You are on page 1of 167

Arenberg Doctoral School of Science, Engineering & Technology

Faculty of Science
Department of Mathematics
Centre for Plasma Astrophysics

On regular shock refraction in hydro-


and magnetohydrodynamics

Peter Delmont

Dissertation presented in partial


fulfillment of the requirements for
the degree of Doctor
in the Sciences

Leuven, September 2010


On regular shock refraction in hydro-
and magnetohydrodynamics

Peter Delmont

Supervisor: Dissertation presented in partial


Prof. Dr. R. Keppens fulfillment of the requirements for
Jury: the degree of Doctor
Prof. Dr. Ir. G. Lapenta (Chairman) in the Sciences
Prof. Dr. S. Poedts (Secretary)
Prof. Dr. M. Goossens
Prof. Dr. W. Van Assche
Prof. Dr. Ir. T. Baelmans
Prof. Dr. Ir. B. Koren (CWI, Amsterdam)

Leuven, September 2010


© Katholieke Universiteit Leuven – Faculty of Science
Celestijnenlaan 200 B, B-3001 Leuven (Belgium)

Alle rechten voorbehouden. Niets uit deze uitgave mag worden vermenigvuldigd
en/of openbaar gemaakt worden door middel van druk, fotocopie, microfilm,
elektronisch of op welke andere wijze ook zonder voorafgaande schriftelijke
toestemming van de uitgever.

All rights reserved. No part of the publication may be reproduced in any form by
print, photoprint, microfilm or any other means without written permission from
the publisher.

Wettelijk depot D/2010/10.705/49


ISBN 978-90-8649-356-2
Preface

Four years ago, Tom Van Doorsselaere invited me for his public PhD defense at
the Arenberg castle in Heverlee. He gave a presentation which I didn’t understand
at all, but which made me conclude that plasma physics is a mysterious but
interesting discipline. A few months later, I applied for a PhD position myself,
with this dissertation as a result. I would like to thank everybody who contributed
to it in any way. More explicitly, my gratefulness goes out to the following people.
First of all, I sincerely thank Prof. Rony Keppens. He has always been available
with good advice and critical suggestions, but at the same time he gave me the
freedom to focus on my own interests. He has shown patience with my specific
talent for creating chaos. I am sure that his guidance has led me to scientific
contributions and insights, which would never have been possible without it.
Also I thank Dr. Tom Van Doorsselaere for being my helpdesk. In the first two
years of this project, I have often annoyed him with technical computer and physics
problems. He has been very patient and was (almost) always willing to help. Next
to him, also Dr. Dries Kimpe, Dr. Bart van der Holst, Dr. Zakaria Meliani and
Dr. Allard Jan van Marle have offered me good technical support.
I have experienced the CPA, and the department of mathematics, as a pleasant
working environment. I appreciate the nice conversations with my former office
mates Prof. Andria Rogava, Dr. Emmanuel Chané, Prof. Giga Gogoberidze,
Giorgi Dalakishvili, Andrey Divin and Alkis Vlasis. I thank the Alma/Coffee
Team for the agreeable lunch/coffee breaks. Further, the CPA deserves a word
of gratefulness for the stimulating international atmosphere.
Next to the research, I have enjoyed being a teaching assistant for the
courses "Hogere Wiskunde I", "Hogere Wiskunde II" and "Wiskunde voor
bedrijfseconomen" in the didactic team led by Prof. Johan Quaegebeur. I
appreciate the didactic insights I gained through this experience. It was a pleasure
teaching mathematics to the bachelor students in economical engineering.

i
ii Contents

Of course I thank the members of the examination committee for investing time
to read this work and giving useful comments, interesting suggestions and critical
feedback. I hope they are pleased by the way I have processed their input.
Since the bow should not be bent too often, I like spending time with my friends
in various ways. I think about the many times we went out drinking and the many
times we shared our thoughts, about the walks and the dinners we shared and the
darts and table games we played.
I thank my family for the mental support along the way.
Finally, Char, you complete this list. I am grateful for the inspiration and energy
you’ve given me during this period by my side. I truly appreciate your way of
looking at life, and I especially love how it balances my own approach.
Contents

Contents ii

1 Introduction 1
1.1 The equations of magnetohydrodynamics . . . . . . . . . . . . . . 1
1.1.1 Plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Fluid mechanics . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.3 Adiabatic equation of state . . . . . . . . . . . . . . . . . . 5
1.1.4 Electromagnetism . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.5 Magnetohydrodynamics . . . . . . . . . . . . . . . . . . . . 7
1.1.6 Scale independence of the MHD equations . . . . . . . . . . 10
1.2 Shock waves in (M)HD . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2.1 The Rankine-Hugoniot jump conditions . . . . . . . . . . . 10
1.2.2 Characteristic speeds . . . . . . . . . . . . . . . . . . . . . . 13
1.2.3 Characteristic speeds in HD . . . . . . . . . . . . . . . . . . 15
1.2.4 Characteristic speeds in ideal MHD . . . . . . . . . . . . . 16
1.2.5 Wave steepening: an HD example . . . . . . . . . . . . . . 17
1.2.6 Wave steepening: an MHD example . . . . . . . . . . . . . 18
1.3 The Richtmyer-Meshkov instability . . . . . . . . . . . . . . . . . . 20
1.3.1 Shock tube problems . . . . . . . . . . . . . . . . . . . . . . 23
1.3.2 Shock refraction: a 1D example . . . . . . . . . . . . . . . . 25

iii
iv CONTENTS

1.3.3 Richtmyer-Meshkov instability . . . . . . . . . . . . . . . . 31


1.3.4 Vorticity evolution equation for inviscid compressible MHD
flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.4 AMRVAC and the VTK file format . . . . . . . . . . . . . . . . . . . 35
1.4.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.4.2 Adaptive Mesh Refinement . . . . . . . . . . . . . . . . . . 36
1.4.3 The VTK file format . . . . . . . . . . . . . . . . . . . . . . 37
1.5 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2 An exact Riemann solver for regular shock refraction 41


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.2 Configuration and governing equations . . . . . . . . . . . . . . . . 43
2.2.1 Problem setup . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.2.2 Stationary MHD equations . . . . . . . . . . . . . . . . . . 44
2.2.3 Planar stationary Rankine-Hugoniot condition . . . . . . . 45
2.3 Riemann Solver based solution strategy . . . . . . . . . . . . . . . 46
2.3.1 Dimensionless representation . . . . . . . . . . . . . . . . . 46
2.3.2 Relations across a contact discontinuity and an expansion fan 48
2.3.3 Relations across a shock . . . . . . . . . . . . . . . . . . . . 51
2.3.4 Shock refraction as a Riemann problem . . . . . . . . . . . 51
2.3.5 Solution inside of an expansion fan . . . . . . . . . . . . . . 52
2.4 Implementation and numerical details . . . . . . . . . . . . . . . . 53
2.4.1 Details on the Newton-Raphson iteration . . . . . . . . . . 53
2.4.2 Details on AMRVAC . . . . . . . . . . . . . . . . . . . . . . 53
2.4.3 Following an interface numerically . . . . . . . . . . . . . . 54
2.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.5.1 Fast-Slow example . . . . . . . . . . . . . . . . . . . . . . . 55
2.5.2 Slow-Fast example . . . . . . . . . . . . . . . . . . . . . . . 57
CONTENTS v

2.5.3 Tracing the critical angle for regular shock refraction . . . . 57


2.5.4 Abd-El-Fattah and Hendersons experiment . . . . . . . . . 59
2.5.5 Connecting slow-fast to fast-slow refraction . . . . . . . . . 60
2.5.6 Effect of a perpendicular magnetic field . . . . . . . . . . . 61
2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3 Shock Refraction in ideal MHD 67


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.3 MHD shocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.4 Riemann Solver based solution strategy . . . . . . . . . . . . . . . 74
3.4.1 Dimensionless representation . . . . . . . . . . . . . . . . . 74
3.4.2 Path variables . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.4.3 Tackling the signals . . . . . . . . . . . . . . . . . . . . . . 80
3.5 Demonstration of result . . . . . . . . . . . . . . . . . . . . . . . . 80
3.6 Suppression of the Richtmyer-Meshkov Instability . . . . . . . . . . 82
3.7 Wave configuration transitions . . . . . . . . . . . . . . . . . . . . 83
3.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

4 Parameter ranges for intermediate shocks 87


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.1.1 Intermediate shocks in MHD . . . . . . . . . . . . . . . . . 87
4.1.2 The Rankine-Hugoniot jump conditions . . . . . . . . . . . 89
4.1.3 MHD shock types: classical 1 − 2 − 3 − 4 classification . . . 90
4.2 Solution to the Rankine-Hugoniot conditions . . . . . . . . . . . . 91
4.3 Physical meaning of Ω . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
4.4.1 (θ, M )-diagrams at fixed β . . . . . . . . . . . . . . . . . . 96
vi CONTENTS

4.4.2 Equivalence classes introduced by the RH conditions . . . . 99


4.4.3 Positive pressure requirement . . . . . . . . . . . . . . . . . 101
4.4.4 Switch-on shocks and switch-off shocks . . . . . . . . . . . . 104
4.4.5 Parameter ranges for 1 → 3 shocks. . . . . . . . . . . . . . . 106
4.4.6 Parameter ranges for 2 → 3 shocks. . . . . . . . . . . . . . . 109
4.4.7 Parameter ranges for 1 → 4 shocks . . . . . . . . . . . . . . 110
4.4.8 Parameter ranges for 2 → 4 shocks . . . . . . . . . . . . . . 111
4.4.9 Relationship between pre- and post-shock magnetic field . . 112
4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

5 Nederlandstalige samenvatting 115

6 Conclusions 121

A Structure of AMRVAC 123

B Compilation 125

C USR-file 127

D Stationary planar Rankine-Hugoniot conditions 129

E Relations across a shock 131

F Solving the cubic analytically 135

G Integration across rarefaction waves 141

Bibliography 145

Curriculum Vitae 153


Education and Research History . . . . . . . . . . . . . . . . . . . . . . 153
List of Scientific Contributions . . . . . . . . . . . . . . . . . . . . . . . 153
Chapter 1

Introduction

1.1 The equations of magnetohydrodynamics

1.1.1 Plasma

In 1927, the term plasma was introduced by Langmuir & Jones [52] to refer to
ionized gases, which are characterized by their two-fluid nature consisting of freely
moving electrons and ions. They were probably inspired by the similarity of this
fluid with the blood plasma, consisting of red and white corpuscles. At first,
we define a plasma as a completely ionized gas. Since the freely moving ions
and electrons are charged particles allowing for electric currents, electromagnetic
forces will play a central role in the description of a plasma, which is therefore
fundamentally different from the description of ordinary gases. It leads us to
consider plasma as a fourth phase, next to the well-known phases of solid, liquid
and gas.
As shown in figure 1.1, plasmas do exist in the physical world under a wide
range of temperatures and number densities. In fact, as Goedbloed & Poedts
[32] put it: "Astronomers agree that, ignoring the more speculative nature of
dark matter, matter in the Universe consists more than 90% of plasma. Hence
plasma is the ordinary state of matter in the Universe." The environment in which
plasmas occur allows us to classify them into two major categories, namely (i)
astrophysical or (ii) laboratory plasmas. At the Centre for Plasma Astrophysics,
research is done on, e.g., (relativistic) astrophysical jets, coronal mass ejections
(CME’s) or space weather. Figure 1.2 shows an example of those typical plasma
environments. The left frame shows a satellite (TRACE) observation of the solar
atmosphere (taken from [6]). The picture was taken on November 9th, 2000, one

1
2 INTRODUCTION

Figure 1.1: Plasmas exist in the physical world under a wide range of plasma
parameters. They mostly occur in astrophysical or laboratory environments
(Taken from [67]).

day after a solar storm. The right panel shows a laboratory plasma, namely
a confined plasma created within the TEXTOR tokamak (taken from [28]). In
such tokamaks, research on controlled nuclear fusion is performed. The typical
values for characteristic parameters in these different environments can differ by
orders of magnitude, but the theory of magnetohydrodynamics (MHD) describes
the dynamical behaviour of any plasma, satisfying certain conditions which we
describe later on.
The Sun is by astrophysicists sometimes refered to as an excellent natural
plasma laboratory. Since its distance to the earth is only one astronomical
unit, it is the star best observed. This human interest in the sun is
nothing recent: next to worshiping it, many cultures before ours have made
observations of the sun and have studied it. Ghezzi & Ruggles [34] discovered
e.g. a 2300-year old solar observatory in coastal Peru. Despite this old
interest in the sun, which mainly focused on large scale phenomena such as
gravity, plasma physics is a relatively young discipline. This of course is not
THE EQUATIONS OF MAGNETOHYDRODYNAMICS 3

Figure 1.2: Two typical plasma environments. Left: A TRACE observation of a


coronal loop on the solar surface. (taken from [6]) Right: A picture taken in the
tokamak TEXTOR in Jülich. (taken from [28])

surprising, since in order to describe the large scale motion of a plasma, one
uses the equations of ideal magnetohydrodynamics. Ideal MHD is in essence
a generalization of hydrodynamics (HD), where one allows for non-vanishing
magnetic fields. Therefore MHD is basically founded on (i) fluid mechanics and
(ii) electromagnetism.

1.1.2 Fluid mechanics

Fluid mechanics is a discipline with a long history. While in the ancient Greek era,
Archimedes mainly focused on equilibria, medieval Arab physicists like Al-Bı̄rūnı̄
and Al-Khazini augmented this work with dynamical aspects. Fluid mechanics
reached the west through da Vinci, and finally, it was Bernoulli who introduced
a mathematical description in his Hydrodynamica (1738). This led Euler to
publish his "Principes generaux du mouvement des fluides" in 1757, where he
introduced one of the first published systems of partial differential equations,
nowadays referred to as the Euler equations. The Euler equations in their original
form describe conservation of mass and momentum, and are therefore not fully
determined. Indeed, there are only two equations in three independent variables.
We therefore need a third equation to close the system. This equation is called the
equation of state. At first, one added the equation of incompressibility to close the
system, but later on also more realistic equations of state for gaseous environments
were introduced. One often exploit the so-called adiabatic equation of state, which
we will briefly describe in the next section. A derivation of these Euler equations
can be found in any textbook on fluid mechanics. When the adiabatic equation is
4 INTRODUCTION

used as an equation of state, the Euler equations take the following form:
∂U ∂F ∂G ∂H
+ + + = 0, (1.1)
∂t ∂x ∂y ∂z
where the vector of conserved variables
!t
ρ(vx2 + vy2 + vz2 ) p
U= ρ, ρvx , ρvy , ρvz , + (1.2)
2 γ−1

and we introduced flux terms:


 
ρvx
2

 ρvx +p


F=
 ρv x vy ,

(1.3)
 ρv x vz 
2
+vy2 +vz2 )
   
ρ(vx γ
vx 2 + γ−1 p

 
ρvy
 ρvx vy 
ρvy2 + p
 
G= , (1.4)
 
 ρvy vz 
2
+vy2 +vz2 )
   
ρ(vx γ
vy 2 + γ−1 p

and
 
ρvz

 ρvx vz 

H=
 ρvy vz .

(1.5)
 ρvz2 + p 
2
+vy2 +vz2 )
   
ρ(vx γ
vz 2 + γ−1 p

These equations should hold at any time over the complete domain. t represents
time and the ratio of specific heat, γ, is a dimensionless equation parameter
(explained in a little more detail in the next section), but the meaning of ρ, v
and p are not that straightforward. All those quantities are defined statistically
by the following reasoning in phase space. Every particle at any time t has an
exact position r = (x, y, z) and velocity w = (wx , wy , wz ) (in any arbitrary
chosen Cartesian frame). In the classical case, it is usual to define the phase
space now as R6 , such that it contains a combination of locations and velocities
(x, y, z, wx , wy , wz ). One then defines a distribution function f : R7 → R+ , such
THE EQUATIONS OF MAGNETOHYDRODYNAMICS 5

that in any infinitesimal volume element, drdv, in phase space, the probable
number of particles, N , in that volume element at time t is given by
Z Z
N (t) ≡ f (r, w, t)drdw. (1.6)

The exact definition of mass density ρ, velocity v and thermal pressure p are now
given by averaging only over the velocity dimensions of phase space:
Z
ρ(r, t) ≡ Mp f (r, w, t)dw, (1.7)
R
wf (r, w, t)dw
v(r, t) ≡ R , (1.8)
f (r, w, t)dw

Mp
Z
p(r, t) ≡ (w − v)2 f (r, w, t)dw, (1.9)
3

where Mp is the mass of a single particle. When the gas consists of multiple
different particles (as is the case in MHD) a generalization has to be made. For
this we refer to any textbook on fluid mechanics. The Euler equations thus express
the conservation of mass (i.e. the first line), the conservation of momentum
(mx , my , mz ) ≡ m ≡ ρv = (ρvx , ρvy , ρvz ) (i.e. the next three lines) and the
2
ρ(vx +vy2 +vz2 ) p
conservation of energy e ≡ 2 + γ−1 (i.e. the fifth line).

1.1.3 Adiabatic equation of state

In 1834, Clapeyron was the first to formulate the ideal gas law, which is valid
under relatively mild conditions. The equation, as it reads nowadays, is given by

pV = nRT. (1.10)

Here p denotes the thermal pressure as before, V is the volume of the gas element,
n is the amount of gas, R is the universal gas constant and T is the absolute
temperature. It follows that under isothermal conditions pV is constant. Now,
taking derivatives of equation (1.10) leads to

pdV + V dp = nRdT. (1.11)

When a gas element expands, it decreases its internal energy and performs work.
Now, the adiabatic equation of state concerns, as the name suggests, adiabatic
processes where no heat is absorbed. The first law of thermodynamics is thus
valid for any volume element of gas, stating that the internal energy of a gas
6 INTRODUCTION

element balances out with the work done on it. The loss of internal energy in the
gas element is proportional to the amount of gas and the change in temperature.
The proportionality constant cv is called the specific heat for constant volume.
The work done by the gas element equals pdV . Therefore the first law of
thermodynamics reads

ncv dT + pdV = dQ = 0. (1.12)

Combining equations (1.10) and (1.11) gives now

cv + R dV dp
+ = 0. (1.13)
cv V p

The proportional constant cvc+Rv


is by definition nothing else than the ratio of
specific heats, γ, which we introduced in the previous section.
We are now ready to define the entropy S. Consider a gas element with initial
pressure p0 and initial volume V0 . Say the gas has expanded and forms a state of
pressure p1 and volume V1 . Then we define the entropy S
p1 V1
S ≡ cv ln + (cv + R) ln (1.14)
p0 V0
It is clear that the adiabatic equation of state now reduces as dS = 0.
Anyhow, this law should clearly only hold when the expansion is continuous.
This text is also concerned about discontinuous expansion, where the second law
of thermodynamics should hold: dS ≥ 0. This last expression is completely
d p p
equivalent with dt ργ ≥ 0, and from now on, we will refer to ργ as being the
entropy S.

1.1.4 Electromagnetism

According to Stringari & Wilson [79], one of the first published experiments
that proved the relation between electricity and magnetism was performed by
Romagnosi in 1802 (and thus not by Oersted, who experimented in the 1820’s).
This led to a cascade of strong discoveries throughout the nineteenth century,
which in 1865 eventually led to Maxwell’s theory of electromagnetism. Maxwell’s
THE EQUATIONS OF MAGNETOHYDRODYNAMICS 7

equations, in their most exploited form, are given by

τ
∇·E = , (1.15)
ǫ0
∇·B = 0, (1.16)

∂B
∇×E = − , (1.17)
∂t
∂E
∇×B = µ0 ǫ0 + µ0 j. (1.18)
∂t

In these equations the constants ǫ0 and µ0 respectively are the permittivity of


vacuum (ǫ0 = 8.85·10−12F ·m−1 ) and the permeability of vacuum (µ0 = 4π10−7 T ·
m · A−1 ). τ denotes the charge density, E is the electric field, j the current density,
and finally B is the magnetic field. The first equation is called Poisson’s law,
while the second equation implies that no magnetic monopoles exist. The third
equation is called Faraday’s law, and the last equation is Ampère’s law, extended
with the displacement current distribution. Since µ0 ǫ0 = c−2l << 1(s2 m−2 ) (where
cl denotes the light speed), one often drops the displacement current ǫ0 ∂E ∂t from
Ampère’s law, which then becomes ∇ × B = µ0 j.

1.1.5 Magnetohydrodynamics

Since a plasma is according to our first definition defined as an ionized gas,


magnetic fields play an important role in plasmas and should not be neglected. In
order to describe the dynamical behavior of such plasmas, one needs to generalize
the Euler equations by including a magnetic field. The mathematical relation
between the magnetic field and the conserved HD variables is given by the ideal
MHD equations: after noting that a current is in direct relation with the speed
of the electrons, one can eliminate the current j from Ampère’s law. Extending
Euler’s equations with the Lorentz force and Maxwell’s equations leads to the
equations of ideal MHD. For details we refer to Goedbloed & Poedts [32], Goossens
[36] or any introductory textbook on MHD. In this dissertation we will consistently
use the conservational form of the ideal MHD equations, which in Cartesian vector
notation becomes:

∂U ∂F ∂G ∂H
+ + + = 0, (1.19)
∂t ∂x ∂y ∂z
8 INTRODUCTION

where the vector of conserved variables

!t
ρ(vx2 + vy2 + vz2 ) p Bx2 + By2 + Bz2
U= ρ, ρvx , ρvy , ρvz , + + , Bx , By , Bz
2 γ−1 2µ0

(1.20)

and we introduced flux terms:


ρvx
 
Bx2 By2 +Bz2
ρvx2 + p − + 2µ0
 
 2µ0 
 Bx By 

 ρvx vy − µ0 

 ρvx vz − Bµx B z 
F=
 ρ(vx2 +vy2 +vz2 ) 2 2
0 , (1.21)
γ B +B (vy By +vz Bz )Bx 
 (
 2 + γ−1 p + y2µ0 z )vx − µ0



 0 

 vy Bx − vx By 
vz Bx − vx Bz

ρvy
 
B B

 ρvx vy − µx 0 y 

B2 B 2 +B 2
ρvy2 + p − 2µy0 + x2µ0 z
 
 
B B
 

G= ρvy vz − µy 0 z 
, (1.22)
 ρ(vx2 +vy2 +vz2 ) γ B 2 +B 2 (v B +v B )B

 (
 2 + γ−1 p + x2µ0 z )vy − x x µ0z z y 


 vx By − vy Bx 

 0 
vz By − vy Bz

and
ρvz
 

 ρvx vz − BxµB0
z 

By Bz

 ρvy vz − µ0 

 B2 B 2 +B 2 
 ρvz2 + p − 2µz0 + x2µ0 y 
H=

2 2 2 2 2
. (1.23)
 ( ρ(vx +vy +vz ) + γ p + By +Bz )vz − (vx Bx +vy By )Bz


 2 γ−1 2µ0 µ0 

 vx Bz − vz Bx 

 vy Bz − vz By 
0

As before, ρ is the mass density, v = (vx , vy , vz ) is the velocity, p is the thermal


pressure and B = (Bx , By , Bz ) denotes the magnetic field. Since these conserved
THE EQUATIONS OF MAGNETOHYDRODYNAMICS 9

variables are often used, they have a particular name and notation. As before, the
ρ(v 2 +v 2 +v 2 )
p B 2 +B 2 +B 2
momentum m = ρv and the total energy density e ≡ x
2
y z
+ γ−1 + x 2y z .
The first line of the MHD equations expresses the conservation of mass. The next
three lines express the conservation of momentum, taking the Lorentz force into
account. The fifth line expresses the conservation of energy, which includes the
magnetic energy, and the last three lines finally express Faraday’s law, where the
electric field is eliminated by use of Ohm’s law, which under ideal (non-resistive)
MHD conditions reduces to E = −v × B. The ideal MHD system consists now of
eight partial differential equations in the eight primitive variables ρ, v, p, B. Once
the values of these variables are found, the pre-Maxwell equations are used to
solve for the electric field E and the current j. Note that when the magnetic field
vanishes, the MHD equations reduce to the Euler equations. Finally also note that
as in the Maxwell equations, the no-monopole equation is still an initial condition.
Thus

∇·B=0 (1.24)

completes the ideal MHD equations.


Finally note that some conditions need to be satisfied, in order to validate this
macroscopic plasma description, namely

• The interaction with neutrals should be small compared to the interaction


of charged particles. Therefore we conclude that typical time scales should
be negligible in comparison with the time scales of collisions with neutrals
or

long-range Coulomb interactions must dominate over short-range


interactions;

• the plasma should be quasi-neutral. When using for the typical lengthqscales
ǫ0 T
of particle interaction the Debye length, which is defined as λD ≡ ne e2
(where ne is the electron density, T is the plasma temperature and e is the
electron charge), we conclude that

the typical length scales should be much larger than the Debye length;

• and finally, the particles should be close enough to each other such that
there are enough charged particle interactions. Defining a Debye sphere as
a sphere with radius λD , this means that

there have to be enough charged particles in a Debye sphere.


10 INTRODUCTION

1.1.6 Scale independence of the MHD equations

Since the MHD equations are scale independent, one can give a dimensionless
representation. We will do so, following Goedbloed & Poedts [32]. They argue
that one needs to make three independent scalar scalings. One can e.g. scale the
magnetic field components to a typical value B̂, the length to a typical length l̂,
and the density to a typical density ρ̂, i.e.:

ρ
ρ̄ ≡ , (1.25)
ρ̂
B
B̄ ≡ , (1.26)

¯ l
l ≡ . (1.27)

In order to get rid of the inconvenient constant µ0 , we now scale the velocities
to v̂ ≡ √B̂µ0 ρ̂
, such that the times are scaled to t̂ ≡ v̂l̂ . We arbitrarily scale p to
2
p̂ ≡ ρ̂v̂ . Finally, from the scaling of length and time, it is now trivial that the
the differential operators ∇ and ∂t are respectively scaled to ∇ ¯ ≡ l̂∇ and t̂∂t . In
other words: if one defines the dimensionless variables p̄ ≡ pp̂ and v¯ι ≡ vv̂ι , the
MHD equations can now be re-written in exactly the same form as 1.19, but now
every symbol ♦ in equations (1.19)-(1.23) should now be replaced by ♦, ¯ and all
the µ0 ’s should be dropped. From now on we will exploit this dimensionless form,
but we will consistently drop the bars from the notation, and exploit the newly
introduced dimensionless variables. Recovering the physical values from the set of
dimensionless variables is now straightforward (taking into account that t = t̂t̄).

1.2 Shock waves in (M)HD

1.2.1 The Rankine-Hugoniot jump conditions

Our research deals with shocks in HD and ideal MHD. Let us briefly discuss
the concept of characteristic speeds of a (hyperbolic) system of conservation laws
(such as HD and ideal MHD), and its connection to so-called weak solutions to
the system. For simplicity, we restrict our analysis to the one-dimensional case.
But first let us mention that shock waves are relevant. Figure 1.3 shows two
shocks. The left frame (taken from [73]) shows the HD shock wave which caused
the Sailor Hat Explosion crater on the Hawaiian island Kahoolawe, in a US army
experiment in 1965. A movie of the experiment can be seen on [74]. The middle
and right pictures of figure 1.3 are taken by the SOHO satellite observatory (taken
SHOCK WAVES IN (M)HD 11

Figure 1.3: The left picture shows a hydrodynamical shock wave during a US army
experiment (taken from [73]). The middle and right picture show a solar MHD
shock wave (taken from [6]).

from [6]) and shows an MHD shock wave on the solar surface. The picture is
taken on September 24, 1997. Solar shock waves are accompanied by violent
events like solar flares or CME’s, which can affect Earth’s magnetosphere and its
orbiting satellites. These are two examples which are meant to justify (M)HD
shock research. However, in this dissertation, we strife for a more mathematical
description of shock waves.
Consider a solution U(x, t0 ) of a hyperbolic system of conservation laws Ut +Fx =
0 at a certain time t0 . Suppose U is constant at both sides of a continuous
transition layer (as shown in the left frame of figure 1.4), i.e.:

 Ul , −δ/2 > x
U= U(x), −δ/2 < x < δ/2 (1.28)
Ur , x > δ/2

In the limit of δ → 0, we call the solution a weak solution to the hyperbolic system.
Note that at |x| > δ, the conservation law is trivially satisfied. in the limit case for
δ → 0, we derive the so-called Rankine-Hugoniot jump conditions (RH) at x = 0,
which describe solutions to the integral form of the MHD equations.
Let ui be a i-th conserved variable, and let Fi be the corresponding flux term.
Take an infinitesimal volume element [x0 , x1 ] × [t0 , t1 ] in the (x, t)-plane such that
at time t0 the shock is located at x0 , and at time t1 the shock is located at x1 , as
shown in the right frame of figure 1.4. The i-th line of the one dimensional MHD
equations now reads as ∂U ∂Fi
∂t + ∂x = 0. Now: denoting x̃ = (t, x), F̃i = (Ui , Fi )
i

and ∇˜ = (∂t , ∇), the integral form of these equations becomes


ZZ
˜ F̃i dx̃ = 0,
∇ (1.29)
(x,t)
12 INTRODUCTION

Ur

Ul

−δ/2 δ/2 x

Figure 1.4: Left: Two constant states connected through a transition layer of
thickness δ. In the limit case δ → 0, a discontinuity appears. Right: Integration on
an infinitesimal volume element [x0 , x1 ] × [t0 , t1 ] leads to the RH jump conditions.

According to Gauss’ theorem, this equation reduces as


I
F̃i · ndS = 0. (1.30)

Here S is the boundary of the volume element and n is the normal vector on
S pointing outwards. It is now straightforward to evaluate the left hand side of
equation (1.30), and find that

(x1 − x0 )(Ui,r − Ui,l ) − (t1 − t0 )(Fi,r − Fi,l ) = 0, (1.31)

where the index r refers to the right region and the index l refers to the left region.
Rewriting equation (1.31) and collecting all the values of i, we arrive at

[[F]] = s [[U]] , (1.32)


x1 −x0
where the jump [[·]] = ·r − ·l , and s ≡ t1 −t0 is the shock speed.
For now, it is sufficient to note that in (M)HD, these discontinuities are called
shocks whenever the normal velocity jumps across the discontinuity do not vanish:
[[ux ]] 6= 0. Otherwise, the discontinuity is called a linear discontinuity. An example
of such a linear discontinuity, both in HD and ideal MHD, is a contact discontinuity
(CD), where the only primitive variable which jumps across it is the density ρ.
Discontinuous solutions, which thus only satisfy the RH conditions, are called weak
solutions.
SHOCK WAVES IN (M)HD 13

1.2.2 Characteristic speeds

Let Ut + Fx = 0 be a n-dimensional hyperbolic system of conservation laws. We


can rewrite this system in quasilinear form as

Ut + FU Ux = 0. (1.33)

One can introduce the variable ξ ≡ xt and suppose that U : R × R+ → Rn :


(x, t) 7→ U(x, t) is a homogeneous function, i.e. ∀µ ≥ 0 : U(µx, µt) = U(x, t).
these solutions are in fluid mechanics often referred to as self-similar solutions:
the vector of conserved variables U(ξ) is a function of ξ only.
we can rewrite this quasilinear form again, now as an eigenvalue problem

FU Uξ = ξUξ , (1.34)

from which it is clear that wherever U is not locally constant (Uξ 6= 0), ξ must
be an eigenvalue of the flux Jacobian FU . Since the system is hyperbolic, all n
eigenvalues must be real, by definition of hyperbolicity.
We can find such self-similar solutions if at t = 0, U is constant everywhere, except
at x = 0. This initial setup is called a Riemann problem. A first observation is
that the eigenvalues λi (1 ≤ i ≤ n, and λ1 ≤ λ2 ≤ ...λi ≤ λi+1 ≤ ... ≤ λn ) are a
function of space and time. When solving such a Riemann problem, the appearing
signals are located where ξ = xt = λi (x, t) and that is why we call the eigenvalues
of the Jacobian of the flux matrix with respect to the conserved variables (or any
other set of independent variables) the characteristic speeds. These characteristic
speeds can be used to define characteristic curves σi,κ : si (x, t) = κ in the (x, t)-
plane as follows:
∂si ∂si
+ λi = 0. (1.35)
∂t ∂x
From the implicit function theorem, it now follows that (x0 , t0 ) ∈ σi,κ can be
locally expressed as x(t), such that x(t0 ) = x0 and

dx
(t0 ) = λi . (1.36)
dt
More compactly, these characteristic curves are also called characteristics, and
from equation (1.36), these characteristics describe the path at which information
travels at the i-th characteristic speed λi , while from equation (1.35) we know
that along this i-th characteristic, the value of si is invariant. si is said to be a
Riemann invariant.
The characteristics of the Riemann problem are schematically shown in figure 1.5.
The signals separate two regions of constant U, and can (for a non-linear system)
14 INTRODUCTION

t
λ1 λi λn

Figure 1.5: The characteristic speeds of a Riemann problem. Up to n real


characteristic speeds divide the (x, t)-plane in up to n + 1 regions of constant
U. The transitions can both be continuous or discontinuous. The continuous case
is represented by multiple lines, while the discontinuous signals are represented by
a single line. The signals travel at speed dx x
dt = t . The regions separated by the
signals have constant speed.

be both continuous or discontinuous, where the discontinuous case can be formed


form a continuous case by wave steepening, which we will explain in more detail in
Section 1.2.5. Consider now two (x, t)-regions in the space-time plane, separated
by a single (the i-th) signal which is traveling at speed s, it can be shown (and is
shown by e.g. Lax [53]) that for discontinuous signals we must have

λi (Ur ) ≤ s ≤ λi (Ul ). (1.37)

When one of the inequalities is a strict inequality, also the other inequality is a
strict inequality. Those discontinuous solutions are called shocks. On the other
hand, when λi (Ul ) = s = λi (Ur ), we say that the solution a linear discontinuity.
Across continuous signals, s varies from λi (Ul ) to λi (Ur ).
Figure 1.6 shows the characteristics in both the discontinuous and the continuous
case. In the left panel, λi (Ul ) > λi (Ur ). This means that the characteristics
SHOCK WAVES IN (M)HD 15

t t

x x

Figure 1.6: Left: The i-th characteristics at both sides of a shock. Since
characteristics can not cross, wave steepening leads to shock formation. Right:
The i-th characteristics at both sides of an expansion fan.

in Ul and Ur should meet and a shock is formed. The exact value of s is not
a priori known from λi (Ul ) and λi (Ur ), but we do know that it is bounded by
λi (Ul ) > s > λi (Ur ). The right panel shows the situation where λi (Ul ) < λi (Ur ).
Therefore the characteristics never meet. The dotted characteristics do not exist
at t = 0.

1.2.3 Characteristic speeds in HD

Let us now derive the characteristic speeds of the Euler system. We follow the
approach of Courant & Friedrichs [14], which can also be found in textbooks as
e.g. Leveque [54], Gombosi [35] or Batchelor [7]. First of all, the one-dimensional
version of these equations becomes in conservational form Ut + Fx = 0 with
2
U = (ρ, ρv, ρ v2 + γ−1 1
p)t and the flux term becomes

ρv
 

F= ρv 2 + p . (1.38)
ρv 2 γ
v( 2 + γ−1 p)
To find the eigenvalues of FU , we solve
λ−v −ρ 0


−v 2 λ − 2ρv −1

= 0. (1.39)
3 γ γ

− v2 − γ−1 p − 32 ρv 2 λ − γ−1 v

and find that the HD system is (strictly) hyperbolic, namely λi ∈ {v − c, v, v + c},


where we introduced the sound speed
γp
r
c≡ . (1.40)
ρ
16 INTRODUCTION

Note that c > 0 such that v − c < v < v + c. Since the system is hyperbolic, we
can diagonalize the Jacobian matrix as FU = RΛR−1 , where the i-th column of
R is the right eigenvector belonging to λi , and Λ the diagonal matrix with λi on
the (i, i)-th entry. The quasilinear form simplifies thus as Ut + RΛR−1 Ux = 0, or
by denoting that the p-th column of R−1 (i.e. the p-th left eigenvector of the FU )
as lp , this statement is equivalent to (lp · dU)dx=λp dt = 0. Rewriting this in terms
of total derivatives dρ, dv and dp, we find the characteristic equations

(dp − ρcdv)dx=(v−c)dt = 0, (1.41)

(dp − c2 dρ)dx=vdt = 0, (1.42)

(dp + ρcdv)dx=(v+c)dt = 0. (1.43)

This is equivalent to st + Λsx = 0, where the vector of the Riemann invariants s


is given by
 
2 p 2
s = (s1 , s2 , s3 ) = v − c, S(= γ ), v + c . (1.44)
γ−1 ρ γ−1

We have just shown that sp is invariant along the curves dx = λp dt.


Denoting ξ = xt , we can even derive relations across continuous signals. We will
do so in chapter 2, but now it suffices to mention the relations holding across
continuous signals: [li · dU]dx=λj dt = 0, when i 6= j. This means that sj is
invariant across the i-th signal, whenever i 6= j.

1.2.4 Characteristic speeds in ideal MHD

The exact same procedure, performed on the one-dimensional version of equation


(1.19) can be performed in one-dimensional MHD (see e.g. Courant & Hilbert
[16] or Jeffrey & Taniuti [46]). This leads to the conclusion that the characteristic
speeds of 1D ideal MHD are given by {vx ± vf,x , vx ± va,x , vx ± vs,x , vx } where we
introduced the normal and total Alfvén speed, respectively given by
s
Bx2
va,x ≡ , (1.45)
ρ
s
Bx2 + By2 + Bz2
va ≡ , (1.46)
ρ
SHOCK WAVES IN (M)HD 17

and the fast and slow magnetosonic speeds, respectively given by


r 
1 2 q 
vf,x ≡ c + va2 + (c2 + va2 )2 − 4c2 va,x
2 , (1.47)
2
r 
1 2 q 
vs,x ≡ c + va2 − (c2 + va2 )2 − 4c2 va,x
2 . (1.48)
2

Here, again c denotes the sound speed as introduced in equation (1.40). The
MHD system is thus a non-strictly hyperbolic system with three different highly
anisotropic wave speeds. Here the non-strictness of the hyperbolicity means that
the characteristic speeds are not necessarily distinct. Indeed, when for example
(p, Bx , By , Bz ) = (γ −1 , 1, 0, 0), it follows that vf,x = va,x = vs,x . This anisotropy
makes the MHD system much richer than the Euler system, where the only wave
speed is the isotropic sound speed.
Also, if the initial setup of a certain problem is planar, in the sense that the
streamlines and magnetic field lines lay in a single plane, one can again perform
the same procedure. Now, the spectrum of this system has only 5 characteristic
speeds, namely {vx ± vf,x , vx ± vs,x , vx }. Note that this is a subset of the full 1D
MHD spectrum.

1.2.5 Wave steepening: an HD example

As mentioned above, the HD equations are nonlinear, and thus allow for
large amplitude waves. In the wave steepening limit, these waves can become
discontinuous and a shock can be formed. The necessary conditions are stated
in equation (1.37) and shown in the left panel of figure 1.6: since crossing
characteristics are impossible, they lead to shock formation through wave
steepening.
We illustrate this by a 1-dimensional numerical simulation where we connect two
constant states continuously through a transition region. We assume that γ = 35 .
As initial conditions, we set

ρ = γ, (1.49)

v = 3/2, (1.50)

 1, ∀x < 1,
p = x2 , ∀x ∈ [1, 2], (1.51)
4, ∀x > 2.

18 INTRODUCTION

Now, this leads to the following characteristic speeds:



 1/2 ∀x < 1,
v−c = 3/2 − x ∀x ∈ [1, 2], (1.52)
−1/2 ∀x > 2.

v = 3/2, (1.53)

 5/2 ∀x < 1,
v+c = 3/2 + x ∀x ∈ [1, 2], (1.54)
7/2 ∀x > 2.

The upper left panel of figure 1.7 shows the initial primitive variables and the
upper right panel shows the initial characteristic speeds over the domain. Since
the characteristic speed u − c is decreasing in function of x, it is clear that it leads
to crossing 1-characteristics, and we expect one shock to be formed. The lower left
and right signals show respectively the primitive variables at t = 0.5 and t = 1. As
predicted, the most left signal, located at x ≈ 1.5 has steepened, and a stationary
shock will form. Furthermore, notice that the sign changing of v − c is equivalent
to a transition from Ms ≡ v/c > 1 to Ms < 1. Here Ms is called the sonic Mach
number. Also, at x ≈ 2.2 a linearly degenerate signal is being formed. Since only
the mass density jumps across it, it is a CD.

1.2.6 Wave steepening: an MHD example

When a magnetic field is applied, the system has 3 anisotropic characteristic speeds,
instead of one single isotropic sound speed. Just as in HD, the MHD equations are
nonlinear, and thus allow for large amplitude waves. In the wave steepening limit,
these waves can become discontinuous and a shock can be formed. The necessary
conditions are stated in equation (1.37) and shown in the left panel of figure 1.6:
since crossing characteristics are impossible, they lead to shock formation through
wave steepening.
We illustrate this by a 1-dimensional numerical simulation where we connect two
constant states continuously through a transition region. Again, we assume that
γ = 53 and as initial conditions we set on the x-axis:
SHOCK WAVES IN (M)HD 19

Figure 1.7: Upper Left: The initial configuration for the primitive variables for the
HD wave steepening AMRVAC simulation. Note that all variables are continuous.
Only the pressure is non-constant in the transition region. Upper Right: The
initial characteristic speeds. Only v − c is decreasing and will thus lead to crossing
characteristics. Therefore the left signal will steepen and in the limit case a shock
will be formed. Lower Left: The primitive variables at t = 0.5. Lower Right: The
primitive variables at t = 1.
20 INTRODUCTION

ρ = γ, (1.55)
 q √
4+ γ2 + γ2 4γ+1+1
∀x < 1,


2

 r
 q
2

vx = 3−2x+ 2γ+2 2
γ x +2+2 ( γ+1
γ ) x4 − γ−1 2
γ x +1 (1.56)
∀x ∈ [1, 2],
q2

 r
10+ γ8 +2 9+ 40 16

γ + γ 2 −1


∀x > 2.

2

vy = 0 (1.57)

 1, ∀x < 1,
p = x2 , ∀x ∈ [1, 2], (1.58)
4, ∀x > 2.


Bx = γ (1.59)

 1, ∀x < 1,
By = x, ∀x ∈ [1, 2], (1.60)
2, ∀x > 2.

(1.61)

In the upper left frame of figure 1.8, the initial distribution of ρ, vx , p and By
are plotted, while in the upper right frame the initial characteristic speeds of the
planar 1D MHD system are plotted. Note that vx − vf,x is decreasing and will
lead to crossing characteristics. Therefore we expect that the most left signal will
steepen and a MHD shock will be formed. This will be a fast MHD shock, which
we will introduce in Section 3.3.
The two lower panels of figure 1.8 show ρ, vx , p and By at respectively t = 1 and
t = 2. We indeed notice that the most left signal has become discontinuous.

1.3 The Richtmyer-Meshkov instability

Any instability of a contact discontinuity (CD) separating two different gases,


which is caused by the interaction of this CD with a hydrodynamical shock is
referred to as a Richtmyer-Meshkov instability (RMI). We will give a brief literature
review of the RMI.
THE RICHTMYER-MESHKOV INSTABILITY 21

Figure 1.8: The initial configuration for the HD wave steepening AMRVAC
simulation. Upper Left: The initial configuration for the primitive variables for
the MHD wave steepening AMRVAC simulation. Note that all variables are
continuous. Upper Right: The initial characteristic speeds. Only vx − vf,x is
decreasing and will thus lead to crossing characteristics. Therefore the left signal
will steepen and in the limit case a shock will be formed. Lower left: The AMRVAC
snapshot at t = 1. Lower Right: The AMRVAC snapshot at t = 2.
22 INTRODUCTION

Figure 1.9: Snapshots of the density in an AMRVAC RMI simulation. The first
snapshot shows the initial setup. The second frame shows the vorticity deposition
phase, and the latter snapshot shows the vorticity evolution phase. Here the
interface has become RM-unstable.

(M)HD shocks appear, and are extensively studied, in a wide range of astrophysical
environments. van Eerten et al.[25] studied the encounter between the relativistic
blast wave from a gamma-ray burster and a stellar wind termination shock,
De Sterck et al.[23] performed 3D simulations of stationary slow shocks in the
magnetosheath for solar wind conditions, Chané et al.[15], Jacobs et al.[44] and
many others study the interaction of the background wind and interplanetary
shocks. However, astrophysical experiments are very difficult to perform, one
has to restrict oneself to observations. Therefore shock experiments are usually
performed in a laboratory environment, such as a shock tube. In this dissertation,
we will not perform those kind of experiments ourselves. We will analyse these
experiments mathematically and perform numerical simulations. Our goal is in
the first place to understand shocks from a more mathematical perspective.
Since the numerical experiments we perform are all shock tube experiments,
we will first briefly review the literature on shock tubes and shock refraction.
Shock refraction happens whenever a shock interacts with an interface. When a
hydrodynamical shock refracts at an oblique density discontinuity, typically three
signals will arise and the interface will become unstable. Physical experiments can
be performed in a so-called shock tube, where a shock is generated and the set-up
THE RICHTMYER-MESHKOV INSTABILITY 23

is such that it will impinge on density discontinuity. Figure 1.11 shows a schematic
representation of such a shock tube. We performed numerical experiments with
the numerical code AMRVAC (see Section 1.4) and figure 1.9 shows the typical
behavior of the shock refraction process .in three density snapshots Shown is the
density of a numerical simulation for The first snapshot shows the initial setup of
the experiment. Left and right sides of the domain are modeled open, while the
upper and lower boundaries are rigid walls. On the left side of the domain, there
is a genuine right-moving shock, moving towards an inclined density discontinuity,
separating two gases at rest. When the shock and the CD touch each other at
the bottom wall, the shock will refract and disturb the CD. Now the first vorticity
deposition phase has started. Vorticity is deposited on the interface, causing it
to roll up. The refraction is typically centered around one single point, called
the triple point, as indicated in the middle snapshot of figure 1.9. When this
point reaches the upper wall, it will reflect, and after this reflection the vorticity
evolution phase begins. Now the vorticity changes sign, and the interface rolls up.

1.3.1 Shock tube problems

In the HD case, the shock will disturb the CD interface, and refract in a reflected
(R) and a transmitted (T) signal. The pattern created by these three signals
can vary, and it is far from trivial to predict which pattern will arise under a
given set of gas parameters. An overview of different refraction patterns at a slow-
fast CO2 /CH4 interface is given in figure 1.10 (based on Nouragliev et al.[64]).
Regular shock refraction means that the three signals meet at a single point, called
the triple point. Since we can solve those problems exactly, we focus on those
patterns. A more detailed description of irregular shock refractions is given by,
e.g., Henderson [41] or Nouragliev et al.[64]. The upper frame of figure 2.9 shows a
numerical simulation Schlieren plot of a regular shock refraction pattern, while the
lower frame shown an irregular refraction pattern. Research on hydrodynamical
shock refraction has been performed since the 1940’s. The highly analytical work
of von Neumann [62] was pioneering in this field. He was for example the first
one to predict critical angles for regular shock refraction. In 1947, Taub [81]
suggests a solution strategy in the same highly analytical tradition: for the regular
refraction with reflected shock (RRR) pattern he rewrites the Rankine-Hugoniot
shock conditions as a polynomial of degree 12, and solving this polynomial is then
done numerically. Later on, shock tube experiments started to be performed. A
schematic representation of a shock tube is shown in figure 1.11. In such a shock
tube, a shock is generated, often by a piston system, and it propagates through
the tube at a certain sonic (upstream/downstream) Mach number Ms . Here the
shocks (upstream/downstream) sonic Mach number is defined as the ratio of the
shock speed s, and (upstream/downstream) sound speed. For shock refraction
experiments, inside of the tube, a membrane mimics a CD separating 2 gases at
rest. Such simulations where performed by Jahn [45] to study regular refraction
24 INTRODUCTION

→ Increasing α →
Very Weak F N R → F P R → BP R → RRR → RRE
Weak T N R → T RR → F N R → BP R → RRR → RRE
Strong T M R → BP R → RRE
Notation:
RRR Regular Refraction with Reflected shock
RRE Regular Refraction with reflected Expansion fan
BPR Bound Precursor Refraction
FPR Free Precursor Refraction
FNR Free precursor von Neumann Refraction
TRR Twin Regular Refraction
TNR Twin von Neumann Refraction

Figure 1.10: An overview of possible shock refraction patterns at a CD and possible


pattern transitions at a slow-fast CO2 /CH4 interface. All these patterns and
transitions have been observed in shock tube experiments by Abd-El Fattah &
Henderson [1, 2, 3]. We focus on the regular shock refraction patterns RRR and
RRE, since these can be solved exactly. The figure is based on Nouragliev et al.[64].

Figure 1.11: A schematic representation of a shock tube. Taken from [90].

of a plane shock, and later on also by, e.g., Henderson [40], and Abd-El Fattah
et al.[1, 2, 3] for irregular shock refraction or by, e.g., Haas en Sturtevant [37] and
more recently by Kreeft & Koren [50] for shock-bubble interaction.
THE RICHTMYER-MESHKOV INSTABILITY 25

t
R CD T

U2 U3

M0−1

U1 U4

U0

S CD x

0 1

Figure 1.12: Setup of a 1D shock refraction experiment. At time t = M0−1 a


Riemann problem occurs.

1.3.2 Shock refraction: a 1D example

Let us first consider the most trivial interaction of a shock and a CD, and therefore
the most trivial shock tube problem. Consider therefore a 1D domain, with a
stationary shock located at x = 0. At x = 1, we place a genuine left-moving CD
(see figure 1.12). Therefore, we have initially three region, i.e.,

• u1 ≡ (ρ1 , v1 , p1 ) at x < 0;
• u0 ≡ (ρ0 , v0 , p0 ) at x ∈ [0, 1];
• u4 ≡ (ρ4 , v4 , p4 ) at x > 1.

In general, one has three degrees of freedom for nondimensionalization (see e.g.
Goedbloed & Poedts [32]) but since we will look for a self-similar solution the
nondimensionalization is uniquely determined by fixing two variables in a certain
region. We choose to set

• ρ0 = γ;
• p0 = 1.

Now, the speeds are scaled with respect to the downstream sound speed, i.e. c0 ≡
q
γp0
ρ0 = 1.
26 INTRODUCTION

The CD is completely determined by the density ratio η ≡ ρρ40 across it. Since all
speeds are scaled with respect to the sound speed in region u0 , this region moves
with its sonic Mach number towards the shock. Therefore, we now know that

• (ρ0 , v0 , p0 ) = (γ, −M0 , 1) and


• (ρ4 , v4 , p4 ) = (ηγ, −M0 , 1).

At t = 0 the upstream state u1 = (ρ1 , v1 , p1 ) is related to the downstream state u0


through the Rankine-Hugoniot conditions. At time t = M0−1 , the shock and the
CD interact, and a Riemann problem occurs. The shock will refract in a reflected
(R) and a transmitted (T) shock, separated by the shocked contact discontinuity.
Across any signal, the 1D HD Rankine-Hugoniot (RH) conditions should hold. We
discuss these conditions in more depth later on, but for now it suffices to know
that in any stationary shock frame, these conditions reduce to

(γ + 1)Mk2
ρu = ρk , (1.62)
(γ − 1)Mk2 + 2

(γ − 1)Mk2 + 2
vu = uk , (1.63)
(γ + 1)Mk2

2γMk2 + 1 − γ
pu = pk . (1.64)
γ+1

Here the indices u and k refer respectively to the unknown and the known state.
Therefore we know that the initial upstream state is given by
(γ + 1)M02 (γ − 1)M02 + 2 2γM02 + 1 − γ
 
(ρ1 , v1 , p1 ) = ρ 0 , − , .
(γ − 1)M02 + 2 (γ + 1)M0 γ+1

We can express the known sonic Mach number at both the reflected R (Ms,l ) and
transmitted T (Ms,r ) signals respectively as

v1 − ξl
Ml = q , (1.65)
γp1
ρ1

v4 − ξr
Mr = q . (1.66)
γp4
ρ4

Here the newly introduced variables ξl and ξr respectively denote the speeds at
which R and T are traveling. When we apply relations (1.63-1.64) both across the
reflected and transmitted signal we arrive at the following solution:
THE RICHTMYER-MESHKOV INSTABILITY 27

(4γ 2 − 4γ)ξl M03 + (γ 2 − 1) + (2γ 2 + 2γ)ξl M02 + 8γξl M0 + 2γ + 2


` ´
p2 = ` ´ , (1.67)
(γ + 1) (γ − 2)M02 M02 + 2

2ηγ(M0 + ξr )2 − γ + 1
p3 = , (1.68)
γ+1

(1 − γ 2 )M03 − (γ − 1)2 ξl M02 + (2 − 2γ + (1 − γ 2 )ξl2 )M0 + 4(γ − 1)ξl


v2 = + ξl , (1.69)
(γ + 1)((γ − 1)M02 + (γ + 1)ξl M0 + 2)

(1 − γ)ηM02 + (3 − γ)ηξr M0 + 2ηξr2 − 2


v3 = . (1.70)
η(γ + 1)(M0 + ξr )

Both the pressure and the velocity remain invariant across a CD. Therefore,
solving

p2 = p3 ,
(1.71)
v2 = v3 .

will lead to a solution for ξl and ξr . For γ = 53 , the solution is given by

ζ
ξl = , (1.72)
M0

16ζ 2 M02 + M02 + 12ζM0 + 8M04 ζ 2 − 3M04 + 4M03 ζ + 8ξl3 M03 − M06 + 3
ξr = ` ´` ´ , (1.73)
M0 M02 + 2ζM0 + 3 M02 + 3 + 4ζM0

where ζ should satisfy the quartic equation Σ4i=0 ti ζ i , where the coefficients are
given by

t4 = 48η + (16η − 64)M02 , (1.74)

4 2
t3 = (56η − 96)M0 + (208η − 288)M0 + 120η, (1.75)

6 4 2
t2 = (60η − 52)M0 + (288η − 312)M0 + (364η − 468)M0 + 120η, (1.76)

t1 = (18η − 12)M08 + (132η − 108)M06 + (312η − 324)M04 + (252η − 324)M02 + 54η, (1.77)

t0 = −M010 + (9η − 12)M08 + (60η − 54)M06 + (118η − 108)M04 + (60η − 81)M02 + 9η.(1.78)

For Ms,0 = 2, the solution for ξl and ξr is plotted in figure 1.13. Also the speed of
the shocked contact, ξc , is plotted. First of all we notice that the shock will slow
down the CD. The lower the density ratio, the bigger the slowdown of the CD.
28 INTRODUCTION

Figure 1.13: Plotted are ξl , ξr and ξc in function of the density ratio η across the
unshocked CD. The sonic Mach number is kept constant, Ms,0 = 2. The black
line shows ξl , the red line represents ξr and the blue line shows ξc , the speed at
which the shocked contact is traveling.

9 0 6
ρ u p
8
5
7 -0.5

6 4
-1
5
3
4
-1.5
3 2

2 -2
1
1

0 -2.5 0
-3 -2.5 -2 -1.5 -1 -0.5 0 0.5 -3 -2.5 -2 -1.5 -1 -0.5 0 0.5 -3 -2.5 -2 -1.5 -1 -0.5 0 0.5
x x x

Figure 1.14: The exact solution at t = 3/2 to the simple shock tube problem.
Left: ρ(x). Center: v(x). Right: p(x). Note that indeed velocity and pressure are
invariant across the contact, as we claimed before and will show later.

From ξl and ξr , solving the RH conditions across R and T respectively, leads to


the values of U2 and U3 . For η = 2, we plot the solution at t = 3 in figure
1.14. Note that this solution is exact. The analysis above is performed using the
mathematical software package MAPLE 11.0.
THE RICHTMYER-MESHKOV INSTABILITY 29

Figure 1.15: The AMRVAC solution to the introductory shock tube problem. The
effective resolution in this simulation is 30720. Note that indeed this results
confirms our exact solution. The exact solution is overplotted, but not visible.
The next figure show zoomed in plots of the mass density profile of the exact
solution, and the mass density profile of the numerical solution.

Let us compare this exact result with the result of a numerical simulation
performed by the code AMRVAC, which is described in more detail in the next
section. Figure 1.15 shows that this numerical solution agrees with the exact
solution.
Note that, according to figure 1.13 for η < 1 no exact solution exists. This
is because we only looked for RRR refraction patterns, but the output of the
numerical experiment with η = 0.5 shows that no RRR solution is possible in this
case. Indeed the reflected signal is then continuous. This kind of signal is called
an expansion fan (see figure 1.17).
Comparing the slow-fast case (η < 1) to the fast-slow case (η > 1), we note the
following three transitions at η = 1:

• The initial shock was in both cases located at x = 0. Note that in the fast-
slow case the transmitted signal is decelerated (and thus located at x > 0)
while in the slow-fast case this signal is accelerated (and thus located at
x < 0);
• The nature of the shocked contact remains unchanged. By this we mean
that the shocked contact is slow-fast if the initial contact is slow-fast, and
the shocked contact is fast-slow only if the initial contact is fast-slow. Also
note that the pressure p and velocity v do not jump across the shocked
contact;
30 INTRODUCTION

Figure 1.16: Upper : The mass density profile of the exact solution compared to the
mass density profile of the numerical solution across the reflected signal. Middle:
The mass density profile of the exact solution compared to the mass density profile
of the numerical solution across the CD. Lower : The mass density profile of the
exact solution compared to the mass density profile of the numerical solution across
the reflected signal.
THE RICHTMYER-MESHKOV INSTABILITY 31

Figure 1.17: The numerical solution to the introductory shock tube problem with
η = 0.5, with the reflected signal becoming an expansion fan. The effective
resolution in this simulation is again 30720.

• In the fast-slow case, the reflected signal is a shock, and the numerical
solution coincides with our exact solution. In the slow-fast case, on the
other hand, the reflected signal is not a shock, but a continuous rarefaction
fan. Note that the jumps in p, ρ and v all change sign.

These observations are schematically presented in a (x, t)-plot of the involved


signals in figure 1.18. The left panel presents the fast-slow case, while the right
panel presents the slow-fast case. R and T stand for respectively the reflected
and the transmitted shock, while S is the initial shock and CD is the contact
discontinuity. In Chapter 2 of this dissertation we will generalize these facts to
the case where the CD is inclined, and the problem is thus 2D. In this section we
have solved the 1D shock refraction. We will also need to generalize our solution
strategy. We will not be able to give a closed form solution, but a numerical
iteration will converge to the exact solution of the problem. We will also allow for
out-of-plane magnetic fields. In Chapter 3, we will allow for magnetic fields which
are aligned with the shock normal, such that 5 signals will arise.

1.3.3 Richtmyer-Meshkov instability

In the later phase of the shock-interface interaction process, the interface can
become unstable. This instability is called the Richtmyer-Meshkov instability
(RMI), and in contrast to the Rayleigh-Taylor instability (RTI), where an interface
becomes unstable due to an external force (e.g., gravity), the RMI is caused by
the interaction of an interface with a shock. Nonetheless, in 1960, Richtmyer
[69] generalized Taylor’s [82] analysis of the RTI to an impulsive acceleration,
32 INTRODUCTION

t t
R CD T R CD T

S CD x S CD x

Figure 1.18: The involved signals in 2 shock refraction cases in an (x, t)-plane.
Left: Fast-Slow refraction, Right: Slow-Fast refraction.

as appears in the RMI. He predicted theoretically that the interface would


be Richtmyer-Meshkov unstable at all wavelengths, and independent of the
orientation of the impulse. Moreover, in contrast to most surface instabilities
(as the RTI), the RMI was predicted to grow linearly with time, Atwood number
( ρρ22 −ρ
+ρ1 , where the indices 1 and 2 refer respectively to the post and the pre-shock
1

regions), and with wave number k. This means that the interface is unstable for
all wavelengths, at least whenever the density ratio η ≡ ρρ12 > 1. By use of shock
tubes, Meshkov [58] experimentally validated Richtmyers predictions and the RMI
itself became a topic of extensive theoretical, numerical and experimental research.
The first phase of the RMI process can be considered the vorticity (ω ≡ ∇ × v)
deposition phase. The magnitude of the deposited vorticity can be found through
shock polar analysis, as extensively described by Henderson [40] in 1966. After the
shock has passed the interface, the vorticity evolution phase begins. As Hawley &
Zabusky [38] described it in 1989: "The vortex layer diffuses laterally as it rotates
globally. The ends of the vortex layer begin to roll up. (...) The roll up of the lower
interface region proceeds as the vorticity binds with its mirror image. This is the
dominant mechanism for the formation of the wall vortex". This observation is in
agreement with various experimental results. Amongst many others, Sturtevant
[80] gives a good overview of these experiments in 1987. Figure 1.19 plots the
integrated vorticity of the AMRVAC RMI simulation presented by van der Holst
& Keppens [93] and confirms that the instability grows linearly in the first phase of
the process, while after reflection of the top wall, the vorticity deposition changes
sign. In what follows, we derive the MHD generalization of the vorticity evolution
THE RICHTMYER-MESHKOV INSTABILITY 33

Figure 1.19: The total vorticity evolution during the HD shock refraction process
described in van der Holst & Keppens [93]. The vorticity deposition is linear in
the vorticity deposition phase. When the triple point reaches the top wall, the
vorticity evolution phase starts and the vorticity deposition changes sign, causing
the interface to become RM-unstable.

equation for 2D inviscid compressible flows.

1.3.4 Vorticity evolution equation for inviscid compressible MHD


flows

The evolution of the vorticity in ideal MHD can be quantified as follows. Let us
therefore rewrite the momentum equation as
∂ ∇p (∇ × B) × B
v + (v · ∇)v + − = 0. (1.79)
∂t ρ ρ
Then taking the curl of this expression leads to
∂ ∇p (∇ × B) × B
(∇ × v) + ∇ × ((v · ∇)v) + ∇ × −∇× = 0, (1.80)
∂t ρ ρ
or we find

∂ v·v
ω + ∇ × (∇ − v × (∇ × v))
∂t 2
∇p (∇ × B) × B
+∇ × −∇× = 0, (1.81)
ρ ρ

which simplifies to
∂ ∇p (∇ × B) × B
ω + ∇ × (ω × v) + ∇ × −∇× = 0. (1.82)
∂t ρ ρ
34 INTRODUCTION

This can be rewritten as

− v · ∇ω + ω · ∇v − (∇ · v)ω

∇p (∇ × B) × B ∂
−∇ × +∇× = ω. (1.83)
ρ ρ ∂t

The latter two terms of the left hand side (LHS) can be simplified. Indeed,

∇p 1 1
−∇× = −∇ × ∇p − ∇ × ∇p
ρ ρ ρ
1
= (∇ρ × ∇p) , (1.84)
ρ2

and

(∇ × B) × B 1 1
∇× = ∇ × ((∇ × B) × B) + (∇ × [(∇ × B) × B])
ρ ρ ρ
1 1
= − (∇ρ × ((∇ × B) × B)) + (∇ × [(∇ × B) × B])
ρ2 ρ

1 1 B2
= − 2
(∇ρ × ((∇ × B) × B)) + (∇ × [∇ · BB − ∇ ])
ρ ρ 2
1 1
= − (∇ρ × ((∇ × B) × B)) + (∇ × (∇ · BB)). (1.85)
ρ2 ρ

Therefore equation (1.83) simplifies as


ω = −v · ∇ω + ω · ∇v − (∇ · v)ω
∂t
1 1
+ (∇ρ × (∇p − (∇ × B) × B)) + (∇ × (∇ · BB)). (1.86)
ρ2 ρ

On the other hand, note that when using Lagrangian derivatives

D ω 1 D ω D
= ω− 2 ρ
Dt ρ ρ Dt ρ Dt
1 ∂ 1 ω
= ω + (v · ∇)ω + (∇ · v). (1.87)
ρ ∂t ρ ρ
AMRVAC AND THE VTK FILE FORMAT 35

Combining equations (1.87) and (1.83), together with Dω ∂ω


Dt = ∂t + (v · ∇)ω leads
now to the vorticity evolution equations for inviscid compressible MHD flows:
D ω ∇ρ × (∇p − (∇ × B) × B) ∇ × (∇ · BB) (ω · ∇)v
= + + . (1.88)
Dt ρ ρ3 ρ2 ρ

All shock refraction simulations and analyses performed in this dissertation are
2D. The latter term, (ω·∇)v
ρ , vanishes in that case, and therefore the vorticity
evolution equation in 2D compressible flows reduces as
D ω ∇ρ × (∇p − (∇ × B) × B) ∇ × (∇ · BB)
= + . (1.89)
Dt ρ ρ3 ρ2
One could thus conclude that (i) the misalignment of pressure and density
gradients, (ii) the misalignment of the Lorentz force and the density gradient
and (iii) the non-uniformity of the magnetic field are responsible for vorticity
deposition in 2D flows.
More information on the RMI can be found in review articles as Rupert [72],
Zabusky [104] or Bruillette [10].

1.4 AMRVAC and the VTK file format

Since astrophysical "experiments" are only passively observed and laboratory


experiments are expensive to perform, numerical codes gained popularity to
perform computer experiments. Interpreting and analyzing data produced by
a numerical experiment is also easier and accurate. All numerical simulation
discussed in this dissertation have been performed by (MPI)AMRVAC on the
K.U.Leuven HPC cluster VIC. Let us briefly discuss the history and philosophy of
this code.

1.4.1 History

The Versatile Advection Code (VAC) is a parallel any-dimensional multi-physics


open source code for solving hyperbolic systems of conservation laws such as
advection, HD and MHD problems, initially developed by Tóth [87]. Often
astrophysical problems involve super-Alfvénic speeds (as explained later), which
give rise to discontinuous shock solutions to the MHD equations. A numerical
MHD tool should thus exploit appropriate shock capturing schemes. VAC
discretizes the physical domain in a regular recti- or curvilinear grid, and includes
shock-capturing schemes such as the Flux Corrected Transport (FCT), a Total
Variation Diminishing (TVD) scheme with a Roe-type approximate Riemann
36 INTRODUCTION

solver (see e.g. Roe [70] or Roe & Balsara [71], a Total Variation Diminishing
Lax-Friedrich scheme (TVDLF) and implicit and semi-implicit schemes (see Tóth
et al.[89], Keppens et al.[48]). VAC is written in Fortran, using a pre-processor,
which allows to program in any-dimensional manner. This Loop Annotation
Syntax is called the LASY-syntax (as introduced in Tóth [88]). More information
on VAC can be found in [92].

1.4.2 Adaptive Mesh Refinement

The higher the resolution in a numerical simulation, the more accurate the physics
are represented. Especially when discontinuities are involved, not all regions in the
complete domain require the same resolution in order to accurately represent the
physics. Berger & Colella [9] proposed the Adaptive Mesh Refinement (AMR)
algorithm, and this principle is used in AMRVAC, as presented by Keppens
et al.[49]. Initially, the type of AMR exploited in AMRVAC was a patch-based
refinement. Later on, it has been changed to a hybrid-block based refinement (as
explained in van der Holst & Keppens [93]), and still later evolved to the present
form of the octree based block refinement.
Historically, numerical simulations were first performed on a regular uniform
equidistant grid. The main advantage of this approach is that inter- and
extrapolations are straightforward to perform. A major disadvantage on the other
hand, is that the resolution is equal on the whole domain, while the main physics
of a certain problem is often centered in certain "interesting regions". The idea
of AMR is now to tackle this by increasing the resolution in these interesting
regions. One therefore first chooses a number n of AMR levels. On the first level
one employs a regular grid structure. Every coarsest level grid again consists of
regularly distributed cells. On each of these cells, the conserved variables have a
certain value at each time step. One can then exploit various refinement algorithms
to decide whether to flag a certain grid for refinement, but the basic idea is that
large gradients should lead to refinement. On the second AMR level, one then
divides every grid in every dimension into 2, such that the grid is divided into 2d
sub-grids, as shown in figure 1.20. Here d is the number of dimensions. Exactly
the same refinement happens to the cells in these flagged grids. An interpolation
then allows to calculate the new values of the conserved variables in these new
cells. One recursively repeats this procedure n − 1 times.
In this AMR approach inter- and extrapolation are not as straightforward as on a
regular grid. In fact, there is a restriction to this refinement: the proper nesting
requirement should not be violated. The proper nesting requirement states the
following. Let grids g1 and g2 be neighboring grids, with level respectively l1 and
l2 . Then l2 can differ from l1 with at most 1. If this is not the case then, the
coarser of grids g1 and g2 is checked for refinement, until the criterion is satisfied.
OUTLINE OF THE THESIS 37

Figure 1.20: The AMR grids generated at a Cartesian domain. Each of these grids
is divided into a number of uniformly regularly structured cells.

The exploited timestep, ∆t, is constant on each level. Moreover, the Courant
condition should be satisfied on all levels:
∆xi
≥ vmax , (1.90)
∆t
where ∆xi is the thickness of a cell in dimension i, ∆t is the timestep, and vmax
is the maximum speed reached in the simulation. The Courant condition ensures
that information travels at most one cell during one timestep, and is needed for
stability of explicit time integration schemes.
In Appendix A we describe the structure of AMRVAC in more detail. In Appendix
B we give more information on the compilation procedure, and in Appendix C we
show an example of a USR-file, as introduced in Appendices A and B.

1.4.3 The VTK file format

Astrophysical simulations often lead to huge data files, especially in 3-dimensional


simulations, and a major issue is the data visualization, which is preferably
done by parallel software. ParaView [66] is exactly such an open source parallel
visualization application. ParaView was developed by Kit ware, inc., and is able to
process a wide range of data types, amongst which VTK [95], which was especially
developed for ParaView. VTK provides several file formats, which are described
in [96]. We decided to use the VTK file format for unstructured grid, called VTU.
Therefore, we developed a Fortran subroutine which converts the AMRVAC output
to VTU output. It can also be read by several other visualization applications,
among which VisIt [94]. The output is saved in a binary format.

1.5 Outline of the thesis

The following 3 chapters are peer-reviewed publications with minor modifications


applied.
38 INTRODUCTION

Chapter 2 is based on Delmont et al.[19]. In this chapter we study the classical


problem of planar shock refraction at an oblique density discontinuity, separating
two gases at rest. When the shock impinges on the density discontinuity, it
refracts and in the hydrodynamical case 3 signals arise. Regular refraction
means that these signals meet at a single point, called the triple point. After
reflection from the top wall, the contact discontinuity becomes unstable due to
local Kelvin-Helmholtz instability, causing the contact surface to roll up and
develop the Richtmyer-Meshkov instability. We present an exact Riemann solver
based solution strategy to describe the initial self similar refraction phase, by
which we can quantify the vorticity deposited on the contact interface. We
investigate the effect of a perpendicular magnetic field and quantify how addition of
a perpendicular magnetic field increases the deposition of vorticity on the contact
interface slightly under constant Atwood Number. We predict wave pattern
transitions, in agreement with experiments, von Neumann shock refraction theory,
and numerical simulations performed with the grid-adaptive code AMRVAC. These
simulations also describe the later phase of the Richtmyer-Meshkov instability.
Early results on the purely HD case were published in Delmont & Keppens [18].
Chapter 3 is based on Delmont et al.[20]. In this chapter we generalize our
results presented in Chapter 2 to planar ideal MHD. As mentioned earlier, in
the hydrodynamical case, 3 signals arise and the interface becomes Richtmyer-
Meshkov unstable due to vorticity deposition on the shocked contact, in ideal
MHD, on the other hand, when the normal component of the magnetic field does
not vanish, 5 signals will arise. The interface then typically remains stable, since
the Rankine-Hugoniot jump conditions in ideal MHD do not allow for vorticity
deposition on a contact discontinuity. Again, we present an exact Riemann solver
based solution strategy to describe the initial self similar refraction phase. Using
grid-adaptive MHD simulations, we show that after reflection from the top wall,
the interface remains stable.
Chapter 4, is based on Delmont & Keppens [21] and focuses on the mathematical
description of MHD shocks. Due to the existence of three anisotropic characteristic
speeds, the MHD shock classification is much richer than in hydrodynamics,
where only the isotropic sound speed enters. One distinguishes slow, intermediate
and fast shocks, where intermediate shocks connect a sub-Alfvénic state to a
super-Alfvénic state. We investigate under which parameter regimes the MHD
Rankine-Hugoniot conditions, which describe discontinuous solutions to the MHD
equations, allow for certain types of intermediate MHD shocks. We derive limiting
values for the upstream and downstream shock parameters for which shocks of a
given shock type can occur. We revisit this classical topic in nonlinear MHD
dynamics, augmenting the recent time reversal duality finding by Goedbloed [33]
in the usual shock frame parametrization. Our results generalize known limiting
values for certain shock types, such as switch-on or switch-off shocks.
OUTLINE OF THE THESIS 39

Chapter 5 finally gives a popularized dutch summary of the work presented in the
other chapters.
Chapter 2

An exact Riemann solver for


regular shock refraction

In this chapter we analyze the process of regular shock refraction at an inclined


density discontinuity in hydrodynamics. When a shock refracts, three signals
will arrive. We develop an exact Riemann solver to predict the position of the
new-formed signals, and find the values of the conserved variables in the new-
formed regions. We derive critical angles for regular shock refraction, which agree
with numerical AMRVAC simulations and shock tube experiments. Our approach
connects slow-fast and fast-slow refraction in a natural way. Finally, we investigate
the effect of an out-of-plane magnetic field.
After reflection from the top wall, the interface becomes unstable due to local
Kelvin-Helmholtz instability. This instability is called the Richtmyer-Meshkov
instability.
This work was published in Delmont et al.[19].

2.1 Introduction

We study the classical problem of regular refraction of a shock at an oblique


density discontinuity. Long ago, von Neumann [62] deduced the critical angles for
regularity of the refraction, while Taub [81] found relations between the angles
of refraction. Later on, Henderson [40] extended this work to irregular refraction
by use of polar diagrams. An example of an early shock tube experiment was
performed by Jahn [45]. Amongst many others, Abd-El-Fattah & Henderson [2, 3]
performed experiments in which also irregular refraction occurred.

41
42 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

In 1960, Richtmyer performed the linear stability analysis of the interaction of


shock waves with density discontinuities, and concluded that the shock-accelerated
contact is unstable to perturbations of all wavelenghts, for fast-slow interfaces
(Richtmyer [69]). In hydrodynamics (HD) an interface is said to be fast-slow
if η > 1, and slow-fast otherwise, where η is the density ratio across the
interface (figure 2.1). The instability is not a classical fluid instability in the
sense that the perturbations grow linearly and not exponentially. The first
experimental validation was performed by Meshkov [58]. On the other hand,
according to linear analysis the interface remains stable for slow-fast interfaces.
This misleading result is only valid in the linear phase of the process and near
the triple point: a wide range of experimental (e.g. Abd-El-Fattah & Henderson
[3]) and numerical (e.g. Nouragliev et al.[64]) results show that also in this case
the interface becomes unstable. The growth rates obtained by linear theory
compare poorly to experimentally determined growth rates (Sturtevant [80]). The
governing instability is referred to as the Richtmyer-Meshkov instability (RMI)
and is nowadays a topic of research in e.g. inertial confinement fusion ( e.g. Oron
et al.[65]), astrophysics (e.g. Kifonidis et al.[51]), and it is a common test problem
for numerical codes (e.g. van der Holst & Keppens [93]).
In essence, the RMI is a local Kelvin-Helmholtz instability, due to the deposition of
vorticity on the shocked contact. Hawley & Zabusky [38] formulate an interesting
vortex paradigm, which describes the process of shock refraction, using vorticity as
a central concept. Later on, Samtaney et al.[77] performed an extensive analysis
of the baroclinic circulation generation on shocked slow-fast interfaces.
A wide range of fields where the RMI occurs, involves ionized, quasi-neutral
plasmas, where the magnetic field plays an important role. Therefore, more
recently there has been some research done on the RMI in magnetohydrodynamics
(MHD). Samtaney [78] proved by numerical simulations, exploiting Adaptive Mesh
Refinement (AMR), that the RMI is suppressed in planar MHD, when the initial
magnetic field is normal to the shock. Wheatley et al.[98] solved the problem of
planar shock refraction analytically, making initial guesses for the refracted angles.
The basic idea is that ideal MHD does not allow for a jump in tangential velocity, if
the magnetic field component normal to the contact discontinuity (CD), does not
vanish (see e.g. Goedbloed & Poedts [32]). The solution of the Riemann problem
in ideal MHD is well-studied in the literature (e.g. Lax [53]), and due to the
existence of three (slow, Alfvén, fast) wave signals instead of one (sound) signal,
it is much richer than the HD case. The Riemann problem usually considers the
self similar temporal evolution of an initial discontinuity, while we will consider
stationary two dimensional conditions. The interaction of small perturbations
with MHD (switch-on and switch-off) shocks was studied both analytically by
Todd [84] and numerically by Chu & Taussig [17]. Later on, the evolutionarity of
intermediate shocks, which cross the Alfvén speed, has been studied extensively.
Intermediate shocks are unstable under small perturbations, and are thus not
CONFIGURATION AND GOVERNING EQUATIONS 43

evolutionary. Brio & Wu [11] and De Sterck et al.[22] found intermediate shocks
in respectively one and two dimensional simulations. The evolutionary condition
became controversial and amongst others Myong & Roe [60, 61] argue that the
evolutionary condition is not relevant in dissipative MHD. Chao et al.[12] reported
a 2 → 4 intermediate shock observed by Voyager 1 in 1980 and Feng & Wang [29]
recognized a 2 → 3 intermediate shock, which was observed by Voyager 2 in 1979.
On the other hand, Barmin et al.[8] argue that if the full set of MHD equations
is used to solve planar MHD, a small tangential disturbance on the magnetic field
vector splits the rotational jump from the compound wave, transforming it into a
slow shock. They investigate the reconstruction process of the non-evolutionary
compound wave into evolutionary shocks. Also Falle & Komissarov [26, 27] do
not reject the evolutionary condition, and develop a shock capturing scheme for
evolutionary solutions in MHD, However, since all the signals in this paper are
essentially hydrodynamical, we do not have to worry about evolutionarity for the
setup considered here.
In this chapter, we solve the problem of regular shock refraction exactly,
by developing a stationary two-dimensional Riemann solver. Since a normal
component of the magnetic field suppresses the RMI, we investigate the effect of a
perpendicular magnetic field. The transition from slow-fast to fast-slow refraction
is described in a natural way and the method can predict wave pattern transitions.
We also perform numerical simulations using the grid-adaptive code AMRVAC
(van der Holst & Keppens [93]; Keppens et al.[49]).
In section 2, we formulate the problem and introduce the governing MHD
equations. In section 3, we present our Riemann solver based solution strategy
and in section 4, more details on the numerical implementation are described.
Finally, in section 5, we present our results, including a case study, the prediction
of wave pattern transitions, comparison to experiments and numerical simulations,
and the effect of a perpendicular magnetic field on the stability of the CD.

2.2 Configuration and governing equations

2.2.1 Problem setup

As indicated in figure 2.1, the hydrodynamical problem of regular shock refraction


is parametrized by 5 independent initial parameters: the angle α between the
shock normal and the initial density discontinuity CD, the sonic Mach number
M of the impinging shock, the density ratio η across the CD and the ratios of
specific heat γl and γr on both sides of the CD. The shock refracts in 3 signals: a
reflected signal (R), a transmitted signal (T) and a shocked contact discontinuity
(CD), where we allow both R and T to be expansion fans or shocks. Adding a
44 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

(ρ, v = 0, p, B = (0, 0, Bz ), γl ) (ηρ, v = 0, p, B = (0, 0, Bz ), γr )

M α

Figure 2.1: Initial configuration: a shock moves with shock speed M to an inclined
density discontinuity. Both the upper and lower boundary are solid walls, while
the left and the right boundaries are open.

perpendicular magnetic field, B, also introduces the plasma-β in the pre-shock


region,
2p
β= , (2.1)
B2
which is in our setup a sixth independent parameter. As argued later, the shock
then still refracts in 3 signals (see figure 2.3): a reflected signal (R), a transmitted
signal (T) and a shocked contact discontinuity (CD), where we allow both R and
T to be expansion fans or shocks.

2.2.2 Stationary MHD equations

In order to describe the dynamical behavior of ionized, quasi-neutral plasmas, we


use the framework of ideal MHD. We thereby neglect viscosity and resistivity, and
suppose that the length scales of interest are much larger than the Debye length
and there are enough particles in a Debye sphere (see e.g. Goedbloed & Poedts
[32]). Written out in stationary, conservative form and for our planar problem, the
MHD equations (1.19) are reduced to

∂ ∂
F+ G = 0, (2.2)
∂x ∂y
where we introduced the flux terms
!t
B2 γ vx2 + vy2
F= ρvx , ρvx2 + p + , ρvx vy , vx ( p+ρ + B 2 ), vx B, vx γρ , (2.3)
2 γ −1 2
CONFIGURATION AND GOVERNING EQUATIONS 45

y shocked 2
y
1
unshocked
v2
x

v1
atan(λi )
φ
x

n = (sinφ, −cosφ)

Figure 2.2: Left: A stationary shock, separating two constant states across an
inclined planar discontinuity. Right: The eigenvalues of the matrix A from (3.6)
correspond to the refracted signals.

and
!t
B2 γ vx2 + vy2
G= ρvy , ρvx vy , ρvy2 +p+ , vy ( p+ρ + B 2 ), vy B, vy γρ . (2.4)
2 γ−1 2

The applied magnetic field B = (0, 0, B) is assumed purely perpendicular to the


flow and the velocity v = (vx , vy , 0). Note that the ratio of specific heats, γ, is
interpreted as a variable, rather than as an equation parameter, which is done
to treat gases and plasmas in a simple analytical and numerical way. The latter
equation of the system expresses that ∇ · (γρv) = 0. Also note that ∇ · B = 0 is
trivially satisfied.

2.2.3 Planar stationary Rankine-Hugoniot condition

We allow weak (i.e. discontinuous) solutions of the system, which are solutions
of the integral form of the MHD equations. The shock occurring in the problem
setup, as well as those that later on may appear as R or T signals obey the
Rankine-Hugoniot conditions. In the case of two dimensional stationary flows
(see figure 2.2), where the shock speed s = 0, the Rankine-Hugoniot conditions
follow from equation (2.2). When considering a thin continuous transition layer
in between the two regions, withR thickness δ, solutions of the integral form of
2 ∂ ∂
equation (2.2) should satisfy lim 1 ( ∂x F + ∂y G)dl = 0. For vanishing thickness
δ→0
of the transition layer this yields the Rankine-Hugoniot conditions as
46 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

u1

u2 u5

φR

R CD T

u3 u4

Figure 2.3: The wave pattern during interaction of the shock with the CD. The
upper and lower boundaries are rigid walls, while the left and right boundaries are
open.

2  
1 ∂ 1 ∂
Z
− lim F− G dl = 0 (2.5)
δ→0 1 sin φ ∂l cos φ ∂l

[[F]] = ξ [[G]] , (2.6)

where ξ = tan φ and φ is the angle between the x-axis and the shock as indicated
in figure 2.2. The symbol [[ ]] indicates the jump across the interface.

2.3 Riemann Solver based solution strategy

2.3.1 Dimensionless representation

In this section we present how we initialize the problem in a dimensionless manner.


In the initial refraction phase, the shock will introduce 3 wave signals (R, CD, T),
and 2 new constant states develop, as schematically shown in figure 3.3. We choose
a representation in which the initial shock speed s equals its sonic Mach number
M . We determine the value of the primitive variables in the post-shock region by
applying the stationary Rankine-Hugoniot conditions in the shock rest frame. In
absence of a magnetic field, we use a slightly different way to nondimensionalize
the problem. Note ui = (ρi , vx,i , vy,i , ptot,i , Bi , γi ), where the index i refers to the
RIEMANN SOLVER BASED SOLUTION STRATEGY 47

value taken in the i−th region (figure 3.3) and the total pressure

B2
ptot = p + . (2.7)
2
In the HD case, we define p = 1 and ρ = γl in u1 . Now all velocity components
are scaled with respect to the sound speed in this region between the impinging
shock and the initial CD. Since this region is initially at rest, the sonic Mach
number M of the shock equals its shock speed s. When the shock intersects
the CD, the triple point follows the unshocked contact slip line. It does so at a
speed vtp = (M, M tan α), therefore we will solve the problem in the frame of
the stationary triple point. We will look for self-similar solutions in this frame,
u = u(φ), where all signals are stationary. We now have that ṽx = vx − M and
ṽy = vy − M tan α, where ṽ refers to this new frame. From now on we will drop the
tilde and only use this new frame. We now have u1 = (γl , −M, −M tan α, 1, 0, γl )t
and u5 = (ηγl , −M, −M tan α, 1, 0, γr )t . The Rankine-Hugoniot relations now
immediately give a unique solution for u2 , namely

t
(γl2 + γl )M 2 (γl − 1)M 2 + 2 2γl M 2 − γl + 1

u2 = , − , −M tan α, , 0, γl .
(γl − 1)M 2 + 2 (γl + 1)M γl + 1

(2.8)

In MHD, we nondimensionalize by defining B = 1 and ρ = γ2l β , in region 1. Again


all velocity components are scaled with respect to the sound speed in this region.
 t
We now have that u1 = γ2l β , −M, −M tan α, β+1 2 , 1, γl and from the definition
 t
of η, u5 = ηγ2l β , −M, −M tan α, β+12 , 1, γr . The Rankine-Hugoniot relations
now give the following non-trivial solutions for u2 :

t
M 2 −M

−γl βM
u2 = , ω, −M tan α, p2 + , , γl , (2.9)
2ω 2ω 2 ω

where
C1 ω + C2
p2 = , (2.10)
C3 ω + C4
48 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

is the thermal pressure in the post shock region. We introduced the coefficients

C1 = γl β 2 (4γl2 M 4 − 2γl M 2 − γl − 1)

+β (γl2 + 4γl − 5)M 2 − 2 − γl + 2 ,


 
(2.11)

(γl − 1)M β(M 2 (γl2 + 7γl ) − 2γl + 4) − 2γl + 4 ,



C2 = (2.12)

2γl (γl + 1) β((γl − 1)M 2 + 2) + 2 ,



C3 = (2.13)

C4 = 4(γl + 1)(γl − 2)M. (2.14)

The quantity

γl (γl − 1)βM 2 + 2γl (β + 1) ± W
ω = ω± ≡ − , (2.15)
2γl (γl + 1)βM

is the normal post-shock velocity relative to the shock, with

= β 2 M 2 (γl3 − γl2 ) M 2 (γl − 1) + 4



W

+βγl (4M 2 (4 + γl − γl2 ) + 8γl ) + 4γl2 . (2.16)

Note that ω must satisfy −M < ω < 0 to represent a genuine right moving shock.
We choose the solution where ω = ω+ , since the alternative, ω = ω− is a degenerate
solution in the sense that the hydrodynamical limit lim ω− = 0, which does not
β→+∞
represent a right-moving shock.

2.3.2 Relations across a contact discontinuity and an expansion


fan

Rewriting equation (2.2) in quasilinear form leads to

ux + Fu −1 · Gu uy = 0.

(2.17)

In the frame moving with the triple point, we are searching for self-similar solutions
and we can introduce ξ = xy = tan φ, so that u = u(ξ). Assuming that ξ 7→
u(ξ) is differentiable, manipulating equation (2.17) leads to Auξ = ξuξ . So the
RIEMANN SOLVER BASED SOLUTION STRATEGY 49

eigenvalues λi of A represent tan φ, where φ is the angle between the refracted


signals and the negative x-axis. The matrix A is given by

vy ρvy vy
− v2 ρvx 1
0 0
 
vx 2 −v 2
vx −v 2 vx vx2 −v 2
ms x ms ms
vx vy v2 v 1

 0 2 2
vx −vms − v2 −v
ms
2 − ρy v2 −v 2 0 0 

x ms x ms
vy 1
0 0 0 0
 
vx ρvx
A ≡ F−1
 
u Gu =  2
ρvms vy 2
ρvms vx vx vy
 . (2.18)

 0 − v2 −v 2 2 −v 2
vx 2 −v 2
vx 0 0 

x ms ms ms
Bv vy B vy
− v2Bv 1
 
 0 − v2 −vy2 x
2 2 −v 2 0 
x ms x −vms vx ρ vx ms vx
vy
0 0 0 0 0 vx

and its eigenvalues are

p p
vx vy + vms v 2 − vms
2 vy vy vy vy vx vy − vms v 2 − vms
2
λ1,2,3,4,5,6 ={ 2 2
, , , , , 2 2
},
vx − vms vx vx vx vx vx − vms

(2.19)
p
where the magnetosonic speed vms ≡ 2 2
q q c + va and as before the sound speed
c = γpρ and the Alfvén speed va =
B2
ρ . Since A has 3 different eigenvalues, 3
different signals will arise. When uξ exists and uξ 6= 0, i.e. inside of expansion
fans, uξ is proportional to a right eigenvector ri of A. Derivation of ξ = λi with
respect to ξ gives (∇u λi ) · uλ = 1 and thus we find the proportionality constant,
giving

ri
uξ = . (2.20)
∇u λi · ri

While this result assumed continuous functions, we can also mention relations that
hold even across discontinuities like the CD. Denoting the ratio du ri = κ, it follows
i

that [li · du]dx=λj dy = (li · rj )κ = κδi,j , where li and ri are respectively left and
right eigenvectors corresponding to λi . Therefore, if i 6= j,
[li · du]dx=λj dy = 0. (2.21)
From these general considerations the following relations hold across the contact
dy v
or shear wave where the ratio dx = vxy :
 √
 v dv − v dv + vms v2 −vms 2
dptot = 0,
y x x y
√ρc22 2 (2.22)
 v dv − v dv − vms v −vms dp = 0.
y x x y ρc2 tot
50 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

Since v 6= vms , otherwise all signals would coincide, it follows immediately that the
v
total pressure ptot and the direction of the streamlines vxy remain constant across
the shocked contact discontinuity.
These relations across the CD allow to solve the full problem using an iterative
procedure. Inspired by the exact Riemann solver described in Toro [85], we first
guess the total pressure p∗ across the CD. R is a shock when p∗ is larger than
the post-shock total pressure and T is a shock, only if p∗ is larger than the pre-
shock total pressure. Note that the jump in tangential velocity across the CD is
a function of p∗ and it must vanish. A simple Newton-Raphson iteration on this
v
function [[ vxy ]](p∗ ), finds the correct p∗ . We explain further in section 3.5 how we
find the functional expression and iterate to eventually quantify φR , φT , φCD and
the full solution u(x, y, t). From now on p∗ represents the constant total pressure
across the CD.
Similarly, from√the general considerations above, equation (2.21) gives that along
dy vx vy ±vms v 2 −vms
2

dx = 2 −v 2
vx ms
the following relations connect two states across expansion
fans:
dρ − v21 dptot = 0,


 ms
v2

vx dvx + vy dvy + ρcms2 dptot = 0,




−ρdptot + ptot ρdγ + ptot  γdρ = 0, (2.23)
2
−Bdp + γp + B dB = 0,

tot √



 v dv − v dv ± vms v2 −vms 2


y x x y ρc2 dp = 0.
tot

These can be written in a form which we exploit to numerically integrate the


solution through expansion fans, namely
 R p∗
ρi = ρe + ptot,e v21 dptot ,
R p∗ ±vy √v2 −vms

ms


 2 −v v
 x ms

 vx,i = vx,e + ptot,e ρv 2 vms dptot ,



 √
R p∗ ∓vx v2 −vms 2 −v v
y ms
vy,i = vy,e + ptot,e ρv 2 vms dptot , (2.24)

R p B
B = B + dp ,


i e 2
ρvms tot
R pptot,e


 ∗
c2

pi = pe + ptot,e v2 dptot ,



 ms

γi = γe .

The indices i and e stand respectively for internal and external, the states at both
sides of the expansion fans. The upper signs hold for reflected expansion fans (i.e.
of type R), while the lower sign holds for transmitted expansion fans (i.e. of type
T).
RIEMANN SOLVER BASED SOLUTION STRATEGY 51

2.3.3 Relations across a shock

Since the system is nonlinear and allows for large-amplitude shock waves, the
analysis given thus far is not sufficient. We must include the possibility of one or
both of the R and T signals to be solutions of the stationary Rankine-Hugoniot
conditions (equation (2.6)). The solution is given by

p∗
 γ−1
γ+1 + ptot,e
ρi = γ−1 p∗ ρe ,



γ+1 ptot,e +1


 ∗
 ξ (p −p tot,e )
vx,i = vx,e − ρe ∓(vx,e ξ∓ −vy,e ) ,




p∗ −p

 vy,i = vy,e + ρe (vx,e ξ∓tot,e
−vy,e ) ,


γ−1 p∗
γ+1 + ptot,e
(2.25)

 Bi = γ−1 p∗ Be ,
γ+1 ptot,e +1



γi = γe ,




B2

pi = p∗ − 2i ,




φR/T = arctan(ξ+/− ),

where
p
ve,x ve,y ± ĉe ve2 − ĉ2e
ξ± = 2 − ĉ2
, (2.26)
ve,x e

and
(γ − 1)ptot,e + (γ + 1)p∗
ĉ2e = . (2.27)
2ρe
Again the indices i and e stand respectively for internal and external, the states
at both sides of the shocks. The upper signs holds for reflected shocks, while the
lower sign holds for transmitted shocks.

2.3.4 Shock refraction as a Riemann problem

We are now ready to formulate our iterative solution strategy. Since there exist 2
invariants across the CD, it follows that we can do an iteration, if we are able to
express one invariant in function of the other. As mentioned earlier, we choose to
iterate on p∗ = ptot,3 = ptot,4 . This is the only state variable;e in the solution, and
it controls both R and T. We will write φR = φR (u2 , p∗) and φT = φT (u5 , p∗ ),
u3 = u3 (u2 , p∗ ) and u4 = u4 (u5 , p∗ ). The other invariant should match too,
v v
i.e. vx,3
y,3
− vx,4
y,4
= 0. Since u2 and u5 only depend on the input parameters, this
last expression is a function of p∗ . Iteration on p∗ gives p∗ and φR = φR (p∗ ),
vy,3 v
φT = φT (p∗ ), u3 = u3 (p∗ ) and u4 = u4 (p∗ ) give φCD = arctan vx,3 = arctan vy,4
x,4
,
which solves the problem.
52 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

2.3.5 Solution inside of an expansion fan

The only ingredient not yet fully specified by our description above is how to
determine the variation through possible expansion fans. This can be done once
the solution for p∗ is iteratively found, by integrating equations (2.24) till the
appropriate value of ptot . Notice that
√ the location of the tail of the expansion fan
vy,i vx,i ±ci vi2 −c2i
is found by tan(φtail ) = 2 −c2
vx,i
and the position of φhead is uniquely
i
√ 2 2
vy,e vx,e ±vms,e ve −vms,e
determined by tan(φhead ) = 2 −v 2
vx,e . Inside an expansion fan we
ms,e
know u(ptot ), so now we need to find ptot (φ), in order to find a solution for u(φ).
We decompose vectors locally in the normal and tangential directions, which are
respectively referred to with the indices n and t. We denote taking derivatives
with respect to φ as ′ . Inside of the expansion fans we have some invariants given
by equations (2.23). The fourth of these immediately leads to Bpγ as an invariant.
Eliminating ptot from dρ − c12 dptot = 0 and −Bdptot + (γp + B 2 )dB = 0 yields the
invariant Bρ , and combining these 2 invariants tells us that the entropy S ≡ ρpγ
is invariant. The stationary MHD equations (2.2) can then be written in a 4 × 4-
system for vn′ , vt′ , p′tot and ρ′ as:
 ′

 vn′ + vt + vn ρρ = 0,

p′
vn vt + vn vn′ + tot

ρ = 0, (2.28)
2 ′


 vn − vn vt = 0,
2
vms ρ′ − p′tot = 0,

where we dropped B ′ from the system, since it is proportional to ρ′ . Note that γ ′


vanishes. The system leads to the dispersion relation

vn4 − vms
2
vn2 = 0, (2.29)

which in differential form becomes:

4ρvn3 vn′ +vn4 ρ′ −γvn2 p′tot −2γptot vn vn′ −(2−γ)Bvn2 B ′ −(2−γ)B 2 vn vn′ = 0. (2.30)

Elimination of vn′ , ρ′ and B ′ gives


2
dptot vt vms − 2vn2 2
=2 ρvms . (2.31)
dφ vn 3vn2 + (γ − 2)vms
2

This expression allows us to then complete the exact solution as a function of φ.


IMPLEMENTATION AND NUMERICAL DETAILS 53

2.4 Implementation and numerical details

2.4.1 Details on the Newton-Raphson iteration

We can generally note that ptot,pre < ptot,post , where the indeces pre and post
refer to respectively the pre- and the post-shock region. This implies that the
refraction has 3 possible wave configurations: 2 shocks, a reflected rarefaction
fan and a transmitted shock, or 2 expansion fans. Before starting the iteration
v v
on [[ vyx ]](p∗ ), we determine the governing wave configuration. If [[ vyx ]](ǫ) and
vy v
[[ vx ]](ptot,5 −ǫ) differ in sign, the solution has two rarefaction waves. If [[ vyx ]](ptot,5 +
vy
ǫ) and [[ vx ]](ptot,2 − ǫ) differ in sign, the solution has a transmitted shock and a
reflected rarefaction wave. In the other case, the solution contains two shocks in
its configuration. If R is an expansion fan, we take the guess
2
2ρe vx,e −(γe −1)ptot,e
min{ γ+1 |e ∈ {2, 5}} + ptot,5
p∗0 = (2.32)
2
as a starting value of the iteration. This guess is the mean of the critical value
ptot,crit , which satisfies
2
ve,x − ĉ2 (ptot,crit ) = 0, (2.33)
and p5 , which is the minimal value for a transmitted shock. As we explain in section
2
5.3, v2,x − ĉ2 (ptot,crit ) = 0 is equivalent to v52 − ĉ2 = 0 and v5,x
2
− ĉ2 (ptot,crit ) = 0
2 2
is equivalent to v2 − ĉ = 0, and is thus a maximal value for the existence of a
regular solution. If R is a shock, we take (1 + ǫ̂)ppost as a starting value for the
f (p∗ )
iteration, where ǫ̂ is 10−6 . We use a Newton-Raphson iteration: p∗i+1 = p∗i − f ′ (pi∗ ) ,
i
f (p∗ ∗
i +δ)−f (pi )
where f ′ (p∗ ) is approximated numerically by δ , where δ = 10−8 . The
p∗ ∗
i+1 −pi −8
iteration stops when p∗ < ǫ, where ǫ = 10 .
i

2.4.2 Details on AMRVAC

AMRVAC (van der Holst & Keppens [93]; Keppens et al.[49]) is an AMR code,
solving equations of the general form ut +∇·F(u) = S(u, x, t) in any dimensionality.
The applications cover multi-dimensional HD, MHD, up to special relativistic
magnetohydrodynamic computations. In regions of interests, the AMR code
dynamically refines the grid. The initial grid of our simulation is shown in figure 2.4.
The refinement strategy is done by quantifying and comparing gradients. A Löhner
estimator is used, tolerance ǫl = 0.025, with refinement triggered on ρ and auxiliary
variable ωz , in equal weights.
The AMR in AMRVAC is of a block-based nature, where every refined grid
has 2D children, and D is the dimensionality of the problem. Parallelization is
54 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

Figure 2.4: The initial AMR grid at t = 0, for the example in section 5.1.

implemented, using MPI. In all the simulations we use 5 refinement levels, starting
with a resolution of 24 × 120 on the domain [0, 1] × [0, 5], leading to an effective
resolution of 384 × 1940. The shock is initially located at x = 0.1, while the
contact discontinuity is located at y = (x − 1) tan α. We used the fourth order
Runge-Kutta timestepping, together with a TVDLF-scheme (see Tóth & Odstrc̆il
[86]; Yee [102]) with Woodward-limiter on the primitive variables. The obtained
numerical results were compared to and in agreement with simulations using other
schemes, such as a Roe scheme and the TVD-Muscl scheme. The calculations were
performed on 4 processors.

2.4.3 Following an interface numerically

The AMRVAC implementation contains slight differences with the theoretical


approach. Implementing the equations as we introduced them here would lead
to excessive numerical diffusion on γ. According to equations 2.24 and 2.25 the
value of γ does not change through an expansion fan or across a shock. Therefore
γ is a discrete variable in the sense that it only takes a finite number of values (in
our case only 2, namely γ ∈ {γl , γr }), we know γ(x, y, t) exactly, if we are able to
follow the contact discontinuity in time.
Suppose thus that initially a surface, separates 2 regions with different values of γ.
Define a function χ : D × R+ → R : (x, y, t) 7→ χ(x, y, t), where D is the physical
domain of (x, y). Writing χ̃(x, y) = χ(x, y, 0), we ask χ̃ to vanish on the initial
contact and to be a smooth function obeying

• γ = γl ⇔ χ̃(x, y) < 0,
• γ = γr ⇔ χ̃(x, y) > 0.

We take in particular ±χ̃ to quantify the shortest distance from the point (x, y) to
the initial contact, taking the sign into account. Now we only have to note that
(χρ)t = χρt + ρχt = −χ∇ · (ρv) − (ρv · ∇)χ = −∇ · (χρv).
RESULTS 55

a) b)
all shock solver right shock solver
6 6

4 4

2 2
[[vy/vx]]

[[vy/vx]]
0 0

-2 -2

-4 -4

-6 -6

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
p* p*

c) d)
no shock solver combined solver
6 6

4 4

2 2
[[vy/vx]]

[[vy/vx]]

0 0

-2 -2

-4 -4

-6 -6

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
p* p*

hh ii
vy
Figure 2.5: vx (p∗ ) for the reference case from Samtaney [78]: a) all shock
solver; b) right shock solver; c) no shock solver; d) shock ⇔ p∗ > pi . The all shock
solver is selected.

The implemented system is thus (2.2), but the last equation is replaced by
(χρvx )x +(χρvy )y = 0. Every timestep we reset the value of γ to γl or γr , depending
on the sign of χ.
It is now straightforward to show that we did not introduce any new signal. In
essence, this is the level set approach, as presented in e.g. Mulder et al.[59].

2.5 Results

2.5.1 Fast-Slow example

As a first hydrodynamical example, we set α, β −1 , γl , γr , η, M = π4 , 0, 75 , 75 , 3, 2 ,


 

as originally presented in Samtaney [78]. In figure 2.5, the first 3 plots show
v
[[ vxy ]](p∗ ), when assuming a prescribed wave configuration, for all 3 possible
v
configurations. The last plot shows the actual function [[ vyx ]](p∗ ), which consists of
56 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

0
14

12 -0.5

10
-1
8

vx
ρ

6 -1.5

4
-2
2

0 -2.5
0 1 2 3 4 5 6 0 1 2 3 4 5 6
φ φ
0
14

-0.5 12

10
-1
8
vy

-1.5 6

4
-2
2

-2.5 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
φ φ
0.8
1.4
0.7
1.2
0.6
1
0.5
0.8
vx/vy

0.4
S

0.6
0.3

0.2 0.4

0.1 0.2

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
φ φ

Figure 2.6: Solution to the fast-slow refraction problem, for the reference case from
Samtaney [78]. Notice that p and vvxy remain constant across the shocked contact.
RESULTS 57

120 50

100
40

80
30

60

S
p

20
40

10
20

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
φ φ

Figure 2.7: Solution to the slow-fast refraction problem from van der Holst &
Keppens [93]. Notice that S remains constant across R.

piecewise copies from the 3 possible configurations in the previous plots. The initial
guess is p∗0 = 4.111, the all shock solver is selected, and the iteration converges
after 6 iterations with p∗ = 6.078. The full solution of the Riemann problem is
shown in figure 2.6.

2.5.2 Slow-Fast example

In figure 2.7 we show the full solution of the HD Riemann problem, in which the
reflected signal is an  expansion fan,1 connected to the refraction with parameters
α, β −1 , γl , γr , η, M = π3 , 0, 57 , 75 , 10

, 10 from van der Holst & Keppens [93]. The
v
refraction is slow-fast, and R is an expansion fan. Note that p and vxy remain
constant across the CD, and the entropy S is an invariant across R.

2.5.3 Tracing the critical angle for regular shock refraction

Let us examine what the effect of the angle of incidence,α, is. Therefore  we get
−1 7 7
back to the example from Section 2.5.1, β , γl , γr , η, M = 0, ,
5 5 , 3, 2 and let
α vary: α ∈ 0, π2 . Note that α = π2 corresponds to a 1-dimensional Riemann
 

problem. The results are shown in figure 2.8. Note that for regular √ refraction
2 ve,x ve,y ±ĉe v 2 −ĉ2
vy,5 > ĉ25 . We can understand this by noting that ξ± = 2 −ĉ2
ve,x
e e
=
e
 √ 2 2 −1
ve,x ve,y ∓ĉe ve −ĉe
v 2 −ĉ2 = ξˆ∓ , which are the eigenvalues of Gu −1 · Fu = (Fu −1 ·
e,y e

Gu )−1 . Note that we could have started our theory from the quasilinear form uy +
(Gu −1 · Fu )ux = 0 instead of equation (2.17). If we would have done so, we would
1
have found eigenvalues ξ̂, which would correspond to arctan φ . Moreover, both the
eigenvalues, ξ+ and ξ− , have 4 singularities, namely ĉ2 ∈ {−vx,2 , vx,2 , −vy,5 , vy,5 }
58 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

6.4
φR
1.4 φCD
6.35 φT

6.3 1.2

6.25 1

0.8
p*

6.2

φ
6.15 0.6

6.1 0.4

6.05 0.2

6 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4
α α
2
0.7 |vy,5|
c5
0.6
1.8

0.5
1.6
0.4
[[vt]]

0.3 1.4

0.2
1.2
0.1

0 1
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4
α α

Figure 2.8: Upper Left : p∗ (α). Note that for α < 0.61, there are no solutions for p∗ :
the refraction is irregular; Upper Right: the wave pattern for regular refraction;
Lower Left : For α = π2 , the problem is 1-dimensional and there is no vorticity
deposited on the interface. For decreasing α, the vorticity increases. Lower right:
For regular refraction, |vy,5 | > ĉ5 .

for ξ− and ĉ5 ∈ {−vx,5 , vx,5 , −vy,2 , vy,2 } for ξ+ , where thus ĉ25 = v5,y
2
⇔ ĉ22 = v22
2 2 2 2
and ĉ2 = vy,2 ⇔ ĉ5 = v5 . It is now clear that it is one of the latter conditions
that will be met for αcrit . In the example, the transition to irregular refraction
2 2
occurs at −vy,5 = ĉ5 and lim p∗ = 2γr ηM tanγl +1 (αcrit )−γl +1
= 6.67. Figure 2.9
α→αcrit
shows Schlieren plots for density from AMRVAC simulations for the reference case
α = π4 , and the irregular case where α = 0.3. In the regular case, all signals meet
at the triple point, while for α < αcrit = 0.61, the signals do not meet at one triple
point, the triple point forms a more complex structure and becomes irregular. The
CD, originated at the Mach stem, reaches the triple point through an evanescent
wave, which is visible by the contour lines. This pattern is called Mach Reflection-
Refraction. Decreasing α even more, the reflected wave transforms in a sequence
of weak wavelets (see e.g. Nouragliev et al.[64]). This pattern, of which the case
α = 0.3 is an example, is called Concave-Forwards irregular Refraction. These
results are in agreement with our predictions.
RESULTS 59

Figure 2.9: Schlieren plots of the density for β −1 , γl , γr , η, M = 0, 57 , 75 , 3, 2


 

with varying α. Upper: α = π4 : a regular reference case. Lower: α = 0.3: an


irregular case.

2.5.4 Abd-El-Fattah and Hendersons experiment

In 1978, a shock tube experiment was performed by Abd-El-Fattah & Henderson


[3]. It became a typical test problem for simulations (see e.g. Nouragliev et al.[64])
and refraction theory (see e.g. Henderson [42]). The experiment concerns a slow-
fast shock refraction at a CO2/CH4 interface. The gas constants are γCO2 =
µ
1.288, γCH4 = 1.303, µCO2 = 44.01 and µCH4 = 16.04. Thus η = µCH 4
CO2
= 0.3645.
A very weak shock, M = 1.12 is refracted at the interface under various angles.
von Neumann [62] theory predicts the critical angle αcrit = 0.97 and the transition
angle αtrans = 1.01, where the reflected signal is irregular if α < αcrit , a shock if
αcrit < α < αtrans and an expansion fan if αtrans < α. This is in perfect agreement
with the results of our solution strategy as illustrated in figure 2.10. There we show
60 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

1.38
p* 3 φR
ppost φCD
φT
1.36 2.5

1.34 2
p

φ
1.32 1.5

1
1.3

0.5
1.28
0
0.96 0.97 0.98 0.99 1 1.01 1.02 1.03 0.9 1 1.1 1.2 1.3 1.4 1.5
α α

Figure 2.10: Exact solution for the Abd-El-Fattah experiment. Left: p∗ (α)
confirms αcrit = 0.97 and αtrans = 1.01. Right: φ(α).

the pressure p∗ compared to the post shock pressure ppost , as well as the angles
φR , φCD and φT for varying angle of incidence α. Irregular refraction means that
not all signals meet at a single point. The transition at αcrit is one between a
regular shock-shock pattern and an irregular Bound Precursor Refraction, where
the transmitted signal is ahead of the shocked contact and moves along the contact
at nearly the same velocity. This is also confirmed by AMRVAC simulations. If
the angle of incidence, α, is decreased even further, the irregular pattern becomes
a Free Precursor Refraction, where the transmitted signal moves faster than the
shocked contact, and reflects itself, introducing a side-wave, connecting T to CD.
When decreasing α even further, another transition to the Free Precursor von
Neumann Refraction occurs.

2.5.5 Connecting slow-fast to fast-slow refraction

Another example of how to trace transitions by the use of our solver is done by
changing the density ratio η across the CD. Let us start from the example given
in section 5.1 and let us vary the value of η.
Here we have α, β −1 , γl , γr , M = π4 , 0, 75 , 75 , 2 . The results are shown in
 

figure 2.11. Note that, since ppost = 4.5, we have a reflected expansion fan for fast-
slow refraction, and a reflected shock for slow-fast refraction. The transmitted
signal plays a crucial role in the nature of the reflected signal: for fast-slow
refraction φT < π2 , but for slow-fast refraction, φT > π2 and the transmitted
signal bends forwards. We ran our solver for varying values of M and α, and for
all HD experiments with γl = γr , we came to the conclusion that a transition from
fast-slow to slow-fast refraction, coincides with a transition from a reflected shock
to a reflected expansion fan, with φT = π2 . This result agrees with AMRVAC
simulations. In figure 2.12, a density plot is shown for η = 1.2 and η = 0.8.
RESULTS 61

6 2
p* φR
ppost φCD
φT
5.5 π/2
1.5
5

4.5 1
p

φ
4
0.5
3.5

3 0
0.6 0.8 1 1.2 1.4 1.6 1.8 2 0.6 0.8 1 1.2 1.4 1.6 1.8 2
η η

Figure 2.11: Exact solution for α, β −1 , γl , γr , M = π4 , 0, 75 , 75 , 2 and a varying


 
range of the density ratio η. Left: for η < 1 we have p∗ < ppost = 4.5 and thus a
reflected expansion fan, for η > 1 we have p∗ > ppost = 4.5 and thus a reflected
shock. Right: for η < 1: φT < π2 and for η > 1: φT > π2 .

Figure 2.12: Density plots for α, β −1 , γl , γr , M = π4 , 0, 75 , 75 , 2 . Left: A


 

slow/fast refraction with η = 0.8. Note that φT > π2 and R is an expansion


fan. Right: A fast/slow refraction with η = 1.2. Note that φT < π2 and R is a
shock.

2.5.6 Effect of a perpendicular magnetic field

So far, none of our results include magnetic fields. We now consider the case where
a purely out-of-plane magnetic field, where the field is perpendicular to the shock
front and thus acts to increase the total pressure and the according flux terms, is
added.
Note that it follows from equation 2.24 and 2.25 that Bρ is invariant across
shocks and rarefaction fans. Therefore, Bρ can only jump across the shocked and
unshocked contact discontinuity and B cannot change sign.
Revisiting the example from section 5.1, we now let the magnetic field vary.
62 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

0.35 0
[[vt]] [[vt]]

0.3
-0.02
0.25

-0.04
0.2
[[vt]]

[[vt]]
0.15
-0.06

0.1
-0.08
0.05

0 -0.1
0.6 0.8 1 1.2 1.4 1.6 1.8 2 0.6 0.8 1 1.2 1.4 1.6 1.8 2
β β

Figure 2.13: Left: Solution for the fast-slow problem: strong perpendicular
magnetic fields decrease the instability of the CD. Right: Solution for the slow-fast
problem: strong perpendicular magnetic fields decrease the instability of the CD.

Figure 2.13 shows [[vt ]](β) across the CD. Also for η = 0.8, making it a slow-fast
problem, [[vt ]](β) is shown. First notice that no shocks are possible for β < 0.476,
since ω+ would not satisfy ω+ > −M . Manipulating equation (2.15), we know
that this is equivalent to
2
β > βmin ≡ . (2.34)
γl (M 2 − 1)

This relation is also equivalent to c1 > M , which means that the shock is
submagnetosonic, compared to the pre-shock region. Figure 2.14 shows density
plots from AMRVAC simulations at t = 2.0, for (α, γl , γr , η, M ) = π4 , 57 , 75 , 3, 2
with varying β −1 . First note that the interface is unstable for the HD case.
Increasing β −1 decreases the shock strength. For β −1 = 1 the shock is very
weak: the Atwood number At = 0.17, and the interface remains stable. here we
introduced the Atwood number, which is defined in 2.35.
Shown in figure 2.13, is the vorticity across the CD. In the limit case of this minimal
plasma-β the interface is stable, both for fast-slow and slow-fast refraction. As
expected, in the fast-slow case, the reflected signal is an expansion fan, while it is
a shock in the fast-slow case. Also note that the signs of the vorticity differ, causing
the interface to roll up clockwise in the slow-fast regime, and counterclockwise in
the fast-slow regime. When decreasing the magnetic field, the vorticity on the
interface increases in absolute value. This can be understood by noticing that the
limit case of minimal plasma-β is also the limit case of very weak shocks. This can
for example be understood by noting that lim φCD = α (see figure 2.15). A
β→βmin
convenient way to measure the strength of a shock is by use of its Atwood number.
This is often done in the classical HD case (see, e.g., Sadot et al.[75])
ρ2 − ρ1
At = . (2.35)
ρ2 + ρ1
RESULTS 63

Figure 2.14: Density plots at t = 2.0 for (α, γl , γr , η, M ) = π4 , 57 , 75 , 3, 2




with varying β −1 . Upper: β −1 = 0. The hydrodynamical Richtmyer-Meshkov


instability causes the interface to roll up. Center: β −1 = 12 . Although the initial
amount of vorticity deposited on the interface is smaller than in the HD case, the
wall reflected signals pass the wall-vortex and interact with the CD, causing the
RM I to appear. Lower: β −1 = 1. The shock is very weak and the interface
remains stable.
64 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

2.4
1.4
2.2
1.2

2
1

[[vt]]/At
1.8
φCD

0.8

0.6 1.6

0.4 1.4

0.2 1.2

0 1
0.6 0.8 1 1.2 1.4 1.6 1.8 2 0.6 0.8 1 1.2 1.4 1.6 1.8 2
β β

Figure 2.15: The reference problem from Samtaney [78] with varying β. Left: The
dependence of φCD on β. Note that lim φCD = π4 = α, since this is the limit
β→βmin
to infinitely weak shocks: lim At = 0. Right: The vorticity deposition in the
β→βmin
shocked contact scales as the Atwood number and lim [[vt ]] = 1.
β→βmin At

Figure 2.15 shows the jump across the shocked contact [[vt ]], scaled to the shocks
Atwood number. Note that in the limit case of very weak shocks the Atwood
number equals the jump in tangential velocity across the CD, in dimensional
notation:
[[vt ]]
c1
lim = 1. (2.36)
β→βmin At

When keeping the Atwood number constant, the shocks sonic Mach number is
given by
s
1 + At (2 − 2γ − γβ)At2 + (2γβ + 2γ)At − γβ − 2
M = (2.37)
1 − At (γ 2 β)At2 + (γ 2 β − γβ)At − γβ
s
(At + 1)((γβ + 2γ − 2)At − (γβ + 2))
= . (2.38)
γβ(1 − At)(γAt + 1)

Note that in the limit for weak shocks


s
γβ + 2
lim M = , (2.39)
At→0 γβ

which is equivalent to equation 2.34, and in the limit for strong shocks, M →
∞. Figure 2.16 shows the deposition of vorticity on the shocked contact, for a
constant Atwood number. We conclude that under constant Atwood number, the
RESULTS 65

1 6

5
0.9

4
0.8
[[vt]]

M
3

0.7
2

0.6
1

0.5 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
β β

Figure 2.16: Left: The dependence of the vorticity deposition on the shocked
contact in function of β, under constant Atwood number At = 0.17. Right: The
shocks Mach number in function of the plasma-β, under constant Atwood number
At = 0.17.

B 5
Figure 2.17: AMRVAC plots of ρ for At = 11 , with varying beta. Upper: β = 16.
Lower: β = 0.25.

effect of a perpendicular magnetic field is small: Stronger perpendicular magnetic


field increase the deposition of vorticity on the shocked contact slightly. This is
confirmed by AMRVAC simulations (see figure 2.17).
66 AN EXACT RIEMANN SOLVER FOR REGULAR SHOCK REFRACTION

2.6 Conclusions

We developed an exact Riemann solver-based solution strategy for shock refraction


at an inclined density discontinuity. Our self-similar solutions agree with the early
stages of nonlinear AMRVAC simulations. We predict the critical angle αcrit for
regular refraction, and the results fit with numerical and experimental results.
Our solution strategy is complementary to von Neumann theory, and can be used
to predict full solutions of refraction experiments, and we have shown various
transitions possible through specific parameter variations. For perpendicular fields,
the stability of the contact decreases slightly with decreasing β under constant
Atwood number. We will generalize our results for planar uniform magnetic fields,
where up to 5 signals arise. In this case we search for non-evolutionary solutions,
involving intermediate shocks.
Chapter 3

Shock Refraction in ideal MHD

In this chapter we generalize our solution strategy of chapter 2 to regular shock


refraction in ideal MHD. While in the HD case three signals arise, in the planar
MHD case five signals arise. Our Riemann solver finds the exact position of these
five signals, and determines the values of the conserved variables in the new-formed
regions.
After reflection from the top wall the interface remains stable since MHD does not
allow for vorticity deposition on a contact discontinuity.
This Chapter is published as Delmont & Keppens [20].

3.1 Introduction

The interaction of a shock wave with an inclined density discontinuity is a classical


hydrodynamical (HD) shock tube problem. When the impinging shock refracts
at the density discontinuity, 3 signals arise: a reflected signal, a transmitted
signal and a shocked contact in between. Both the reflected and transmitted
signal can be continuous (rarefaction fans) or discontinuous (shocks). Due to
local Kelvin-Helmholtz instability, the shocked contact becomes unstable, and the
Richtmyer-Meshkov instability (RMI) forms. In Chapter 2 (or Delmont et al.
[19]), we solved the initial self similar phase of the problem exactly, exploiting
exact Riemann solver methods. In this chapter, we extend that study to planar
ideal magnetohydrodynamics (MHD).
For that planar case, Samtaney [78] showed by numerical grid-adaptive simulations
that addition of a uniform magnetic field, aligned with the shock normal,

67
68 SHOCK REFRACTION IN IDEAL MHD

suppresses the RMI. The setup of the MHD problem is shown in figure 3.1. The
rectangular domain mimics a shock tube, thus the left and the right boundaries
are modeled as open, while the upper and the lower boundaries are rigid walls.
Completely to the left of the domain, there is a genuine right-moving shock, which
moves with sonic Mach number M . It moves towards a contact discontinuity,
which forms an inclination angle α with the shock normal and separates two gases
at rest, with a density ratio η across it. The ratio of specific heats, γ, is considered
a constant equation parameter, and the initially uniform applied magnetic field
introduces a plasma-β (in the pre-shock region). These 5 parameters define the
problem uniquely. In the MHD case up to 5 signals arise: a fast (FR) and a slow
(SR) reflected signal, a contact discontinuity (CD), and a slow (ST) and a fast
(FT) transmitted signal, separating the 4 new formed states, as shown in figure
3.3.
Figure 3.2 shows simulation snapshots for the case where (α, γ, M, η) = ( π4 , 75 , 2, 3),
after reflection from the top wall. The upper snapshot shows the hydrodynamical
case (β −1 = 0) and the lower snapshot shows an MHD case where β = 2. In the
first case the RMI has formed on the interface, while it is clearly suppressed in
the magnetohydrodynamical case. We refer to Delmont et al.[19] for an exact
solution of the HD case. In this paper we generalize the presented solution
strategy to planar ideal MHD. Our results are in agreement with results from
Wheatley et al.[98]. Our approach is inspired by Wheatley et al.[98]. However,
our solution strategy for the RH jump conditions is essentially different and the
iteration method used is more detailed.
We also compare our results with numerical grid adaptive simulations. These
simulations are performed by the Adaptive Mesh Refinement code AMRVAC (see
e.g. Keppens et al.[49]; van der Holst & Keppens [93]).
The suppression of the RMI in ideal MHD can be explained as a direct consequence
of the Rankine-Hugoniot jump conditions across a shocked contact. It is well-
known that contact discontinuities do not allow for a jump in tangential velocity,
when the normal magnetic field component does not vanish (see e.g. Goedbloed
& Poedts [32]).
In section 3.2, we introduce the equations employed in the analytical solution
strategy, namely the planar stationary MHD equations and the Rankine-Hugoniot
conditions from their integral form. In section 3.3, we summarize some important
features of MHD shocks. In section 3.4, we describe our solution strategy, which
is based on the Riemann solver for ideal MHD presented by Torrilhon [91] and
in section 3.5 we demonstrate the algorithm by solving a case presented first
by Samtaney in [78]. In section 6 finally, we compare our results to AMRVAC
simulations.
INTRODUCTION 69

(ρ, v = 0, p, B) (ηρ, v = 0, p, B)

M α

Figure 3.1: Initial configuration: a shock moves with shock speed M to an inclined
density discontinuity. Both the upper and lower boundary are solid walls, while
the left and the right boundaries are open.

Figure 3.2: In both cases (α, γ, M, η) = ( π4 , 75 , 2, 3). Up: β −1 = 0. When


a hydrodynamical shock impinges on a contact discontinuity, the discontinuity
becomes unstable due to local Kelvin-Helmholtz instability. Down: β −1 = 12 .
When a normal magnetic field is applied, the Rankine-Hugoniot conditions do not
allow for vorticity deposition on a contact discontinuity. Therefore, the interface
remains stable and the RMI is suppressed.
70 SHOCK REFRACTION IN IDEAL MHD

3.2 Governing equations

In order to describe the dynamical behavior of ionized, quasi-neutral plasmas, we


use the framework of ideal MHD. The initial configuration sketched in figure 3.1
will lead to a refraction pattern as sketched in figure 3.3. Regular refraction means
that all signals meet at a single quintuple point. This quintuple point moves along
the unshocked contact at speed vqp = (M, M tan α). We will solve the problem
in a co-moving frame with the quintuple point. We can then assume that the
solution is self similar and time independent: ∂/∂t = 0. Therefore, we can exploit
the stationary MHD equations, which written out in conservative form and for our
planar problem, are given by

∂ ∂
F+ G = 0, (3.1)
∂x ∂y
where we introduced the flux terms for the orthogonal x and y directions given by

ρvx
 
2
B2 B

 ρvx2 + p − 2x + 2y 

F=
 ρvx vy − Bx By 
 (3.2)
2 2
v +v
 vx ( γ p + ρ x y + By2 ) − Bx By vy 
 
γ−1 2
Bx

and
ρvy
 
 ρvx vy − Bx By 
B2 B2
 
G=
 ρvy2 + p + 2x − 2y .

(3.3)
2 2
 vy ( γ p + ρ vx +vy + Bx2 ) − Bx By vx
 

γ−1 2
By

This set of equations expresses conservation of mass density, momentum and


energy. The conserved variables and the components of the flux terms are written
in terms of the mass density ρ, the velocity components vx and vy , the thermal
pressure p and the magnetic field components Bx and By . Faraday’s law ensures
conservation of magnetic flux, which in the stationary case becomes ∇ × E = 0.
The electric field Ez is therefore a global Riemann invariant:

Ez = vy Bx − vx By . (3.4)
v B −E
This allows us to eliminate By = y vxx z from the system, which is reduced to a
5 × 5-system. We hereby exploit the fact that vx (x, t) < 0 in the co-moving frame
GOVERNING EQUATIONS 71

with the quintuple point. In the rest frame this means that the velocity in the
x-direction is bounded by the shock speed, during the refraction process.
We allow weak (i.e., discontinuous) solutions of the system, which are solutions
of the integral form of the MHD equations. The discontinuities occurring in the
problem setup (both the impinging shock and the initial contact discontinuity),
as well as those that may appear as F R, SR, ST or F T signals later on obey
the Rankine-Hugoniot conditions. In Appendix D we give the planar stationary
Rankine-Hugoniot conditions and rewrite them as a 4 × 4-system in [[Bt ]], [[p]], [[ρ]]
and [[vt ]], where [[Q]] refers to the jump in the quantity Q across the CD, and
tangential vector components are referred to by a subscript t.
We now discuss the characteristics of the system (3.1), by writing it out in
quasilinear form

ux + Auy = 0, (3.5)
∂F
where the matrix A is computed from the flux Jacobian matrices Fu ≡ ∂u and
Gu ≡ ∂G
∂u as follows:

A = F−1
u · Gu . (3.6)

Here u can be any arbitrary vector for which Fu is invertible. We set u =


(ρ, vx , vy , p, Bx ). As argued in Delmont et al.[19], the eigenvalues of A coincide
with the tangent of the refraction angles φF R , φSR , φCD , φST and φF T (see figure
3.3). Since the system has 5 distinct eigenvalues, the shock indeed refracts in 5
distinct signals.
Computing the eigenvalues λ of A, leads to the characteristic equation

det(A − λI) = (vx λ − vy ) Σi=0,4 ti λi = 0.



(3.7)
v
This is a quintic with 5 solutions. One eigenvalue λ = vxy is linearly degenerate
and corresponds to the CD. The 4 other eigenvalues are found as roots of a quartic,
with coefficients given by

t4 = vx4 + vx2 (4c2 − 2va2 ) + 2c2 va,x


2
, (3.8)

t3 = 4vx vy (c2 + va2 − vx2 ) − 8c2 va,x va,y , (3.9)

t2 = 6vx2 vy2 − 2va2 (v 2 + c2 ) + 2v 2 c2 , (3.10)

t1 = 4vx vy (c2 + va2 − vy2 ) − 8c2 va,x va,y , (3.11)

t0 = vy4 + vy2 (4c2 − 2va2 ) + 2c2 va,y


2
. (3.12)
72 SHOCK REFRACTION IN IDEAL MHD

u1

u2 u7

φF R = atan(λF R )

u3 u4 u5 u6

Figure 3.3: The wave pattern during interaction of the shock with the CD. The
eigenvalues of the matrix A coincide with the tangent of the refraction angles.
Since A has 5 distinct eigenvalues, 5 signals arise.

q
In these expressions we denote the sound speed c = γp ρ , the directional Alfvén
q 2 q 2
By
q
Bx 2 + v2 .
speeds va,x = ρ , va,y = ρ and the total Alfvén speed va = va,x a,y

We can rewrite the characteristic equation (3.7) as


2 2 2 2
v⊥ (v⊥ − vf,⊥ )(v⊥ − vs,⊥ ) = 0, (3.13)

where the (squared) normal speed is given by

2 (vx λ − vy )2
v⊥ = , (3.14)
1 + λ2
the (squared) fast speed by
s !
2 1 2 2 2 2 2 2
(Bx λ − By )2
vf,⊥ = c + va + (c + va ) − 4c , (3.15)
2 ρ(1 + λ2 )

and the (squared) slow speed by


s !
2 1 2 2 2 2 2 2
(Bx λ − By )2
vs,⊥ = c + va − (c + va ) − 4c . (3.16)
2 ρ(1 + λ2 )

This then clearly shows that the MHD system has the fast and the slow
magnetosonic speeds in the direction perpendicular to the shock front as its basic
characteristic speeds.
All the refracted magnetosonic signals can be expansion fans or shocks.
MHD SHOCKS 73

3.3 MHD shocks

MHD shocks are a topic of extensive research (see e.g. Liberman &qVelikhovich
Bn2
[56]; Goedbloed [33]). Introducing the normal Alfvén speed va,n ≡ ρ , which
is an eigenvalue
q of the full
q three dimensional stationary MHD equations, and
Bt2 2 + v 2 . The full set of eigenvalues of the full set
va,t ≡ ρ leads to va = va,n a,t
of MHD equations is given by {±vf,n , ±va,n , ±vs,n , 0}, where the fast speed vf,n
and the slow speed vs,n are respectively defined as in equations (1.48) and (1.48).
Since 0 ≤ vs,n ≤ va,n ≤ vf,n , the full set of MHD equations is non-strictly
hyperbolic. Those speeds define the up- and downstream states into 4 types (and
3 transition types plus 1 stationary type for completeness), which in order of
increasing entropy S ≡ ρpγ :

• (1) superfast: vf,n < |vn |;


• (1=2) fast: |vn | = vf,n ;
• (2) subfast: va,n < |vn | < vf,n ;

• (2=3) Alfvén: va,n = vn ;


• (3) superslow: vs,n < |vn | < va,n ;
• (3=4) slow: |vn | = vs,n ;
• (4) subslow: |vn | < vs,n ;
• (∞) stationary: vn = 0.

The second law of thermodynamics enforces [[S]] > 0 across a shock. If this
inequality is satisfied, we call a shock admissible. When the upstream state is of
type i and the downstream state is of type j, then we define the shock to be of
type i → j. The RH conditions (D.1) now allow for the following types of shocks
(see also Falle & Komissarov[27], De Sterck et al.[22], Liberman & Velikhovich [56]
and many others):

• (1 → 2)-shocks are called fast shocks. They satisfy Bt,2 > Bt,1 > 0 or
Bt,2 < Bt,1 < 0, i.e. the magnetic field gets refracted away from the shock
normal.
• (3 → 4)-shocks are called slow shocks. They satisfy Bt,1 > Bt,2 > 0 or
Bt,1 < Bt,2 < 0, i.e. the magnetic field gets refracted towards the shock
normal.
74 SHOCK REFRACTION IN IDEAL MHD

• (1 → 2 = 3)-shocks are called switch-on shocks, since they satisfy Bt,1 = 0,


i.e. the upstream magnetic field is aligned with the shock normal.
• (2 = 3 → 4)-shocks are called switch-off shocks, since they satisfy Bt,2 = 0,
i.e. the downstream magnetic field is aligned with the shock normal.
• (1 → 3), (1 → 4), (2 → 3) and (2 → 4)-shocks are called intermediate shocks.
They satisfy Bt,2 ≥ 0 ≥ Bt,1 or Bt,2 ≤ 0 ≤ Bt,1 , i.e. the upstream magnetic
field is aligned with the shock normal.
• (1 → 4)-shocks for which Bt,1 = Bt,2 = 0 also satisfy vt,1 = vt,2 = 0 and
are essentially 1-dimensional. In this case both the u- and downstream
magnetic field are aligned with the shock normal. These shocks are called
hydrodynamical shocks;
• (2 = 3 → 2 = 3)-discontinuities are called Alfvén discontinuities (or,
alternatively, rotational discontinuities). They satisfy Bt,1 + Bt,2 = 0 such
that the upstream and the downstream state are equal, except for a sign
change of Bt ;
• (∞ → ∞)-discontinuities can be contact discontinuities, where only ρ jumps
across them, or tangential discontinuities, where Bn and vn vanish.

The latter two cases are called linear discontinuities, since [[vn ]] vanishes. The
other cases are called MHD shocks. The set of planar MHD equations reduces
the number of possibilities, since we now only have characteristic speeds given by
{±vf , ±vs , 0}, making the planar system strictly hyperbolic.
The existence of intermediate shocks is controversial since they are believed to be
unstable under small perturbations. This topic is widely discussed in the literature
(theoretically in e.g. Barmin et al.[8]; Falle & Komissarov [26, 27]; Myong & Roe
[60, 61], observationally in e.g. Chao et al.[12]; Feng & Wang [29] and numerically
in e.g. Brio & Wu [11], De Sterck et al.[22]). We allow for intermediate shocks
in our solution strategy, since they form the central Alfvénic transition for MHD
shocks (Goedbloed [33]), and as we show later, they naturally emerge in high
resolution numerical simulations.

3.4 Riemann Solver based solution strategy

3.4.1 Dimensionless representation

We now present how we tackle the problem in a dimensionless manner. We choose


a representation in which the initial shock speed s equals its sonic Mach number
M . We determine the value of the primitive variables in the post-shock region
RIEMANN SOLVER BASED SOLUTION STRATEGY 75

vn > vf,n vf,n > vn > va,n va,n > vn > vs,n vs,n > vn

(a) (b)

vn > va,n va,n > vn vn = va,n vn > vf,n = va,n

(c) (d)

vn = an vn > va,n = vs,n vn = an vn = an

(e) (f)
76 SHOCK REFRACTION IN IDEAL MHD

vn = 0 vn = 0 vn = 0 vn = 0

(g) (h)

vn > vf = an an = vs > vn

(i)

Figure 3.4: The classical 1 − 2 − 3 − 4 classification of MHD discontinuities: a)


fast shock; b) slow shock; c) intermediate shock; d) switch-on shock; e) switch-off
shock; f)Alfvén discontinuity; g) contact discontinuity; h) tangential discontinuity;
i) hydrodynamical shock.

by applying the stationary Rankine-Hugoniot conditions in the shock rest frame.


Note ui = (ρi , vx,i , vy,i , pi , Bx,i ), where the index i refers to the value taken in the
i−th region (figure 3.3). Note that we did not include By,i , since it is completely
determined by equation (3.4).
We arbitrarily scale by setting p1 = 1 and ρ1 = γ. Now all velocity components
are scaled with respect to the sound speed in this region between the impinging
shock and the initial CD. Since this region is initially at rest, the sonic Mach
number M of the shock equals its shock speed s. When the shock intersects
the CD, the quintuple point follows the unshocked contact slip line. It does so
at a fixed speed vqp = (M, M tan α), therefore we will solve the problem in the
frame of the stationary quintuple point. We will look for self similar solutions
in this frame, u = u(φ), where all signals are stationary. We now have that
ṽx = vx − M and ṽy = vy − M tan α, where ṽ refers to this new frame. From
now on we will drop the tilde and only use this new frame. We now have u1 =
RIEMANN SOLVER BASED SOLUTION STRATEGY 77

q q
(γ, −M, −M tan α, 1, β2 )t and u7 = (ηγ, −M, −M tan α, 1, β2 )t . The Rankine-
Hugoniot relations now immediately give a unique planar solution for u2 , namely

r t
(γ 2 + γ)M 2 (γ − 1)M 2 + 2 2γM 2 − γ + 1

2
u2 = , − , −M tan α, , .
(γ − 1)M 2 + 2 (γ + 1)M γ+1 β

(3.17)

This then completely determines the initial condition, containing only states u1 , u2
and u7 in terms of the dimensionless parameters α, β, γ, M and η.

3.4.2 Path variables

We know that v, B and p are continuous across the CD (see Appendix D). Since
By is uniquely determined by the other 4 (independent) variables, we have 4
independent scalars which should vanish. Hence we can guess 4 path variables,
1 across each magneto-acoustic signal, and express the vanishing quantities in
function of those path variables χ ≡ (χF R , χSR , χST , χF T ). The fast magneto-
acoustic signals are postulated to be shocks, while the slow magneto-acoustic
signals can be both shocks and expansion fans.
In the case of a shock, fast signals are postulated to be fast shocks, while slow
signals can be slow or intermediate shocks. We then select the correct solution to
the Rankine-Hugoniot jump conditions as explained in Appendices D and E. In
the case of a rarefaction wave, we numerically integrate the MHD equations as
explained in Appendix G. In this manner we reduce the problem to solving

[[vx , vy , Bx , p]](u4 (u3 (u2 , χF R ), χSR ), u5 (u6 (u7 , χF T ), χST )) = 0. (3.18)

The function χ 7→ [[vx , vy , Bx , p]] is not partially differentiable in all points of its
domain. The price to pay is that we will postulate the wave configuration.
As said before, every magneto-acoustic signal is controlled by one path variable,
such that uu (uk , χsignal ), where signal is the signal separating uu from uk . The
path variables should have a one-to-one relationship with the refraction angles
78 SHOCK REFRACTION IN IDEAL MHD

φsignal . The actually used path variables are given by

χF R = tan(φF R ), (3.19)

χF T = tan(φF T ), (3.20)
8 “ “ ” ”
tan φSR −tan φa,R
> tanh −1
2 − 1 , ∀φSR ∈ [φcr,R , φa,R ]
“ “ tan φcr,R −tan φa,R ”
>
>
< ”
tan φSR −tan φsl,R
χSR = tanh −1
2 −1 , ∀φSR ∈ [φa,R , φsl,R [ (3.21)
> “ “ tan φa,R −tan φsl,R ” ”
> tan φSR −tan φst,R
tanh −1
2 tan φsl,R −tan φst,R − 1 , ∀φSR ∈ [φsl,R , φst,R [
>
:

8 “ “ ” ”
tan φST −tan φa,T
>
>
> tanh−1 2 tan φcr,T −tan φ a,T ”
− 1 , ∀φSR ∈ [φa,T , φcr,T ]
< “ “ ”
tan φST −tan φsl,T
χST = tanh −1
2 −1 , ∀φSR ∈ [φsl,T , φa,T [ (3.22)
> “ “ tan φa,T −tan φsl,T ” ”
> tan φST −tan φst,T
tanh −1
2 tan φsl,T −tan φst,T − 1 , ∀φSR ∈ [φst,T , φsl,T [
>
:

where we introduced the Alfvénic angles, the critical angles, the slow angles and
the stationary angles. The Alfvénic angles are those for which the upstream state
is exactly Alfvénic and these are defined by

B3,y − ρ3 v3,y
 
φa,R = arctan √ , (3.23)
B3,x − ρ3 v3,x

B6,y + ρ6 v6,y
 
φa,T = arctan √ , (3.24)
B6,x + ρ6 v6,x

and the critical angles at which the transition from 1 to 3 real solutions to the RH
conditions across the slow signal takes place, by

φcr,R ≡ max{φ < φa,R |(N 2 − D3 )(φ) = 0}, (3.25)

φcr,T ≡ min{φ > φa,T |(N 2 − D3 )(φ) = 0}. (3.26)

Finally, the slow angles φsl,R and φsl,T , are those for which the upstream state is
exactly slow (and are found numerically) and the stationary angles are defined by

vy,3
φst,R ≡ arctan , (3.27)
vx,3
vy,6
φst,T ≡ arctan . (3.28)
vx,6

The functions N and D are defined in Appendix F, and these critical angles are
dependent on the location of the fast signal.
RIEMANN SOLVER BASED SOLUTION STRATEGY 79

Figure 3.5: The wave postulation for the slow transmitted signal. Just as in
the HD case, the signal is postulated to be a shock, whenever there exists an
entropy increasing shock solution. This is the case whenever vn > vs,n . When the
known state is a 3-state, only a slow shock is possible, and the signal is trivially
postulated to be a slow shock. When the upstream state is super-Alfvénic, the
signal is postulated to be an intermediate shock.

Once the wave configuration is determined, χ determines the solution to the


vy,4
problem. A secant method iteration solves equation (3.18) and φSC = arctan vx,4
closes the procedure.
Once the wave configuration is postulated, we know if the slow signal angles are
bigger or smaller than the Alfvénic angles. The initial values of the path variables
controlling the fast signals are taken such that the upstream states u2 and u7
are exactly fast. For the slow signals, initially we put χSR = χST = 0. In the
case of an intermediate shock, this means that our starting angles φSR and φST
respectively satisfy 2φSR = φa,R + φcr,R and 2φST = φa,T + φcr,T . The procedure
also ensures that in every iteration step φa,T < φST < φcr,T and φcr,R < φSR <
φa,R . Similarly, in the case of a slow shock, the initial guesses φSR and φST
respectively satisfy 2φSR = φsl,R + φa,R and 2φST = φsl,T + φa,T . Finally, when
the slow signal is located at φSR > φsl,R there is no shock solution possible which
satisfies the entropy condition, and in the case the signal is a slow rarefaction fan.
The initial iteration angles φSR and φST respectively satisfy 2φSR = φst,R + φsl,R
and 2φST = φst,T +φsl,T . We hereby postulated that the slow signals are expansion
2 2
fans when vn,3/6 (φSR/ST ) < vs,n,3/6 . In this case, the expansion fan is located
between φSR/ST and φsl,R/T . This criterion is equivalent to the criterion that a
rarefaction fan will only occur at a given position if no shock solutions satisfying
the entropy condition are possible, and is in this sense a generalization of the
criterion exploited in Delmont et al.[19]. For the slow transmitted signal, this
wave postulation is schematically presented in figure 3.5.
80 SHOCK REFRACTION IN IDEAL MHD

3.4.3 Tackling the signals

The solution to the stationary Rankine-Hugoniot conditions is given in Appendix


E. One problem is that the uniqueness of its solution is not guaranteed. We express
the tangential component of the magnetic field in the downstream state, Bt,u , as
the root of a cubic, which coefficients are expressed in terms of the known upstream
state. The unknown state is then expressed in terms of Bt,u .
It can be shown that if there is a unique solution to the RH conditions, it is a
fast or a slow shock. If on the other hand there are 3 solutions, they are all of a
different shock type. Thus when the shock types of all magneto-acoustic signals
are known, the Rankine-Hugoniot conditions can be solved uniquely. We solve the
cubic analytically in Appendix F, noting that roots can be real or complex. In
principle this completes the solution algorithm for the RH conditions of a genuine
shock. However, the complications are that

• We need to select the appropriate root from the (up to) 3 mathematical
possibilities, at each magnetosonic signal;
• The analytical expressions contain a square root and a cube root in the
complex plane. These expressions are discontinuous in the negative real
numbers.

In Appendix F, we also provide additional technical information on how


permutation of the root indices can make the analytical expressions for the roots
(F.6-F.8) continuous and differentiable, and hence allows for a secant iteration
method applied on equation (3.18). The only remaining discontinuities are reached,
when the refraction angles cross the critical angles φcr,R or φcr,T , where the
transition from a unique solution to 3 solutions for the RH conditions takes place.
Also [[p]] is discontinuous across the Alfvénic angles. This leads to the restriction
that we cannot cross those critical angles in subsequent iteration steps. This is
taken care of by our choice of path variables.
In Appendix G finally, we describe which relations hold across expansion fans and
how the numerical integration through the fan is performed.

3.5 Demonstration of result

We study the case where (α, β, γ, η, M ) = π4 , 2, 75 , 3, 2 , a case which was studied




in detail before by Samtaney [78] and Wheatley et al. [98]. For this case we have
that
DEMONSTRATION OF RESULT 81

u1 = (1.4000, −2.0000, −2.0000, 1.0000, 1.0000), (3.29)

u2 = (3.7333, −0.7500, −2.0000, 4.5000, 1.0000), (3.30)

u7 = (4.2000, −2.0000, −2.0000, 1.0000, 1.0000). (3.31)

The initial guesses for the iteration procedure are χF R , χSR , χST , χF T =
(0.5314, 0, 0, 3.7306). The iteration converges to the exact solution for the angles
φF R , φSR , φST , φF T = (0.40569, 0.91702, 1.19426, 1.27673). The corresponding
cubics whose roots need to be properly selected are shown in figure 3.6.
Discussing the solution of the refraction pattern in the order of integration, we
encounter:

• the fast reflected (FR) signal, located at φF R = 0.40569. The cubic across the
interface is then evaluated and found to be 0.0346B̂t3 −0.5184B̂t2 +7.2999B̂t +
3.9011 = 0, which has, as predicted, only one real solution B̂t = −0.5149,
such that
u3 = (4.424, −0.834, −1.774, 5.710, 1.159) ,
and By,3 = 0.0685. Note that F R is a fast shock;
• the slow reflected (SR) signal, located at φSR = 0.91702. The cu-
bic relation connected to this interface has 3 real solutions: B̂t ∈
{0.13075 10−3 , 0.34361, −0.02656}. Further physical argumentation is
needed to select the correct root. The first possibility, B̂t = 0.13075 10−3 ,
leads to negative pressure. The second option, B̂t = 0.34361, corresponds to
an intermediate shock, and yields
u4,intermediate = (4.395, −1.290, 2.374, 5.656, 0.198) ,
 γ
together with By,4 = 1.186. The entropy condition pp34 < ρρ43 is satisfied.
The last possible solution, B̂t = −0.02656 corresponds to a slow shock. In
this case
u4,slow = (4.596, −1.048, −2.028, 6.025, 0.736),
 γ
and By = 0.484, which also satisfies the entropy condition pp43 < ρρ43 ;

• the fast transmitted (FT) signal, located at φF T = 1.27673. The cubic


relation across this fast interface again has a unique real solution B̂t =
−1.10819. This solution corresponds to a fast shock and leads to
u6 = (11.932, −1.213, −2.384, 5.275, 1.237),
and By,6 = 0.783;
82 SHOCK REFRACTION IN IDEAL MHD

80
12
0.006
0.001

10
0.004
60

8
0.002

F(B)
6
K
0.2 K 0.1 0 0.1 0.2 0.3 0.4
K K K
F(B) 40

B 0.3 0.2 0.1 0 0.1 0.2 0.3 0.4 0.5 0.6


4
K
0.002
B

20
2
K 0.001
K
0.004

K
1.0 K
0.5 0 0.5 1.0
F(B)
F(B) K
0.006

K B
K K 0

K
2 2 1 1 2
0.008 B
K K
K
0.002

K
4
0.010 20

K6

Figure 3.6: The cubic for respectively FR, SR, ST, FT.

• the slow transmitted (ST) signal, located at φST = 1.19426. The cubic
relation controlling this interface has 3 different real solutions B̂t ∈
{0.44121, −0.70666 10−3 , 0.07837}. The first possibility, B̂t = 0.44121 can
be eliminated since it does not lead to a solution which satisfies the entropy
condition. The second possibility, B̂t = −0.70666 10−3 would lead to
negative pressure. The remaining possibility, B̂t = 0.07837 corresponds to
an intermediate (2 → 4)-shock from u6 to u5 , and yields

u5 = (13.092, −1.048, −2.028, 6.025, 0.736),

with By,5 = −0.484, which satisfies the entropy condition.

We now notice that the correct reflected slow signal is the slow shock and uSR =
u4,slow , since the shocked contact must satisfy the matching conditions (D.7).

3.6 Suppression of the Richtmyer-Meshkov Instability

Since vn vanishes along a contact discontinuity, the Rankine-Hugoniot jump


conditions simplify as
[[vt , p, Bt ]] = 0.
Therefore, they do not allow for vorticity deposition on the contact, and the
interface must remain stable. Figure 3.7 shows snapshots  of a numerical simulation
of the parameter regime (α, β, γ, M, η) = π4 , 2, 75 , 2, 3 performed by the Adaptive
Mesh Refinement code AMRVAC. The first frame shows the density, ρ, during the
shock refraction. Note that the CD remains stable, since the jump in tangential
velocity, [[vt ]], vanishes. The second frame shows the tangential magnetic field
component, Bt . The density is discontinuous in every new-formed signal, while
the CD is not visible in the Bt frame. SR is a slow shock, therefore Bt does not
change sign across this signal, while ST on the other hand is an intermediate shock
and Bt changes sign across it.
WAVE CONFIGURATION TRANSITIONS 83

The lower left panel of figure 3.7 shows the  streamlines


 in the stationary shock
frame, over a Schlieren plot of density (exp 200−|∇ρ|
250 is shown). The streamlines
are coloured by the density profile. One sees that almost all the vorticity is
deposited on FT. The upper right panel of figure 3.8 Shows the magnetic field
lines of this reference case. The lines are coloured by the magnetic field strength.
One can see that Bt only changes sign at ST, which is the only intermediate signal
in the solution. Also one sees that the magnetic field does not jump across the
CD: [[B]]CD = 0.

3.7 Wave configuration transitions

We now wonder under which values of the plasma-β the same wave configuration
will be possible. We therefore ran our solver for varying values of the plasma-
β, postulating that SR is a slow shock, and ST is an intermediate shock. The
solver converges for β ∈ [1.0; 2.3]. For β ≈ 1.0, ST will be a switch-off shock, i.e.
φST = φa,T , and when we decrease β even more, the slow transmitted signal will
be located even further away from the CD, i.e. φST > φa,T , such that ST now
becomes a slow shock. On the other hand, when we decrease the magnetic field
till β ≈ 2.3, SR will be located at φSR = φa,R , such that SR is a switch-off shock.
Decreasing the magnetic field even more such that β > 2.3 will result in a solution
where both SR and ST are intermediate shocks.
We compare the reference case ((α, β, γ, η, M ) = π4 , 2, 75 , 3, 2 , see figure 3.7) to a


case where β = 0.5, and a case where β = 4.


The magnetic field lines of the case where β = 0.5 are plotted in the left frame of
figure 3.8. Here both the reflected and transmitted signals are slow shocks, as Bt
doesn’t change sign across them. Finally, the case where β = 4 is shown in the
right panel. As predicted, not only ST, but also SR is an intermediate shock now,
as Bt changes sign across both the slow signals. In both
 cases the
 magnetic field
lines are plotted over a Schlieren plot of density (exp 200−|∇ρ|
250 is shown), and
the field lines are coloured by |B|.

3.8 Conclusion

We developed an exact Riemann solver based solution for regular shock refraction
in planar ideal MHD. In this case, five signals arise: a CD separates two reflected
signals from two transmitted signals. We postulate both the fast signals to
be shocks, while the slow signals can be slow shocks, intermediate shocks or
rarefaction fans. In the presented case, the transmitted slow signal turns out to
84 SHOCK REFRACTION IN IDEAL MHD

Figure 3.7: Up: Density plot for (α, β, γ, η, M ) = π4 , 2, 57 , 3, 2 . Five signals arise


and their location is found by the iteration procedure. Iterating one more time
yields u(x) and solves the problem. Note that the CD remains stable. Middle: A
plot of the transverse magnetic field, Bt . Note that the transverse magnetic field
is continuous across the CD: [[Bt ]] = 0. Also note that Bt changes sign across ST.
Indeed, ST is an intermediate shock. Lower left: Streamlines in the stationary
shock frame, over a Schlieren plot of density. Lower Right: The magnetic field
lines show ST is an intermediate shock, while SR is a slow shock.
CONCLUSION 85

Figure 3.8: Left: Magnetic field lines fot the case (α, β, γ, η, M ) = π4 , 0.5, 75 , 3, 2 .


Both the slow signals are slow shocks. Right: Magnetic field lines for the case
(α, β, γ, η, M ) = π4 , 4, 57 , 3, 2 . Both the slow signals are intermediate shocks.


be an intermediate shock, which agrees with our numerical AMRVAC simulation.


In the hydrodynamical case, the Richtmyer-Meshkov instability will occur after
the shock reflects from the top wall. Since the MHD equations do not allow for
vorticity deposition on a contact discontinuity, this instability is suppressed in the
presence of magnetic fields (with a non-vanishing normal component).
Chapter 4

Parameter ranges for


intermediate shocks

We investigate under which parameter regimes the MHD Rankine-Hugoniot


conditions, which describe discontinuous solutions to the MHD equations, allow
for slow, intermediate and fast shocks. We derive limiting values for the upstream
and downstream shock parameters for which shocks of a given shock type can
occur. We revisit this classical topic in nonlinear MHD dynamics, augmenting the
recent time reversal duality finding by Goedbloed [33] in the usual shock frame
parametrization.
This work is published in Delmont & Keppens [21].

4.1 Introduction

4.1.1 Intermediate shocks in MHD

The dynamic behavior of plasmas is described by the magnetohydrodynamic


(MHD) equations, where a central role is played by the Alfvén speed. Discontinu-
ous solutions only satisfy the integral form of the MHD equations, i.e. the Rankine-
Hugoniot conditions (RH). These RH conditions have been studied extensively in
established as well as more recent literature (see e.g. Germain [30]; Anderson [5];
Jeffrey & Taniuti [46]; Liberman & Velikhovich [56]; Sturtevant [80]; Gombosi [35];
Goedbloed [33]) and just express the basic nonlinear conservation laws across a
discontinuity. Many authors have since paid attention to MHD shock stability as
well (amongst others Akhiezer et al.[4]). The interaction of small perturbations

87
88 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

with MHD (switch-on and switch-off) shocks was studied both analytically by
Todd [84] and numerically by Chu & Taussig [17]. Later on, the evolutionarity
of intermediate shocks, which cross the Alfvén speed, has been addressed. A
discontinuity is said to be evolutionary when small perturbations imposed on it
lead to an evolution that remains close to the initial discontinuity. According to
classical stability analysis, intermediate shocks are not evolutionary so one should
not obtain such shocks in physically realizable situations. On the other hand,
Wu [99], De Sterck et al.[22] and others found intermediate shocks in one and
two dimensional numerical simulations respectively. De Sterck & Poedts [23] were
the first to find intermediate shocks in three dimensional numerical simulations.
The evolutionary condition became controversial and Myong & Roe [60, 61],
amongst others, argue that the evolutionary condition is not relevant in dissipative
MHD. Wu [100, 101] also argued that intermediate shocks are admissible. Shock
observations in the heliosphere can be found in favor of their existence as well:
Chao et al.[12] reported a 2 → 4 intermediate shock observed by Voyager 1 in
1980 and Feng & Wang [29] recognized a 2 → 3 intermediate shock, which was
observed by Voyager 2 in 1979.
Intermediate shocks are not the only way to connect a sub-alfvénic state to a
super-alfvénic state. One can also encounter compound waves in ideal MHD.
These compound waves can consist of a slow shock which travels with its maximal
propagation speed and a rarefaction fan directly attached to it. Brio & Wu
[11] detected those compound waves in numerical simulations which have become
classical test problems for numerical codes. Another type of compound signal
consists of a slow shock layer, immediately followed by a rotational discontinuity
(Wheatley et al.[97]). The importance of compound waves is that they can be an
alternative to intermediate shocks to cross the Alfvén speed. Barmin et al.[8]
argue that if the full set of MHD equations is used to solve planar MHD, a
small tangential perturbation on the magnetic field vector splits the rotational
jump from the non-evolutionary compound wave. This transforms the latter
one into a slow shock, such that the compound wave is non-evolutionary. They
investigate the reconstruction process of the non-evolutionary compound wave into
evolutionary shocks. Falle & Komissarov [26, 27] also reject intermediate shocks
on evolutionary grounds, and suggest a shock capturing scheme based on Glimm’s
method (Glimm [31]) for numerically obtaining purely evolutionary solutions in
MHD. Our goal in this chapter is to determine in which regions of parameter space
one might encounter intermediate shocks, which play such a prominent role in all
evolutionarity argumentation.

Recently, Goedbloed [33] classified the MHD shocks by rewriting the RH equations
in the de Hoffmann-Teller frame (De Hoffmann & Teller [39]) introducing the
existence of a distinct time reversal duality between entropy-allowed and entropy-
forbidden solutions. This work encouraged us to revisit the classical RH conditions
INTRODUCTION 89

and augment these results in terms of the commonly exploited shock parameters
in the shock frame. Therefore, another goal in this chapter is to determine in
which regions of parameter space slow, intermediate and fast shocks can appear
and how this relates to Goedbloed’s analysis, in particular regarding the duality
between solutions.

4.1.2 The Rankine-Hugoniot jump conditions

The ideal MHD equations are a system of partial differential equations. When
allowing for large amplitude waves, which in the limit case become discontinuous,
these equations are replaced by their (weak) discontinuous form: the RH relations.
These RH relations express jump conditions across the discontinuity. In any frame
where the shock is stationary, the MHD RH conditions become
 
ρvn
B2

 ρvn2 + p + 2t 

 ρvn vt − Bn Bt 
 = 0, (4.1)
 
2 2
 γ v +v 2
 vn (
 γ−1 p + ρ n
2
t
+ Bn ) − Bn Bt vt 

 vn Bt − vt Bn 
Bn

where [[G]] expresses the jump G2 − G1 across the shock. In this chapter, index
2 refers to the downstream state, and index 1 refers to the upstream state, while
index n refers to the direction of the shock normal, and index t refers to the
tangential vector components in the plane spanned up by B1 and B2 . Further, ρ is
the mass density, v is the velocity, p the thermal pressure and B the magnetic field.
The ratio of specific heats, γ, is considered a constant parameter, as we will assume
an ideal gas equation of state. For a derivation of these well-known expressions, we
refer to De Hoffmann & Teller [39]; Jeffrey & Taniuti [46]; Liberman & Velikhovich
[56], Goedbloed & Poedts [32]. The eight governing MHD equations have been
reduced to six jump conditions in equation (D.1). Three equations have been
dropped from the fact that tangential magnetic field components are forced to
lie in the same plane perpendicular to the shock front itself: the conservation
of momentum reduced to two equations and Faraday’s law reduces to a single
equation. On the other hand, in the stationary case, the ∇ · B = 0 constraint
becomes a full equation and is added to equation (D.1).
These six jump conditions can be further restricted by noting that Bn is invariant
across the shock. Since the RH conditions are translation invariant, the exact
values of vt,1 and vt,2 can be eliminated for [[vt ]], which is completely determined
by [[Bt ]]. Therefore only four quantities really matter. Hence we now suppose
that one state, uk = (ρ, vn , p, Bt )k is known, where from now on we consistently
drop the subscript k from known state quantities. We want to express the other
90 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

unknown state, uu , in function of this known state. In order to do so, we introduce


three dimensionless parameters connected to the known state. First, the plasma-β,
which expresses the ratio of thermal and magnetic pressure. Then, the tangent θ
of the angle between the shock normal and the magnetic field in the known state.
Finally the Alfvén Mach number M , which is the ratio of the normal velocity
component in the known state and the normal Alfvén speed in that region, thus

2p
β = , (4.2)
Bn2 + Bt2

Bt
θ ≡ , (4.3)
Bn
|vn |
M ≡ . (4.4)
va,n
q 2
Bn
Here, the normal Alfvén speed va,n ≡ is defined as earlier (in section 3.2).
ρ
q 2
Bt
Analogously, we define the tangential Alfvén speed va,t ≡ ρ and the total
q
Alfvén speed va ≡ va,n 2 + v 2 as in section 3.2. The full set of characteristic
a,t
speeds of the full set of MHD equations is given by {vn ±vf,n , vn ±va,n , vn ±vs,n , vn },
where the fast speed vf,n and the slow speed vs,n are respectively defined as in
equations (1.48) and (1.48). On the other hand, the restricted set of planar MHD
equations, describing MHD dynamics with all vector quantities restricted to lie
in the same plane, has only five characteristic speeds {vn ± vf,n , vn ± vs,n , vn }.
Ignoring cold MHD, where the thermal pressure p vanishes, and assuming that
the normal component of the magnetic field Bn (which can be seen as an equation
parameter now) does not vanish, we conclude that the planar system is strictly
hyperbolic, since vf,n = vs,n would imply that both va and c would vanish.
Obviously c cannot vanish, and for the same reason also vs,n 6= 0.

4.1.3 MHD shock types: classical 1 − 2 − 3 − 4 classification

Since the RH conditions do not necessarily lead to a unique solution, it is common


to introduce another discrete characterization: the shock type i → j. It is well-
known (see e.g. Liberman & Velikhovich [56]) that once the shock type and one
state is given, if a solution exists, it must be unique. Since 0 ≤ vs,n ≤ va,n ≤ vf,n ,
the full set of MHD equations is hyperbolic, but non-strictly hyperbolic. Those
characteristic speeds categorize the up- and downstream states into the classical
1 − 2 − 3 − 4-classification as explained in section 3.3.
We call a shock admissible if it satisfies the second law of thermodynamics: [[S]] >
0 and admissible versus inadmissible shocks can be related through the time duality
SOLUTION TO THE RANKINE-HUGONIOT CONDITIONS 91

principle from Goedbloed [33]. When the upstream state is of type i and the
downstream state is of type j, then the shock type is i → j. Furthermore, in
terms of these shock types, the admissibility condition translates as i < j. The
RH conditions (D.1) together with the admissibility condition (i < j) now lead to
the MHD shock classification presented in section 3.3. All these different shock
types are shown in figure 3.4.
The classification given in section 3.3 is well-known and features in many classical
textbooks, such as Liberman & Velikhovich [56]. Recently, a contribution to this
classical theory was described by Goedbloed [33], where the RH conditions for
MHD shocks were rewritten in an insightful, symmetric form by changing to the
de Hoffmann-Teller frame, where [[vt ]] = [[Bt ]] (see De Hoffmann & Teller [39]). In
that work, the central role played by intermediate shocks was emphasized, and time
reversal duality arguments were introduced, linking admissible and inadmissible
shocks. Here we augment this recent study by paying attention to how the MHD
shocks appear in the shock reference frame, and analyzing under which conditions
intermediate shocks occur.

4.2 Solution to the Rankine-Hugoniot conditions

We scale densities to the known density ρ and magnetic fields to Bn (Bn is constant
and equal for known and unknown state, as seen in equation (D.1)). Since no
distances or times are involved, the problem is now uniquely scaled. Note that
velocities are scaled to the known normal Alfvén speed va,n . The known state is
now given by

β(1 + θ2 )
 
u ≡ (ρ, vn , p, Bt ) = 1, σM, ,θ , (4.5)
2

where σ ≡ −1 when the known state is upstream and σ ≡ 1 when the known state
is downstream, and σ just gives the proper sign to the known normal velocity
component.
92 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

Under the assumption that M 6= 1 and θ 6= 0, solving the RH equations leads to


the unknown state quantities

M 2ψ
ρu = , (4.6)
θ(M 2 − 1) + ψ

θ(M 2 − 1) + ψ
vn,u = σ , (4.7)

ψ 2 + θψ + 2 γ − 1 β(1 + θ2 ) + (ψ − θ)2
 
2
pu = + 2
M2 − , (4.8)
γ+1 M −1 γ+1 2

Bt,u = ψ, (4.9)

where ψ satisfies the cubic equation

C(ψ) ≡ ψ 3 + τ2 ψ 2 + τ1 ψ + τ0 = 0, (4.10)

and its coefficients are given by

−θ (γ − 1)(M 2 − 1) − M 2 ,

τ2 = (4.11)

(M 2 − 1) (γ − 1)(M 2 − 1) + γ(β(θ2 + 1) + θ2 ) − 2 ,

τ1 = (4.12)

τ0 = −(γ + 1)θ(M 2 − 1)2 . (4.13)

The tangential velocity is then found from


θ−ψ
[[vt ]] = . (4.14)
M

We can now define the dimensionless parameters referring to the unknown state
in function of those in the known state. In our first view, we will keep ψ in the
expressions, noting that the cubic equation (4.10) can be seen as a continuously
partially differentiable function of the known dimensionless parameters and ψ, thus
C(M, θ, β, ψ), such that the implicit function theorem ensures that locally ψ can be
seen as a ψ(M, θ, β). Moreover ψ(M, θ, β) is continuously partially differentiable
whenever ∂C∂ψ (M, θ, β, ψ(M, θ, β)) does not vanish. This latter restriction means
that ψ cannot be a double root of the cubic. As we will show later, this condition is
equivalent to the condition that none of the characteristic speeds in the unknown
region vanishes.
PHYSICAL MEANING OF Ω 93

After straightforward algebra, the expressions for (Mu , θu , βu ) now become


s
(M 2 − 1)θ + ψ
Mu = , (4.15)
ψ

θu = ψ, (4.16)

(γ − 1)((θ − ψ)2 + (1 + θ2 )β) − 4M (M 2 − 1) + 2M 2 (ψθ + ψ 2 )


2
` ´
βu = (4.17)
.
(M 2 − 1) (γ + 1) (1 + ψ 2 )

In terms of the dimensionless parameters, these equations lead to the following


well-known (see e.g. Goedbloed [33]) invariants across a shock:

[[(M 2 − 1)θ]] = 0, (4.18)

2M 2 + β(1 + θ2 ) + θ2
 
= 0, (4.19)
 
γ 2 2 2
( β + M )(1 + θ )M = 0. (4.20)
γ −1

When we define the new quantity Ω, computed from the coefficients of the
governing cubic equation (4.10), as

Ω ≡ 27τ02 + 4τ13 + 4τ22 τ0 − τ22 τ12 − 18τ2 τ1 τ0 , (4.21)

it is well-known that the cubic C(ψ) has three real solutions when
Ω < 0. (4.22)
Note that this criterion is equivalent to equation (F.3).
Mathematically speaking, the RH conditions are thus governed by the existence
and multiplicity of the real roots of a cubic, hence Ω = 0 will play a crucial
role, as will be explained in section 3. It should be noted, however, that most
analytical expressions in this choice of frame and parametrization are complicated
expressions, e.g. one can note that Ω is a polynomial of degree 6 in M 2 , where
(M 2 − 1) appears as a double factor. The same governing equations, as expressed
in the de Hoffmann-Teller frame exploited by Goedbloed [33] give a much more
compact algebraic form along with an easier classification.

4.3 Physical meaning of Ω

Note that Ω = 0 is exactly the condition for having a real solution with multiplicity
two to C(ψ) = 0. Therefore, whenever Ω 6= 0, (Mu , θu , βu ) is a smooth function of
94 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

1
Figure 4.1: Left: The (θ, M ) state plane for β = 10 and γ = 35 . Shown are the
curves fast: vn = vf,n , Alfvén: vn = va,n and slow: vn = vs,n . These curves
separate the (θ, M ) plane in the classical 1 − 2 − 3 − 4 state regions. Right: Shown
is the curve Ω = 0. Note that states with M = 1 appear as a double zero of Ω.
Where Ω > 0, only one real solution to the RH conditions exist. On the other
hand, where Ω < 0, there exist three real solutions to the RH conditions. On
Ω = 0, the RH solutions allow for 2 distinct real solutions.

(M, θ, β). A related observation is that equation (4.22) is the analytical condition
for the existence of an intermediate shock, since when only one solution exists to
the RH jump conditions it must be a fast or a slow shock (see e.g. Liberman &
Velikhovich [56]).
We will argue that all the states, which can be connected to a state which is
exactly fast, slow or Alfvénic, satisfy Ω = 0. In other words: the surface
ω ≡ {(M, θ, β) |Ω(M, θ, β) = 0} in our 3D parameter space will be shown to
correspond to all states which can be connected by the RH conditions to known
states satisfying (vn − va,n )(vn2 − vf,n
2
)(vn2 − vs,n
2
) = 0. One can visualize these
concepts by drawing the corresponding curves in the (θ, M )-plane, for fixed β and
γ. This is done in figure (4.1), where the plot at left concentrates on the 1−2−3−4
state regions, and the panel at right shows Ω = 0 curves.
First, we make some general observations. 1-states can be connected to a 2-state, a
3-state and a 4-state. Therefore, they can only appear as an upstream state. There
are at most three different real solutions with a 1-state as an known upstream
state. When there is only one solution, this solution is a fast 1 → 2 shock. In the
transition case where two different real solutions exist, these solutions include one
fast (1 → 2) shock, and two coinciding intermediate 1 → 3 = 4 shocks. Finally,
when there are three different real solutions, one of these is a fast (1 → 2) shock and
PHYSICAL MEANING OF Ω 95

the 2 other solutions are respectively an intermediate 1 → 3 and an intermediate


1 → 4 shock.
Similarly, 2-states can be connected to a 1-state, a 3-state and a 4-state. Therefore,
the admissibility condition which demands that i < j for an i → j shock, ensures
that a 2-state can, at most, occur in one downstream state solution and two
upstream state solutions to the RH conditions. Again, when only one real solution
exists, it is a fast (1 → 2) shock. In the transition case where two different
real solutions exist, these solutions include one fast shock and two coinciding
intermediate 2 → 3 = 4 shocks. Finally, when there are three different real
solutions, one of these is a fast shock, and the 2 other solutions are respectively
an intermediate 2 → 3 and an intermediate 2 → 4 shock.
The same reasoning states that generally speaking, 3-states can be connected to
a 1-state, a 2-state and a 4-state. The admissibility condition ensures that they
can, at most, occur in two downstream state solutions and one upstream state
solution to the RH conditions. Now, when only one real solution exists, it is a
slow (3 → 4) shock. In the transition case where two different real solutions exist,
these solutions include one slow (3 → 4) shock and two coinciding intermediate
1 = 2 → 3 shocks. Finally, when there are three different real solutions, one of
these is a slow shock, and the 2 other solutions are respectively an intermediate
1 → 3 and an intermediate 2 → 3 shock.
Generally speaking, 4-states can be connected to a 1-state, a 2-state and a 3-state.
Therefore, the admissibility condition ensures that they can, at most, occur in
three downstream state solutions to the RH conditions. Again, when only one
real solution exists, it is a slow (3 → 4) shock. In the transition case where two
different real solutions exist, these solutions include one slow (3 → 4) shock and
two coinciding intermediate 1 = 2 → 4 shocks. Finally, when there are three
different real solutions, one of these is a slow (3 → 4) shock and the 2 other
solutions are respectively an intermediate 1 → 4 and an intermediate 2 → 4 shock.
These general observations now allow us to consider the transition cases. A 1 = 2-
state, where vn = vf,n or a 3 = 4-state, where vn = vs,n , will always have the
trivial solution. Therefore they can only appear in none or two real non-trivial
intermediate solutions, depending on the sign of Ω. Both these solutions, labeled
as unknown state a and b, have uk as a double solution, thereby satisfying Ωu,a =
Ωu,b = 0, and as we will show in section 4.2 mutually satisfy the RH conditions.
This is argued in what follows.
Consider an upstream state which satisfies vn = vf,n . Straightforward algebra
shows that C(θ) = 0, so one of the solutions to the RH jump conditions is the
trivial one. By solving C(ψ)
ψ−θ = 0, one finds 2 solutions. Those solutions are real
only when Ω ≤ 0. In this case they lead to a 1 = 2 → 3 and a 1 = 2 → 4 solution,
which we respectively call uu,a and uu,b . These uu,a and uu,b mutually satisfy the
96 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

RH conditions and can therefore be separated by a slow 3 → 4 shock. Both these


solutions have our specific known upstream state uk , where vn = vf,n , as a double
solution, and therefore satisfy Ωu,a = Ωu,b = 0.
Completely analogously, consider now a downstream state with vn = vs,n .
Straightforward algebra shows that C(θ) = 0, so one of the solutions to the RH
jump conditions is the trivial one. By solving C(ψ)ψ−θ = 0, one finds 2 solutions.
Those solutions are real only when Ω ≤ 0. In this case they lead to a 1 → 3 = 4 and
a 2 → 3 = 4 solution, which we respectively call uu,a and uu,b . We again find that
uu,a and uu,b mutually satisfy the RH conditions and can therefore be separated
by a fast shock. Both these solutions have our specific known downstream state
uk , where vn = vs,n , as a double solution, and therefore satisfy Ωu,a = Ωu,b = 0.
Finally, consider a state where vn = va,n . Now C(ψ) simplifies as ψ 3 + θi ψ 2 =
0, where i = 1, 2 selects respectively the up- or the downstream solution. The
solutions to the RH conditions are now a switch-on shock, a switch-off shock and
a rotational wave. The switch-on and the switch-off solution also mutually satisfy
the RH jump condition. This can be easily checked by straightforward algebra.
Note that in this case Ωu = 0 automatically.

4.4 Results

Since the governing expressions in the shock frame are complicated, all calculations
reported here are performed by a symbolic computational software package,
namely MAPLE 11.0. We search for critical values for the parameters in both
up- and downstream states, for which different shock types can occur. These
critical values are found by simple geometrical arguments using the knowledge
obtained so far: namely that in the (θ, M )-plane, 2 families of curves as shown
in figure 4.1 split up the state plane in various regions. Unless explicitly stated
otherwise, we assume from now on that γ = 53 .

4.4.1 (θ, M)-diagrams at fixed β

For varying β, we plot both the curves where the known upstream normal velocity
component vn is exactly fast, Alfvénic or slow (i.e. vn = vf,n , vn = va,n , vn =
vs,n ) and the curves ω : Ω = 0 in a (θ, M )-diagram in figure 4.2. Those curves
divide the (θ, M )-parameter space in regions, where certain types of shocks are
mathematically possible. An important transition occurs at β = γ2 . Here the
sound speed equals the Alfvén speed and the thermal pressure equals the magnetic
pressure. When β < γ2 , we call the plasma magnetically dominated. In this case,
the curves vn = vf,n and vn = va,n have a common point at (θ, M ) = (0, 1). When
RESULTS 97

2
the plasma is thermally dominated, β > γ, the curves vn = vs,n and vn = va,n
have a common point at (θ, M ) = (0, 1).
Every region is then coded with a latin number code for (θ, M ), indicative of an
upstream state, i.e. (θ1 , M1 ). This means:

• (I) One fast shock and two intermediate shocks of type 1 → 3 and 1 → 4 are
possible;
• (II) Only a fast shock is possible;
• (III) Two intermediate shocks of type 2 → 3 and 2 → 4 are possible;
• (IV) No shocks are possible;
• (V) Only a slow shock is possible.

We can interpret the graphs in terms of the downstream state in a similar manner,
where (θ, M ) = (θ2 , M2 ). Every region is also coded with a letter code, meaning
the following:

• (A) One slow shock and two intermediate shocks of type 2 → 4 and 3 → 4
are possible;
• (B) Only a slow shock is possible;
• (C) Two intermediate shocks of type 1 → 4 and 2 → 4 are possible;
• (D) No shocks are possible;
• (E) Only a fast shock is possible.

These regions are shown in figure 4.2, where the axes labels must be interpreted
as (θ2 , M2 ) instead of (θ1 , M1 ) from above.
The 1 − 2 − 3 − 4 classification divides the (β, M, θ)-space into four regions. The
surface ω : Ω = 0 divides each of those regions into two: one region where Ω > 0
and one where Ω < 0, such that the (β, M, θ) space is divided into eight regions
by these curves.
The time reversal duality principle from Goedbloed [33] is hereby made visible
in the fact that every region corresponds to 1 or 3 mathematical solutions, but
the coding tells whether the state can appear as an upstream or as a downstream
region.
98 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

(a) (b)

(c) (d)
RESULTS 99

(a) (b)

Figure 4.2: Parameter space divided into various regions. Panels differ in their
β-value: a) β = 0.1, b) β = 1.0, c) β = 1.2 = γ2 , d) β = 1.3, e) β = 1.3637, f)
β = 1.5. The latin cypher-letter code is explained in the text, and related to the
time reversal duality argument put forth by Goedbloed [33].

4.4.2 Equivalence classes introduced by the RH conditions

The RH conditions are equivalent to equations (4.18-4.20), which express the


existence of three shock invariants. Therefore two states can be connected
through the stationary RH conditions if and only if they have the same value
for the expression ζ1 ≡ (M 2 − 1)θ, ζ2 ≡ 2M 2 + β(1 + θ2 ) + θ2 and ζ3 ≡
γ
( γ−1 β + M 2 )(1 + θ2 )M 2 . Denoting the relation "state A can be connected to
state B through the stationary RH conditions" as A 7→RH B, this relation 7→RH is
an equivalence. Indeed:

• 7→RH is reflexive: A 7→RH A. Every state can be connected to itself through


the stationary RH conditions.
• 7→RH is symmetric: A 7→RH B ⇒ B 7→RH A. If state A can be connected
to state B through the stationary RH conditions, then also state B can
be connected to state A by these conditions. Of course only one of these
connections satisfies the entropy condition.
• 7→RH is transitive: A 7→RH B ∧ B 7→RH C ⇒ A 7→RH C. Indeed: if A 7→RH B,
then ζi (A) = ζi (B), and if B 7→RH C, then ζi (B) = ζi (C). Hence A 7→RH
B ∧ B 7→RH C implies ζi (A) = ζi (B) = ζi (C), which means that A 7→RH C.
100 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

We conclude that (ζ1 , ζ2 , ζ3 ) defines equivalence classes on the parameter space.


All these equivalence classes contain one, two, three or four states.
A state A is in an equivalence class with one element in the following cases:

• If M = 1 and θ = 0. In this degenerate case the switch-on solution, the


switch off-solution and the rotational solution all coincide.
• If Ω > 0, and (vn2 − vs,n
2
)(vn2 − vf,n
2
) = 0. In this case, the only real solution
to the RH conditions is A itself.

A state A is in an equivalence class with two elements, in the following cases.

• If Ω > 0 and (vn2 − vs,n


2
)(vn2 − vf,n
2
) 6= 0. In this case the equivalence class
contains A itself, and the state introduced by the unique (non-trivial) real
solution to the cubic equation (4.10).
• If Ω = 0 and (vn2 − vs,n
2
)(vn2 − vf,n
2
) = 0. Since Ω = 0, the solutions to the
RH conditions crossing the Alfvén speed are coinciding. Since vn is exactly
fast or exactly slow, it appears itself as a solution to the cubic.

A state A is in an equivalence class with three elements in the following cases.

2
• if Ω = 0, but (vn2 − vs,n 2
)(vn2 − vf,n ) 6= 0. As explained earlier, in this case the
RH conditions have an exactly slow or fast state as a double solution, and a
single solution which also satisfies Ω = 0.

• if Ω < 0 and vn = vf,n or vn = vs,n . As explained earlier, in this case the RH


conditions have 2 different solutions with Ω = 0, and the considered state
itself as a third solution.

A state is in an equivalence class with four elements in the following case:

• if Ω < 0 and vn 6= vf,n and vn 6= vs,n . As explained above, in this case the RH
conditions have three different (non-trivial) solutions. Also, the considered
state itself is in the equivalence class.
• if vn = va,n . The RH conditions now have a switch-on shock, a switch-
off shock and a rotational wave as a solution. The considered state itself
completes the equivalence class.

Hence, when A is a superfast state, with Ω > 0, it is in an equivalence class with


exactly one other state. This state should be subfast, since we know that if only
RESULTS 101

one solution exists, it does not cross the Alfvén speed. Also, the solution state
has only one solution, therefore it also satisfies Ω > 0. Interpreting this result
on figure 4.2, we conclude that the II − C-region can only be connected to the
IV − E-region and the other way around.
Completely analogously we find that a subslow state with Ω > 0 can only be
connected with a superslow state, also satisfying Ω > 0. Interpreting this again
on figure 4.2, we conclude that the IV − B-region can only be connected to the
V −D-region, and vice versa. Further, we conclude that for any state in the regions
I − D, III − E, V − C and IV − A, there exist states in the regions I − D, III − E,
V − C and IV − A respectively, which can be mutually connected through the
RH jump condition. The initial eight regions of parameter space have thus been
divided into two groups of two mutually connectable regions and one group of four
mutually connectable regions.
If we do not take into account that βu should be positive, these connections are
mappings. We know the entropy condition in an i → j shock reduces to j > i.
Thus, when we consider one equivalence class we have four or two states, such that
the entropy increases with state type. Therefore, we know that when one of the
mathematical solutions has negative pressure, it must be the 1-states. When two
of those states have negative pressure, it must be the 1-state and the 2-state, and
so on. In the next section, we will additionally consider the physical restriction
that all pressures should be positive.

4.4.3 Positive pressure requirement

Until now we have only used the nonlinear relations expressed by RH, as well
as the admissibility condition, to count the number of solutions in the (θ, M, β)
state space. The admissibility restriction is that [[S]] > 0, which is satisfied for an
i → j-shock if and only if j > i. Further restrictions are that the solution should
have positive thermal pressure, pu > 0 and density ρu > 0. The latter restriction
is trivially satisfied. In fact, we can only encounter problems when the known
state is a downstream state, because when the upstream pressure is positive the
upstream entropy is also positive. Hence, the downstream entropy is positive and
the downstream pressure too. Therefore, we can expect that for a given 1-state,
the positive pressure requirement is trivially satisfied. When the known state is
a 2-state, we expect to encounter one critical surface dividing the 2-state regions
into sub-regions where the fast (1 → 2) shock solution has positive versus negative
thermal pressure. When the given state is a 3-state, we expect to encounter two of
such critical surfaces. And when the given state is a 4-state we expect to encounter
three of such critical surfaces. Therefore, the parameter space is now divided in
16 regions, as summarized in table (4.1).
102 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

Ω<0 Ω>0
1-state 1 1
2-state 2 2
3-state 3 1
4-state 4 2

Table 4.1: The positive pressure requirement divides the eight regions in parameter
space in even more regions. The number of sub-regions in which the original regions
are divided are shown above. Since the region is only important when the given
state is downstream, it is not surprising that the total number of regions is doubled
from 8 to 16.

Figure 4.3 shows the regions in which pu > 0 for β = 0.1 and β = 2. The figure
now plots: (i) the three curves defining vn = vs,n , vn = va,n and vn = vf,n ; (ii) the
curves where Ω = 0; and (iii) the lines defining pu = 0. Finding these regions is
straightforward: pick the correct root of the cubic, fill it out in the expression for
pu and make it vanish. The governing expressions can even be found analytically.
As mentioned above, the addition of these pu = 0 curves now divides the parameter
space into 16 regions, namely:

• (i) This state can occur as the upstream state of a fast shock;
• (ii) This state can occur as the downstream state of a fast shock;
• (iii) This state cannot occur, since the only solution has negative pressure;
• (iv) This state can occur as the upstream state of a fast shock, and as the
upstream state of an intermediate 1 → 3 and 1 → 4 shock;
• (v) This state can occur as the upstream state of an intermediate 2 → 3 and
2 → 4 shock, and as the downstream state of a fast shock;
• (vi) This state can occur as the upstream state of an intermediate 2 → 3
and 2 → 4 shock, but not as the downstream state of a fast shock, since the
solution would have a negative thermal pressure;
• (vii) This state can occur as the upstream state of a slow shock, or as the
downstream state of an intermediate 1 → 3 and 2 → 3 shock;
• (viii) In this region, one of the three solutions has negative pressure. This
will always be the 1 → 3-solution. Therefore this region can occur as the
upstream state of a slow shock, or as a downstream state of an intermediate
2 → 3 shock;
RESULTS 103

Figure 4.3: The 16 regions in parameter space where pu > 0, respectively for
β = 0.1 and β = 2.
104 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

• (ix) This region can occur as the upstream state of a slow shock;
• (x) This state has two solutions with negative pressure: namely both
intermediate solutions. Therefore it can only occur as the upstream state of
a slow shock;
• (xi) This state can occur as the downstream region of a slow shock;
• (xii) This state has two solutions with negative pressure: namely both
intermediate solutions. Therefore it can only occur as the downstream state
of a slow shock;
• (xiii) This state cannot occur, since the single real solution has negative
pressure.
• (xiv) This state cannot occur, since all 3 solutions have negative pressure.
• (xv) This state can occur as the downstream state of a slow shock, and
as the downstream state of an intermediate 2 → 4 shock, but not as the
downstream state of an intermediate 1 → 4 shock, since the solution would
have a negative thermal pressure;
• (xvi) This state can occur as the downstream state of a slow shock, and as
the downstream state of an intermediate 1 → 4 and 2 → 4 shock;

The combination of graphing the surface defined by Ω = 0, together with the


surfaces defined by vn = vs,n , vn = va,n , vn = vf,n and the positive pressure
requirement hence provides a complete, but admittedly non-trivial, graphical
means to the many possibilities for the MHD shock transitions. We now continue
to exploit this knowledge to delimit possible parameter ranges for certain shock
types. As a matter of fact, in what follows, we will use figure 4.3 and how this
figure varies with β, to find limiting values of the parameters at which different
shock types can occur. By doing so, we will actually find the equations of the
curves traced out in (θ, M, β) parameter space that are labeled with P, Q, R, S, T,
U and V as indicated in these plots. These points identify the intersections of the
various regions, and thus act to delimit realizable parameter ranges for shocks.

4.4.4 Switch-on shocks and switch-off shocks

Switch-on shocks are possible where θ = 0. An extra condition for switch-on


shocks to be possible is Ω < 0, since otherwise only fast shocks are possible. Note
that switch-on shocks are only possible when β < γ2 . In this case the plasma is
said to be magnetically dominated. Therefore the maximum value at which we can
RESULTS 105

find switch-on shocks, is found by filling θ = 0 out in the expression for Ω. This
works out to be
s
γ(1 − β) + 1
1 < Ms-on < . (4.23)
γ−1

For the same reason, switch-off shocks can be found in a completely similar manner,
yielding
s
γ(1 − β) + 1
< Ms-off < 1, (4.24)
γ−1

whenever γ(1−β)+1
γ−1 > 0, i.e. when β < γ+1γ . This agrees with the well-known
expressions for these degenerate shock cases (e.g. Kennel et al.[47]).
When solving the RH conditions for θ = 0, when β < γ+1 γ , two switch-on or
5
switch-off solutions exist. For γ = 3 , these solutions are given by

1 − 2M 2 − β
βu = −1 (4.25)
(γ − 1)M 4 + (γ(β − 2))M 2 + γ(1 − β)

Mu = 1 (4.26)
p
θu = ± ((γβ − 2) − (γ − 1)(M 2 − 1))(M 2 − 1). (4.27)

Finally, there is also another solution, which is a hydrodynamical shock. This


solution is given by
r !
6M 2 − β 2M 2 + 5β
(βu , Mu , θu ) = , ,0 , (4.28)
4 8

which only has positive pressure for β < 6M 2 .


Figure 4.4 shows the plane given by θ = 0 in parameter space. The graphs plotted
are (i) the curves given by Ω = 0; (ii) the curve vn = va,n ; (iii) the curve
given by pHD ≡ 6M 2 − β = 0, which is the limiting curve for the existence of
a hydrodynamical shock solution; and (iv) the curve given by ps-off ≡ −4 M 2 −
2 β + 2 + 5 M 2β + 2 M 4 = 0, which is a limiting curve for the existence of a switch-
off solution. The entropy condition ensures that there is no equivalent limiting
curve for switch-on solutions. Note that the known state of a switch-on shock,
where θ = 0 is always magnetically dominated, whereas the known state of a
switch-off shock, where θ = 0 is always thermally dominated. Also shown are the
curves where the hydrodynamical shock solutions and the switch-on and switch-off
106 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

solution have pu = 0. Again, the intersection point, separating different regions,


are labeled with A, B, C and D and used in what follows.
We conclude that the maximum Mach number at which switch-on solutions exist,
is reached in point A, for which (θ, M, β) = (0, 2, 0), as indicated in figure 4.4,
while the maximum plasma-β for switch-on shocks is reached in point B, where
(θ, M, β) = (0, 1, γ2 = 1.2). Therefore the Mach number for the existence of a
switch on-shock is bounded by 1 < M < 2 and the plasma-β at which these
shocks can occur must satisfy 0 < β < γ2 = 1.2. For switch-off shocks, the
minimum plasma-β is reached in point B and is therefore β = 65 = 1.2. The
minimum value of the Mach number M is reached in point C whose coordinates
are found by solving
(
q pu = 0,
(4.29)
γ(1−β)+1
γ−1 = M.
q
γ−1 4
Therefore C satisfies (θ, M, β) = (0, γ+1 , γ+1 ) = (0, 0.5, 1.5), and switch-off
shocks satisfy
r
1 γ−1
= ≤ Ms-off < 1. (4.30)
2 γ+1
In fact, it is straightforward to show that in C, both the curves pHD = 0 and
ps-off = 0 touch, while ω : Ω = 0 also contains C.
Regarding HD shocks, the maximum plasma-β at which they can exist is reached
4
in D, where (θ, M, β) = (0, 1, γ−1 ) = (0, 1, 6), such that all HD shocks satisfy
β < 6.
q
Finally note that ps-off = 0 has M = γ−1 γ = 0.6325 as a horizontal asymptotic,
such that for θ = 0, the RH conditions always lead to at least one solution whenever
M > 0.6325.

4.4.5 Parameter ranges for 1 → 3 shocks.

The upstream state.


We now search for critical values for the upstream parameters, for which the RH
conditions allow for 1 → 3 shocks. First, note that no 1 → 3 shocks are possible
for β > γ2 , since region (iv), as shown in figure 4.3 then no longer exists.
When, at fixed β, implicitly taking derivatives of M to θ on the boundary Ω = 0,
it can be shown that ∂M
∂θ < 0, for θ > 0 and M > 1, and the other way around:
∂M
∂θ < 0, for θ < 0 and M > 1. Therefore the maximum value of M on Ω = 0
RESULTS 107

Figure 4.4: The regions where switch-on or switch-off shocks can occur are colored
in greyscale. By finding the coordinates of points A, B and C, we know the
limiting values for which these shocks can occur. More details are given in the
text.

is reached when θ = 0 (which in figure 4.3 coincides with point P). Hence the
maximum value of the upstream Mach number at which intermediate 1 → 3 shocks
can be found is reached at θ = 0. For varying β, the left panel of figure 4.5 shows
the maximum value of M1,1-3 . Therefore 1 < M1,1-3 < 2.
The maximum value of θ1 for which intermediate 1 → 3 shocks can be found, is
reached on the curve traced out by point Q for varying β in figure 4.3, and thus
requires solving the system

vn = vf,n ,
(4.31)
Ω = 0.

The analytical expression of the solution is again complicated, but the right panel
of figure 4.5 plots the solution in function of β. Since this critical value is decreasing
for increasing β, and the limit value for β = 0, equals 0.65633, we conclude that
−0.65633 < θ1,1→3 < 0.65633.
108 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

2.0

1.8

1.6

1.4

1.2

1.0

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Figure 4.5: Left: The critical upstream Alfvén Mach number for the existence of
intermediate 1 → 3 shocks in function of the upstream β. Right: The critical
upstream θ number for the existence of intermediate 1 → 3 shocks in function of
the upstream β.

Figure 4.6: The critical downstream θ for which 1 → 3 shocks can occur, in
function of the downstream β.

The downstream state.


We now search for critical values for the downstream parameters, for which the
RH conditions allow for 1 → 3 shocks.
RESULTS 109

The downstream region in which 1 → 3 shocks can occur, is region (vii) (as also
shown in figure 4.3. Hence, we find the maximum downstream θ for 1 → 3 shocks
to be reached on the curve traced out by point S for varying β (as also shown in
figure 4.3). Therefore this value can be found by solving

M = 1,
(4.32)
pu = 0.
We find this critical θ to be located at
s p
(γ − 1)β 2 + (γ − 3)β − 2 β(γβ + γ − 1)
θ= ,
(γ − 1)(β + 1)2
4 √1
for β ∈]0, γ−1 ]. Note that the maximum value γ−1
is reached for β = γ −1 .
−1/2
Therefore no 1 → 3 shocks are possible for θ > (γ − 1) = 0.77460.
4
As a bonus, we derived that for β > γ−1 = 0, no 1 → 3 shocks can occur. Figure
4.6 plots this critical value of θ in function of β.
We find the maximum downstream M for 1 → 3 shocks to be reached on the
curve related to the intersections labeled as T or U (as also shown in figure 4.3),
depending on the value of β. At β = 0.1 this maximum downstream value is found
to be located at M = 0.94943, as seen in the top panel of figure 4.3..

4.4.6 Parameter ranges for 2 → 3 shocks.

The upstream state.


The maximum value of M at which the RH conditions allow for 2 → 3 shocks, is
reached on the curve related to point Q, as shown in figure 4.3, and can thus be
found by solving equations (4.31). The left panel of figure 4.7 also shows a plot
of the Alfvén Mach number on Q for varying β. Since this value is decreasing
for increasing β, and the limit value for β = 0, equals 1.19615, we find that
1 < M1,2→3 < 1.19615.
The maximum upstream value of θ at which 2 → 3 shocks can occur is reached on
the curve related to point R, as shown in figure 4.3. Hence, we need to solve the
following system:


pu = 0,
(4.33)
Ω = 0.

A straightforward iteration on β shows that the maximum value is reached at


β = 0.44, and equals θ2 = 1.34283, hence −1.34283 < θ2,2→3 < 1.34283. The
110 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

1.4

1.2

0.8

θ
0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2
β

Figure 4.7: Left: The critical upstream Alfvén Mach number for the existence of
intermediate 2 → 3 shocks in function of the upstream β. Right: The critical value
of θ, for the upstream state of an intermediate 2 → 3 shock for varying β.

variation with β is shown in the right panel of figure 4.7.

The downstream state.


Since the downstream state of a 2 → 3 shock is located in region (vii) or (viii),
as also shown in figure 4.3, and the lower branch of the pu = 0 surface, starting
at T does not cross M = 1, it follows that there are no limiting values for the
downstream θ of an intermediate 2 → 3 shock.
We find the maximum downstream M for 2 → 3 shocks to be reached on the curve
T or V (as also shown in figure 4.3), depending on the value of β. Therefore, at
β = 0.1 this critical value is found to be located at M = 0.94943 again.

4.4.7 Parameter ranges for 1 → 4 shocks

The upstream state.


The exact same reasoning we made for the upstream state of an intermediate 1 → 3
shock can be repeated, thus the limiting values for the upstream state of a 1 → 4
shock are exactly the same.

The downstream state.


We search for critical values for the downstream parameters, for which the Rankine-
Hugoniot conditions allow for 1 → 4 shocks. The downstream state of a 1 → 4
RESULTS 111

shock, must be located in region (xvi).


We first find the minimal value of the Mach number M at which 1 → 4 can occur.
A first important observation is that for all β > 1.2, the (θ, M )-parameter space
contains a region (xiv), since it can be shown that for all β, the pu = 0 curve
crosses the vn = vs,n .
At β = 32 , the pu = 0 curve crosses the Ω = 0 curve in θ = 0. Therefore, when
β > 1.5, the minimum value of M in region (xiv) is reached at θ = 0. When
1.2 < β < 1.5, this minimum value can be located at point T , as labeled in the
bottom panel of figure 4.3).
When β > 1.5, this value is reached on curve T, hence we need to solve

pu = 0;
(4.34)
θ = 0,
in order to find the minimum Mach number forq1 → 4 √ intermediate shocks. For
4−5β+ 25β 2 −24β
fixed β, this minimum value is found to be M = 4 . This function
reaches its minimum at β = 1.5, and the minimal value is 0.5. It can be shown
that for β < 1.5, this minimum value of M is bigger. Therefore 0.5 < M2,1→4 < 1.
For fixed β2 , we can also find the maximum value for θ2 . Therefore we solve the
system

pu = 0;
(4.35)
vn = vs,n .
For β = 2, we find this critical value to be θ = 0.75604, as seen in figure 4.3.

4.4.8 Parameter ranges for 2 → 4 shocks

The upstream state.


The upstream state for an intermediate 2 → 4 shock should be located in region
(v). The limiting upstream values for 2 → 3 shocks are exactly the same as the
limiting upstream values for 2 → 3 shocks.

The downstream state.


The downstream state of a 2 → 4 shock, must be located in region (xv) or (xvi).
Since on pu = 0, it can be shown that, for fixed β, ∂M ∂θ > 0. Therefore M2,2,→4
will reach its minimum value in region (xvi), and it equals the minimum value for
M2,1→4 .
The maximum value of θ2,2→4 is reached on the curve corresponding to point V
in figure 4.3, and for β = 2, θ2,2→4 = 1.65654, as also shown in figure 4.3.
112 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

4.4.9 Relationship between pre- and post-shock magnetic field

In hydrodynamical shock refraction, von Neumann [63], Courant & Friedrichs


[14], Liepmann & Roshko [55], Anderson [5] and many others have studied the
relationship between the pre- and the post-shock streamlines in several initial
setups. Since in the MHD case, these are both dependent on the pre- and
post-shock magnetic field, we will generalize these studies by investigating the
relationship between the pre- and the post-shock magnetic fields.
Therefore we plot (θ, ψ)-diagrams for fixed (β, M ). We assume that the known
state is the upstream state. Figure 4.8 plots these relations for β = 0.1 and varying
M . In the first plot, we have set M ∈ {1.1; 1.2; 1.3; 1.4}. In this range, both fast
and intermediate solutions are possible. In the second panel of figure 4.8, we show
the (θ, ψ)-diagram for M = 1.1, and divide it into certain regions. These regions
mean the following:

• Region A: The solution is a fast shock.


• Region B: There are no solutions with (M, β, θ) = (1.1, 0.1, θ) as upstream
state.
• Region C: There are no solutions with (M, β, θ) = (1.1, 0.1, θ) as upstream
state. The solution with (M, β, θ) = (1.1, 0.1, θ) even has negative pressure.
• Region D: There is an intermediate 1 → 4 solution.
• Region E: There is an intermediate 2 → 4 solution.
• Region F: There is an intermediate 2 → 3 solution.
• Region G: There is an intermediate 1 → 3 solution.

On the other hand, we can also assume that at fixed (β, M ), the known state is the
downstream state. The third panel of figure 4.8 plots for β = 2 and varying M , the
curves connecting the known and unknown magnetic fields. The curves are shown
for M ∈ {0.3, 0.5, 0.7, 0.9}. Here we did not worry about the admissibility of the
solution. In the fourth panel of figure 4.8, we only plot the physically admissible
parts of these curves.

4.5 Conclusion

Magnetohydrodynamical shocks are governed by the Rankine-Hugoniot jump


conditions. These equations can be solved analytically, and doing so essentially
reduces to solving a cubic equation. This solution can have one or three real
CONCLUSION 113

Figure 4.8: The relation between up- and downstream Bt . In the upper panels,
the known state corresponds to the upstream state, and in the lower panels, the
known state corresponds to the downstream state. Upper left: β = 0.1 and M ∈
1.1, 1.2, 1.3, 1.4. According to figure 4.3, three solutions exist in this case. For
each M , the curve reaches its maximum value at Ω = 0. Upper right: Transitions
between certain shock types. In the fast branch, only region A leads to an upstream
state solution. In the intermediate branch, the solution changes shock type: region
D contains 1 → 4 solutions, region E contains 2 → 4 solutions, region F contains
2 → 3 solutions and finally region G leads to 1 → 3 shocks. Lower left: β = 2.
For M ∈ {0.3, 0.5, 0.7, 0.9}, the relation between the known and unknown Bt
are plotted. The physical admissibility is not taken into account. Lower right:
β = 2. For M ∈ {0.3, 0.5, 0.7, 0.9}, the relation between the known and unknown
Bt are plotted. The physical admissibility is taken into account now. The figure
corresponds to the lower panel of figure 4.3.
114 PARAMETER RANGES FOR INTERMEDIATE SHOCKS

solutions. When there is only one real solution, it corresponds to a fast or a slow
magneto-acoustic shock, but when there are three real solutions, also intermediate
shock solutions, which cross the Alfvén speed, can be found. Inspired by the time
reversal principle from Goedbloed [33], we revisited the RH shock relations in the
frequently employed shock frame, and made the duality visible in the (θ, M, β)
state space. Using the thus obtained graphical classification of the state space,
augmented with a positive pressure requirement, we derived limiting values for
parameters in the shock rest frame at which intermediate shocks can be found.
Chapter 5

Nederlandstalige samenvatting

"Over schokbreking in hydro- en magnetohydrodynamica", dat is de letterlijke


vertaling van de titel van dit proefschrift. Eerst leggen we de begrippen
hydrodynamica (HD) en magnetohydrodynamica (MHD) uit, daarna kunnen we
dieper ingaan op schokken, en schokbreking.
HD, ook wel vloeistofmechanica genoemd, is een wiskundig model dat dient om
het gedrag van fluïda te beschrijven. Hoewel zowel gassen als vloeistoffen fluïda
zijn, gebruiken we in deze thesis HD enkel om het gedrag van gassen te beschrijven.
Fluïda bestaan uit vele neutrale deeltjes die onderling interageren. In principe is
het mogelijk om Newtoniaanse mechanica te gebruiken om de snelheid en positie
van ieder deeltje te berekenen doorheen de tijd. Dit zou echter leiden tot een
enorm stelsel van vergelijkingen. Zelfs voor zeer kleine systemen is het oplossen
van zo’n stelsel onbegonnen werk. We moeten dus verstandiger te werk gaan, willen
we het gedrag van fluïda op een bruikbare manier beschrijven. In plaats van de
beweging van elk deeltje afzonderlijk te beschrijven, wordt het fluïdum beschouwd
als een continuüm. In deze continuümbeschrijving spreken we over massadichtheid,
momentum en energiedichtheid, wat essentieel statistische begrippen zijn. De
aannames die HD dan mogelijk maken zijn het behoud van massa, het behoud
van impuls en het behoud van energie.
Een fenomeen dat geïntroduceerd wordt door deze continuümbeschrijving van
gassen is de golf. Een golf is een verstoring die zich voortplant in tijd en
ruimte. Een verstoring is hier gedefinieerd als een kleine afwijking van een
achtergrondstoestand, in om het even welke grootheden (zoals dichtheid, snelheid,
temperatuur). Hierbij is het belangrijk op te merken dat enkel de verstoring zich
verplaatst, niet noodzakelijk de individuele deeltjes. Denk daarbij bijvoorbeeld aan
een watergolf, een microgolf, een radiogolf of een geluidsgolf, of zelfs een mexican
wave.

115
116 NEDERLANDSTALIGE SAMENVATTING

Elk gas heeft een karakteristieke snelheid. Deze karakteristieke snelheid wordt
de geluidssnelheid van het gas genoemd. Het karakteristieke aan deze snelheid
is dat er een schokgolf onstaat wanneer het medium lokaal de geluidssnelheid
overschrijdt. Dit is bijvoorbeeld het geval wanneer een vliegtuig door de
geluidsmuur vliegt. Exacter uitgedrukt: een stationaire HD schok scheidt een
supersonische toestand van een subsonische toestand. Deze eigenschap wordt de
Prandtl-Meyer -eigenschap genoemd. Wiskundig valt dit te begrijpen door op te
merken dat de HD vergelijkingen in dit geval enkel discontinue oplossingen hebben,
die exact deze schokgolven beschrijven. Verder kan aangetoond worden dat de
geluidssnelheid isotroop is: deze snelheid is namelijk even groot in elke richting.
Daarom verplaatst een golf zich sferisch.
De temperatuur van een gas wordt feitelijk bepaald door de random snelheid van
de elektronen: hoe hoger die snelheid, hoe hoger de temperatuur. Wanneer een
gas opgewarmd wordt, stijgt de snelheid van de elektronen rond de protonen.
Wordt het gas voldoende opgewarmd, zodat de elektronen genoeg kinetische
energie bezitten opdat hun momentum de aantrekkingskracht van de protonen
kan overwinnen, dan komen de elektronen los van de ionen, zodat het gas bestaat
uit ionen en vrij bewegende elektronen. Een gas in deze toestand wordt een plasma
genoemd. Gezien elektronen en ionen respectievelijk een negatieve en een positieve
elektrische lading hebben, induceren ze een elektrisch veld, dat op zijn beurt een
magneetveld creëert. Dit magnetisch veld induceert een magnetische druk zodat de
fysica van het probleem essentieel verandert. Het wiskundig systeem dat plasma’s
beschrijft heet MHD. Dit wiskundig systeem is nu zevendimensionaal, waar het HD
systeem driedimensionaal was. Dit heeft verregaande gevolgen. Terwijl HD enkel
de geluidssnelheid als karakteristieke snelheid heeft, heeft MHD 3 karakteristieke
snelheden: de trage magnetosonische snelheid, de Alfvénsnelheid en de snelle
magnetosonische snelheid. Verder kan aangetoond worden dat geen van deze
drie snelheden isotroop is. Informatie plant zich dus niet in elke richting even
snel voort. De veralgemening van de Prandtl-Meyer-eigenschap is nu verre van
triviaal. In plaats van een sub- en een supersonische toestand, kan men met de
drie karakteristieke MHD snelheden nu vier verschillende toestanden definiëren.
Men noemt deze toestanden respectievelijk supersnel (1), subsnel (2), supertraag
(3) of subtraag (4). Een schok die een supersnelle toestand van een subsnelle
toestand scheidt wordt een snelle MHD-schok genoemd. Een schok die een
subtrage toestand van een supertrage toestand scheidt wordt een trage MHD-
schok genoemd, en alle andere schokken worden intermediaire schokken genoemd.
Intermediaire schokken zijn zeker en vast oplossingen van de vergelijkingen die
schokgolven beschrijven (de Rankine-Hugoniot sprongvoorwaarden), maar er is
geen eensgezindheid over het bestaan van deze schokken in de fysische wereld.
Alleszins, in hoofdstuk 4 van deze thesis leiden we af welke wiskundige voorwaarden
het bestaan van zo’n intermediaire schokken toelaten. Ik beschrijf kort onze
aanpak. De Rankine-Hugoniot sprongvoorwaarden (RHV) moeten op elk punt
NEDERLANDSTALIGE SAMENVATTING 117

in de ruimte en op elk tijdstip gelden. Dit zijn feitelijk 8 vergelijkingen in 8


onbekenden, maar gezien een MHD-schok lokaal bekeken een 2D fenomeen is,
is het mogelijk om de RHV te reduceren tot 6 vergelijkingen, in 6 onbekenden:
de massadichtheid ρ, de thermische druk p, twee componenten van de snelheid
(vn , vt ) en de twee componenten van het magneetveld (Bn , Bt ). Nu kunnen
we een schok volledig bepalen door drie dimensieloze parameters, zodat we de
RHV verder willen vereenvoudigen in essentieel drie vergelijkingen. Uit één van
de Maxwell vergelijkingen, ∇ · B = 0, weten we dat Bn constant is aan beide
kanten van de schok, zodat we Bn van de vergelijkingen kunnen elimineren.
Uit het behoud van momentum kunnen we ρ van de vergelijkingen elimineren,
en tenslotte: gezien de RHV translatieinvariant zijn kunnen we vt van de
vergelijkingen elimineren zodat we inderdaad tot drie vergelijkingen komen in
drie onbekenden. We zetten dit stelsel nu om tot een stelsel in de dimensieloze
parameters, en versimpelen het verder tot 1 vergelijking in 1 onbekende. Dit is
een derdegraadsvergelijking. Verder wordt de redenering eenvoudig. Het oplossen
van een derdegraadsvergelijking is gekend. Het wordt dikwijls verkeerdelijk
toegeschreven aan Cardano, een Italiaanse wiskundige uit de Renaissance, maar
feitelijk was de Perzische wiskundige Khayyam (1048 - 1131) hem voor. In elk geval,
een derdegraadsvergelijking kan één, twee of drie verschillende (reële) oplossingen
hebben, en we hebben een nodige en voldoende voorwaarde voor de uniciteit van de
oplossing. We gaan dat zelf ook na in Appendix F. Daarnaast is het niet zo moeilijk
aan te tonen dat een oplossing van deze vergelijking die tot een intermediaire schok
leidt enkel mogelijk is wanneer de derdegraadsvergelijking geen unieke oplossing
heeft. Dit leidt ons tot een wiskundige voorwaarde voor de mogelijkheid tot
intermediaire schokken. Uit deze voorwaarde kunnen we bijvoorbeeld afleiden
dat geen intermediaire schokken mogelijk zijn wanneer de normale snelheid groter
is dan tweemaal de normale Alfvénsnelheid. Tot zover hoofdstuk 4.
Hoofdstuk 2 en 3 behandelen schokbreking in respectievelijk het HD- en het
MHD-geval. Een contactdiscontinuïteit (CD) is het oppervlak dat twee gassen
of plasmas scheidt. In schoktubes worden CD’s nagebootst met behulp van een
heel dun membraan. In deze schoktubes kan men schokken creëren, en deze laten
interageren met een CD. Onze studie is echter niet experimenteel, maar semi-
analytisch. Dit betekent dat de oplossing exact is, maar er zijn iteraties nodig om
deze exacte oplossing te benaderen. We voeren ook computerexperimenten uit met
de numerieke code AMRVAC en vergelijken de exacte oplossing met de numerieke
oplossing. De oplossingen komen goed overeen.
In het HD-geval breekt de schok op de CD en drie signalen onstaan: een
gereflecteerd signaal, een overgebracht signaal en het geschokte contact daartussen.
In het MHD geval onstaan (in het pure planaire geval) vijf signalen: twee
gereflecteerde signalen, twee overgebrachte signalen, en de geschokte CD er
tussenin. Onze bedoeling is de exacte locatie van deze nieuwe signalen te
voorspellen. Onze methode steunt feitelijk op een methode die vooral in
118 NEDERLANDSTALIGE SAMENVATTING

computersimulaties gebruikt wordt, namelijk een exacte Riemannoplosser. Ik leg


eerst het idee uit in het HD geval, en daarna leg ik kort uit welke veralgemeningen
nodig zijn om de Riemannoplosser te laten werken in MHD.
Eerst en vooral moeten we een assenstelsel kiezen waarin we het probleem zullen
oplossen. Als we dit slim doen, en een assenstelsel kiezen dat mee met de schok
reist, wordt het probleem sterk vereenvoudigd: Alle signalen zijn nu immers
stationair. Tenminste wanneer we veronderstellen dat het meest eenvoudige
refractiepatroon gevormd wordt, namelijk dat waarbij alle signalen samenkomen
in één enkel punt. Dit wordt reguliere refractie genoemd. Gezien de CD twee
verschillende gassen scheidt, moet de snelheid er lokaal gelijk zijn aan 0. Ook het
drukverschil tussen de twee kanten van de CD moet er 0 zijn. Het idee is nu dat
we een gok wagen voor deze druk rond de CD. Na wat niet-triviale algebra weten
we tot welke fout op de snelheden rond de CD onze gok voor de druk leidt, en met
deze kennis kunnen we onze gok verbeteren, tot op de gewenste nauwkeurigheid.
Eens we de exacte druk kennen, kunnen we de posities van de signalen bepalen.
Wanneer de schok van een licht naar een zwaar medium reist, zal het overgebrachte
signaal trager bewegen dat de initiële schok. In dit geval zal het gereflecteerde
signaal tevens een schok zijn. Het geschokte contact zal instabiel worden en
oprollen naar rechts. Deze instabiliteit wordt de Richtmyer-Meshkov instabiliteit
genoemd. Wanneer de schok daarentegen van een zwaar naar een licht medium
reist, zal het overgebrachte signaal sneller bewegen dan de initiële schok. Het
gereflecteerde signaal zal in dit geval geen schok meer zijn, maar een continu
signaal dat een verdunningsgolf genoemd wordt. De geschokte CD zal ook in dit
geval Richtmyer-Meshkov instabiel worden, maar nu zal het contact oprollen naar
links. Figuur 2.12 toont een AMRVAC snapshot van deze beide gevallen.
Niet alle parameterkeuzes leiden tot reguliere schokrefractie. We tonen aan dat
wanneer de hoek tussen het schokfront en de CD een bepaalde kritische hoek
overschreidt, zo’n regulier refractiepatroon onmogelijk is. Onze Riemannoplosser
maakt het mogelijk deze kritische hoek te voorspellen, en dus op voorhand te
weten of de refractie al dan niet regulier zal zijn. Figuur 2.9 toont een reguliere
en een irreguliere schokrefractie.
Het basisidee waarop de MHD Riemannoplosser steunt is in essentie hetzelfde. Er
zijn enkel enkele technische complicaties die de implementatie van deze oplosser
moeilijker maken. Waar we in het HD geval enkel een gok voor de druk rond
de CD maken, maken we hier gokken voor de magneetveldcomponenten, de
snelheidscomponenten en de thermische druk. Daardoor zal de numerieke iteratie
nu op een vierdimensionele functie uitgevoerd moeten worden. Bovendien is deze
functie niet overal continu, zodat het verbeteren van onze gokken niet triviaal is.
Erger nog: zoals we in hoofdstuk vier aantonen kunnen de RH sprongvoorwaarden
in MHD tot drie verschillende (reële) oplossingen hebben. Gezien we deze
vergelijkingen tot viermaal oplossen leidt dit in elke tijdstap tot hoogstens 81
NEDERLANDSTALIGE SAMENVATTING 119

verschillende oplossingen. We verhelpen dit euvel door een kleine toegeving te


doen: we postuleren de aard van de verschillende signalen vooraleer we de oplosser
aan het werk laten. Hierbij postuleren we steeds dat de snelle signalen snelle
schokken zijn. De trage signalen daarentegen kunnen zowel verdunningsgolven,
trage schokken als intermediaire schokken zijn. Tevens kunnen we transities
vaststellen tussen verschillende gevallen. We hebben parameterregimes gevonden
waaronder de oplosser zowel een overgebrachte als gereflecteerde intermediaire golf
voorspelt. De AMRVAC simulaties bevestigen dit resultaat. Figuur 3.8 toont de
magneetvelden in verschillende MHD gevallen.
Verder kan beargumenteerd worden dat de Richtmyer-Meshkov instabiliteit in het
MHD geval onderdrukt is, en de CD stabiel blijft. Figuur 3.2 toont het verschil
tussen het HD en het MHD geval.
Chapter 6

Conclusions

We developed an exact Riemann solver-based solution strategy for shock refraction


at an inclined contact discontinuity (CD) in HD and ideal MHD.
In the HD case, our self-similar solutions agree with the early stages of nonlinear
AMRVAC simulations. We predict the critical angle αcrit for regular refraction,
and the results fit with numerical and experimental results. Our solution strategy
is complementary to von Neumann theory, and can be used to predict full
solutions of refraction experiments, and we have shown various transitions possible
through specific parameter variations. After reflection from the top wall, the CD
becomes RM-unstable. Adding perpendicular magnetic fields leaves the contact
RM-unstable.
In planar ideal MHD, the shock refracts in five signals instead of three. After
reflection from the top wall, the CD remains RM-stable, since the ideal MHD
equations do not allow for vorticity deposition on a CD. We are able to reproduce
results from the literature and results by numerical simulations performed by
AMRVAC.
Magnetohydrodynamical shocks are governed by the Rankine-Hugoniot jump
conditions. These equations can be solved analytically, and doing so essentially
reduces to solving a cubic equation. This solution can have one or three real
solutions. When there is only one real solution, it corresponds to a fast or a slow
magnetoacoustic shock, but when there are three real solutions, also intermediate
shock solutions, which cross the Alfvén speed, can be found. Inspired by the time
reversal principle from Goedbloed [33], we revisited the RH shock relations in the
frequently employed shock frame, and made the duality visible in the (θ, M, β)
state space. Using the thus obtained graphical classification of the state space,
augmented with a positive pressure requirement, we derived limiting values for

121
122 CONCLUSIONS

parameters in the shock rest frame at which intermediate shocks can be found.
We are generalizing this approach to shock refraction in relativistic hydrodynamics.
Next to this shock refraction research, we have also added a data conversion
subroutine to AMRVAC, such that the scientific data produced by the numerical
code AMRVAC (van der Holst & Keppens [93]; Keppens et al.[49]) can be read in
by visualization software as ParaView and VisIt.
Appendix A

Structure of AMRVAC

The AMRVAC code exists of several directories:

• src: contains the source code, namely:


the main program: after compilation, this program is adapted to the
physics, numerical methods, dimensionality, etc.;
the physics subroutines: AMRVAC contains physics modules for
advection (for testing purposes) and classical and (special) relativistic hydro-
and magnetohydrodynamics. There is also a radiative cooling routine
implemented. We have developed classical HD and MHD modules which
allow for an interface separating two different gases, i.e. gases with different
values of γ (as described in Delmont et al.[19]);
the numerical methods: including TVD, TVDLF, HLLC, TVDMU,
approximate Roe solvers etc. and limiters including minmod, Woodward,
etc.;
IO subroutines;
Conversion subroutines: Conversion subroutines to idl [43], dx [24],
tecplot [83], vtu, ascii are included. I developed the VTU-conversion
subroutine as part of my PhD research.
• par: contains simulation dependent choices such as:
Input and output files. AMRVAC produces two kinds of output files.
The OUT-file saves all the conserved variables, and optionally also other
user defined variables, at certain times. The log file saves global integrated
values of conserved variables at certain times, plus the number of cells at
each AMR-level.

123
124 STRUCTURE OF AMRVAC

Saving times. The user can decide at which moment an OUT file is
produced, and at which moment global integrated data are saved in the log
file.
The stop criterion.
The numerical methods. The user decides which numerical methods
should be used. This choice can be level dependent. Also the limiters and
∇ · B-control algorithm is chosen by the user.
Boundary conditions. The number of ghost cells is defined here. Also
the physics dependent boundaries are defined, for every boundary, for every
conserved variable. Pre-defined choices include continuous to mimic open
boundaries, symmetric or antisymmetric to mimic rigid walls or periodic
for periodic problems (e.g., in cylindrical coordinates). The user has the
possibility to define other boundaries too.
AMR related choices, such as the number of AMR levels, the resolution
on the coarsest level, the size of the domain, the tolerance for the refinement
criterion.
The Courant parameter.
• usr: contains the initial conditions. Appendix C gives the usr-file for the
HD shock tube problem presented in Delmont et al.[19], and in Chapter 2.
Appendix B

Compilation

Before creating the usr- and the par-file, the AMRVAC user knows the geometry,
dimensionality and physics of the problem. He also needs to define the values ci
that define the numbers of cells (per dimension i) in one grid block. Once this is
done, the compilation can be done as follows:
make c l e a n
seta mr va c −p=mhd −d=23 −g =16 ,16 −u=rimmhd23problem1
make amrvac

The setamrvac command selects the correct physics, dimensionality and ci . In


the example given above, the MHD module will be used for a 2.5D problem,
which means that vectors have three components on a two-dimensional -by default
Cartesian- domain. Every grid will consist of 16 × 16 cells, including ghost cells.
The usr-file used will be rimmhd23problem1.t, and is given in Appendix C. This
file is related to the RMI problem from Chapter 2.

125
Appendix C

USR-file

This appendix shows the USR file used for the simulation of the shock tube problem
presented in Delmont et al.[19], or in Chapter 2. It is written in LASY syntax, as
introduced in Tóth [88].
s u b r o u t i n e i n i t o n e g r i d _ u s r (w, ixG^L , x )

! i n i t i a l i z e one g r i d

include ’ amrvacdef . f ’

! scalars
integer : : ixG^L

! arrays
d o u b l e p r e c i s i o n : : w( ixG^S , 1 : nw ) , x ( ixG^T , 1 : ndim )

! local scalars
d o u b l e p r e c i s i o n : : xpi , vpost , r h o p o s t , ppost , xshock , xpi1 , xbound , tang

!−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−

o k t e s t = i n d e x ( t e s t s t r , ’ i n i t o n e g r i d _ u s r ’)>=1
i f ( o k t e s t ) w r i t e ( unitterm , ∗ ) ’−> i n i t o n e g r i d _ u s r ( in ): ’ ,&
’ ixG^L : ’ , ixG^L

{^IFONED stop ’ This i s not a 1D problem ’ }


{^IFTHREED stop ’ This i s not a 3D problem ’ }
{^IFTWOD

! parameters

127
128 USR-FILE

eqpar (gamma_)=1.4 d0
M=10.0 d0
e t a =0.1 d0
tang=one

! l o c a t i o n shock and CD
xshock =0.1 d0
xbound=1.5 d0

v p o s t=M−(( eqpar (gamma_)−one ) ∗M∗∗2+2.0 d0 ) / ( ( eqpar (gamma_)+one ) ∗M)


r h o p o s t=−one ∗ ( v p o s t+M) ∗ eqpar (gamma_) ∗M
p p o s t =(2∗ eqpar (gamma_) ∗M∗∗2− eqpar (gamma_)+one ) / ( eqpar (gamma_)+one )

where ( x ( ixG^S ,1) > xshock . and . ( x ( ixG^S ,1) > x ( ixG^S , 2 ) / tang+xbound ) )
! p r e shock r e g i o n
w( ixG^S , rho_)= eqpar (gamma_) ∗ e t a
w( ixG^S , m1_)= z e r o
w( ixG^S , m2_)= z e r o
w( ixG^S , e_)=one / ( eqpar (gamma_)−one )
endwhere
where ( x ( ixG^S ,1) > xshock . and . ( x ( ixG^S,1)<=x ( ixG^S , 2 ) / tang+xbound ) )
! p r e shock r e g i o n
w( ixG^S , rho_)= eqpar (gamma_)
w( ixG^S , m1_)= z e r o
w( ixG^S , m2_)= z e r o
w( ixG^S , e_)=one / ( eqpar (gamma_)−one )
endwhere
where ( x ( ixG^S,1)<= xshock )
! p o s t shock r e g i o n
w( ixG^S , rho_)= r h o p o s t
w( ixG^S , m1_)= v p o s t ∗ r h o p o s t
w( ixG^S , m2_)= z e r o
w( ixG^S , e_)=one / ( eqpar (gamma_)−one )+0.5 d0∗ r h o p o s t ∗ v p o s t ∗∗2
endwhere
}

return
end s u b r o u t i n e i n i t o n e g r i d _ u s r
Appendix D

Stationary planar
Rankine-Hugoniot conditions

We allow weak solutions of the system, which are solutions of the integral form
of the MHD equations, that may contain discontinuities. The shock occuring in
the problem setup, as well as those that may appear as F R, SR, ST or F T
signals later on obey the Rankine-Hugoniot conditions. In the case where the
shock speed s = 0, the Rankine-Hugoniot conditions follow from equation (3.1).
When considering a thin continuous transition layer in between the two regions,
with thickness δ, solutions for the integral form of equation (3.1) should satisfy
R2 ∂ ∂
lim 1 ( ∂x F + ∂y G)dl = 0. For vanishing thickness of the transition layer, this
δ→0
yields
 
ρvn
2 B2 B2

 ρvn + p − 2n + 2t 

 ρvn vt − Bn Bt 
 = 0, (D.1)
 
γ v 2 +v 2
vn ( γ−1 p + ρ n 2 t + Bn2 ) − Bn Bt vt

 
 
 vn Bt − vt Bn 
Bn

where the index n refers to the direction normal to the shock front and the index t
refers to the direction tangential to the shock front (see e.g. Goedbloed & Poedts
[32]). From equations (D.1) we know that m ≡ ρvn and Bn are constant across a

129
130 STATIONARY PLANAR RANKINE-HUGONIOT CONDITIONS

shock. Inserting these constants in the other equations of (D.1) yields

1 1
m2 [[ ]] + [[p]] + [[Bt2 ]] = 0, (D.2)
ρ 2

m[[vt ]] − Bn [[Bt ]] = 0, (D.3)

Bt
m[[ ]] − Bn [[vt ]] = 0, (D.4)
ρ

γ p m2 1 B2 B2
[[ ]] + [[ 2 ]] + [[ t ]] − n2 [[Bt2 ]] = 0, (D.5)
γ−1 ρ 2 ρ ρ 2m

2
γ vn +vt2
where we used equation (D.4) to eliminate vt from [[vn ( γ−1 p+ρ 2 + Bn2 ) −
Bn Bt vt ]] = 0 to arrive at

γ p m3 1 B2 B2
m[[ ]] + [[ 2 ]] + m[[ t ]] − n [[Bt2 ]] = 0, (D.6)
γ−1 ρ 2 ρ ρ 2m

which is equivalent to equation (D.5), under the assumption that m 6= 0, which is


true for all magnetoacoustic signals.
dy v
Since the signal located at dx = vyx is a contact discontinuity, it clearly obeys
m = 0. This simplifies equations (D.2-D.4) drastically:
t
[[(vt , Bt , p) ]] = 0. (D.7)

Hence, in terms of the primitive variables, this means that vx , vy , p and Bx , and
thus also By , remain constant across a contact discontinuity.
Appendix E

Relations across a shock

Suppose we know the primitive variables uk at one side of a stationary shock, and
denote the unknown primitive variables at the other side of the shock uu .
We will discuss the consequences of the RH condition relating up- and downstream
states, for the case where Bt,k 6= 0. After elimination of [[vt ]] from equation (D.4),
we arrive at the following 3 × 3-system (see also Torrilhon [91]):

1 1
p̂ − 1 + C( − 1) + (B̂t2 − A2 ) = 0, (E.1)
ρ̂ 2

B̂t
C( − A) − B 2 (B̂t − A) = 0, (E.2)
ρ̂
1 p̂ 1 1 1 1
( − 1) + ( − 1)(p̂ + 1) + ( − 1)(B̂t − A)2 = 0, (E.3)
γ − 1 ρ̂ 2 ρ̂ 4 ρ̂

where we introduced the dimensionless quantities connecting the up-and down-


v B
stream values given by ρ̂ = ρρuk , p̂ = ppuk , vˆt = ct,u
k
and B̂t = √t,u
pk and the
Bt,k
dimensionless parameters, quantifying the known values in state uk by A ≡ √ ,
pk
2
Bn,k ρk vn,k
B ≡ √
pk and C = pk . These parameters allow for simple criteria for fast,
B̂t
slow or intermediate shocks: fast shocks are characterised by A > 1, slow shocks
B̂t
are characterised by 0 < < 1 and intermediate shocks satisfy B̂At < 0, since
A
the known state is the upstream state and the unknown state is the downstream
state.

131
132 RELATIONS ACROSS A SHOCK

Note that super-Alfvénic flow in the direction normal to the shock front in the
state uk now implies

C > B 2 ⇔ vn,k
2
> a2n,k . (E.4)

The solution to equations (E.1-E.3) is given by

C B̂t
ρ̂ = , (E.5)
AC + B 2 (B̂t − A)
!
2C γ−1 (A − B̂t )2 B̂t C(A + B̂t )
p̂ = − +1 − , (E.6)
γ+1 γ+1 2 (γ + 1)(C − B 2 )

where B̂t must satisfy a cubic relation

Σi=0,3 τi B̂ti = 0, (E.7)

with its coefficients given by

τ3 = B2, (E.8)

(γ − 1)(B 2 − C) + C A,

τ2 = (E.9)

τ1 = ((γ + 1)B 2 − (γ − 1)C − (2 + A2 )γ)(B 2 − C), (E.10)

τ0 = −(γ + 1)A(B 2 − C)2 . (E.11)

Once the solution B̂t from equation (E.7) is determined, and used to calculate
p̂ and ρ̂ from equations (E.5)-(E.6), we find the upstream state from ρu = ρ̂ρk ,

pu = p̂pk and Bt,u = B̂t pk . Finally equation (D.1) delivers vn,u and Bn,u , while
from equation (D.4), we know vt,u :

m
vn,u = , (E.12)
ρ̂ρk
Bn,k
vt,u = vt,k + (Bt,n − Bt,u ), (E.13)
ρk vn,k

Bn,u = Bn,k . (E.14)

Finally, note that the regular solution discussed above is only valid when B 2 6= C,
i.e., when the normal velocity in the known region does not equal the Alfvén
RELATIONS ACROSS A SHOCK 133

velocity, or equivalently when Bt,k does not vanish. In this special case the system
(E.1-E.3) has four mathematical solutions possible, one of which is the trivial
solution, one of which is a rotational shock (which is the limit case of solution
(E.5-E.6)) and finally the following non-trivial switch-off shock:

B̂t = 0, (E.15)

2(γ + 1)C
ρ̂ = q , (E.16)
2
γ(A2 + 2C + 2) ± (γ(A2 + 2)) + 4C(C + A2 − 2γ)
q
2
2 + 2C + A2 ∓ (γ(A2 + 2)) + 4C(C + A2 − 2γ)
p̂ = . (E.17)
2(γ + 1)
Appendix F

Solving the cubic analytically

The cubic (E.7) is solved analytically by the following procedure. Defining the
real quantities

D = 4(τ22 − 3τ3 τ1 ), (F.1)

N = 4(2τ23 − 9τ3 τ2 τ1 + 27τ32 τ0 ), (F.2)

one notes that D is the discriminant of the derivative of the cubic. When D > 0,
the inequality N 2 − D3 < 0 gives limiting values for τ0 , for which the 2 extrema
of the cubic have opposite signs. Since N 2 − D3 < 0 implies D > 0, the criterium
on the coefficient of the cubic to have 3 different real solutions is

N 2 − D3 < 0. (F.3)

From these real-valued quantities, we define the following, possibly complex,


quantities:

p  13
H = N 2 − D3 − N , (F.4)

D
J = . (F.5)
H

135
136 SOLVING THE CUBIC ANALYTICALLY

In terms of these introduced quantities, the 3 solutions of the cubic equation (E.7)
are given by

J + H − 2τ2
B̂t,0 = , (F.6)
6τ3

J + H + 4τ2 − 3(H − J)i
B̂t,1 = − , (F.7)
12τ3

J + H + 4τ2 + 3(H − J)i
B̂t,2 = − . (F.8)
12τ3

The evaluation of expressions (F.6 - F.8) requires the evaluation of N 2 − D3 and
√ 1 √
N 2 − D3 − N 3 . The principal square root : R → R is continuous, but the

principal cube root 3 : C → C has a branch cut along the negative real axis. It
√ 1
follows that h : R2 → C : (N, D) 7→ N 2 − D3 − N 3 is discontinuous when

N 2 − D3 − N is a negative (and thus real) number.
Before we explain how to permute the indices of the roots (F.6-F.8), let us first
rewrite these expressions as

B̂ι = Ĥι + Jˆι , (F.9)


“ √ ”ι “ √ ”−ι
− 12 + 23 i H − 21 + 3
2 i J
where Ĥι = Hι − τ2
6τ3 , Jˆι = Jι − τ2
6τ3 , Hι = 6τ3 , Jι = 6τ3 and
ι ∈ {0, 1, 2}.
Let us make some small technical remarks.
 √ ι 3
• Note that − 12 + 3
2 i = 1, for each integer value of ι;

• Note that τ3 > 0. Therefore, equation (F.9) is well-defined. Also, when we


know that one of those expressions is real, then its sign equals the sign of its
numerator;
• It turns out that dividing the complex plane into 6 sextants is useful for our
analysis. We therefore define the sextants
π π
Sj ≡ {r(cos ϕ + i sin ϕ)|(j − 1) < ϕ < j , r > 0}, (F.10)
3 3
and Li , the lines separating those sextants, i.e.,
π
Lj ≡ {r(cos ϕ + i sin ϕ)|ϕ = j , r > 0}. (F.11)
3
These regions are shown in figure F.1.
SOLVING THE CUBIC ANALYTICALLY 137

Figure F.1: The sextants Si and their separators, Li .

One can distinguish the following cases.

• Case I: D > 0, N > 0 and N 2 − D3 > 0.


Note that H ∈ L1 and since D > 0, J ∈ L5 . Therefore, H1 , J1 ∈ √L3 , J2 ∈ L1
and H2 ∈ L5 . It follows that B̂t,1 is a real number. Note that N 2 − D3 <
√ √ 2
N , thus 2(N 2 − D3 ) < 2N N 2 − D3 or N 2 − D3 − N < D3 , thus

H < D or |H| < |J|. Therefore, B̂t,2 ∈ S1 and B0 ∈ S6 .

• Case II: D > 0, N < 0 and N 2 − D3 > 0.


Note that H ∈ L6 and since D > 0 also J ∈ L6 . Therefore, √ H2 , J1 ∈ L4 and
H1 , J2 ∈ L2 . It follows that B0 is a real number. Note that N 2 − D3 > N ,
√ √ 2 √
thus 2(N 2 −D3 ) > 2N N 2 − D3 or N 2 − D3 − N > D3 , thus H > D
or |H| > |J|. Therefore, B̂t,1 ∈ S3 and B̂t,2 ∈ S4 .

• Case III: D > 0, N > 0 and N 2 − D3 < 0.


Note that H ∈ S1 and since D >√0, J ∈ S6 . Therefore, H1 ∈ S3 , H √2 ∈ S5 ,
J1 ∈ S4 and J2 ∈ S2 . Since | N 2 − D3 − N |2 = D3 , |H| = D and
|H| = |J|. Hence √
all of the roots are

real. After some

algebra, we know that

τ2 D τ2 D τ2 D τ2 D
B̂t,0 > − 3τ 3
+ 2 , B̂t,1 < − 3τ3 − 2 and − 3τ3 − 2 < B̂t,2 < − 3τ3 − 2 .
138 SOLVING THE CUBIC ANALYTICALLY

• Case IV: D > 0, N < 0 and N 2 − D3 < 0.


Note that H ∈ S6 and since D >√0, J ∈ S1 . Therefore, H1 ∈ S2 , H √2 ∈ S4 ,
J1 ∈ S3 and J2 ∈ S5 . Since | N 2 − D3 − N |2 = D3 , |H| = D and √
τ2
|H| = |J|. Hence all the roots are real. We deduce that B̂t,0 > − 3τ + 2D ,
√ √ √ 3
τ2 D τ2 D τ2 D
B̂t,2 < − 3τ 3
− 2 and − 3τ 3
− 2 < B̂t,1 < − 3τ 3
− 2 .

• Case V: D < 0 ⇒ N 2 − D3 > 0.


Note that H ∈ L6 and thus J ∈ L3 . Hence B̂t,0 is a real number. Since D is
the discriminant of the derivative of the cubic, we know that the cubic only
has one real root. Note that J1 ∈ L1 , H1 ∈ L2 , H2 ∈ L4 and J2 ∈ L5 , which
tells us that ImB̂t,1 > 0 and ImB̂t,2 < 0.

The Rankine-Hugoniot conditions indeed allow for a unique real solution if and
only if N 2 − D3 > 0.

When N , D or N 2 − D3 change sign, we might need to permute the indices. We


will not describe all the possible permutations in detail, but instead will illustrate
this in one example. Suppose that in 2 subsequent iteration steps, D > 0, but in
the first iteration step N > 0, while in the following step N < 0. Comparing case
1 and case 2, we come to the conclusion that we need to permute the indices of
B̂t,0 and B̂t,1 .

So far, we considered B̂t,ι just as a finction of the cubics coefficients τκ . These


coefficients are actually functions of the known state parameters, which are on
their turn functions of the guessed wave refraction angles. The question remains
at which angles B̂t,ι (φguess ) are continuous (and differentiable). Here φguess refers
to the guessed wave refraction angle.
Let us illustrate these findings by a simplified example. Take the initial guess
φF T = 1.27678, which is the exact solution, and let the guess for φST vary. In figure
F.2 we show the real part of the unpermuted roots of the cubic (E.7), together
with N 2 − D3 and B 2 − C across the corresponding signal.
Note that the Alfvénic angle φa,T = 1.1936, since at this position B 2 − C vanishes.
Therefore, when φST < φa,T , the slow transmitted signal is a slow shock. On the
other hand, when φST > φa,T , the signal is an intermediate shock. Also note that
φa,T is a double root of N 2 − D3 . Since B 2 − C changes sign, the roots B̂t,1 and
B̂t,2 are permuted.
The critical angle φcr,T = 1.1948 is actually the smallest root of N 2 − D3 , bigger
than φa,T . For φ > φcr,T , the cubic (E.7) has only one real solution and three real
solutions for φ < φCr,T .
SOLVING THE CUBIC ANALYTICALLY 139

0.2 0.15
Re(B1) N2-D3
Re(B2) B2-C
0 Re(B3)
0.1
-0.2

parameters
0.05
-0.4
Re(Bi)/A

-0.6
0

-0.8
-0.05
-1

-1.2 -0.1
1.193 1.1935 1.194 1.1945 1.195 1.1955 1.196 1.193 1.1935 1.194 1.1945 1.195 1.1955 1.196
φ φ

Figure F.2: Left: the real parts of the roots of the cubic (E.7) for ST. Right: The
graphs of (B 2 − C)(φ) and (N 2 − D3 )(φ). Note that the Alfvénic angle is a root
of B 2 − C and a double root of N 2 − D3 .

N changes sign at φ = 1.1953, therefore, B̂t,0 and B̂t,1 are permuted here, since
D > 0. The exact solution is located at φ = 1.19283, and the root we select is
B̂t,1 .
Appendix G

Integration across rarefaction


waves

When the slow signal is located at a position where no shock solution is possible
which satisfies the entropy condition, we postulate the solution to be a rarefaction
fan. In this case, we rewrite the stationary MHD equations (3.1) in cylindrical form,
by local decomposition in normal and tangential components. We additionally

assume self similarity, i.e. ∂t = 0, where the index t again refers to the tangential
direction. Since the entropy S can be shown to be invariant in expansion fans
(see e.g. De Sterck et al.[22]), we replace the energy equation by the isentropic
equation. Doing so, we obtain

ρ′
  
vn ρ 0 0 0

 0 ρvn4 − (vt Bn − Ez )2 vn Bn (vt Bn − Ez ) vn3 0 
 vn′ 

 0 −Bn (vt Bn − Ez ) −ρvn3 + vn Bn2 0 0  vt′ =
p′
  
 γp 0 0 ρ 0  
0 0 0 0 vn Bn′
 
−ρvt
 vt (vt Bn − Ez )2 
2vn2 Bn2 − ρvn4 + vt Bn (vt Bn − Ez )
 
 (G.1)
 
 0 
−vt Bn + Ez

where the prime ′ denotes derivation in the normal direction. Also note that Bt is
eliminated from the system since it is completely determined by the other variables.

141
142 INTEGRATION ACROSS RAREFACTION WAVES

Only when the system (G.1) is singular it has a solution, in other words
 2
(vt Bn − Ez )2
  
Bn γp γp Bn
vn vn4 − + + vn2 + = 0, (G.2)
ρ ρvn ρ ρ ρ

m
2 2 2 2
v⊥ (v⊥ − vf,⊥ )(v⊥ − vs,⊥ ) = 0, (G.3)

which is equivalent to characteristic equation (3.7). This relation should hold inside
of the expansion fans, since u is differentiable there. The solution to equations
(G.1) is given by

vn vt (2vn2 − a2 − c2 ) + c2 a˜n a˜t


ρ′ (φ) = 2ρ , (G.4)
(4 + 2γ)vn4 − ((2γ + 1)a2 + (3 + γ)c2 )vn2 + γa2n c2

vn vt (2vn2 − a2 − c2 ) + c2 a˜n a˜t


p′ (φ) = 2γp , (G.5)
(4 + 2γ)vn4− ((2γ + 1)a2 + (3 + γ)c2 )vn2 + γa2n c2

Bn′ (φ) = −Bt , (G.6)

vt (vn2 ((2γ − 1)a2 + (γ + 1)c2 − 2γvn2 ) − γa2n c2 ) + 2vn c2 a˜n a˜t


vn′ (φ) = (G.7)
,
(4 + 2γ)vn4 − ((2γ + 1)a2 + (3 + γ)c2 )vn2 + γa2n c2

Σ6i=0 τi vni
vt′ (φ) = . (G.8)
((4 + 2γ)vn4 − ((2γ + 1)a2 + (3 + γ)c2 )vn2 + γa2n c2 )vn a˜t

Bn Bt
Here we introduced the symbols a˜n ≡ √
2p
and a˜t ≡ √
2p
, and the coefficients τi
are given by

τ0 = −γa4n c3 , (G.9)

τ1 = 0, (G.10)

τ2 = ((γ − 2)a˜n a˜t + (γ + 3)c2 + (2γ + 1)a2 )a˜n c2 , (G.11)

τ3 = 2(a2 + c2 )a˜n vt , (G.12)

τ4 = −(2γ + 1)a˜t a2 − ((γ + 3)a˜t + (2γ + 4)a˜n )c2 , (G.13)

τ5 = −4a˜n vt , (G.14)

τ6 = (2γ + 4)a˜t . (G.15)


INTEGRATION ACROSS RAREFACTION WAVES 143

These expressions are valid for slow rarefaction fans. The integration has to be
performed, starting from φsl,R/T till φSR/ST .
Bibliography

[1] Abd-El-Fattah, A. M., Henderson, L. F. & Lozzi, A. 1976 Precursor


shock waves at a slow-fast gas interface. J. Fluid Mech. 76 157–176.
[2] Abd-El-Fattah, A. M. & Henderson, L. F. 1978 Shock waves at a fast-
slow gas interface. J. Fluid Mech. 86 15–32.
[3] Abd-El-Fattah, A. M. & Henderson, L. F. 1978 Shock waves at a slow-
fast gas interface . J. Fluid Mech. 89 79–95.
[4] Akhiezer, A. I., Lyubarskii, G. Ya., Polovin, R. V. 1959 On the stability
of shock waves in magnetohydrodynamics. Soviet Phys. -JETP 8, 507–511.
[5] Anderson, J. E. 1963 Magnetohydrodynamical Shock Waves. MIT Press,
Cambridge, Massachusetts.
[6] http://apod.nasa.gov/apod/ap001115.html
[7] Batchelor, G. K. 1967 An Introduction to Fluid Dynamics Cambridge
University Press.
[8] Barmin, A. A., Kulikovskiy, A. G. & Pogorelov, N. V. 1996 Shock-
capturing approach and nonevolutionary solutions in magnetohydrodynamics.
J. Comp. Phys. 126, 77–90.
[9] Berger, M. J. & Colella. 1989 Adaptive mesh refinement for shock
hydrodynamics J. Comp. Phys. 126, 77–90.
[10] Brouillette, M.. 2002 The Richtmyer-Meshkov Instability Annu. Rev.
Fluid Mech. 34, 445–468.
[11] Brio, M. & Wu, C. C. 1988 An upwind differencing scheme for the
equations of ideal magnetohydrodynamics. J. Comp. Phys. 75, 400–422.
[12] Chao, J. K., Lyu, L. H., Wu, B. H., Lazarus, A. J., Chang, T. S. &
Lepping, R.P. 1993 Observation of an intermediate shock in interplanetary
space. J. Geophys. Res. 98, 17443–17450.

145
146 BIBLIOGRAPHY

[13] Coppi, P. S., Blandford, R. D. & Kennel, C. F. 1988 On the existance


and stability of intermediate shocks ESA SP-285, 381-384
[14] Courant, R. & Friedrichs, K. O. 1948 Supersonic flow and shock waves
. Interscience Publishers, New York.
[15] Chané, E., van der Holst, B., Poedts, S. & Kimpe, D. 2005 On
the effect of the initial magnetic polarity and of the background wind on the
evolution of CME shocks. Astron. Astrophys. 432(1), 331–339.
[16] Courant, R & Hilbert, D. 1962 Methods of mathematical physics, Vol. 2
. Wiley.
[17] Chu, C. K. & Taussig, R. T. 1967 Numerical Experiments of
Magnetohydrodynamic Shocks and the Stability of Switch-On Shocks. Phys.
Fluids 10, 249–256.
[18] Delmont, P. & Keppens, R. 2008 An exact solution strategy for regular
shock refraction at a density discontinuity. ECA 32D D2.008.
[19] Delmont, P., Keppens, R. & van der Holst, B. 2009 An exact Riemann-
solver-based solution for regular shock refraction. J. Fluid Mech. 627, 33–53.
[20] Delmont, P. & Keppens, R. 2010 Shock refraction in ideal MHD. Journal
of Physics: Conference Series, accepted.
[21] Delmont, P. & Keppens, R. 2010 Parameter regimes for slow, intermedi-
ate and fast MHD shocks. J. Plasma Phys., DOI 10.1017/S0022377810000115
[22] De Sterck, H., Low, B. C. & Poedts, S. 1998 Complex
magnetohydrodynamic bow shock topology in field-aligned low-β flow around
a perfectly conducting cylinder. Phys. of Plasmas 11, 4015–4027.
[23] De Sterck, H. & Poedts, S. 2000 Intermediate Shocks in Three-
Dimensional Magnetohydrodynamic Bow-Shock Flows with Multiple Interact-
ing Shock Fronts Phys. Rev. Lett. 84, 5524–5527.
[24] www.opendx.org
[25] van Eerten, H. J, Meliani, Z., Wijers, R.A.M.J. & Keppens, R. 2009
No visible optical variability from a relativistic blast wave encountering a wind
termination shock MNRAS 398, L63-L67.
[26] Falle, S. A. E. G. & Komissarov, S. S. 1997 On the existence of
intermediate shocks. Mon. Not. R. Astron. Soc. 123, 265–277.
[27] Falle, S. A. E. G. & Komissarov, S. S. 2001 On the inadmissibility of
non-evolutionary shocks. J. Plasma Phys. 65, 29–58.
BIBLIOGRAPHY 147

[28] http://www.fom.nl/live/imgnew.db?31619
[29] Feng, H. & Wang, J. M. 2008 Observations of a 2 → 3 Type Interplanetary
Intermediate Shock. Solar Phys. 247, 195–201.
[30] Germain, P. 1960 Shock Waves and Shock-wave structure in magneto-fluid
dynamics. Rev.Mod.Phys. 32, 951–958.
[31] Glimm, J. 1965 Solutions in the large for nonlinear hyperbolic systems of
equations. Comm. Pure Appl. Math 41, 569–590
[32] Goedbloed, H. & Poedts, S. 2004 Principles of Magnetohydrodynamics
With Applications to Laboratory and Astrophysical Plasmas. Cambridge
University Press.
[33] Goedbloed, J. P. 2009 Time reversal duality of magnetohydrodnamical
shocks. Phys. Plasmas 16, 1–14
[34] Ghezzi, I. & Ruggles, C. 2007 Chankillo: A 2300-Year-Old Solar
Observatory in Coastal Peru. Science 315 1239–1243.
[35] Gombosi, T. I. 1998 Physics of the Space Environment . Cambridge
University Press.
[36] Goossens, M. 2003 An introduction to plasma astrophysics and magnetohy-
drodynamics. Kluwer Academic Publishers.
[37] Haas, J. F. & Sturtevant, B. 1987 The refraction of a plane shock wave
at a gas interface. J. Fluid Mech. 26 607.
[38] Hawley, J. F. and Zabusky, N. J. 1989 Vortex Paradigm for Shock-
Accelerated Density-Stratified Interfaces. Phys. Rev. Lett. 63 1241–1245
[39] de Hoffmann, F. and Teller, E. 1950 Magneto-Hydrodynamic Shocks.
Phys. Rev. 80, 692–703.
[40] Henderson, L. F. 1966 The refraction of a plane shock wave at a gas
interface. J. Fluid Mech. 26 607–637.
[41] Henderson, L. F. 1989 On the refraction of shock waves. J. Fluid Mech.
198 365–386.
[42] Henderson, L. F. 1991 On the refraction of shock waves at a slow-fast gas
interface. J. Fluid Mech. 224 1–27.
[43] http://www.ittvis.com/ProductServices/IDL/RecentReleases.aspx.
[44] Jacobs, C, Poedts, S., van der Holst, B., Chané, E. 2005 On the
effect of the background wind on the evolution of interplanetary shock waves.
Astron. Astrophys. 430 1099–1107.
148 BIBLIOGRAPHY

[45] Jahn, R. G. 1956 The refraction of shock waves at a gaseous interface.


J. Comp. Phys. 1 457–489.
[46] Jeffrey, A. & Taniuti, T. 1964 Nonlinear Wave Propagation. Academic
Press, New York.
[47] Kennel, C. F., Blanford, R. D. & Coppi, P. 1989 MHD intermediate
shock discontinuities. I - Rankine-Hugoniot conditions. J. Plasma Phys. 42,
219–319.
[48] Keppens, R., Tóth, G. & Botchev, M. A., van den Ploeg, A. 1999
Implicit and Semi-Implicit Schemes: algorithms Intern. Journ. for Numer.
Meth. in Fluids 30, 335–352.
[49] Keppens, R., Nool, M., Tóth, G. & Goedbloed, H. 2003 Adaptive
Mesh Refinement for conservative systems: multi-dimensional efficiency
evaluation. J. Comp. Phys. 153, 317–339.
[50] Kreeft, J. J. & Koren, B. 2010 A new formulation of Kapila’s five-
equation model for compressible two-fluid flow and its numerical treatment.
J. Comp. Phys. 229, 6220–6242.
[51] Kifonidis, K., Plewa, T., Scheck, L., Janka, H.-Th. & Müller, E.
2006 Nonspherical core collapse supernovae. Astr. Astroph. 453, 661–678.
[52] Langmuir & Jones 1927 The Characteristics of Tungsten Filaments as
Functions of Temperature. G. E. Rev. 30 310.
[53] Lax, P. D. 1957 Hyperbolic System of Conservation Laws II. Comm. Pure
Appl. Math. 10, 537–566.
[54] Leveque, R. J. 1992 Numerical Methods for conservation laws. Springer
[55] Liepmann, H. W. & Roshko, A. 1957 Elements of Gasdynamics. John
Wiley & Sons, Inc.
[56] Liberman, M. A. & Velikhovich, A. L. 1986 Physics of Shock Waves in
Gases and Plasmas. Springer
[57] Maxwell, J. C. 1865, A dynamical theory of the electromagnetic field
Philosophical Transactions of the Royal Society of London 155 459–512.
[58] Meshkov, E. E. 1969 Instability of the interface of two gases accelerated by
a shock wave. Fluid Dynamics 4 101–104.
[59] Mulder, W., Osher, S. & Sethian, J. A. 1992 Computing interface
motion in compressible gas dynamics. J. Comp. Phys. 100 209–228.
BIBLIOGRAPHY 149

[60] Myong, R. S. & Roe, P. L. 1997 Shock waves and rarefaction waves in
magnetohydrodynamics. Part 1. A model system. J. Plasma. Phys. 58, 485.
[61] Myong, R. S. & Roe, P. L. 1997 Shock waves and rarefaction waves in
magnetohydrodynamics. Part 2. The MHD system. J. Plasma. Phys. 58, 521.
[62] von Neumann, J. 1943 Collected Works, Vol. 6. Permagon (1963)
[63] von Neumann, J. 1943 Progress report on the theory of shock waves.
National Defence Research Committee, Division 8 1140.
[64] Nouragliev, R. R., Sushchikh, S. Y., Dinh, T. N. & Theofanous,
T. G. 2005 Shock wave refraction patterns at interfaces. Internat. J.
Multiphase Flow 31, 969–995.
[65] Oron, D., Sadot, O., Srebro, Y., Rikanati, A., Yedvab, Y., Alon,
U., Erez, L., Erez, G., Ben-Dor, G, Levin, L. A., Ofer, D &
Shvarts, D. 1999 Studies in the nonlinear evolution of the Rayleigh-Taylor
and Richtmyer-Meshkov instabilities and their role in inertial confinement
fusion. Laser and Particle Beams 17 465–475.
[66] http://www.paraview.org.
[67] http://www.plasma-universe.com/Plasma-Universe.com.
[68] Powell, K. G., Roe, P. L., Linde, T. J., Gombosi, T. I.,
De Zeeuw, D. L. 1999 A solution-adaptive upwind scheme for ideal
magnetohydrodynamics. J. Comput. Phys. 154 284.
[69] Richtmyer, R. D. 1960 Taylor instability in shock acceleration of
compressible fluids. Commun. Pure Appl. Maths. 13 297–319.
[70] Roe, P. L. 1981 Approximate Riemann solvers, parameter vectors, and
difference schemes. J. Comput. Phys. 43 357.
[71] Roe, P. L. & Balsara, D. S. 1996 Notes on the eigensystem of
magnetohydrodynamics. J. Appl. Math. 56 57.
[72] Rupert, V. 1992 Shock-Interface interaction: current research on the
Richtmyer-Meshkov Problem. Proc. Intl. Symp. Shock Waves 18 83–94.
[73] http://www.globalsecurity.org/wmd/ops/sailor-hat.htm.
[74] http://kauscience.k12.hi.us/˜ted/Craters/SailorHat.
[75] Sadot, O., Yosef-Hai, A., Oron, D. & Rikanati, A. 2001
Dependence of the Richtmyer-Meshkov instability on the Atwood number ond
dimensionality: theory and experiments. Prc. SPIE 4183, 798
150 BIBLIOGRAPHY

[76] Saliba, G. 1987 Theory and observation in islamic astronomy: the work of
ibn al-shāţir of Damascus. Journ. history of astronomy 18, 35
[77] Samtaney, R., Ray, J. & Zabusky, N. J. 1998 Baroclinic Circulation
Generation on Shock Accelerated Slow/fast Gas Interfaces. 1998 Phys. Fluids
10 1217–1230.
[78] Samtaney, R. 2003 Suppression of the Richtmyer-Meshkov instability in the
presence of a magnetic field. Phys. Fluids 15, L53–L56.
[79] Stringari, S. & Wilson, R. R. 2000 Romagnosi and the discovery of
electromagnetism. ( Rend. Fis. Acc. Lincei) 9 115–136
[80] Sturtevant, B. 1987 Shock Tubes and Waves. VCH Verlag, Berlin
[81] Taub, A. H. 1947 Refraction of plane shock waves. Phys. Review 72 51–59
[82] Taylor, G. I. 1950 The instability of liquid surfaces when accelerated in a
direction perpendicular to their plane. Proc. Roy. Soc. A210 192–196
[83] www.tecplot.com
[84] Todd, L. 1965 Evolution of switch-on and switch-off shocks in a gas of finite
electrical conductivity. J. Fluid Mech. 24, 597–608.
[85] Toro, E. F. 1999 Riemann Solvers and Numerical Methods for Fluid
Dynamics. Berlin: Springer Verlag.
[86] Tóth, G. & Odstrčil, D. 1996 Comparison of some flux corrected transport
and total variation diminishing numerical schemes for hydrodynamic and
magnetohydrodynamic problems. J. Comput. Phys. 128 (1), 82–100.
[87] Tóth, G. 1996 General Code for Modeling MHD flows on Parallel Computers:
Versatile Advection Code. Astroph. Lett.& Comm. 34, 245.
[88] Tóth, G. 1997 The LASY Preprocessor and its Application to General Multi-
Dimensional Codes. J. Comput. Phys. 138, 981.
[89] Tóth, G., Keppens, R., Botchev, M. A. 1998 Implicit and semi-implicit
schemes in the Versatile Advection Code: numerical tests. A. & A. 332, 1159–
1170.
[90] http://www.ipc.uni-karlsruhe.de/mol/405.php
[91] Torrilhon, M. 2002 Exact Solver and Uniqueness Conditions for Riemann
Problems of Ideal Magnetohydrodynamics, Technical Report 2002-06, SAM,
ETH Zurich.
[92] http://grid.engin.umich.edu/˜gtoth/VAC/.
BIBLIOGRAPHY 151

[93] van der Holst, B. & Keppens, R. 2007 Hybrid block-AMR in cartesian
and curvilinear coordinates: MHD applications. J. Comp. Phys. 26, 925–946.
[94] http://www.llnl.gov/VisIt.
[95] http://www.vtk.org.
[96] Kitware, Inc. 2006 VTK User’s Guide. Kitware, Inc.
[97] Whang, Y. C., Zhou J., Lepping, R. P., Szabo, A., Fairfield, D.,
Kukobun, S., Ogilvie, K. W., Fitzenreiter R. 1998 Double discontinuity
: A compound structure of slow shock and rotational discontinuity J. Geophys.
Res. 103 6513–6520.
[98] Wheatley, V., Pullin, D. I. & Samtaney, R. 2005 Suppression of the
Richtmyer-Meshkov instability in the presence of a magnetic field. J. Fluid
Mech. 552, 179–217.
[99] Wu, C. C. 1987 On MHD intermediate shocks. Geophys. Res. Lett. 14, 668-
671.
[100] Wu, C. C. 1988 The MHD intermediate shock interaction with an
intermediate wave: Are intermediate shocks physical?. J. Geophys. Res.
93(A2), 987–990.
[101] Wu, C. C. 1990 Formation, structure, and stability of MHD intermediate
shocks. J. Geophys. Res. 95(A6), 8149–8175
[102] Yee, H.C. 1989 A class of high-resolution explicit and implicit shock-
capturing methods. NASA TM-101088
[103] Yee, H. C. & Sjögreen, B. 2007 Simulations of Richtmyer-Meshkov
instability by sixth-order filter methods. Shock Waves 17, 185–193.
[104] Zabusky, N. J. 1999 Vortex paradigm for accelerated inhomogeneous flows:
visiometrics for the Rayleigh-Taylor and the Richtmyer-Meshkov environments.
Annu. Re. Fluid Mech. 31, 495–536.
Curriculum Vitae

Education
• 1991-1997
Wetenschappen-wiskunde (8u), VIA Tienen
• 1997-2001
Licentiaat Wiskunde, K.U.Leuven

List of Scientific Contributions

Publications
• P. Delmont & R. Keppens, ‘An exact solution strategy for regular shock
refraction at a density discontinuity’, 2008, ECA 32D, D2.008
• P. Delmont, R. Keppens, & B. van der Holst, ‘An exact Riemann solver
based solution for regular shock refraction’, 2009, J. Fluid Mech. 627, 33-53.
• P. Delmont & R. Keppens, ’Shock refraction in ideal MHD’, 2010, Journal
of Physics: Conference Series, 216, 012007
• P. Delmont & R. Keppens, ’Parameter regimes for slow, intermediate and
fast MHD shocks’, 2010, J. Plasma Phys., doi:10.1017/S0022377810000115
• P. Delmont & R. Keppens, ’Limiting parameter values for switch-on
and switch-off shocks in ideal magnetohydrodynamics’, 2010, acta technica,
accepted
• P. Delmont & R. Keppens, ’Rankine-Hugoniot conditions and intermdiate
MHD shocks’, 2010, ECA, submitted

153
154 SCIENTIFIC CONTRIBUTIONS

Poster Contributions
• P. Delmont, R. Keppens, B. van der Holst & Z. Meliani, ’Grid-adaptive
approaches for computing magnetized plasma dynamics’, poster at ‘JETSET
school on Numerical MHD and instabilities’, Torino, Italy, 8-13 January
2007.
• P. Delmont & R. Keppens, ’Planar Richtmyer Meshkov instabilities in
MHD’, poster at ’32th conference of the Dutch and Flemish Numerical
Analysis Communities’, Woudschoten, The Netherlands, 3-5 october 2007.
• P. Delmont & R. Keppens, ’Supression of the Richtmeyer-Meshkov instabil-
ity in MHD’, poster at ’internal kick-off meeting for the Leuven Mathematical
Modeling and Computational Science Centre (LMCC)’, Leuven, Belgium, 24
april 2008.
• P. Delmont & R. Keppens, ’An exact Riemann solver based solution for
regular shock refraction’, poster at ’35th EPS plasma physics conference’,
Hersonissos, Crete, Greece, 9-13 june 2008.
• P. Delmont & R. Keppens, ’Supression of the Richtmeyer-Meshkov
instability in MHD’, poster at ’33th conference of the Dutch and Flemish
Numerical Analysis Communities’, Woudschoten, The Netherlands, 8-10
october 2008. (Winner of the Poster Prize)
• R. Keppens, Z. Meliani, A. J. van Marle, P. Delmont & A. Vlasis, ’
Multi-scale simulations with MPIAMRVAC’, Poster at ’LMCC workshop on
Modeling and simulations of multi-scale and multi-physics systems’, Leuven,
Belgium, 8-9 september 2009.
• P. Delmont & R. Keppens, ’Parameter ranges for intermediate MHD
shocks’, poster at ’34th conference of the Dutch and Flemish Numerical
Analysis Communities’, Woudschoten, The Netherlands, 7-9 october 2009
• P. Delmont & R. Keppens, ’Parameter regimes for intermediate MHD
shocks’, poster at ’37th EPS plasma physics conference’, Dublin, Ireland,
21-25 june 2010

Lectures & Seminars


• P. Delmont, ’The Richtmyer-Meshkov Instability in 2D hydrodynamical
flows’, seminar at Centre for Plasma Astrophysics, Leuven, Belgium, 13
december 2007
• P. Delmont & R Keppens, ’An exact Riemann-solver-based solution for
regular shock refraction’, oral at Belgian Physical Society General scientific
meeting 2009, Hasselt, Belgium, 1 april 2009.
BIBLIOGRAPHY 155

• P. Delmont & R. Keppens, ’An exact Riemann-solver strategy for regular


shock refraction’, oral at BIFD 2009, Nottingham, UK, 10-13 augustus 2009.
• P. Delmont, ’Parameter ranges for intermediate MHD shocks’, seminar at
Centre for Plasma Astrophysics, Leuven, Belgium, 16 december 2009.
• P. Delmont & R. Keppens, ’Parameter regimes for intermediate MHD
shocks’, oral at ’24th Symposium on Plasma Physics and Technology’,
Prague, Czech Republic, 14-17 june 2010.

You might also like