• No results found

Particle Acceleration in CME-driven Shocks: Study of Solar Energetic Particle Events as a Function of Shock Mach Number

N/A
N/A
Protected

Academic year: 2021

Share "Particle Acceleration in CME-driven Shocks: Study of Solar Energetic Particle Events as a Function of Shock Mach Number"

Copied!
94
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

U

NIVERSITY OF

A

MSTERDAM

MSc Physics and Astronomy

Track: Astronomy & Astrophysics

M

ASTER

T

HESIS

Particle Acceleration in CME-driven Shocks

Study of Solar Energetic Particle Events

as a Function of Shock Mach Number

by

Monika Pikhartov´a

12301469 (UVA)

60 ECTS

2nd of September 2019 - 31st of July 2020

Supervisors:

Jacco Vink

Dimitrios Kantzas

Examiners:

Jacco Vink

Shin’ichiro Ando

Anton Pannekoek

Institute

(2)
(3)

Contents

Popular Abstract

1

Abstract

3

Acknowledgements

5

1

Introduction

7

1.1

Cosmic Rays

. . . .

7

1.1.1

Solar Energetic Particles . . . .

9

1.2

Coronal Mass Ejections . . . .

10

1.3

Collisionless Shocks . . . .

13

1.4

Particle Acceleration . . . .

15

1.4.1

Second Order Fermi Acceleration Theory . . . .

15

1.4.2

Diffusive Shock Acceleration

. . . .

17

1.4.3

DSA Theory . . . .

17

1.4.4

Non-Linear Diffusive Shock Acceleration . . . .

20

1.4.5

Low Mach Number Shocks

. . . .

20

1.5

Project Goals . . . .

21

2

Advanced Composition Explorer

22

2.1

EPAM . . . .

25

2.2

SWEPAM . . . .

26

2.3

MAG

. . . .

26

3

Analysis

27

3.1

Data Selection and Preparation . . . .

27

3.2

Shock Notation . . . .

28

3.3

Compression Ratios and Mach Numbers . . . .

29

3.4

Power-Law Slopes of ACE Data . . . .

30

3.5

Thermal and Non-thermal Pressures . . . .

30

3.6

Acceleration Efficiency and Theoretical Non-Thermal Pressure . . . .

31

3.7

Theoretical Non-Thermal Pressure and Magnetic Field Pressure

. . . .

32

4

Results

34

4.1

Classification . . . .

35

(4)

Contents

4.2

Particle distributions

. . . .

41

4.2.1

Electron contribution . . . .

45

4.3

Acceleration efficiency . . . .

46

4.4

Pre-Existing Particles in Momentum Conservation Law . . . .

48

4.5

Magnetic Field Importance - Plasma Beta . . . .

51

4.6

Magnetic Field in Momentum Conservation Law . . . .

53

5

Discussion and Conclusions

56

5.1

General Remarks . . . .

56

5.2

Caveats . . . .

59

5.3

Highlights . . . .

60

Appendix

65

A Supplemental Figures

65

A.1 Slopes of Particle Distributions . . . .

65

A.2 Including Pre-Existing Particles and Magnetic Field

in Momentum Conservation Law . . . .

68

(5)
(6)

Popular Abstract

The nearest star to Earth, the Sun, affects us directly in various manners. It influences the life

on Earth, as well as in space. Despite the detailed studies for decades, many properties of

the Sun are still poorly understood. In particular, processes that are related to the activity of

the Sun, such as the coronal-mass ejections (CMEs; ejections of material that create shock

waves, disturbances, traveling through space), their launching mechanism, their influence

on the medium CME travel through, the acceleration of solar energetic particles (SEPs)

in the aforementioned shock waves and others, are complex and the basic models do not

sufficiently explain the underlying physics. SEPs arising from solar activity play a crucial

role in space technology and influence space missions. If these particles have very high

energy and penetrate into Earth’s atmosphere, they can even disable navigation systems or

cause storms on Earth that drop power grids. In this work, we concentrate on SEP events

and in particular, how are SEPs accelerated by the shock waves coming from the Sun. There

are two acceleration mechanisms that are generally considered. We study 19 years of data

from a spacecraft that orbits between the Sun and the Earth to distinguish which of these two

fundamental mechanisms dominates. We find that the data shows no preference for either

of the acceleration mechanisms we consider. We suggest including pre-existing accelerated

particles and magnetic field in the equations of conservation of mass and momentum that

describe the shock. We see that including pre-existing particles in the conservation equations

could improve the theoretical predictions of our data while including magnetic field would

actually diminish them.

(7)
(8)

Abstract

The nearest star to Earth, the Sun, affects us directly in various manners. It influences the life

on Earth, as well as in space. Despite the detailed studies for decades, many properties of the

Sun are still poorly understood. In particular, processes that are related to the activity of the

Sun, such as the coronal-mass ejections (CMEs), their launching mechanism, their influence

on the interplanetary medium, the acceleration of solar energetic particles (SEPs) in

CME-driven shocks and others, are complex and the basic models do not sufficiently explain the

underlying physics. SEP events play a crucial role in space technology and influence space

missions. In this work, we focus on the SEP events and in particular, on the acceleration

mechanisms of SEPs in CME-driven shocks. We consider the particle acceleration to occur

due to multiple shock crossings as explained by the diffusive shock acceleration (DSA). We

focus on the non-linear regime of this mechanism (NLDSA) and discuss its feasibility based

on data obtained by the Advanced Composition Explorer (ACE) spacecraft. We examine 19

years of data and discuss the connection between the macroscopic quantities of the

CME-driven shocks (e.g., flow temperature, velocity, pressure and magnetic field) and the

micro-scopical ones (e.g., the accelerated particle distributions and the acceleration efficiency). We

suggest that models of (NL)DSA should take into account macroscopic quantities like the

pressure of the pre-existing particles but it is a valid assumption to neglect the magnetic field.

Our data do not show any preference for DSA or NLDSA. We see no propensity of shock

acceleration efficiency or particle distribution on the velocity of the shock.

(9)
(10)

Acknowledgements

First and foremost, I would like to express my sincere gratitude to my advisor, Dr. Jacco

Vink. Throughout the whole year his advice, knowledge and ardor about the topic bolstered

me continuously.

I would like to offer a very special thank you to my daily supervisor, Dimitrios Kantzas

for always being there for me, supporting me when I needed or wanted it and even when I

did not. His endless enthusiasm and encouragement was what kept me going no matter the

hardships I was facing.

Thank you to my fellow classmates and Master room office mates for sharing your

knowl-edge and struggles with me and for the comic relief that was solely needed and appreciated.

I would like to specifically thank Bart van Baal for being my coding hero.

Lastly, I would like to thank to the rest of the Supernova Remnant research group for their

input and assistance throughout my research project.

(11)
(12)

Chapter 1

Introduction

1.1

Cosmic Rays

Cosmic rays (CRs) have been thoroughly studied ever since their discovery in 1912 by Victor

Hess. Victor Hess used balloon experiments of the atmosphere and found that the penetrating

radiation, whose origin has been long debated, had to be of an extraterrestrial origin. He

concluded that the Sun could not be its source due to lack of radiation decrease during night

and solar eclipses (

Rigden 2003

;

Hess 2018

).

CRs are a population of elementary particles and nuclei with energies that range from a

few MeV up to more than ∼ 10

10

GeV. The CR energy spectrum is a steep broken

power-law (

Tanabashi et al. 2018

). In Figure

1.1

adopted from

Evoli

(

2018

), we show the CR

energy flux as a function of the total energy of the particles. The gray shaded region

with a turnover shows the solar contribution to CRs. Protons dominate the low GeV

en-ergies. The contributions of other CR particles, such as protons, antiprotons, electrons, etc.,

are shown as well for comparison. The flux of these particles ranges between 10

−7

and

∼ 10

3

GeV m

−2

s

−1

sr

−1

.The proton flux is of the order of 1 cm

−2

s

−1

around GeV range.

Particles with energy more than ∼ 1 GeV are of extrasolar but still Galactic origin. These

Galactic CRs go up to the so called ‘knee’, which lies at ∼ 3 PeV. Up to this energy, the

spectrum follows a power-law with a spectral index of -2.7. Around the ‘knee’, the flux is

of the order of 1 m

−2

yr

−1

. Above the knee, the origin of the CRs is unclear but most likely

extragalactic. The spectral index of the power-law in this energy range is -3.1. The energy

range of the Large Hadron Collider is in this region. Around the so called ‘ankle’ (from ∼

3 PeV to ∼ 5 EeV), the flux is of the order of 1 km

−2

yr

−1

and the spectrum hardens

(flat-tens). The spectrum shows a cutoff at around ∼ 10

20

eV. The data were was collected by

various facilities both on the surface of Earth and in orbit (see

Evoli 2018

, and references

therein).

(13)

Cosmic Rays

F

IG

. 1.1

The CR energy spectrum as a function of the CRs’ total energy. The famous CR-spectrum features,

the ‘knee’ and the ‘ankle’ are indicated. The different colors indicate the different instruments collecting these

data in various energy ranges as shown in the legend. Adopted from

Evoli

(

2018

).

the CR spectrum (

Forbush 1946

). Supernovae (SN) and Supernova Remnants (SNRs;

Baade

& Zwicky 1934

), Pulsar Wind Nebulae (

Manconi et al. 2020

), X-Ray Binaries (

Cooper et al.

2020

) and clusters of massive stars (

Aharonian et al. 2018

) are considered the dominant

candidate sources of Galactic CRs. Active Galactic Nuclei, Gamma-Ray Bursts, clusters of

galaxies are among the dominant sources for extragalactic CRs (see e.g.,

Hillas 1984

;

Kotera

& Olinto 2011

;

Moskalenko & Seo 2018

;

Boezio et al. 2020

, and references therein).

Further open questions regarding CRs are: the CR origin; the CR composition, especially in

the high-energy regime where one experiment favors protons (

Hanlon 2019a

,

b

) while other

favors heavier elements (

Deligny 2019

;

The Pierre Auger Collaboration et al. 2019

;

Perrone

& the Pierre Auger Collaboration 2020

); the potential role that CRs play in the evolution

and the dynamics of stars and galaxies. The mechanism that drives the CR acceleration

(14)

Cosmic Rays

up to more than ∼ 10

10

GeV is another open question (

Drury 1983

). In this work we are

investigating the mechanism behind the acceleration of solar cosmic rays.

1.1.1

Solar Energetic Particles

Solar Energetic Particles (SEPs) are electrons, protons, and heavier nuclei associated with

Coronal Mass Ejections (CMEs; see below) and solar flares (see

Cane & Lario 2006

, and

references therein). In the CR spectrum, SEPs lay in the low energy region (up to GeV).

SEPs were discovered by

Forbush

(

1946

) in 1942 who noticed an increase of CR intensity

related to solar activity. SEP events are either impulsive or gradual. The impulsive events

are associated with solar flares. The gradual events are associated with CMEs and can last

up to days (

Li et al. 2005

).

We show the typical particle intensity of protons and electrons over time in Figure

1.2

adopted from (

Reames 1999

). The left panel shows a gradual event on 1981 December 5

in which a filament erupts from the Sun as a CME, without any accompanying solar flare.

The right panel shows events on 1982 August 13 and 14 that are associated with

impul-sive solar flares with no accompanying CMEs. The difference in the time scales of these

SEP events are clearly visible in the Figure. We see that the gradual event is dominated by

protons while the impulsive events are dominated by electrons.

F

IG

. 1.2

Particle intensity profile over time for a gradual SEP event on 1981 December 5 (left) and an

impulsive SEP event on 1982 August 13, 14 (right). The different points correspond to different particles as

depicted in the legend. Figure adopted from

Reames

(

1999

).

(15)

Coronal Mass Ejections

SEPs play a crucial role to the physical processes related to the space weather as they have

been considered a notorious space hazard. If these energetic particles hit a spacecraft or

penetrate astronauts’ spacesuit, they damage the equipment and can be hazardous to human

health (

Feynman & Gabriel 2000

;

Reames et al. 2001

). Shielding needs to be incorporated

into spacesuits and instrumental setups in order to safeguard against SEPs. SEPs can cause

storms in the Earth’s atmosphere that can disrupt power grids (see e.g.,

Stauning 2002

). All

these reasons are why we need to study, in particular: their origin, their composition, their

connection to CME-driven shocks and how they are accelerated.

1.2

Coronal Mass Ejections

Most strong SEP events are accelerated by CME shock waves (see e.g.,

Gargat´e et al. 2014

,

and references therein). During a CME, plasma carrying magnetic field is ejected from the

Sun’s corona into the interplanetary medium (IPM). About 10

15

g of plasma gets expelled

during a typical CME. The kinetic energy of this plasma can reach up to ∼ 10

32

erg with

plasma velocities in the range of 20–2500 km s

−1

(

Manchester et al. 2005

). CME-driven

shocks are considered efficient SEP accelerators. Type II radio bursts that often accompany

a CME shock are regarded as a sign for particle acceleration (

Gopalswamy et al. 2010

). Up

to 10% of the kinetic energy of a CME could go into energetic particles (

Mewaldt 2006

).

CMEs can significantly vary in time and can last from hours up to days (

Manchester et al.

2005

;

Webb & Howard 2012

). The frequency of CME occurrence follows the solar cycle

(

Webb & Howard 1994

; more CMEs during maximum solar activity and vice versa).

CMEs form as a result of an eruption of a magnetic flux rope through magnetic reconnection

in the Sun’s corona. A magnetic flux rope is a coherent magnetic structure with all magnetic

field lines twisting around a central, usually helical, axis. This magnetic flux rope can be

established during or even before to a solar eruption (see

Song & Yao 2020

;

Xing et al.

2020

, and references therein). We show a schematic of a magnetic flux rope in Figure

1.3

that was adopted from Professor Lang of the Tufts University. The CME ejection starts

as a prominence-like loop that stretches further and further away from the Sun. A shock is

formed at the front and is driven by the CME that carries the trapped hot plasma (grey shaded

region). As the magnetic field lines stretch with the shock travelling further away from the

Sun, lines going in opposite directions come closer and closer until finally they snap and

magnetic field reconnects. Magnetic reconnection takes place at the thick black vertical line.

The closed field region above the prominence (black shaded region) becomes the flux rope.

We show two images of a CME in Figure

1.4

. The images were taken by two of the

corona-graphs (C2 and C3) of the Large Angle and Spectrometric Coronograph (LASCO) on board

of the Solar and Heliospheric Observatory (SOHO). The images in the visible spectrum were

taken roughly ∼ 6 hours apart. We can see a “lightbulb” shape of the the blast in white color.

The field of view of LASCO C2 ranges from 1.5 to 6 solar radii, while LASCO C3 ranges

1

(16)

Coronal Mass Ejections

F

IG

. 1.3

An illustration of a CME formation adopted from Lang (2010)

1

.

from 3.5 to 30 solar radii (

Domingo et al. 1995

).

CME shock waves may interact with the Earth’s magnetosphere when travelling through

IPM towards Earth. When the CME shock carries high energetic SEPs, the interaction with

Earth’s magnetosphere could cause geomagnetic storms that affect for example electricity

grids, e.g. transportation, satellite navigation and others (see e.g.

Gosling et al. 1991

). In

Figure

1.5

we show an illustration of a CME and its subsequent interaction with the Earth’s

magnetic field. The left side of the Figure is a superimposed Extreme ultraviolet Imaging

Telescope (EIT) image, taken at the 30.4 nm wavelength, on a LASCO C2 optical image.

Both are on board of SOHO. The right side of the Figure is an artist’s impression of the

shock’s interaction with the Earth’s magnetic field. Magnetic field lines are shown in blue

color. Due to the impact these CME-driven shock waves may have on Earth, we are

moti-vated to study the acceleration of the SEPs that are driven by CME shocks.

2

https://sohowww.nascom.nasa.gov/gallery/images/las02.html

3

https://www.esa.int/ESA Multimedia/Images/2003/04/Coronal mass ejection CME blast

(17)

Coronal Mass Ejections

F

IG

. 1.4

CME observed on 2000 February 27 by LASCO C2 and C3 ∼ 6 hours apart. The white circle

denotes the optical sun, while the larger filled circle blocks the direct light. Credits: SOHO/LASCO consortium

2

.

(18)

Collisionless Shocks

1.3

Collisionless Shocks

Shocks are ubiquitous phenomena in the Universe. They can be described as transition layers

in which the flow and thermodynamic properties of the plasma, such as the bulk velocity or

the number density, rapidly change from one moment to another. They form when a

distur-bance propagates through a medium at a supersonic speed; i.e. supernova explosions, stellar

bow shock, Gamma Ray Bursts, Active Galactic Nuclei and other explosive astrophysical

phenomena.

The CME-driven shocks described above are collisionless shocks. Collisionless means that

energy is not dissipated through particle-particle collisions but by wave-particle interactions

instead (

Treumann 2009

;

Burgess & Scholer 2015

). The interaction cross-section between

particles is so low that collisions do not have an effect on the system.

In this work, we study non-relativistic collisionless CME-driven shocks. In Figure

1.6

, we

show six examples of CME-driven collisionless shocks detected by the Advanced

Composi-tion Explorer (ACE) spacecraft and studied by

Giacalone

(

2012

). In each of these six plots

the panels from top to bottom are: ion flux, magnetic field, flow velocity and bulk number

density. The vertical line indicates the shock crossing the spacecraft and clearly separates

the unshocked medium on the left side from the shocked medium on the right. The magnetic

field, flow velocity and number density of each of the shocks have the same profiles. Notable

is, however, how varied the shocks are in the flux (top panel). All of the shocks have an

in-crease in flux already before the shock, but the inin-crease can vary in length. After the shock

crossing, the flux either stays high or drops down. These drops in flux also vary in length.

(19)

Collisionless Shocks

F

IG

. 1.6

Examples of ion flux at two different energies (47 – 65 keV solid line, 65 – 112 keV dashed line),

magnetic field, flow velocity and number density (each panel from top to bottom) of CME-driven shocks. Figure

adopted from

Giacalone

(

2012

), data taken by ACE.

(20)

Particle Acceleration

1.4

Particle Acceleration

CRs gain high energies that reach up to 10

20

eV via various mechanisms.

Fermi

(

1949

) first

postulated a way of accelerating charged particles to such high energies. Particle acceleration

is the inevitable outcome of charged particles interacting with turbulent motions. This initial

idea, nowadays known as the second order Fermi acceleration, states that the net particle

en-ergy gain scales as (v/c)

2

, where v is the velocity of the perturbations and c is approximately

the velocity of the accelerated particles. The value of the exponent drives the name of this

acceleration mechanism. The resulting particle energy spectrum follows a power-law. Since

the velocity of the shock is much less than the speed of light (v  c), this mechanism is not

very efficient.

1.4.1

Second Order Fermi Acceleration Theory

Let us first discuss the second order Fermi acceleration, which was also historically presented

first.

Fermi

(

1949

) postulated that CRs are accelerated primarily in the interstellar space by

collisions against moving magnetic fields. These moving magnetic fields are interstellar

clouds. Nowadays, these clouds are interpreted as shock waves.

Following

Longair

(

2011

) and

Uroˇsevi´c et al.

(

2019

), we consider a particle with mass m,

kinetic energy E and total energy ε = m c

2

+E moving through a medium.When the particle

collides with a cloud that moves along the x-axis with speed v

s

the particle is reflected on

the pitch angle θ. Right before the collision, the energy and the momentum of the particle in

the shock (moving) frame of reference are:

ε

0

= Γ (ε + v

s

p

x

) ,

(1.1)

p

0

x

= Γ (p

x

+ v

s

ε/c

2

) ,

(1.2)

where Γ ≡ (1 − v

s

2

/c

2

)

−1/2

is the Lorentz factor of the shock. The collision is elastic

and therefore, the initial total energy of the particle in the shock frame is conserved (ε

0

i

=

ε

0

f

). The particle moves in the opposite direction, thus p

0

x,i

= −p

0

x,f

. Using the Lorentz

transformation, the total energy of the particle after the collision in the lab frame is:

ε

f

= Γ (ε

0

+ v

s

p

0

x

) = Γ

2

(ε + 2 v

s

p

x

+ v

2

s

ε/c

2

) .

(1.3)

And since p = Γ m v and ε = Γ m c

2

, we can write that:

p

x

ε

=

p cos θ

ε

=

v cos θ

c

2

.

(1.4)

(21)

Particle Acceleration

Γ = (1 − v

s

2

/c

2

)

−1/2

≈ 1 + (v

s

/c)

2

, we can write:

ε

f

ε

=

1 + 2 (v

s

/c) (v/c) cos θ + (v

s

/c)

2

1 − (v

s

/c)

2

≈ 1 + 2 (v

s

/c) (v/c) cos θ + (v

s

/c)

2

.

(1.5)

Assuming a random distribution of angles θ and neglecting the rest energy in the relativistic

regime of accelerated particles, the average kinetic energy gain per collision is:

 ∆E

E



=

 E

f

− E

E



'

8

3

 v

s

c



2

.

(1.6)

The energy gain of the particle scales as the square of the velocity of the shock (v

s

/c)

2

and

that is why this process is called “second-order” Fermi acceleration. If the average time

between collisions is 2L/c, where L is the mean free path of the particle, the rate of energy

increase is:

dE

dt

=

4

3

 v

2

s

cL



E ≡ α E .

(1.7)

Since the particle propagates away from the shock-front, the actual acceleration can only

continue for time τ , while the particle stays within the accelerating region.

To derive the resulting energy spectrum of the accelerated particles, we solve the kinetic

equation (for derivation see Chapter 17.3 of

Longair 2011

):

dN (E)

dE

= −



1 +

1

α τ

 N (E)

E

,

(1.8)

which leads to

n(E) = C

2

× E

 α τ + 1

α τ



,

(1.9)

where n is the number of particles at a certain energy and C

2

is a constant. The particle

energy spectrum of the second order Fermi acceleration mechanism follows a power-law.

The second order Fermi acceleration mechanism is inadequate for explaining multiple

as-pects. The theory does not properly explain why the energy spectrum has the observed slope

of around -2 for strong shocks (see e.g.,

Vink 2012

, and references therein). The random

velocities of the postulated clouds and the mean free path for CR scatterings are

compara-tively low and that leads to a slow, almost impossible particle acceleration (

Longair 2011

).

This acceleration mechanism served instead as a foundation for later studies of particle

ac-celeration, and for establishing the first order Fermi acceleration theory (

Axford et al. 1977

;

Krymskii 1977

;

Bell 1978a

,

b

;

Blandford & Ostriker 1978

;

Drury 1983

).

(22)

Particle Acceleration

1.4.2

Diffusive Shock Acceleration

The essence of first order Fermi acceleration, or Diffusive Shock Acceleration (DSA), is that

charged particles of the shock-heated plasma move diffusively and consecutively cross the

shock front head-on due to magnetic field irregularities. This repeating process ultimately

results in an energy gain that scales as v/c, making DSA more efficient than the second order

Fermi acceleration (

Bell 1978a

). The resulting particle energy spectrum follows a power-law

that scales as E

−q

.

1.4.3

DSA Theory

A more efficient mechanism for acceleration of CRs is DSA or first order Fermi acceleration.

Essential to further understanding of DSA is to describe the conditions around the

CME-driven shock. The physical properties and the structure of the shock, both in the unshocked

and shocked region, can be described by the Rankine-Hugoniot relations (

Landau & Lifshitz

1987

;

Landau & Sykes 1987

;

Shu 1992

;

Nieuwenhuijzen et al. 1993

). These relations treat

the shock macroscopically as a single discontinuous jump and are:

ρ

0

v

0

= ρ

2

v

2

,

(1.10)

ρ

0

v

0

2

+ P

0

= ρ

2

v

2

2

+ P

2

,

(1.11)

h

0

+

1

2

v

2

0

= h

2

+

1

2

v

2

2

.

(1.12)

where ρ is the mass density of the fluid, v is the velocity of the flow, P is the total pressure,

and h is the enthalpy. The equations are in the frame co-moving with the shock. Plasma

enters the shock with v

0

= v

sh

and moves away from the shock with v

2

(shocked). These

equations describe the conservation of mass, momentum and energy across the shock,

re-spectively. The equations show that no mass accumulates in the shock layer and the kinetic

energy converts to enthalpy. The enthalpy for a perfect gas satisfies the following relation:

h =

γ

γ − 1

P

ρ

=

γ

γ − 1

kT

m

= E +

P

ρ

.

(1.13)

γ is the adiabatic index, k is the Boltzmann constant, T is the temperature, m is the

elemen-tary mass of the crossing particles, and E is the internal energy. By taking into account the

magnetic field of the upstream and the downstream regions, we can include extra terms for

the magnetic field pressure and energy density in the above equations (

Kennel et al. 1989

;

see Section

3.7

for more detail).

A schematic diagram from

Longair

(

2011

) shows the dynamics of the high energy particles

in proximity of a shock (see Figure

1.7

). The top left depiction shows a strong shock wave

propagating at a velocity U that is faster than the speed of the medium. The density, pressure

and temperature (ρ, P, and T, respectively) differ between the unshocked (subscript 1) and the

(23)

Particle Acceleration

shocked medium (subscript 2). The conditions can be described by the Rankine-Hugoniot

re-lations (see Equations

1.10

1.12

). The top right depiction shows the shock reference frame.

Equation

1.10

of the Rankine-Hugoniot relations, the ratio of unshocked to shocked region

velocity is (γ + 1)/(γ − 1). For a fully ionized non-relativistic plasma, γ = 5/3, that gives

a ratio of 4. In the bottom left depiction is the frame of reference where the upstream gas

is stationary, showing how the velocity in the shocked region equals to 3U /4 (we later call

that v

s

). The bottom right depiction is the frame of reference where the shocked plasma is

stationary.

F

IG

. 1.7

Top left: A shock moving with a speed U . Pressure, density and temperature in the unshocked

medium (p

1

, ρ

1

, T

1

) and in the shocked plasma (p

2

, ρ

2

, T

2

). Top right: Using the Rankine-Hugoniot relations,

the ratio between the unshocked and shocked region velocities is v

1

/v

2

= 4 because γ = 5/3. Bottom left:

Reference frame where the unshocked gas is stationary. The velocity of the downstream flow is 3/4 U. Bottom

right: Reference frame where the shocked gas is stationary. The velocity of the upstream flow is then 3/4 U.

Figure adopted from

Longair

(

2011

).

We assume that high energy particles are in both unshocked and shocked regions of the

shock. When such a particle crosses the shock front from the unshocked to the shocked

region, it encounter the gas behind the shock that moves at speed of 3v

s

/4. The particle

scatters due to magnetic field irregularities and gains energy. This gain scales as ∝ v

s

/c.

When on the other hand, a particle crosses from the shocked to the unshocked region, it

again encounters a gas there that moves at speed of V = 3v

s

/4 and again gains the same

amount of energy. This process can repeat multiple times allowing for the particles to attain

high enough energies.

(24)

Particle Acceleration

crossing into the shocked region in the shock frame is:

ε

0

= Γ (ε + V p

x

) ,

(1.14)

where V is the shocked plasma approaching the particle (V = 3/4v

s

), the x-axis is

perpen-dicular to the shock front, the shock is non-relativistic (V  c, Γ = 1). The particles are

relativistic, therefore, E = p c, p

x

= (E/c) cos θ and:

∆E = p V cos θ and

∆E

E

=

V

c

cos θ ,

(1.15)

where ∆E is the difference in energy in each shock crossing. The average increase in energy

can be written as:

 ∆E

E



=

2

3

V

c

.

(1.16)

From classical kinetic theory (see Chapter 17.4 of

Longair

(

2011

) for details), n c/4 is the

flux of particles crossing the shock in either direction, where n is the number density of the

particles. Particles in the shocked region are advected or swept away from the shock at a rate

nV = nv

s

/4 creating a fraction of particles that escape per unit time, v

s

/c. Since only a

small fraction of particles is lost per cycle, the probability that a particle remains within the

accelerating region after one collision is P

e

= 1 − (v/c). After some number of collisions

k, there are n

f

= n

i

P

e

k

particles with energies E

k

= E

i

β

k

, where β is a constant connected

to the energy gain (β = 1 + 4V /(3c) in one round trip). We can eliminate k by taking the

logarithm of both n

f

/n

i

and E

k

/E

i

:

ln(n

f

/n

i

)

ln(E

k

/E

i

)

=

ln P

e

ln β

,

(1.17)

which results in:

n

f

n

i

=

 E

k

E

i



ln P

e

/ ln β

.

(1.18)

For infinite number of collisions (k → ∞) we neglect the subscript and write E

k

= E. From

the above equation we can obtain the energy spectrum of the accelerated particles:

n(E) dE = C

1

× E

−1+(ln P

e

/ ln β)

dE .

(1.19)

If we now use the values of P

e

and E

f

we derived earlier, we can get the differential

power-law spectrum of:

n(E) dE = E

−q

dE ,

(1.20)

where q = (χ + 2)/(χ − 1) (q = 2 for a strong shock χ = 4) and we considered that v

is also a function of compression ratio. We have to stress that the above derivation is for

relativistic particles (

Bell 1978a

). For the non-relativistic particle limit the spectrum gets as

flat as q = 1.5 (

Asvarov et al. 1990

;

Caprioli 2014

).

(25)

Particle Acceleration

1.4.4

Non-Linear Diffusive Shock Acceleration

When the accelerated particles carry a sizable fractions of pressure and energy density of

the shock they start affecting the dynamics of the system. This is the non-linear regime of

DSA (NLDSA;

Drury 1983

;

Caprioli & Haggerty 2019

). This theory suggests that particles

with higher energies diffuse further away from the shock forming the so called precursor

and slowing down the upstream plasma. This precursor results in an additional pressure in

the unshocked region which compresses the plasma and makes the shock-jump smaller (less

strong). Accelerated particles also enhance the magnetic field strength and turbulence that

have a back reaction to the accelerated particles.The spectrum of the accelerated particles

does not follow a power-law anymore but is concave instead. In particular, at low energies,

the spectrum becomes steeper and at higher energies the spectrum becomes flatter than the

DSA’s prediction of q = 2 (for strong shocks and relativistic particles).

Nevertheless, the local interstellar spectra (Galactic CRs) measured by Voyager 1 shows a

power-law with ∼ E

−1.45

at low energies (kinetic energy below GeV; see e.g.,

Stone et al.

2013

;

Potgieter 2014

, and references therein). This is in agreement with simulations of

ac-celerated non-relativistic Galactic CRs that follow a distribution of E

−1.5

(see e.g.,

Caprioli

& Spitkovsky 2014

). Observations of electrons (see e.g.,

Lin et al. 1982

) and simulations

of solar CR spectra (see e.g.,

Kong et al. 2019

, and references therein) show a variability in

the power-law spectrum between ∼ E

−0.6

and ∼ E

−4.3

. The above-mentioned power-law

slopes are a hint for NLDSA.

When magnetic field amplification is not important, the high energy particles stream further

upstream, experiencing larger compression, thus having a flatter distribution than DSA. On

the other hand, when magnetic field amplification becomes important, the particle energy

spectrum experiences the opposite effect. The lower energies become flatter, while higher

energies become steeper (

Caprioli 2012

).

1.4.5

Low Mach Number Shocks

Shock acceleration is tightly connected to the properties of the shock where particles

accel-erate. A fundamental quantity that describes the shock is the Mach number that is defined

as the ratio of the flow velocity v with respect to the local speed of sound c

s

. The larger

the value of Mach number, the “stronger” the shock. When studying interstellar shocks

de-tected in situ, it is useful to derive the Mach number from measurable quantities. We show

in Section

3.3

that a good approximation for the Mach number is

M =

s

γ + 1 − χ(γ − 1)

,

(1.21)

where γ is the adiabatic index (5/3). The compression ratio χ is the ratio of the number

density in the shocked over the unshocked medium.

(26)

Project Goals

An open question is whether shocks with low Mach number are associated with efficient

par-ticle acceleration.

Gopalswamy et al.

(

2010

) indicates that there is a possible critical Mach

number under which particle acceleration does not happen. It is the Alfv´en Mach number

M

A

≈ 2.7, where M

A

is defined as the ratio of the flow velocity with respect to the Alfv´en

velocity (v

A

= B (4 π ρ)

−1/2

, where B is magnetic field and ρ is the bulk density).

Lee et al.

(

2009

) found a critical Mach number of 2.46 with particle-in-cell simulations for the solar

wind termination shock (at the point in the heliosphere where the solar wind slows down

to subsonic speed because of its interactions with the local ISM). The termination shock is

also associated with particle acceleration (see e.g.,

Florinski et al. 2013

).

Giacalone

(

2012

)

presented that, from observations, CME-driven shocks with χ > 2.5 show evidence of

parti-cle acceleration.

Vink & Yamazaki

(

2014

) analytically derived a similar critical lower Mach

number that should exist for DSA, below which shocks do not accelerate particles. This

critical value of Mach number is

5 (that corresponds to χ = 2.5) if magnetic fields are not

dynamically important. If magnetic field pressure dominates the pressure in the unshocked

medium, the critical Mach number would then be 2.5. More recently,

van Marle

(

2020

)

performed a series of hybrid simulations (magnetohydrodynamics and particle-in-cell) for

collisions of two galaxy clusters and the subsequent collisionless shock that is formed.

van

Marle

(

2020

) found that no DSA occurs in shocks with a Mach number below 2.25.

1.5

Project Goals

Life on Earth is directly affected by the Sun. Accelerated SEPs propagating towards Earth

can depreciate equipment or instruments in space. They can also cause geomagnetic storms

on Earth that can disrupt global electrical power grids or navigation systems. The physical

processes that are related to these particles, e.g. the CME launching mechanism, the

influ-ence of CME-driven shocks to the ISM, the acceleration of SEPs in these shocks and other,

are complex and not yet fully understood. Nevertheless, it is crucial to study them further

so as to be able to tackle SEP events better in the future. The analytical models describing

these processes are still fairly basic using many assumptions and approximations.

The main objective of this work is to investigate the mechanisms behind acceleration of SEPs

due to CME-induced non-relativistic collisionless shocks using real data from the Advanced

Composition Explorer (ACE) spacecraft (see Chapter

2

for details). We investigate SEP

ac-celeration, its relationship to the shock Mach number and the disposition toward (NL)DSA.

We consider the connection between the microscopical and macroscopical quantities of the

shock and test whether including microscopic quantities into analytical models could

im-prove their reliability. We specifically examine whether including pre-existing particles and

magnetic field into the shock conservation relations (Equations

1.10

1.12

) increases the

ac-curacy of theoretical predictions. We inspect the data for a “critical” Mach number as we

introduced in the above section. In the absence of such a “critical” Mach number, we

inves-tigate whether shocks of smaller Mach number accelerate particles less efficiently.

(27)

Chapter 2

Advanced Composition Explorer

F

IG

. 2.1

Illustration of the ACE spacecraft. Figure adopted from the ACE Photo Gallery

1

.

The Advanced Composition Explorer (ACE) (see Figure

2.1

), being part of the NASA

Ex-plorer program, was launched in 1997 and is still active in 2020, exceeding the initial

per-spective of a 5 year mission. ACE is in orbit at the L1 Lagrangian point of the Sun-Earth

sys-tem, ∼ 0.01 au sun-ward of the Earth. The instruments on board the spacecraft include six

high-resolution spectrometers and three additional monitoring instruments. Together they

study the elemental, isotopic, and ionic composition of energetic nuclei in interplanetary

space, at energies ranging from ∼ 1 keV/nucleon (solar wind) to ∼ 0.5 GeV/nucleon

(Cos-mic Radiation). This includes ions accelerated at different places: the Sun, ISM, the edge of

the heliosphere (bow shock), the Galaxy (for Galactic sources see Section

1.1

). The list of

all the instruments and their positions on board ACE is in Table

2.1

and Figure

2.2

.

We use data from ACE because the spacecraft is in a unique position for us to study particles

in situ. ACE also measures in a uniquely low energy regime which is what our work

con-centrates on. For this work, we combine data from three of the instruments on board ACE:

1

http://www.srl.caltech.edu/ACE/Gallery/gallery.html

2

http://www.srl.caltech.edu/ACE/ASC/level2/index.html

3

(28)

Chapter 2. Advanced Composition Explorer

Acronym

Full Name

Type of Investigation

Energy

Range

CRIS

Cosmic Ray Isotope

Spectrometer

Isotope composition of cosmic rays

100–500

MeV/nuc.

EPAM

Electron, Proton and

Alpha Monitor

Energetic particles across a broad

range of energies

0.04-93

MeV

MAG

Magnetic Field

Exper-iment

Local magnetic field

SEPICA

Solar Energetic

Parti-cle Charge Analyser

Ionic charge states of energetic

parti-cles

0.1–5

MeV/nuc.

SIS

Solar

Isotope

Spec-trometer

Isotopic composition of energetic

nuclei from He to Ni

5–150

MeV/nuc.

SWICS

Solar Wind Ion

Com-position Spectrometer

Chemical and ionic charge state

composition of the solar wind, H and

He isotopes only

0.5–100

keV/Q

SWIMS

Solar Wind Ion Mass

Spectrometer

Chemical and isotopic composition

of the solar wind from He to Ni

0.5–20

keV/Q

SWEPAM

Solar Wind Electron,

Proton,

and

Alpha

Monitor

Solar wind plasma electron and ion

fluxes

0.001–36

keV

ULEIS

Ultra Low Energy

Iso-tope Spectrometer

Ion fluxes of suprathermal and

ener-getic particle ranges of H through Ni

0.02–14

MeV/nuc.

T

ABLE

2.1 List of scientific payload on board the ACE spacecraft, their purpose and the energy range where

they operate. Q is the charge and ‘nuc.” stands for nucleon. For this project, we use EPAM, SWEPAM and

MAG. Information taken from the ACE Science Center (ASC) website

2

.

EPAM, SWEPAM and MAG. All the data have been obtained from the ACE Science Center

(ASC) public archives (

Garrard 1997

;

Garrard et al. 1998

). In this Chapter, we discuss the

the instruments we use in more depth.

(29)

Chapter 2. Advanced Composition Explorer

F

IG

. 2.2

Exploded model of the ACE spacecraft and all its instruments on board. Figure adopted from the

ACE Photo Gallery

3

.

(30)

EPAM

2.1

EPAM

The Electron, Proton and Alpha Monitor (EPAM) is composed of two Low Energy Foil

Spec-trometers (LEFS), LEFS60/LEFS150, two Low Energy Magnetic SpecSpec-trometers (LEMS),

LEMS30/LEMS120, and Composition Aperture (CA), CA60. Each of the numbers after the

acronyms denote the orientation of the instrument with respect to the spacecraft spin-axis in

degrees. As the spacecraft spins, these five instruments sweep out the area around the

tele-scope, presenting nearly full coverage for ions and approximately 40% coverage for

elec-trons, where the electronics automatically break out the sample evenly (

Gold et al. 1998

).

The different time resolutions of EPAM are 12 s, 5 min, hourly and daily averages. LEFS

measure the flux and directions of electrons between ∼44 keV and ∼ 4.9 MeV. LEMS

mea-sure the flux and directions of ions between approximately 46 keV and 4.8 MeV. There is

also a Deflected Electrons (DE) detector, or DE30, measuring electrons that are being

de-flected from the LEMS30 detector by a rare-earth magnet. CA measures ion composition

between ∼ 0.4 and 92.7 MeV. Each of these energy ranges is split up between two or more

energy channels (

Stone et al. 1998

;

Hawkins 1999

). The large dynamic range is crucial for

understanding the dynamics of solar flares, interplanetary shock acceleration and

propaga-tion through heliosphere, and helps support the other instruments on board of ACE as well

(

Gold et al. 1998

).

LEMS30

LEMS120

Channel

Energy Range (keV)

Mean Energy (keV)

Energy Range (keV)

Mean Energy (keV)

1

46-67

56

47-86

56

2

67-115

88

68-115

88

3

115-193

149

115-195

150

4

193-315

247

195-321

250

5

315-580

427

310-580

424

6

580-1060

784

587-1060

789

7

1060-1880

1412

1060-1900

1419

8

1880-4700

2973

1900-4800

3020

T

ABLE

2.2 EPAM channels and their specific energy ranges. Information from

Hawkins

(

1999

).

DE30

Channel

Energy Range (keV)

Mean Energy (keV)

1

38-53

45

2

53-103

74

3

103-175

134

4

175-315

235

T

ABLE

2.3 DE30 channels and their specific energy ranges. Information from

Hawkins

(

1999

).

For this work, we use all the 16 LEMS30/LEMS120 and all the 4 DE30 channels (see Tables

2.2

2.3

for specific energy ranges) at the best possible resolution the instrument provides,

which is 12 seconds. The data is given in average particle intensities.

(31)

MAG

2.2

SWEPAM

The Solar Wind Electron, Proton, and Alpha Monitor (SWEPAM) measures the bulk flow

and kinetic properties of the solar wind. Some of the primary objectives of SWEPAM are

determining coronal composition, studying solar plasma conditions and solar wind

accelera-tion, and constraining particle acceleration models. SWEPAM has two electrostatic

analyz-ers that measure electrons and ions for studying these phenomena. Both analyzanalyz-ers measure

the energy per charge of a particle by bending their flight paths, where the ions are detected

between 0.26 and 36 keV and electrons between 0.001 and 1.35 keV (

McComas et al. 1998

;

Stone et al. 1998

).

SWEPAM data is available in 64 s, hourly or daily averages. We select the best possible

resolution, which is 64 s. We use the provided bulk proton number density, velocity and

temperature. The proton velocity SWEPAM measures is also determined with respect to

the GSE (Geocentric Solar Ecliptic)

4

, GSM (Geocentric Solar Magnetospheric)

5

, and RTN

(Radial Tangential Normal)

6

coordinates (

Stone et al. 1998

). We neglect the direction

infor-mation in lieu of making the simplest model possible.

2.3

MAG

The Magnetic Field Experiment (MAG) measures the local magnetic field in the IPM. MAG’s

observations facilitate studies of propagation and evolution of thermal and energetic particle

populations because MAG measures the large-scale structure of the interplanetary magnetic

field (IMF), its fluctuation and turbulence. MAG consists of two duplicate fluxgate

magne-tometers that are attached to booms extending outwards from opposite sides of the spacecraft

(

Smith et al. 1998

).

The time resolution of the data is in 1 s, 16 s, 64 s, 4 min, hourly, and daily averages. MAG

also measures the magnetic field with respect to the GSE, GSM and RTN coordinates. We

again neglect these measurements. Since we combine the data with those of SWEPAM,

which has a 64-second resolution, we select 64 s for MAG as well. We download the data

in bulk from the ACE Science Center website as MAGSWE merged datasets of MAG and

SWEPAM(

Stone et al. 1998

). From the MAG part of the dataset we use the average magnetic

field estimates.

4

GSE coordinate system is defined as having an origin at the Earth, positively increasing towards the Sun

(x-axis). The z-axis has a direction of the North Ecliptic Pole. The y-axis is the cross product of the x- and

y-axes.

5

GSM coordinate system has the same x-axis as the GSE coordinate system. The z-axis is the projection of

the Earth’s magnetic dipole axis onto the yz plane. The y-axis is again the cross product of the x- and y-axes.

6

RTN is the spacecraft centered coordinate system. The r-vector (x-axis) goes from the Sun towards the

spacecraft. The y-axis (t-vector) is the cross product of the solar rotational axis and the x-axis (r-vector). The

z-axis (n-vector) is once more the cross product of the x- and y-axes

(32)

Chapter 3

Analysis

3.1

Data Selection and Preparation

The selection of CME-driven shocks we survey in this work is based on the ACE Lists of

Disturbances and Transients

observed by ACE (

Smith 2016

)

1

. The MAG team has reported

the detection of 609 shocks between 1997 and 2016. We use all the identified 609 shocks

and the identified times of shock crossings that are when the ACE spacecraft encountered

each individual shock. The data are acquired from the ACE Science Center (ASC) public

archives (

Garrard 1997

;

Garrard et al. 1998

)

2

in 24-hour blocks of EPAM and MAGSWE

(merged MAG and SWEPAM data) in the time resolution of 12 s for EPAM and 64 s for

MAGSWE.

Whenever ACE was encountering difficulties on taking measurements that led to missing

or invalid data, the collaboration set the values to -9999.9 or -999.9 to distinguish from the

proper data. We replace these values by a python value describing empty measurement,

numpy.nan

or NaN (Not A Number), so as to be able to perform numerical mathematical

analysis with these data.

ACE does not provide instrumental errors for their measurements. All the uncertainties we

provide are statistical errors derived from error propagation or 1σ errors when discussing

mean values.

1

http://www.ssg.sr.unh.edu/mag/ace/ACElists/obs list.html

2

(33)

Shock Notation

F

IG

. 3.1

Pressure profile of a shock. Diagram adapted from Astroparticle Physics Lectures at the University

of Amsterdam 2019.

3.2

Shock Notation

When a CME-driven shock propagates through the IPM it encounters the material of its

medium. The shock signature is evident in the density measurements (see for example the

last panels of Figure

1.6

). The density is mostly unperturbed and when the shock crosses,

it sharply increases creating an abrupt jump in the measurements. The same signature is

present in the temperature, velocity, and the pressure profiles as well. We refer to the

undis-turbed plasma before the shock crossing as the upstream region, and the shocked plasma

after the shock crossing as the downstream region. When particles accelerate and stream

towards the upstream region, a precursor is formed (see Figure

3.1

and Section

1.4

). In this

work, we refer to the upstream medium as subscript 0, the precursor region as subscript 1,

and the downstream medium as subscript 2. Sometimes, the region 2 can be referred to as

the “sub-shock” or “shock”. In this work, we use the term “shock”.

In Figure

3.1

we show the profile of a “perfect” shock. In particular, we plot the pressure as

a function of space. In region 0 (upstream), there is only thermal pressure of the medium.

In region 1 (precursor), the pressure is enhanced because of the shock-driven accelerated

particles. In region 2 (downstream), the pressure is increased and is the sum of the gas and

accelerated particles. The particle intensity follows a similar profile (see Section

4.1

).

When considering the upstream, precursor, shock crossing, and downstream regions, we

select specific time frames from which we take the respective data. We establish the upstream

region from the beginning of each 24-hour block to avoid any precursor “contamination”.

Even in the absence of a precursor (see Section

4.1

), the subscript of the upstream remains

0. The time of the shock crossing t

sh

is provided by ACE. We select the precursor region to

(34)

Compression Ratios and Mach Numbers

be averaged over 60 or 30 min before the shock crossing. In that way we avoid data from

the far upstream region. We choose the downstream region to be averaged over the 60 or

30 min after the shock crossing. The default choice is 60 min, whereas the 30 min option is

exclusive to shocks that are close to either edge of the dataset.

3.3

Compression Ratios and Mach Numbers

In Section

1.4.3

we discussed how the conditions upstream and downstream can be described

by the Rankine-Hugoniot relations. Using the conservation of mass (Equation

1.10

), we

define the compression ratio χ as the ratio of the velocities in the upstream and downstream

region:

χ ≡

ρ

2

ρ

0

=

v

0

v

2

.

(3.1)

The Mach number is usually defined as the velocity of the medium over the local speed of

sound c

s

. We define it as (

Landau & Lifshitz 1987

):

M ≡

v

c

s

=

r

v

γ

P

ρ

.

(3.2)

By using the conservation of mass and momentum, we relate the compression ratio to the

Mach number as:

χ =

(γ + 1)M

2

(γ − 1)M

2

+ 2

,

(3.3)

M =

s

γ + 1 − χ(γ − 1)

,

(3.4)

where the Mach number is the same as in Equation

1.21

. To calculate the compression ratio

based on Equation

3.1

, we used the SWEPAM number density data. We averaged the density

data over one hour before and after the shock crossing to eliminate any fluctuations in our

sample. If the shock crossing was within one hour of the beginning or the end of the 24-hour

block data, we only averaged half-an-hour before and after the shock. From Equation

3.4

we

get the shock Mach numbers, assuming γ = 5/3 for a non-relativistic flow. We discuss all the

results of how the shocks vary with respect to the Mach number or compression ratio and

other quantities in the following Chapter. The compression ratio values range between 1 and

∼10 and Mach numbers between 1 and ∼32.

(35)

Thermal and Non-thermal Pressures

3.4

Power-Law Slopes of ACE Data

To understand how effective the particle acceleration is and the mechanism behind it, we

study the particle distributions in the precursor and the shock. We calculate the differential

number density and energy density of both the thermal (SWEPAM) and non-thermal (EPAM)

data.

From dimensional analysis we extract the differential non-thermal number density to be (see

also

Longair 2011

)

n

i

= 4π

I

i

v

i

,

(3.5)

where I

i

is the particle intensity and v

i

for particle velocity. To get the thermal differential

number density, we divide the number density from SWEPAM by E, which is the mean

energy of the instrument (18.13 keV). To calculate the energy density we multiply the

differ-ential number density by E

2

. For the non-thermal data, all the energy bands and their mean

energies are shown in Table

2.2

.

We use the three moments, the time of the shock crossing t

sh

and 60 or 30 min before and

after t

sh

for precursor and downstream regions, respectively (as described in Section

3.2

), to

study the particle distribution in the differential number density and energy density. The

non-thermal particles follow a power-law (see Equation

1.20

) as described in Section

1.4

. Based

on the non-thermal data, we calculate the power-law slopes q, using a linear least-squares

regression method in python, scipy.stats.linregress().

3.5

Thermal and Non-thermal Pressures

We get the temperature T as well as the number density of the thermal particles n directly

from SWEPAM. We derive the thermal pressure of the gas by using the ideal gas law,

P

th

= nkT ,

(3.6)

where k is the Boltzmann constant.

The calculation of the non-thermal pressure is more complicated. EPAM measures only

particle intensity which we use to derive the non-thermal pressure as

P

nth

= (γ − 1)

X

i

n

i

E

i

dE

i

,

(3.7)

where n

i

is the number density per energy bin, E

i

is the average energy of each energy bin

(36)

Acceleration Efficiency and Theoretical Non-Thermal Pressure

and each of them has a set mean energy and size (Table

2.2

;

Hawkins 1999

). We calculate

the differential number density n

i

from Equation

3.5

. The particles detected by EPAM are

non-relativistic so the velocity for each energy bin is expressed in terms of the kinetic energy

from the well known expression

v

i

=

r

2E

i

m

,

(3.8)

where m is the mass of the particles. We do not study heavy elements in this work so we

take the mass to be the rest mass of the proton. Combining Equations

3.7

3.8

we write the

non-thermal pressure as

P

nth

= (γ − 1)

8

X

i=1

4πI

i

r

mE

i

2

dE

i

.

(3.9)

We calculate both P

th

and P

nth

for all the shocks we study. We compute the pressure values

for the upstream, precursor, t

sh

, and downstream regions (as described in Section

3.2

).

3.6

Acceleration Efficiency and Theoretical Non-Thermal

Pres-sure

By examining both the thermal and accelerated particle distributions we can learn about the

microscopic properties of the CME-driven shocks. We can investigate the macroscopic

quan-tities of the gas pressure and the pressure due to the accelerated particles in the shock. We

can connect the microscopic and macroscopic quantities by studying the particle efficiency

of the shock. We define the acceleration efficiency as the fraction of the non-thermal pressure

over the total pressure (

Vink & Yamazaki 2014

):

ω ≡

P

nth

P

nth

+ P

th

=

P

nth

P

tot

,

(3.10)

where we use P

tot

= P

nth

+ P

th

assuming that the magnetic field pressure is negligible.

The acceleration efficiency expresses how much of the total pressure of the shock is in the

non-thermal population. This expression assumes that there are no pre-existing accelerated

particles (P

nth,0

= 0). If we assume that nearly all non-thermal particles come from particle

acceleration, the acceleration efficiency discerns how effective the shock is at accelerating

particles. To calculate the acceleration efficiency, we first need to interpolate the thermal

pressure, so the resolution of both pressures is the same and the data have the same size

(see Chapter

2

for details). To do that we use the simplest python interpolate function,

numpy.interp(), which does a one-dimensional linear interpolation.

(37)

Theoretical Non-Thermal Pressure and Magnetic Field Pressure

Based on the above definition, we write the acceleration efficiency for both the upstream and

the downstream regions as

ω

0

=

P

nth,0

P

tot,0

and ω

2

=

P

nth,2

P

tot,2

,

(3.11)

respectively.

Based on the conservation of momentum along the shock (Equation

1.11

) and Equation

3.1

we express the total pressure in the downstream in terms of the total pressure in the upstream

as

P

tot,2

= P

tot,0

+



1 −

1

χ



ρ

0

v

2

.

(3.12)

We express the Mach number in terms of the total pressure by combining Equations

3.4

and

3.10

:

M

2

=

1

γ

ρ

0

v

2

(1 − ω

0

) P

tot,0

.

(3.13)

Combining the last Equations

3.10

3.13

we express the predicted non-thermal pressure

downstream P

nth,2

as a function of the non-thermal pressure upstream P

nth,0

, acceleration

efficiency ω for both regions, the compression ratio χ, the mach number M and the adiabatic

index γ:

P

nth,2

= P

nth,0

ω

2

ω

0



1 +



1 −

1

χ



(1 − ω

0

) γ M

2



.

(3.14)

In Section

4.4

, we specifically use the calculated values of upstream, precursor and

down-stream for the acceleration efficiency of each shock. We also compare the result of Equation

3.14

that shows the theoretical predicted value of the downstream non-thermal pressure to

the value we derive from the data based on Equation

3.9

.

3.7

Theoretical Non-Thermal Pressure and Magnetic Field

Pres-sure

In the above we neglected the pressure exerted by the magnetic field on the flow. In the

following we want to examine the importance of this term in the evolution of the

CME-driven shock and the accelerated particles. We first define the magnetic field pressure as

P

mag

=

B

2

Referenties

GERELATEERDE DOCUMENTEN

In dit voorstel wordt een motivering gegeven voor de aanschaf van apparatuur voor het digitaal verwerken van beelden.. Verder wordt een aantal van de belangrijkste eisen

If the laser-produced resonant plasma wave is strongly driven, the energy distribution of accelerated electrons can be very narrow, a process which has an analogy with water waves: in

My contribution will exist in taking resources, institutions, and ideas into account when looking if power imbalances between Northern and Southern NGOs exist within

The most influential and accepted models are the intention-based models and more specifically the Entrepreneurial Event Theory and the Theory of Planned Behavior. Due to

In the final analysis, it seems that educators should, on the one hand, concentrate on helping learners understand the nature of human rights as a universal values

Daarnaast blijkt dat van de vaders die wel eens een vraag hebben over seksualiteit dat vaders die een gesloten seksuele opvoeding hebben gehad vaker een vraag hebben over hoe zij op

Nowadays, there is an intense research activity in designing systems that operate in real life, physical environments. This research is spanned by various ar- eas in computer

To test a traditional image search interface acting as a control interface, an image retrieval system is required that is able to search using text.. To test the RF interface, the