. Geophys. Res. Space Physics, 119, doi:10.1002/2013JA019508.

Pan, Q., M. Ashour-Abdalla, R. J. Walker, and M. El-Alaoui (2014b), Ion energization and transport associated with magnetic dipolarizations, Geophys. Res. Lett., 41, 5717–5726, doi:10.1002/2014GL061209.

180

Pan, Q., M. Ashour-Abdalla, R. J. Walker, M. El-Alaoui (2015a), A Comparison of ion and electron energization and transport mechanisms in the magnetotail during substorms, submitted to J. Geophys. Res..

Pan, Q., M. Ashour-Abdalla, R. J. Walker, G. Lapenta (2015b), Production of power law electrons: magnetic field diffusion or dipolarization, in preparation for publication.

Pfitzer, K.A., J.R. Winckler (1969), Intensity correlations and substorm electron drift effects in the outer radiation belt measured with the Ogo3 and ATS 1 satellites, J. Geophys. Res., 74, 5005.

Pritchett, P. L. (2001), Geospace Environment Modeling magnetic reconnection challenge: Simulations with a full particle electromagnetic code, J. Geophys. Res., 106(A3), 3783-3798.

Pritchett, P. L. (2006a), Relativistic electron production during driven magnetic reconnection, Geophys. Res. Lett., 33(13).

Pritchett, P. L. (2006b), Relativistic electron production during guide field magnetic reconnection, J. Geophys. Res., 111(A10).

Pritchett P.L. (2008), Collisionless magnetic reconnection in an asymmetric current sheet, J. Geophys. Res., 113(A12), A06210, doi:10.1029/2007JA012930.

Pritchett, P. L., F. V. Coroniti (2010), A kinetic ballooning/interchange instability in the magnetotail, J. Geophys. Res., 115(A6).

Raeder, J., J. Berchem, and M. Ashour-Abdalla (1998), The geospace environment modeling grand challenge: Results from a global geospace circulation model, J. Geophys. Res., 103, 14,787- 14,797, doi:10.1029/98JA00014.

Raeder, J., R. L. McPherron, L. A. Frank, S. Kokubun, G. Lu, T. Mukai, W. R. Paterson, J. B. Sigwarth, H. J. Singer, and J. A. Slavin (2001), Global simulation of the Geospace Environment Modeling substorm challenge event, J. Geophys. Res., 106, 381-395, doi:10.1029/2000JA000605.

Reeves, G.D., D.N. Belian, T.A. Fritz (1991), Numerical tracing of energetic particle drifts in a model magnetosphere, J. Geophys. Res., 96, 13997.

Retinò, A., et al. (2008), Cluster observations of energetic electrons and electromagnetic fields within a reconnecting thin current sheet in the Earth’s magnetotail, J. Geophys. Res., 113, A12215, doi:10.1029/2008JA013511.

Ricci, P., G. Lapenta, and J. U. Brackbill (2003), Electron acceleration and heating in collisionless magnetic reconnection, Physics of Plasmas, 10 (9), 3554-3560.

181

Roederer, J. G. (1970), Dynamics of Geomagnetically Trapped Radiation, Cambridge Univ. Press, New York.

Runov, A., V. Angelopoulos, C. Gabrielse, X.-Z. Zhou, D. Turner, and F. Plaschke (2013), Electron fluxes and pitch-angle distributions at dipolarization fronts: THEMIS multipoint observations, J. Geophys. Res., 118, 744–755, doi:10.1002/jgra.50121.

Runov, A., V. Angelopoulos, M. I. Sitnov, V. A. Sergeev, J. Bonnell, J. P. McFadden, D. Larson, K.-H. Glassmeier, and U. Auster (2009), THEMIS observations of an earthward propagating dipolarization front, Geophys. Res. Lett., 36, L14106, doi:10.1029/2009GL038980.

Runov, A., V. Angelopoulos, and X.-Z. Zhou (2012), Multipoint observations of dipolarization front formation by magnetotail reconnection, J. Geophys. Res., 117(A5).

Runov, A., V. Angelopoulos, X.-Z. Zhou, X.-J. Zhang, S. Li, F. Plaschke, and J. Bonnell (2011), A THEMIS multicase study of dipolarization fronts in the magnetotail plasma sheet, J. Geophys. Res., 116, A05216, doi:10.1029/2010JA016316.

Russell, C. T. and R. L. McPherron (1973), The magnetotail and substorms, Space Sci. Rev., 15, 205.

Sagdeev, R.Z., and V.D. Shapiro (1973), Influence of transverse magnetic field on Landau damping. JETP Lett., 17, 279.

Sarris, E. T., S. M. Krimigis, and T. P. Armstrong (1976), Observations of magnetospheric bursts of high-energy protons and electrons at ∼35 RE with Imp 7, J. Geophys. Res., 81(13), 2341- 2355.

Sarris, T. E., X. Li, N. Tsaggas, and N. Paschalidis (2002), Modeling energetic particle injections in dynamic pulse fields with varying propagation speeds, J. Geophys. Res., 107(A3), 1033, doi:10.1029/2001JA900166.

Sauvaud, J.-A., and J. Winckler (1980), Dynamics of plasma, energetic particles, and fields near synchronous orbit in the nighttime sector during magnetospheric substorms, J. Geophys. Res., 85(A5), 2043–2056, doi:10.1029/JA085iA05p02043.

Schmid, D., et al. (2011), A statistical and event study of magnetotail dipolarization fronts, Ann. Geophys., 29, 1537.

Schriver, D., M. Ashour-Abdalla, and R. L. Richard (1998), On the origin of the ion-electron temperature difference in the plasma sheet, J. Geophys. Res., 103(A7), 14879–14895, doi:10.1029/98JA00017.

Schriver, D., et al. (2011), Electron transport and precipitation at mercury during the MESSENGER flybys: Implications for electron-stimulated desorption, Planetary and Space Science., 59 (15), 2026-2036.

182

Schulz, M., and L. Lanzerotti (1974), Particle Diffusion in the Radiation Belts, vol. 7, Springer- Verlag, New York.

Sergeev, V., V. Angelopoulos, S. Apatenkov, J. Bonnell, R. Ergun, R. Nakamura, J. McFadden, D. Larson, and A. Runov (2009), Kinetic structure of the sharp injection/dipolarization front in the flow-braking region, Geophys. Res. Lett., 36, L21105, doi:10.1029/2009GL040658.

Sergeev, V. A., V. Angelopoulos, J. T. Gosling, C. A. Cattell, and C. T. Russell (1996), Detection of localized, plasma-depleted flux tubes or bubbles in the midtail plasma sheet, J. Geophys. Res., 101(A5), 10,817–10,826, doi:10.1029/96JA00460.

Sergeev, V. A., V. Angelopoulos, and R. Nakamura (2012), Recent advances in understanding substorm dynamics, Geophys. Res. Lett., 39, L05101, doi:10.1029/2012GL050859.

Sergeev, V., R. C. Elphic, F. S. Mozer, A. Saint-Marc, and J.-A. Sauvaud (1992), A two-satellite study of nightside flux transfer events in the plasma sheet, Planet. Space Sci., 40, 1551– 1572.

Sergeev, V. A., M. A. Shukhtina, R. Rasinkangas, A. Korth, G. D. Reeves, H. J. Singer, M. F. Thomsen, and L. I. Vagina (1998), Event study of deep energetic particle injections during substorm, J. Geophys. Res., 103(A5), 9217–9234, doi:10.1029/97JA03686.

Sharber, J. R., and W. J. Heikkila (1972), Fermi acceleration of auroral particles, J. Geophys. Res., 77(19), 3397–3410, doi:10.1029/JA077i019p03397.

Sharma, A. S., et al. (2008), Transient and localized processes in the magnetotail: A review. Annales Geophysicae, 26(4), pp. 955-1006, Copernicus GmbH.

Shay, M. A., J. F. Drake, B. N. Rogers, and R. E. Denton (1999), The scaling of collisionless, magnetic reconnection for large systems, Geophys. Res. Lett., 26(14), 2163-2166.

Shay M.A., J.F. Drake, B.N. Rogers, R.E. Denton (2001), Alfvénic collisionless magnetic reconnection and the Hall term, J. Geophys. Res., 106, 3759–3772, doi:10.1029/1999JA001007.

Shinohara, I., T. Nagai, M. Fujimoto, T. Terasawa, T. Mukai, K. Tsuruda, and T. Yamamoto (1998), Low-frequency electromagnetic turbulence observed near the substorm onset site, J. Geophys. Res., 103(A9), 20365-20388.

Shiokawa, K., W. Baumjohann, and G. Haerendel (1997), Braking of high speed flows in the near- Earth tail, Geophys. Res. Lett., 24, 1179– 1182.

Sitnov, M. I., and M. Swisdak (2011), Onset of collisionless magnetic reconnection in two- dimensional current sheets and formation of dipolarization fronts, J. Geophys. Res., 116(A12).

183

Sitnov, M. I., M. Swisdak, A. V. Divin (2009), Dipolarization fronts as a signature of transient reconnection in the magnetotail, J. Geophys. Res., 114(A4).

Slavin, J. A., R. P. Lepping, J. Gjerloev, D. H. Fairfield, M. Hesse, C. J. Owen, M. B. Moldwin, T. Nagai, A. Ieda, and T. Mukai (2003), Geotail observations of magnetic flux ropes in the plasma sheet, J. Geophys. Res., 108(A1), 1015, doi:10.1029/2002JA009557.

Speiser, T.W. (1965), Particle trajectories in model current sheets, 1. Analytical solutions. J. Geophys. Res., 70, 4219–4226.

Speiser, T.W. (1967), Particle trajectories in model current sheets, 2. Applications to auroras using a geomagnetic tail model. J. Geophys. Res., 72, 3919–3932.

Sonnerup, B. U. Ö. (1979). Magnetic field reconnection. In Lanzerotti, L. J., Kennel, C. F., and Parker, E. N., editors, Solar System Plasma Physics, volume 3, pp. 46. North-Holland Publishing Co., Amsterdam.

Summers, D., and R. M. Thorne (1991), The modified plasma dispersion function, Physics of Fluids B: Plasma Physics, 3, 1835.

Terasawa T., A. Nishida (1976), Simultaneous observations of relativistic electron bursts and neutral-line signatures in the magnetotail, Planet. Space Sci., 24, 855-866, doi:10.1016/0032- 0633(76)90076-3.

Thomsen, M. F., J. E. Borovsky, D. J. McComas, and M. R. Collier (1998), Variability of the ring current source population, Geophys. Res. Lett., 25, 3481–3484.

Thomsen, M. F., J. Birn, J. E. Borovsky, K. Morzinski, D. J. McComas, and G. D. Reeves (2001), Two-satellite observations of substorm injections at geosynchronous orbit, J. Geophys. Res., 106(A5), 8405–8416, doi:10.1029/2000JA000080.

Tsyganenko, N. A. (1989), A magnetospheric magnetic field model with a warped tail current sheet, Planet. Space Sci., 37(1), 5-20.

Tverskoy, B.A. (1969), Main mechanisms in the formation of the Earth’s radiation belts, Rev. Geophys. Space Phys., 7, 219–231, doi:10.1029/RG007i001p00219.

Ukhorskiy, A. Y., M. I. Sitnov, V. G. Merkin, and A. V. Artemyev (2013), Rapid acceleration of protons upstream of earthward propagating dipolarization fronts, J. Geophys. Res. Space Physics, 118, 4952-4962, doi:10.1002/jgra.50452.

Vaivads, A., A. Retinò, Y. V. Khotyaintsev, and M. André (2011), Suprathermal electron acceleration during reconnection onset in the magnetotail, Ann. Geophys., 29, 1917-1925, doi:10.5194/angeo-29-1917-2011.

184

Vasyliunas, V. M. (1968). A survey of low-energy electrons in the evening sector of the magnetosphere with OGO 1 and OGO 3. J. Geophys. Res., 73(9), 2839-2884.

Vasyliunas, V. M. (1980), Upper limit on the electric field along a magnetic O line, J. Geophys. Res., 85(A9), 4616–4620, doi:10.1029/JA085iA09p04616.

Vekstein G.E., E.R. Priest (1995), Nonlinear magnetic reconnection with collisionless dissipation, Phys. Plasmas, 2, 3169–3178, doi:10.1063/1.871149.

Vu, H.X. and J.U. Brackbill (1992), CELEST1D: an implicit, fully kinetic model for low- frequency electromagnetic plasma simulation, Comput. Phys. Commun., 69, 253-276.

Wan, W., and G. Lapenta (2008), Evolutions of non-steady-state magnetic reconnection, Physics of Plasmas, 15 (10).

Wan, W., G. Lapenta, G. Delzanno, and J. Egedal (2008), Electron acceleration during guide field magnetic reconnection, Physics of Plasmas, 15 (3).

Wang, R., Q. Lu, X. Li, C. Huang, and S. Wang (2010), Observations of energetic electrons up to 200 keV associated with a secondary island near the center of an ion diffusion region: A Cluster case study, J. Geophys. Res., 115, A11201, doi:10.1029/2010JA015473.

Wilken, B., et al. (2001), First results from the RAPID imaging energetic particle spectrometer on board Cluster, Ann. Geophys., 19, 1355-1366.

Wiltberger, M., T. I. Pulkkinen, J. G. Lyon, and C. C. Goodrich (2000), MHD simulation of the magnetotail during the December 10, 1996, substorm, J. Geophys. Res., 105(A12), 27,649- 27,664, doi:10.1029/1999JA000251.

Wolf, R. A., Yifei Wan, X. Xing, J-C. Zhang, and S. Sazykin (2009), Entropy and plasma sheet transport, J. Geophys. Res., 114(A9).

Wu, M. Y.,Q. M. Lu, M. Volwerk, Z. Vörös, T. L. Zhang, L. C. Shan, and C. Huang (2013), A statistical study of electron acceleration behind the dipolarization fronts in the magnetotail, J. Geophys. Res., 118, 4804–4810, doi:10.1002/jgra.50456.

Zaharia, S., J. Birn, R. H. W. Friedel, G. D. Reeves, M. F. Thomsen, and C. Z. Cheng (2004), Substorm injection modeling with nondipolar, time-dependent background field, J. Geophys. Res., 109, A10211, doi:10.1029/2004JA010464.

Zaharia, S., C. Z. Cheng, and J. R. Johnson (2000), Particle transport and energization associated with substorms, J. Geophys. Res., 105, 18,741-18,752, doi:10.1029/1999JA000407.

Zelenyi L.M., M.S. Dolgonosov, E.E. Grigorenko, J. Sauvaud (2007), Universal properties of the nonadiabatic acceleration of ions in current sheets, JETP Lett., 85, 187–193 doi:10.1134/S0021364007040017.

185

Zeleny, L. M., A. S. Lipatov, D. G. Lominadze, and A. L. Taktakishvili (1984), The dynamics of the energetic proton bursts in the course of the magnetic field topology reconstruction in the earth‘s magnetotail, Planetary and Space Science, 32 (3), 313-324.

Zelenyi, L. M., J. G. Lominadze, and A. L. Taktakishvili (1990a), Generation of the energetic proton and electron bursts in planetary magnetotails, J. Geophys. Res., 95, 3883.

Zelenyi, L.M., D.V. Zogin, J. Büchner (1990b), Quasi-adiabatic dynamics of charged particles in the tail of the magnetosphere, Cosm. Res., 28, 369-380.

Zhang, X.-J., V. Angelopoulos, A. Runov, X.-Z. Zhou, J. Bonnell, J. P. McFadden, D. Larson, and U. Auster (2011), Current carriers near dipolarization fronts in the magnetotail: A THEMIS event study, J. Geophys. Res., 116, A00I20, doi:10.1029/2010JA015885.

Zhou, M., M. Ashour-Abdalla, X. Deng, M. El-Alaoui, R. L. Richard, and R. J. Walker (2011), Modeling substorm ion injection observed by the THEMIS and LANL spacecraft in the near- Earth magnetotail, J. Geophys. Res., 116, A08222, doi:10.1029/2010JA016391.

Zhou, M., M. Ashour-Abdalla, X. Deng, D. Schriver, M. El-Alaoui, and Y. Pang (2009), THEMIS observation of multiple dipolarization fronts and associated wave characteristics in the near- Earth magnetotail, Geophys. Res. Lett., 36, L20107, doi:10.1029/2009GL040663.

Zhou, M., X. Deng, M. Ashour-Abdalla, R. Walker, Y. Pang, C. Tang, S. Huang, M. El-Alaoui, Z. Yuan and H. Li (2013), Cluster observations of kinetic structures and electron acceleration within a dynamic plasma bubble, J. Geophys. Res., Space Physics, 118, 674–684, doi:10.1029/2012JA018323.

186