Download momentum-space dynamics of runaway electrons

Document related concepts

Quantum electrodynamics wikipedia , lookup

Quantum vacuum thruster wikipedia , lookup

Renormalization wikipedia , lookup

Aharonov–Bohm effect wikipedia , lookup

T-symmetry wikipedia , lookup

Electromagnetism wikipedia , lookup

Condensed matter physics wikipedia , lookup

Nuclear physics wikipedia , lookup

Introduction to gauge theory wikipedia , lookup

Electrical resistivity and conductivity wikipedia , lookup

Lepton wikipedia , lookup

Elementary particle wikipedia , lookup

Density of states wikipedia , lookup

History of subatomic physics wikipedia , lookup

Plasma (physics) wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Atomic theory wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Transcript
THESIS FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY
MOMENTUM-SPACE
DYNAMICS OF RUNAWAY
ELECTRONS IN PLASMAS
Adam Stahl
Department of Physics
Chalmers University of Technology
Gothenburg, Sweden, 2017
MOMENTUM-SPACE DYNAMICS OF RUNAWAY ELECTRONS IN PLASMAS
Adam Stahl
© Adam Stahl, 2017
ISBN 978-91-7597-532-0
Doktorsavhandlingar vid Chalmers tekniska högskola
Ny serie nr 4213
ISSN 0346-718X
Subatomic and Plasma Physics
Department of Physics
Chalmers University of Technology
SE-412 96 Gothenburg
Sweden
Telephone +46-(0)31-772 10 00
Printed by
Reproservice
Chalmers tekniska högskola
Gothenburg, Sweden, 2017
MOMENTUM-SPACE DYNAMICS OF RUNAWAY ELECTRONS IN PLASMAS
Adam Stahl
Department of Physics
Chalmers University of Technology
Abstract
Fast electrons in a plasma experience a friction force that decreases with increasing particle speed, and may therefore be continuously accelerated by sufficiently strong electric fields. These so-called runaway electrons may quickly
reach relativistic speeds. This is problematic in tokamaks – devices aimed at
producing sustainable energy through the use of thermonuclear fusion reactions – where runaway-electron beams carrying strong currents may form. If
the runaway electrons deposit their kinetic energy in the plasma-facing components, these may be seriously damaged, leading to long and costly device
shutdowns.
Crucial to the runaway phenomenon is the behavior of the runaway electrons
in two-dimensional momentum space. The interplay between electric-field
acceleration, collisional momentum-space transport, and radiation reaction
determines the dynamics and the growth or decay of the runaway-electron
population. In this thesis, several aspects of this interplay are investigated,
including avalanche multiplication rates, synchrotron radiation reaction, modifications to the critical electric field for runaway generation, rapidly changing
plasma parameters, and electron slide-away. Two numerical tools for studying electron momentum-space dynamics, based on an efficient solution of the
kinetic equation, are presented and used throughout the thesis. The spectrum of the synchrotron radiation emitted by the runaway electrons – a useful
diagnostic for their properties – is also studied.
It is found that taking the electron distribution into account properly is crucial
for the interpretation of synchrotron spectra; that a commonly used numerical avalanche operator may either overestimate or underestimate the runawayelectron growth rate, depending on the scenario; that radiation reaction modifies the critical electric field, but that this modification often is small compared
to other effects; that electron slide-away can occur at significantly weaker electric fields than expected; and that collisional nonlinearities may be significant
for the evolution of runaway-electron populations in disruption scenarios.
Keywords: fusion-plasma physics, tokamak, runaway electrons, synchrotron
radiation, critical electric field, slide-away, non-linear collision operator
i
Publications
This thesis is based on the work contained in the following papers:
A
A. Stahl, M. Landreman, G. Papp, E. Hollmann, and T. Fülöp,
Synchrotron radiation from a runaway electron distribution in tokamaks,
Physics of Plasmas 20, 093302 (2013).
http://dx.doi.org/10.1063/1.4821823
http://arxiv.org/abs/1308.2099
B
M. Landreman, A. Stahl, and T. Fülöp,
Numerical calculation of the runaway electron distribution function and associated synchrotron emission,
Computer Physics Communications 185, 847-855 (2014).
http://dx.doi.org/10.1016/j.cpc.2013.12.004
http://arxiv.org/abs/1305.3518
C
A. Stahl, O. Embréus, G. Papp, M. Landreman, and T. Fülöp,
Kinetic modelling of runaway electrons in dynamic scenarios,
Nuclear Fusion 56, 112009 (2016).
http://dx.doi.org/10.1088/0029-5515/56/11/112009
http://arxiv.org/abs/1601.00898
D
A. Stahl, E. Hirvijoki, J. Decker, O. Embréus, and T. Fülöp,
Effective critical electric field for runaway electron generation,
Physical Review Letters 114, 115002 (2015).
http://dx.doi.org/10.1103/PhysRevLett.114.115002
http://arxiv.org/abs/1412.4608
E
A. Stahl, M. Landreman, O. Embréus, and T. Fülöp,
NORSE: A solver for the relativistic non-linear Fokker-Planck equation for
electrons in a homogeneous plasma,
Computer Physics Communications 212, 269-279 (2017).
http://dx.doi.org/10.1016/j.cpc.2016.10.024
http://arxiv.org/abs/1608.02742
F
A. Stahl, O. Embréus, M. Landreman, G. Papp, and T. Fülöp,
Runaway-electron formation and electron slide-away in an ITER post-disruption
scenario.
Journal of Physics: Conference Series 775, 012013 (2016).
http://dx.doi.org/10.1088/1742-6596/775/1/012013
http://arxiv.org/abs/1610.03249
iii
Statement of contribution
In Papers A, C, E and F, I performed all simulations and associated analysis,
and produced all the figures. I wrote all the text in Papers C, E and F, and
a majority of that in Paper A. In addition I did all of the programming associated with the code SYRUP used in Paper A and the tool NORSE described
in Paper E, as well as most of the programming related to the capabilities
of CODE described in Paper C. I performed the majority of the analytical
calculations required for Papers A, C and E. In Paper B, I was mainly responsible for the work presented in Section 6, but in addition contributed to
the remainder of the text. To a lesser extent, I was also involved in the development of the tool CODE described in the paper. In Paper D, I performed
the numerical simulations, together with the associated implementation and
analysis, produced the majority of the figures and text, and did some of the
analytical calculations.
iv
Additional publications, not included in the thesis
G
O. Embréus, A. Stahl, and T. Fülöp,
Effect of bremsstrahlung radiation emission on fast electrons in plasmas,
New Journal of Physics 18, 093023 (2016).
http://dx.doi.org/10.1088/1367-2630/18/9/093023
http://arxiv.org/abs/1604.03331
H
J. Decker, E. Hirvijoki, O. Embréus, Y. Peysson, A. Stahl, I. Pusztai,
and T. Fülöp,
Numerical characterization of bump formation in the runaway electron
tail,
Plasma Physics and Controlled Fusion 58, 025016 (2016).
http://dx.doi.org/10.1088/0741-3335/58/2/025016
http://arxiv.org/abs/1503.03881
I
E. Hirvijoki, I. Pusztai, J. Decker, O. Embréus, A. Stahl, and T. Fülöp,
Radiation reaction induced non-monotonic features in runaway electron
distributions,
Journal of Plasma Physics 81, 475810502 (2015).
http://dx.doi.org/10.1017/S0022377815000513
http://arxiv.org/abs/1502.03333
J
O. Embréus, S. Newton, A. Stahl, E. Hirvijoki, and T. Fülöp,
Numerical calculation of ion runaway distributions,
Physics of Plasmas 22, 052122 (2015).
http://scitation.aip.org/content/aip/journal/pop/22/5/10.1063/1.4921661
http://arxiv.org/abs/1502.06739
K
G. I. Pokol, A. Kómár, A. Budai, A. Stahl, and T. Fülöp,
Quasi-linear analysis of the extraordinary electron wave destabilized by
runaway electrons,
Physics of Plasmas 21, 102503 (2014).
http://dx.doi.org/10.1063/1.4895513
http://arxiv.org/abs/1407.5788
v
Conference contributions
L
A. Tinguely, R. Granetz, and A. Stahl, Analysis of Runaway Electron
Synchrotron Emission in Alcator C-Mod, Proceedings of the 58th Annual Meeting of the APS Division of Plasma Physics 61, 18, TO4.00007
(2016). http://meetings.aps.org/Meeting/DPP16/Session/TO4.7
M
C. Paz-Soldan, N. Eidietis, D. Pace, C. Cooper, D. Shiraki, N. Commaux, E. Hollmann, R. Moyer, R. Granetz, O. Embréus, T. Fülöp, A.
Stahl, G. Wilkie, P. Aleynikov, D. P. Brennan, and C. Liu,
Synchrotron and collisional damping effects on runaway electron distributions, Proceedings of the 58th Annual Meeting of the APS Division of
Plasma Physics 61, 18, CO4.00010 (2016).
http://meetings.aps.org/Meeting/DPP16/Session/CO4.10
N
O. Ficker, J. Mlynar, M. Vlainic, V. Weinzettl, J. Urban, J. Cavalier,
J. Havlicek, R. Panek, M. Hron, J. Cerovsky, P. Vondracek, R. Paprok,
J. Decker, Y. Peysson, O. Bogar, A. Stahl, and the COMPASS Team,
Long slide-away discharges in the COMPASS tokamak,
Proceedings of the 58th Annual Meeting of the APS Division of Plasma
Physics 61, 18, GP10.00101 (2016).
http://meetings.aps.org/Meeting/DPP16/Session/GP10.101
O
Y. Peysson, G. Anastassiou, J.-F. Artaud, A. Budai, J. Decker, O. Embréus, O. Ficker, T. Fülöp, K. Hizanidis, Y. Kominis, T. Kurki-Suonio,
P. Lauber, R. Lohner, J. Mlynar, E. Nardon, S. Newton, E. Nilsson, G.
Papp, R. Paprok, G. Pokol, F. Saint-Laurent, C. Reux, K. Sarkimaki,
C. Sommariva, A. Stahl, M. Vlainic, and P. Zestanakis,
A European Effort for Kinetic Modelling of Runaway Electron Dynamics, Theory and Simulation of Disruptions Workshop (2016).
http://tsdw.pppl.gov/Talks/2016/Peysson.pdf
P
T. Fülöp, O. Embréus, A. Stahl, S. Newton, I. Pusztai, and G. Wilkie,
Kinetic modelling of runaways in fusion plasmas, Proceedings of the
26th IAEA Fusion Energy Conference, Kyoto, Japan, TH/P4–1 (2016).
Q
O. Embréus, A. Stahl, and T. Fülöp,
Effect of bremsstrahlung radiation emission on fast electrons in plasmas,
Europhysics Conference Abstracts 40A, O2.402 (2016).
http://ocs.ciemat.es/EPS2016PAP/pdf/O2.402.pdf
vi
R
A. Stahl, O. Embréus, E. Hirvijoki, I. Pusztai, J. Decker, S. Newton,
and T. Fülöp, Reaction of runaway electron distributions to radiative
processes, Proceedings of the 57th Annual Meeting of the APS Division
of Plasma Physics 60, 19, PP12.00103 (2015).
http://meetings.aps.org/link/BAPS.2015.DPP.PP12.103
S
O. Embréus, A. Stahl, and T. Fülöp, Conservative large-angle collision
operator for runaway avalanches, Proceedings of the 57th Annual Meeting of the APS Division of Plasma Physics 60, 19, PP12.00107 (2015).
http://meetings.aps.org/link/BAPS.2015.DPP.PP12.107
T
S. Newton, O. Embréus, A. Stahl, E. Hirvijoki, and T. Fülöp, Numerical
calculation of ion runaway distributions, Proceedings of the 57th Annual
Meeting of the APS Division of Plasma Physics 60, 19, CP12.00118
(2015). http://meetings.aps.org/link/BAPS.2015.DPP.CP12.118
U
I. Pusztai, E. Hirvijoki, J. Decker, O. Embréus, A. Stahl, and T. Fülöp,
Non-monotonic features in the runaway electron tail,
Europhysics Conference Abstracts 39E, O3.J105 (2015).
http://ocs.ciemat.es/EPS2015PAP/pdf/O3.J105.pdf
V
G. Papp, A. Stahl, M. Drevlak, T. Fülöp, P. Lauber, and G. Pokol,
Towards self-consistent runaway electron modelling,
Europhysics Conference Abstracts 39E, P1.173 (2015).
http://ocs.ciemat.es/EPS2015PAP/pdf/P1.173.pdf
W
O. Embréus, S. Newton, A. Stahl, E. Hirvijoki, and T. Fülöp,
Numerical calculation of ion runaway distributions,
Europhysics Conference Abstracts 39E, P1.401 (2015).
http://ocs.ciemat.es/EPS2015PAP/pdf/P1.401.pdf
X
A. Stahl, E. Hirvijoki, M. Landreman, J. Decker, G. Papp, and T. Fülöp,
Effective critical electric field for runaway electron generation,
Europhysics Conference Abstracts 38F, P2.049 (2014).
http://ocs.ciemat.es/EPS2014PAP/pdf/P2.049.pdf
Y
G. Pokol, A. Budai, J. Decker, Y. Peysson, E. Nilsson, A. Stahl, A. Kómár,
and T. Fülöp, Interaction of oblique propagation extraordinary electron
waves and runaway electrons in tokamaks,
Europhysics Conference Abstracts 38F, P2.042 (2014).
http://ocs.ciemat.es/EPS2014PAP/pdf/P2.042.pdf
vii
Z
viii
A. Stahl, M. Landreman, T. Fülöp, G. Papp, and E. Hollmann,
Synchrotron radiation from runaway electron distributions in tokamaks,
Europhysics Conference Abstracts 37D, P5.117 (2013).
http://ocs.ciemat.es/EPS2013PAP/pdf/P5.117.pdf
Contents
Abstract
i
Publications
iii
1 Introduction
1
2 Runaway-electron generation and loss
2.1 The runaway region of momentum space . . . . . . . . . . . . .
2.2 Runaway-generation mechanisms . . . . . . . . . . . . . . . . .
2.3 Damping and loss mechanisms for runaways . . . . . . . . . .
7
7
14
17
3 Simulation of runaway-electron momentum-space dynamics
3.1 The kinetic equation . . . . . . . . . . . . . . . . . . . .
3.2 Collision operator . . . . . . . . . . . . . . . . . . . . .
3.3 Avalanche source term . . . . . . . . . . . . . . . . . .
3.4 CODE . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5 NORSE . . . . . . . . . . . . . . . . . . . . . . . . . . .
21
21
24
26
29
30
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
4 Synchrotron radiation
33
4.1 Emission and power spectra . . . . . . . . . . . . . . . . . . . . 34
4.2 Radiation-reaction force . . . . . . . . . . . . . . . . . . . . . . 40
5 Nonlinear effects and slide-away
45
5.1 Ohmic heating . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Electron slide-away . . . . . . . . . . . . . . . . . . . . . . . . 46
6 Concluding remarks
49
6.1 Summary of the included papers . . . . . . . . . . . . . . . . . 49
6.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Bibliography
55
Included papers (A-F)
69
ix
Acknowledgments
The life of a doctoral student has sometimes been described as a stumbling
journey down a seemingly endless dark tunnel. Not so in the Plasma Theory
Group at Chalmers, where the guidance of Professor Tünde Fülöp provides a
map, as well as the light by which to read it. I am fortunate and thankful to
have had her support throughout my doctoral studies. In fact, I wish every
doctoral student the privilege of having Tünde as their supervisor (although
I recognize the logistical nightmare this would entail)!
I also owe a lot to my co-supervisors Dr. Matt Landreman and Dr. István
Pusztai. Matt in particular has been involved on some level in most projects
I have undertaken during these years and have identified (and often solved)
many a problem I didn’t even know I had. István has provided the stability
(and ability) closer to home, always contributing knowledgeable input on all
things plasma physics.
The plasma theory group would be nothing without its current and former
members, of which there are many. Thanks for the marvelous physics and
non-physics discussions, salty lunches, pizza parties and ski trips. I will miss
you all. In particular I would like to thank my long-time roommate and
close collaborator Ola Embréus for his remarkable insight and investigative
journalism, and my short-time roommate and long-time friend Dr. Gergely
Papp for his eagle eye and grounded feet.
Finally; the reason I leave the office in the evening with a smile on my face.
Thank you Hedvig for providing distractions and focus, cookies and salads,
insight and bogus, odd meters and ballads. Without these, no map in the
world would see me through to the end of the tunnel.
Thank you all
xi
1 Introduction
In plasma physics, many interesting phenomena occur that are outside of
our everyday experience. One of these is the generation of so-called runaway
electrons (or simply runaways) – electrons that under certain conditions are
continuously accelerated by electric fields [1, 2]. The dynamics of the process
is such that the runaways quickly reach relativistic energies; they move with
speeds very close to that of light. The study of runaway electrons therefore
combines two fascinating areas of physics: Einstein’s special relativity [3], and
plasma physics; giving rise to interesting dynamics (as well as complicated
mathematics). As we shall see, runaways appear in a variety of atmospheric,
astrophysical and laboratory contexts.
Apart from their intrinsic interest, these highly energetic particles are also a
cause for concern in the context of fusion-energy experiments [4]. Generating
electric power using controlled thermonuclear fusion reactions is a promising
concept for a future sustainable energy source [5–7], but stable and controllable
operating conditions are required for a successful fusion power plant. The
presence of runaway electrons in the plasmas of fusion reactors under certain
circumstances is one of the main remaining hurdles on the road to realization
of fusion power production [8], as the runaways have the potential to severely
damage the machine when they eventually leave the plasma and strike the
wall [9]. In order to accurately assess the frequency of such events, as well
as the resulting damage in a given situation, there is a great need to improve
the understanding of the mechanisms that generate and suppress runaway
electrons, and to better describe their dynamics [10].
In order for runaways to be generated, a comparatively long-lived electric
field is needed. Due to the natural tendency of the plasma particles to rearrange in order to screen out such fields, they are not normally present in
unmagnetized plasmas. However in certain situations, for instance if a current running through the plasma changes quickly or if the magnetic field lines
in a magnetized plasma reconnect, an electric field is induced which may be
sufficient to lead to runaway formation. Runaway electrons do form in atmospheric plasmas – they have been linked to for instance lightning discharges
1
Chapter 1.
Introduction
Figure 1.1: A tokamak plasma (pink), together with various common terms
and concepts.
[11], impulsive radio emissions [12], and terrestrial gamma-ray flashes [13] –
and in the mesosphere [14]. In astrophysical plasmas, they are expected to
form in for instance solar flares [15] and large-scale filamentary structures in
the galactic center [16]. Under certain circumstances, other plasma species
may also run away. Both ion and positron runaway have been investigated in
recent work (see Refs. [17–19], as well as Paper J, not included in the thesis).
Our main interest in this thesis is however electron runaway in the context of
magnetic-confinement thermonuclear fusion.
The most common type of fusion device is called a tokamak (see for instance
[7, 20]). It uses strong magnetic fields to confine a plasma in which the fusion
reactions between hydrogen-isotope ions take place. The charged particles in
the plasma follow helical orbits (spirals) around the magnetic field lines due
to the Lorentz force [21, 22], and are thus (to a first approximation) prevented
from reaching the walls of the device. In a tokamak, the “magnetic cage” (and
thus the plasma) has the form of a torus (a doughnut), as shown in Fig. 1.1.
The torus shape can be thought of as being formed from a cylinder, bent
around so that its two ends connect. The direction along the axis of this
cylinder is referred to as toroidal (and the axis itself the magnetic axis), while
2
Introduction
the direction around the circumference of the cylinder is called poloidal and
its radius is the minor radius. The radius of the circle defined by the magnetic
axis is called the major radius, and the relation between the major and minor
radii is referred to as the aspect ratio. The plasma temperature and density
are maximized close to the magnetic axis, and this is also where the runaways
predominantly form. Due to the loop-like toroidal geometry, the magnetic
field acts as an “infinite racetrack” for the runaways, which make millions of
toroidal revolutions of the tokamak each second.
In order to achieve satisfactory plasma confinement, it is necessary to drive a
strong current in the plasma. The poloidal magnetic field induced by the current introduces a helical twist to the magnetic field lines, and it can be shown
that each field line covers a toroidal surface of constant pressure. The tokamak
plasma can therefore be viewed as being made up of a series of such nested flux
surfaces. The plasma current is generated using transformer action: a changing current is driven through a conducting loop interlocked with the tokamak
vessel, and the change in this current induces a voltage (the so-called loop voltage) which drives a current in the plasma. Since a hot plasma is a very good
conductor, the loop-voltage does not need to be particularly strong during normal operation (it is usually of order 1 V), and tokamak plasmas (discharges
or “shots”) are routinely maintained for several seconds, and sometimes for
several minutes or more. Inductive current drive does not enable continuous
(steady-state) operation, however, and the tokamak is fundamentally a pulsed
device (in the absence of auxiliary current-drive systems).
During the start-up of a tokamak discharge, a plasma is formed by the ionization of a gas. For this process, a strong electric field is usually needed.
Runaways may form in this situation [23–25], however their formation can
usually be avoided by maintaining a high enough gas/plasma density. The
case of a changing current is more problematic. Abrupt changes in plasma
current occur during so-called disruptions, in which the plasma becomes unstable, rapidly cools down due to a loss of confinement, and eventually terminates [8, 26, 27]. As the plasma cools, the resistivity increases drastically
(since it is proportional to T −3/2 ), and a large electric field is induced which
tries to maintain the current (in accordance with Lenz’s law [28]). Near the
magnetic axis, this field is often strong enough to lead to runaway generation,
and runaway beams in the center of the plasma have been observed during
disruptions in many tokamaks (for instance JET [29–31], DIII-D [32, 33], Alcator C-Mod [34], Tore Supra [35], KSTAR [36], COMPASS [37], ASDEX
Upgrade [38] and TCV [39]). Runaways can also be generated in so-called
sawtooth crashes [40], and even during normal stable operation if the density
3
Chapter 1.
Introduction
is low enough [39, 41, 42] (the accelerating field in this case is the normal
loop voltage). Some auxiliary plasma-heating schemes produce an elevated
tail in the electron velocity distribution which can lead to increased runaway
production, should a disruption occur [39, 43].
The runaways predominantly form close to the magnetic axis of the tokamak,
where the flux-surface radius is small compared to the major radius. Many
aspects of the fundamental runaway dynamics can therefore be studied in
the large–aspect-ratio limit where transverse spatial effects can be neglected.
Since the plasma is essentially homogeneous in the direction along the magnetic field, the spatial dependence can be neglected entirely, and for an understanding of the basic mechanisms it is sufficient to treat the runaway process
purely in momentum-space. As discussed in Sec. 3.1, one of the momentumspace coordinates (the angle describing the gyration around the field lines),
can be averaged over, reducing the problem to two momentum-space dimensions. These simplifications are done throughout this thesis, except for parts
of Paper A (where a radial dependence is included). In practice, the situation
is more complicated, however, and spatial effects are often important, as will
be discussed in Sec. 2.3. Several numerical tools that take magnetic-trapping
and radial diffusive-transport effects into account (such as LUKE [44–46] and
CQL3D [47, 48]) also exist.
The main reason for the interest in runaway research is that the runaways pose
a serious threat to tokamaks. During disruptions, a large fraction of the initial
plasma current (which is often several megaamperes) can be converted into
runaway current. The runaways are normally well-confined in the tokamak,
but a variety of mechanisms (such as instabilities or a collective displacement
of the entire runaway beam) can transport them out radially. Unless their
generation is successfully mitigated, or runaway-beam stability can be externally enforced, the runaways eventually escape the plasma and strike the wall
where they can destroy sensitive components or degrade the wall material [49].
In present-day tokamaks, runaways are a nuisance, but usually not a serious
threat (although there are exceptions, see for instance Ref. [31]). However,
the avalanche multiplication (see Sec. 2.2.3) of a primary runaway seed is predicted to scale exponentially with plasma current [50], and it is believed that
in future devices which will have a larger current (such as the International
Thermonuclear Experimental Reactor ITER [51, 52] and eventually commercial fusion reactors), the problem will be much more severe. In these devices,
disruptions can essentially not be tolerated at all and much effort is devoted
to research on runaway and disruption mitigation techniques [10, 33, 49, 53].
4
Introduction
This thesis focuses on the dynamics of the runaway electrons in momentum
space, their generation and loss, and the forces that affect them. Most of
the results presented herein are not specific to fusion plasmas or a tokamak
magnetic geometry, however we will make use of fusion-relevant plasma parameters throughout. The majority of the work described in this thesis was
conducted using two numerical tools developed as part of the thesis work:
CODE, described in Papers B and C; and NORSE, described in Paper E. With
these tools, all three major runaway-generation mechanisms (Dreicer generation, hot-tail generation and avalanche generation through knock-on collisions
– to be discussed in Sec. 2.2), as well as two important energy-loss mechanisms (synchrotron and bremsstrahlung radiation emission) can be studied in
detail. Synchrotron radiation is particularly important, and will be discussed
in Chapter 4. It is also the main subject of Papers A and D – as a diagnostic for the runaway distribution and as a damping mechanism for runaway
growth, respectively.
Let us now turn to discussing the basic mechanisms responsible for the runaway phenomenon, and the quantities characterizing the runaway dynamics.
5
2 Runaway-electron generation
and loss
In short, a runaway electron is an electron in a plasma which experiences a net
accelerating force during a substantial time – enough to give it a momentum
significantly larger than that of the electrons in the thermal population.
The accelerating force on the electron is supplied by an electric field E, so that
FE = −eE, where e is the elementary charge. Meanwhile, Coulomb interaction with the other particles in the plasma (commonly referred to as collisions)
introduces a friction force FC (v). The origin of the runaway phenomenon is
that FC,k , the component of the friction force parallel to the electric (or magnetic, if the plasma is magnetized) field, is a nonmonotonic function of the
particle velocity, with a maximum around the electron thermal speed (vth )
[54], as illustrated in Fig. 2.1. Therefore ∂FC,k /∂v < 0 for particles that are
faster than vth – the friction force on these particles decreases with increasing
particle velocity. The physical origin of this effect is that the faster particles
spend less time in the vicinity of other particles in the plasma; as the particle
speed increases, the impulse delivered to the fast particle in each encounter
decreases more rapidly than the number of encounters increases, leading to
a reduction in the friction [54]. This implies that if the accelerating force is
sufficient to overcome the friction at the current velocity v0 of the particle,
|FE | > FC,k (v0 ), it will be able to accelerate the particle for all v > v0 , i.e.
the particle will be continuously accelerated to relativistic energies – it will
run away – as long as the electric field persists.
2.1 The runaway region of momentum space
For an electric field of intermediate strength (Ec < E < Esa , with the critical
field Ec and slide-away field Esa to be defined in this section), there exist
some velocities for which the acceleration by the electric field overcomes the
7
Chapter 2.
Runaway-electron generation and loss
F
FC,||
vth
c
v
Figure 2.1: Friction force on an electron due to collisions, as a function of
its velocity (schematic).
collisional friction. These velocities constitute the runaway region Sr in velocity space, indicated in Fig. 2.2. If only collisional friction is considered,
this region is semi-infinite in momentum space, extending from some lowest
critical momentum pc to arbitrarily high momenta, due to the dependence of
the friction force on velocity, as discussed above. The particles that are not
in the runaway region are only to a lesser extent affected by the electric field,
since the collisions dominate their dynamics.
Due to the directivity of the electric field, the force balance is not homogeneous in velocity. Several other forces also affect the dynamics, in particular radiation-reaction forces associated with synchrotron and bremsstrahlung
emission. In these cases, the force balance is altered, ultimately preventing
the electrons from reaching arbitrarily high momenta (this is the motivation
for the use of the slightly vague definition of runaway at the beginning of
this chapter). How to determine the runaway region in these more general
situations is discussed in Sec. 2.1.3.
Furthermore, the picture is complicated by the fact that not only friction due
to Coulomb collisions (collisional slowing down) contributes to the dynamics.
The collisions also lead to diffusion of the electrons in velocity space: both parallel to the particle velocity and perpendicular to it (referred to as pitch-angle
scattering). The diffusion is caused by velocity-space gradients in the distribution of particle velocities (see Sec. 3.1), parallel and perpendicular to v for
the two effects, respectively. Pitch-angle scattering does not directly affect the
energy of the electron, but is important for the behavior in two-dimensional
velocity space (see for instance [54] for a comprehensive introduction to collisional phenomena). In general, the full evolution of the runaway population
can only be obtained using numerical simulations.
8
2.1 The runaway region of momentum space
|F|
eEsa
Fee,||
eE
eE>Fee,||
eEc
vth
vc
Sr
c
v
Figure 2.2: Forces corresponding to collisional friction against electrons
(Fee,k ), the critical field (Ec ) and the slide-away field (Esa ). The runaway
region in velocity space (Sr ) associated with the electric field E is shown as
the gray part of the velocity axis.
2.1.1 Critical electric field and critical momentum
The critical electric field for runaway electron generation, Ec , is the weakest
field at which runaway is possible, see Fig. 2.2. The accelerating force due
to Ec is simply equal (and opposite) to the sum of all the forces acting to
slow the
down,
P particle
P at the speed vmin where they are minimized: eEc =
min
F
(v)
=
i f,i
i Ff,i (vmin ). In the simplest case, the only retarding
force is due to collisions with electrons (due to the mass difference, the energy
lost by the electrons in collisions with ions is neglected, as are all other forces):
eEc = Fee,k (v = vmin ). In this case, it is easy to obtain an expression for Ec .
The friction force on a highly energetic electron is given by [54]:
Fee,k (v) =
1
me c3 νrel ,
v2
(2.1)
where v is the speed of the particle, c is the speed of light, me is the electron
rest mass and
νrel =
ne e4 ln Λ
4πε20 m2e c3
(2.2)
is the collision frequency for a highly relativistic electron. Here ne is the number density of electrons, ln Λ is the Coulomb logarithm (see for instance Refs.
9
Chapter 2.
Runaway-electron generation and loss
[54, 55]) and ε0 is the vacuum permittivity. The collision frequency is defined
such that 1/ν is (approximately) the average time for a particle to experience
a 90◦ deflection due to an accumulation of small-angle Coulomb interactions
(which are much more frequent than large-angle collisions in fusion plasmas).
The friction force in Eq. (2.1) is minimized as v → c (we have already mentioned that Fee,k is monotonically decreasing for large velocities)1 . We thus
have that the critical field is Ec = Fee,k (v → c)/e, or
Ec =
ne e3 ln Λ
me c
νrel =
,
e
4πε20 me c2
(2.3)
which was obtained by Connor and Hastie in 1975 [56]. As discussed in Paper D and Sec. 4.2.2, synchrotron radiation reaction leads to an increase in the
critical field as the minimum of the friction force is effectively raised. Since
the synchrotron radiation-reaction force vanishes along the parallel axis, this
is however an effect of dynamics in 2D momentum-space and cannot be easily
accounted for in the simple model considered here.
For any E > Ec , there exists some speed vc above which the electric field
overcomes the friction force. Particles with a velocity greater than this critical
speed will run away, and vc thus marks the lower boundary of the runaway
region (in the parallel direction), as illustrated in Fig. 2.2. It is customary
to study runaways in terms of momentum rather than velocity. The critical
momentum is a simple function of the electric field strength
p if expressed in
terms of the normalized momentum p = γv/c, where γ = 1/ 1 − v 2 /c2 is the
relativistic mass factor:
1
pc = p
,
E/Ec − 1
(2.4)
if the electron is assumed to move parallel
to the electric field [56]. Similarly,
p
the corresponding critical γ is γc = (E/Ec )/(E/Ec − 1).
2.1.2 Dreicer and slide-away fields
The critical field Ec corresponds to the field balancing the minimum of the
collisional friction force. The Dreicer field, ED [1, 2], on the other hand,
1 Like
much of plasma physics, this result assumes that the Coulomb logarithm ln Λ is a
constant. In reality, it is energy dependent and increases logarithmically with p for large
p. The minimum of the friction force is correspondingly found at v somewhat below c,
however the value of Ec is only moderately affected.
10
2.1 The runaway region of momentum space
approximately balances the p
maximum of the friction force, which is located
around v = vth , with vth = 2Te /me the electron thermal speed and Te the
electron temperature2 . The critical and Dreicer fields are related by the ratio
of the thermal energy to the electron rest energy:
ED =
ne e3 ln Λ
me c2
Ec =
.
Te
4πε20 Te
(2.5)
For fields stronger than the slide-away field Esa ' 0.214ED [1], the accelerating
force overcomes the friction at all particle velocities, and the whole electron
population thus runs away. This phenomenon is called slide-away [57].
In practice, the electric field in a fusion plasma is almost always much smaller
than the Dreicer and slide-away fields. Therefore runaway dynamics can usually be studied in the regime where Ec < E Esa is fulfilled, in which case
the runaways can be treated as a small perturbation to a velocity distribution
that is close to local thermal equilibrium (i.e. a Maxwellian). If the electric
field is comparable to Esa , however; the distribution will deviate strongly from
a Maxwellian shape. This will in turn lead to a reduction in the friction in the
bulk and a corresponding decrease in the slide-away field. Thus – as discussed
in Sec. 5.2 – even though E < Esa initially, a transition to slide-away can
quickly occur due to the distortion of the distribution caused by the strong
electric field.
2.1.3 General calculation of the runaway region
Let us now discuss how to define the runaway region in the full two-dimensional
momentum space. The two momentum-space dimensions are conveniently
parametrized by the coordinates (p, ξ), where p = γv/c is the normalized
momentum and ξ = p|| /p is the cosine of the particle pitch angle (which
characterizes the pitch of the helix that describes the particle orbit around
a magnetic field line). These coordinates are suitable for analytical as well
as numerical calculations, and will therefore be used throughout much of the
remainder of this thesis. In the definition of ξ, pk is the component of the
momentum parallel to the magnetic field and similarly p⊥ is the perpendicular component. The coordinates (pk , p⊥ ) are convenient for visualizing the
distribution and will therefore be used in several figures throughout the thesis.
Note also that the relativistic mass factor is related to p through γ 2 = p2 + 1.
2 It
is customary in plasma physics to let Te ≡ kB Te , so that the “temperature” actually is
the thermal energy, and to express it in eV.
11
Chapter 2.
Runaway-electron generation and loss
In general, the object of interest when studying runaway-electron dynamics
is the distribution function f of electron momenta, which will be thoroughly
introduced in Sec. 3.1. For the following discussion, we just note that the
runaways normally form a narrow tail in the distribution function, centered
around the parallel axis (i.e. at ξ ≈ 1), whereas the contours of the equilibrium
(Maxwellian) distribution form concentric half circles in the (pk , p⊥ )-plane, as
illustrated in Fig. 2.3.
The lower boundary of the runaway region in 2D momentum-space is often
called the separatrix, as it separate two regions with distinct dynamical properties. It is well-defined in the parallel direction where it takes the value p = pc
[56], however there are several ways to express its dependence on the pitch
angle. In the following discussion, an unambiguous separatrix in momentum
space is obtained by neglecting the effect of collisional diffusion.
A common definition of the separatrix divides momentum space into two regions based on whether the accelerating electric field or the friction from
Coulomb collisions dominate. The boundary between these regions is described by p = ps , with
−1
p2s (ξ) = (ξE/Ec − 1)
(2.6)
(see for instance Paper H). At the parallel axis (ξ = 1), we find ps (1) = pc , as
expected.
Equation (2.6) does not take into account the fact that in 2D, the electric field
can accelerate a particle into the runaway region, even though it experiences
a net slowing-down force (i.e. is not in the runaway region initially). This is
possible because of the anisotropy of ps , and the use of Eq. (2.6) leads to an
underestimation of the fraction of particles that will run away. A more pertinent definition can be obtained by looking at particle trajectories in phase
space [58]. The trajectory which terminates at ξ = 1 and p = pc is the separatrix, since particles on it neither end up in the bulk population nor reach
arbitrarily high energies. This trajectory is given by [59]
2
ps,traj (ξ) =
−1
ξ+1 E
−1 .
2 Ec
(2.7)
The two separatrices in a typical scenario are shown in Fig. 2.3. In many cases,
the runaways form a narrow beam close to the parallel axis, in which case ps
and ps,traj give similar results. In fact, in these cases an isotropic runaway
region (ps,iso (ξ) = pc ) also is a good approximation (as discussed in Paper C),
12
2.1 The runaway region of momentum space
Figure 2.3: Runaway-region separatrices ps , ps,traj and ps,iso at an electric
field E/Ec = 800, overlayed on contours of a Maxwellian distribution with
Te = 5.1 eV and ne = 5 · 1019 m−3 . F is the distribution f normalized to its
maximum value (F = f / max[f ]).
and such a separatrix has been included in the figure as well. In certain cases,
such as when hot-tail generation dominates (see Sec. 2.2.2), the details of the
separatrix are however of importance for the size of the runaway population.
The separatrices discussed so far are valid in the limit where the bulk of
the distribution is well described by a nonrelativistic Maxwellian, and when
including only friction due to Coulomb collisions. As is suggested by the results
in Paper D, however; synchrotron radiation reaction may have a significant
impact on the separatrix (see also Refs. [60–62]).
In general, the separatrix for an arbitrary electron distribution can be obtained
by considering the forces that affect a test particle:
dp
p
= FEp − FCp − FSyn
,
dt
dξ
ξ
= FEξ − FCξ − FSyn
,
dt
(2.8)
(2.9)
where the expressions for the force associated with the electric field, FEi (with
i
i ∈ {p, ξ}), and the synchrotron radiation-reaction force FSyn
are discussed in
13
Chapter 2.
Runaway-electron generation and loss
Sec. 3.1 and Sec. 4.2, respectively. The expressions for the collisional electronelectron friction FCi are given in Paper E (see also Sec. 5.1 of that paper).
Although other reaction forces (such as bremsstrahlung) may also contribute
to the force balance, here we include only the synchrotron radiation-reaction
force as it is the dominant contribution in most plasmas of interest. The
critical momentum in the parallel direction can be determined from dp/dt = 0
at ξ = 1, since the separatrix becomes purely perpendicular to the parallel axis
as ξ → 1. The separatrix can then be traced out by numerically integrating
the above equations from ξ = 1 to ξ = −1. In the appropriate limit, the result
agrees with ps,traj (ξ). In general, the separatrix depends on the distribution
through the terms FCi and should be updated as the distribution changes.
It should be noted that in certain situations, additional regions in momentum space emerge which cannot be characterized as either bulk or runaway
regions. The example of bump-on-tail formation induced by synchrotron or
bremsstrahlung emission, where relativistic electrons accumulate around a
certain multi-MeV energy, is discussed in Papers G, H and I (not included
in the thesis). The momentum space dynamics leading to the formation of
the bump is complicated and involves the interaction between acceleration,
pitch-angle scattering and subsequent synchrotron or bremsstrahlung emission
back-reaction, forming structures akin to convection cells in the high-energy
electron distribution [Paper H]. Such phenomena are not described by the
above definition for the separatrix (although generalized models exist that do
take them into account, see for instance [60, 63]).
2.2 Runaway-generation mechanisms
There are two main mechanisms for generating runaways, referred to as Dreicer [1, 2] and avalanche [50, 64–66] generation. In the former, initially thermal electrons become runaways by a gradual diffusion through momentum
space until they reach a velocity where they run away. Dreicer generation is
an example of a primary mechanism, as it generates runaways without the
need for a pre-existing fast population. Once some runaways exist, one of
them may impart a large fraction of its momentum to a thermal electron in a
single event, known as a knock-on collision. This generates a second runaway
if both electrons are in the runaway region after the collision. This process
is also called secondary runaway generation, as it requires the presence of a
seed, and leads to an exponential growth of the runaway population (hence
the name avalanche). In this section, we will look more closely at these two
14
2.2 Runaway-generation mechanisms
mechanisms. We will also discuss another primary runaway mechanism: hottail generation, which can dominate if the plasma temperature decreases on
a short timescale.
2.2.1 Dreicer generation
Due to momentum-space transport processes, new particles steadily diffuse
into the runaway region, increasing the runaway density. This is known as
Dreicer generation, and is caused by the gradient ∂f /∂p that develops across
the separatrix as particles in the runaway region are accelerated to higher
energies. An approximate expression [56, 67–69] for the steady-state growth
rate of the runaway population due to this effect is
"
#
r
dnr
1 + Zeff
1
−3(1+Zeff )/16
= Cne νth exp − −
,
(2.10)
dt
4
where = E/ED , nr is the runaway number density,
νth =
ne e4 ln Λ
3
4πε20 m2e vth
(2.11)
is the electron-electron collision frequency of thermal particles, C is a constant
of order unity [4, 56] (not determined by the analytical model), and Zeff is the
effective ion charge, which is a measure of the plasma composition (Zeff = 1
is a plasma consisting of pure hydrogen, or otherwise singly charged ions).
Equation 2.10 is valid when the distribution is close to a Maxwellian, i.e. when
is small; and for E Ec . Closer to the critical field, correction factors are
introduced in the exponents so that the growth rate vanishes for E ≤ Ec [56],
but these are neglected here for simplicity.
Note that the growth rate depends on (and is exponentially small in) E/ED ,
not E/Ec . This means that even if the field is significantly larger than Ec , the
runaway production rate may be very small if E ED . This effect, which is
in essence a temperature dependence, can partly explain recent observations
indicating that E/Ec & 10 is required for runaway acceleration [41, 42]. Its
importance is discussed and quantified in Paper D.
2.2.2 Hot-tail generation
Primary runaways can also be produced by processes other than momentumspace diffusion, for instance by highly energetic γ-rays through pair produc-
15
Chapter 2.
Runaway-electron generation and loss
tion, or in tritium decay (in fusion plasmas). In a typical fusion plasma, these
two processes are usually insignificant, however if the conditions are right they
may provide a sufficient seed for avalanche multiplication [70]. There is however an additional primary-runaway mechanism – hot-tail generation [47, 71]
– which relies on a rapid cooling of the plasma. If the plasma-cooling timescale
is significantly shorter than the collision time at which particles equilibrate,
electrons that initially constituted the high-energy part of the bulk distribution can remain as a drawn-out tail at the new lower temperature. This is
because they take a longer time to equilibrate, as their collision time is significantly longer than that of the slow particles. If an electric field is also present,
some of these tail electrons may belong to the runaway region of momentum
space and will therefore be accelerated.
Under certain circumstances, hot-tail generation can be the dominating runaway-generation mechanism – albeit for a short time – and may provide a
strong seed for multiplication by the avalanche mechanism. This is particularly the case in disruptions in tokamaks, where the plasma quickly loses
essentially all stored thermal energy due to a sudden degradation in confinement. Approximately, hot-tail generation dominates over Dreicer generation
in a disruption if
√
3/2
1 3 π µ0 ene vth,f qa R
(2.12)
νth,0 tT <
3
4
Ba
is fulfilled [59], where νth,0 is the initial thermal collision frequency, tT is the
temperature-decay time, µ0 is the vacuum permeability, vth,f is the thermal
speed at the final temperature, R is the major radius of the tokamak, and Ba
and qa are the magnetic field and so-called safety factor [20] on its magnetic
axis. This estimate was obtained by assuming the temperature drop to be
described by Te (t) = Te,0 (1 − t/tT )2/3 , with Te,0 the initial temperature.
The hot-tail mechanism is discussed in Paper C. See also Refs. [59, 72–74] for a
more in-depth discussion of hot-tail growth rates in various cooling scenarios.
2.2.3 Secondary generation
Secondary runaways are formed when existing runaways collide with thermal
electrons, if the collision imparts enough momentum to the thermal electron to
kick it into the runaway region while the incoming (primary) electron remains
a runaway itself [50, 64–66]. Such events are referred to as close, large-angle or
16
2.3 Damping and loss mechanisms for runaways
knock-on collisions, and are normally rare in a fusion plasma (their contribution to the collisional dynamics is a factor ln Λ smaller than that of small-angle
collisions). In the context of runaway generation they become important due
to the special characteristics of the runaway region, since once a particle is
a runaway it can quickly gain enough energy to cause knock-on collisions of
its own. For it to be able to contribute to the avalanche process, the kinetic
energy of the incoming runaway must be at least twice as large as the critical
energy: γ − 1 > 2(γc − 1).
The avalanche growth rate was calculated by Rosenbluth & Putvinski [50], who
also derived an approximate operator for avalanche generation (see Paper C
and Sec. 3.3 for a detailed discussion of avalanche operators and their influence
on runaway dynamics). In a cylindrical plasma, the growth rate takes the form
−1/2
4(Z + 1)2
(E − 1)
dnr
' nr νrel
1 − E −1 + 2 eff2
,
(2.13)
dt
cz ln Λ
cz (E + 3)
p
where E = E/Ec and cz = 3(Zeff + 5)/π. In the limit where E Ec and
Zeff = 1, this simplifies to
r
(E − 1)
dnr
π
'
nr νrel
.
(2.14)
dt
2
3 ln Λ
The growth rate is proportional to the runaway density nr , meaning that the
growth is exponential (hence the name avalanche). We also note that the
dependence on E is linear in Eq. (2.14), and nearly so in the more general
expression (2.13), whereas it is exponential in Eq. (2.10) for the Dreicer growth
rate. Therefore, avalanche generation tends to dominate for weak fields (as
long as there is some runaway population to start with), but for strong fields
primary generation becomes more important.
2.3 Damping and loss mechanisms for runaways
The discussion so far has focused on the interplay between the electric field and
elastic Coulomb collisions in a quiescent, homogeneous, fully ionized plasma.
In practice, runaway electrons do not reach arbitrarily high energies or persist indefinitely. Many processes contribute to the damping of their growth,
slowing them down, or transporting them out of the plasma.
Of particular importance when it comes to limiting the energy achieved by
the runaways are radiative processes: synchrotron-radiation emission due to
17
Chapter 2.
Runaway-electron generation and loss
(predominantly) the gyro motion, and fast-electron bremsstrahlung due to inelastic scattering off the much heavier ions. The emitted radiation takes away
momentum, and the electron must therefore lose a corresponding amount.
This radiation reaction effectively introduces an additional force which can
counteract the accelerating electric field. Synchrotron radiation is discussed
in more detail in Chapter 4 and the effect of bremsstrahlung emission was
studied in Paper G.
Another factor that can increase the effective friction compared to the classical
estimate is partially ionized atoms. The highly energetic runaway electrons
may penetrate (parts of) the electron cloud surrounding the nucleus and thus
effectively scatter off a charge larger than the net charge of the ion. For
heavy ions such as argon or tungsten, which are often present during or after
disruptions, this can have a significant effect on the runaway slowing-down
[75–79]. The runaways may also lose energy in ionizing collisions.
The above mechanisms are pure momentum-space effects. The runaway beam
will however eventually occupy a sizable fraction of the tokamak cross-section,
and magnetic trapping effects may become important. They typically lead to
a reduction in both the Dreicer and avalanche growth rates, which can be
as large as 50% already at r/R = 0.1, with r and R the minor and major
radii [46]. Additionally, stochastic field-line regions (caused by for instance
overlapping magnetic-island structures) can lead to radial transport of the
runaways towards the edge of the plasma, where they are eventually lost to
the wall [49]. This can be both beneficial (if it occurs early in the acceleration
process, before the runaways have reached high energies) and detrimental
(if a fully formed, substantial runaway beam is transported into the wall).
By applying external magnetic fields with a well-defined periodicity, so-called
Resonant Magnetic Perturbations (RMPs), the stochasticity of the edge region
of the plasma can be purposefully increased. This can lead to a more rapid
radial transport of the runaways, resulting in a reduction in their kinetic
energy upon impact with the wall, however since the runaways predominantly
form in the center of the plasma, efficient mitigation can be hard to achieve
[80–82].
The runaways are also subject to outward radial transport because of another,
more fundamental effect: the acceleration of a runaway particle implies a
change in its angular momentum with respect to the symmetry axis of the
torus. This causes a shift of the runaway orbit away from the flux surface, as
the canonical angular momentum of the particle should be conserved [83, 84].
18
2.3 Damping and loss mechanisms for runaways
At high enough particle energies, the runaways will simply drift out of the
plasma and into the wall.
Another effect not captured by a pure momentum-space treatment is the interaction of the runaways with various waves in the plasma. There is evidence
that existing waves, such as toroidal Alfvén eigenmodes, can disperse the runaway beam [85–88]. Due to their highly anisotropic momentum distribution,
the runaways may also destabilize and act as a drive for plasma waves, such
as the whistler [89–91] and EXtraordinary ELectron (EXEL) waves [92, Paper K], which in turn can affect the runaway distribution and reduce the
runaway growth.
The picture is thus complicated in practice, however even the basic dynamics of
the runaway process are not always well understood. Significant experimental
and theoretical effort is spent on improving that understanding and it is the
aim of this thesis to contribute to this endeavor.
19
3 Simulation of runaway-electron
momentum-space dynamics
Although the single-particle estimates considered in Chapter 2 can be useful
in describing some of the phenomena associated with runaways, a complete
and thorough understanding of their dynamics can only be gained through a
treatment of the full kinetic problem. In some idealized situations the equations can be solved analytically, however in general the interplay between the
various processes involved in the momentum-space transport of electrons must
be studied using numerical tools.
The runaways often comprise a small fraction of the total number of electrons, and features in the distribution of electrons many orders of magnitude
smaller than the bulk population must be accurately resolved. Continuum
discretization methods (i.e. finite difference, element, and volume methods)
are well adapted for problems of this type, whereas Monte Carlo methods become inefficient and have problems with numerical noise. In this thesis, two
finite-difference tools for studying runaway-electron dynamics are described:
CODE (Papers B and C, Sec. 3.4) and NORSE (Paper E, Sec. 3.5). Microscopic
Coulomb interaction between particles (collisions) are very important for the
runaway dynamics and we will discuss the treatment of both small (Sec. 3.2)
and large-angle (Sec. 3.3) collisions. Another important effect is synchrotron
radiation reaction, however we postpone the description of the corresponding
operator to Sec. 4.2.3.
We begin by discussing the equations governing the evolution of the electron
population.
3.1 The kinetic equation
When it comes to describing plasma phenomena, several theoretical frameworks of varying degrees of complexity (and explanatory power), have been
21
Chapter 3.
Simulation of runaway-electron momentum-space dynamics
developed. Fluid theories, although tractable, numerically efficient, and useful in other contexts, are based on the assumption that the plasma particles
are everywhere in local thermal equilibrium and can be described by nearMaxwellian distributions. In order to treat the runaway-electron phenomenon,
such a model is inadequate1 , as the runaways by definition constitute a highenergy (non-thermal) tail of the particle distribution. It is therefore necessary
to use kinetic theory, where the distribution of particle positions and velocities
is the prime object of study.
The so-called kinetic equation describes the evolution of a distribution of
plasma particles of species a, fa (x, p, t), according to
X
∂
∂
∂fa
+
(ẋfa ) +
(ṗfa ) =
Cab {fa , fb } + S,
∂t
∂x
∂p
(3.1)
b
where x and p denote the position and momentum, respectively, and ṗ describes the macroscopic equations of motion (given for instance by the Lorentz
force due to the presence of macroscopic electric and magnetic fields). The
collision operator Ca describes microscopic interactions between the plasma
particles (collisions), which are normally treated separately from the macroscopic equations of motion. In general, the collision operator depends on the
distributions of all the particle species b in the plasma and includes contributions from both elastic and inelastic Coulomb collisions. In the latter (which
are often neglected), photons are emitted and carry away some of the energy
and momentum – this radiation is referred to as bremsstrahlung (see for instance Paper G). S represents any sources or sinks of particles or heat, such as
ionization and recombination of neutral atoms, fueling in laboratory plasmas
or heat lost from the plasma, due to radiative processes.
Under certain conditions, the collisions can be neglected, in which case Eq. (3.1)
(with S = 0) is known as the Vlasov equation. With a two-particle collision
operator valid for arbitrary momentum transfer (or equivalently collision distance), it is called the Boltzmann equation, although in practice several simplifications must be made to be able to treat the collisions. Under the assumption
that the momentum transfer in each collision is small, the Boltzmann collision
operator simplifies to the Fokker-Planck collision operator, and Eq. (3.1) is
correspondingly called the Fokker-Planck equation [93, 94]. This operator is
sufficient to treat primary runaway generation, but is not able to describe the
1 Interestingly,
in his seminal papers on electron runaway, Dreicer derived the basic runaway
dynamics using a two-fluid treatment [1]. He did however recognize the limitations of
this description and the follow-up paper uses a kinetic approach [2].
22
3.1 The kinetic equation
avalanche process in which the momentum transfer to the secondary particle
in a knock-on collision is significant. Avalanche generation is instead treated
by including a special source term Sava , discussed in Sec. 3.3.
In general, the distribution fa is defined on a six-dimensional phase-space,
and is very demanding to treat in its entirety. Various approximations are
routinely employed to reduce the kinetic equation to a manageable number of
dimensions (see for instance Ref. [54]). In many situations, the fundamentals
of the runaway problem can be studied in a homogeneous plasma, so that
the spatial dependence can be ignored. In addition, one of the momentumspace dimensions (describing the rapid gyro motion around the magnetic field
lines) can be averaged over if a sufficiently strong magnetic field is present (so
that the gyro radius is ignorable in comparison to the typical length scale of
the gradients in the plasma and the gyration time is short compared to the
timescales of other processes). The tools developed here therefore solve the
kinetic equation in two momentum-space dimensions only, allowing for fast
calculation while most of the relevant physics is retained.
The kinetic equation implemented in CODE and NORSE can be expressed as
∂fe
eE ∂fe
∂
−
·
+
· (Fsyn fe ) = Cee {fe } +Cei {fe } +Sava +Sp +Sh , (3.2)
∂t me c ∂p ∂p
where the second term describes the acceleration due to the electric field,
the third term describes the effects of synchrotron radiation reaction (see
Sec. 4.2), Cee and Cei describe collisions with electrons and ions, respectively,
and Sp and Sh are sources of particles and heat. The two momentum-space
dimensions are conveniently described by the coordinates (p, ξ) introduced in
Sec. 2.1.3. In these coordinates, the electric-field term becomes
eEk
eE ∂fe
∂fe
1 − ξ 2 ∂fe
·
=
+
ξ
.
(3.3)
me c ∂p
me c
∂p
p
∂ξ
Equation (3.2) is then solved for the electron distribution. Both CODE and
NORSE can calculate the time evolution of fe , starting from some initial (often Maxwellian) distribution, and CODE is also able to determine the (quasi)
steady-state distribution directly (in the absence of an avalanche source). In
general, parameters such as the electric field, effective charge, temperature,
and density may vary in time, and both tools have the ability to model this.
Such capability is necessary in order to describe hot-tail runaway generation
and other dynamic scenarios.
23
Chapter 3.
Simulation of runaway-electron momentum-space dynamics
Throughout the rest of this thesis, we will omit the subscript e, and let f
denote the distribution function of electrons. We will also assume an implicit
minus sign in the electric field, so that the runaway electrons are accelerated
in the positive-p direction.
The evolution of the distribution function in a typical runaway case is shown
in Fig. 3.1. Starting from a Maxwellian distribution, the electric field pulls out
a high-energy tail centered around p⊥ = 0 (but with significant spread in p⊥
due to pitch-angle scattering). In 500 thermal collision times, the tail of the
distribution reaches pk ≈ 8, which corresponds to a kinetic energy of 3.6 MeV.
3.2 Collision operator
As discussed in the previous section, the two tools CODE and NORSE have
many of the same capabilities. The main difference between them lies in the
treatment of the electron-electron Coulomb collisions. CODE uses a collision
operator linearized around a Maxwellian, taking advantage of the fact that the
runaways in many cases constitute a small part of the electron distribution,
so that the collisions between runaways may be neglected. This approach
allows for very efficient numerical evaluation of the problem as long as the
plasma parameters remain constant. NORSE, on the other hand, uses a fully
nonlinear relativistic collision operator [95–97], which makes it possible to
treat distributions of arbitrary shape. Thus, NORSE can be used in situations
where the runaways make up a sizable fraction of the distribution, or where
the electric field is strong enough that the electron population is in the slideaway regime (E > Esa ). For a thorough discussion of collision operators in
general, see Ref. [54].
The generally valid collision operator in NORSE accurately treats the elastic
electron-electron collisions in the Fokker-Planck limit. However, in the linearization procedure used to derive the operator in CODE, some properties of the
full operator are compromised. In particular, the linearized operator is often
written as a sum of so-called test-particle and field-particle terms:
l
tp
fp
Cee {f } ' Cee
{f } = Cee
+ Cee
,
(3.4)
tp
where the test-particle term Cee
= Cee {f1 , fM } describes collisions of the
fp
perturbation f1 with the bulk plasma (fM ), and the field-particle term Cee
=
Cee {fM , f1 } describes the reaction of the bulk to the perturbation. Here,
f = fM + f1 with f1 fM , and thus collisions between particles represented
24
3.2 Collision operator
Figure 3.1: Evolution of the electron distribution under a constant electric
field of E = 0.4 V/m (corresponding to E/Ec = 9.6 and E/ED = 0.056).
Contours of the distribution (F = f / max[f ]) in 2D momentum space are
shown at a) the initial time, b) τth = 167, c) τth = 333, and d) τth = 500
thermal collision times. The parameters were: T = 3 keV, n = 5 · 1019 m3 ,
Zeff = 2 and B = 0, and the results were obtained using CODE with avalanche
generation disabled.
25
Chapter 3.
Simulation of runaway-electron momentum-space dynamics
by the perturbation (Cee {f1 , f1 }) have been neglected as they are second order
in the small quantity f1 .2 It is common in runaway studies to neglect the fieldparticle term, as it only affects the bulk plasma and complicates the problem.
If this is done, however, the conservation properties of the linearized operator
l
are compromised as the test-particle term only conserves particles, not
Cee
momentum or energy. This does not significantly affect the runaway dynamics,
but is important for the accurate determination of properties of the bulk
(such as the conductivity). CODE includes both test-particle and field-particle
terms, as discussed in Paper C, however unlike in NORSE, the latter term is
nonrelativistic (i.e. the bulk temperature is assumed to be small compared to
the electron rest energy). The test-particle term in CODE, which is valid for
arbitrary energies, was derived in Ref. [80].
Electron-ion collisions can be described by a much simpler operator, due to
the mass difference between the species involved in the collision (and assuming
the ions to be immobile on the timescales of interest). Both CODE and NORSE
use an electron-ion collision operator which describes pitch-angle scattering,
but neglects the energy transfer to the much heavier ion [54].
As part of the work on Paper G (not included in this thesis), an operator for
inelastic (bremsstrahlung) Coulomb collisions was developed, and is available
in CODE [98].
3.3 Avalanche source term
The avalanche process due to large-angle Coulomb collisions between existing
runaways and thermal electrons cannot be captured using the Fokker-Planck
formalism, and a special source term – derived from the Boltzmann collision
operator – must be included to treat this process. In a linearized formulation,
several avalanche sources describing the creation of the secondary runaway
particles can be formulated (using various assumptions, as will be discussed
shortly), however in the case of a strongly non-Maxwellian distribution function f , no such easily tractable operator is available. For this reason, the
avalanche operators described below are included in CODE but not in NORSE.
2 Note
that for the runaway problem, it is not required that f1 (p, ξ) fM (p, ξ) for all p
and ξ – only that the perturbation is small in a global sense, so that collisions between
runaway particles can be ignored. In the tail, the perturbation is usually many orders
of magnitude larger than the Maxwellian at the corresponding momentum.
26
3.3 Avalanche source term
The generally-valid Boltzmann collision operator is notoriously difficult to
handle numerically, and an efficient solution of the runaway problem requires
the use of reduced models. Work on a simplified conservative treatment of the
avalanche process is ongoing [Conf. Contrib. S], but no practical such operator
is yet available. From the kinematics of a single large-angle collision, a source
term for the generated secondary particles can however be derived, taking
the energy distribution of the incoming electrons into account by utilizing the
full Møller scattering cross-section [99]. This was demonstrated by Chiu et
al. in 1998 [47]. This operator obeys the kinematics of the problem (in the
sense that the momentum of the generated particle is restricted by that of the
incoming particle), however since no sink of particles is included, it violates
the conservation properties of the full Boltzmann operator. No modification
to the momentum of the incoming particle is made and no particle is removed
from the thermal population. Nevertheless, the operator in Ref. [47] is able to
accurately capture the exponential growth of the avalanche. The source term
at a point (p,ξ) takes the form
SCh (p, ξ) =
1 νrel p4in f˜(pin )Σ(γ, γin )
,
2 ln Λ
γpξ
(3.5)
where pin and γin are the normalized momentum and relativistic mass factor
of the incoming primary runaway, γ is the relativistic mass factor for the
generated secondary runaway, f˜ is the angle-averaged electron distribution
(i.e. all incoming particles are assumed to have vanishing pitch-angle), and Σ
is the Møller cross section. Note that due to the kinematics of the problem,
the coordinates are related in such a way that only primary particles with a
single pin can contribute to the source at a point (p, ξ).
By taking the high-energy limit of the Møller cross section, i.e. assuming
that the incoming primaries are all highly relativistic, the source term can be
simplified further. The resulting operator, first derived by Rosenbluth and
Putvinski the year before [50], is widely used and given by
nr νrel
1 ∂
1
SRP (p, ξ) =
δ(ξ − ξs ) 2
.
(3.6)
4π ln Λ
p ∂p 1 − γ
Due to the assumptions used, the kinematics are restricted further, and all
secondary runaways are generated on a parabola given by ξs = p/(1 + γ).
However, since every runaway is assumed to have very large momentum, secondary particles can be generated with momenta larger than that of any of
the particles in the actual distribution. This is illustrated in Fig. 3.2, where
27
Chapter 3.
Simulation of runaway-electron momentum-space dynamics
Figure 3.2: Electron distribution (F = f / max[f ]) after 10 thermal collision
times with a) the source in Eq. (3.5) and b) the source in Eq. (3.6). This
early in the evolution, the result with only primary generation included is
indistinguishable from that in a). The parameters were: T = 300 eV, n =
5 · 1019 m3 , E = 0.5 V/m (corresponding to E/Ec = 12 and E/ED = 0.07),
Zeff = 1 and B = 0, and the results were obtained using CODE.
an unphysical horn-like structure is created, extending to large momenta. In
some situations, the Rosenbluth-Putvinski model therefore tends to overestimate the avalanche growth rate, compared to the source term of Chiu et al.
However, as is shown in Paper C; in certain parameter regimes, the opposite
tendency is seen. This is due to the non-trivial dependence of the Møller cross
section on the momenta of the colliding particles.
If secondary generation dominates, the quasi-steady-state runaway distribution function can be calculated analytically (assuming a growth rate consistent
with the Rosenbluth–Putvinski source), and is given by
fava (pk , p⊥ ) =
pk
nr Ê
p2
exp −
− Ê ⊥ ,
πcz pk ln Λ
cz ln Λ
pk
(3.7)
p
where Ê = (E/Ec − 1)/2(1 + Zeff ) and cz = 3(Zeff + 5)/π [89]. Equation
(3.7), which is valid when γ 1 and E/Ec 1, was used extensively in the
calculation of synchrotron spectra in Paper A, and also as a benchmark in
Paper B. An example distribution is plotted in Fig. 3.3.
28
3.4 CODE
Figure 3.3: Contour plot of the tail of the analytical avalanche distribution
in Eq. (3.7) for Te = 1 keV, ne = 1 · 1020 m−3 , Zeff = 1.5 and E/Ec = 15.
The distribution is not valid for the bulk plasma, and is therefore cut off at
low momentum (in this case at pk = 5).
3.4 CODE
CODE (COllisional Distribution of Electrons) was developed to be a lightweight
tool dedicated to the study of the properties of runaway electrons. It solves
the kinetic equation (3.2) using a finite-difference discretization of p together
with a Legendre-mode decomposition of ξ, and an implicit time-advancement
scheme. The discretization is advantageous as the Legendre polynomials are
eigenfunctions of the collision operator, allowing for a straight-forward implementation and an efficient numerical treatment. In particular, time advancement can be performed at low computational cost, as it is sufficient
to build and invert the matrix representing the system only once, provided
that the plasma parameters are independent of time. The system can then
be advanced in time using just a few matrix operations in each time step.
For small to moderately-sized problems, such as the scenario considered in
Fig. 3.1, CODE runs in a couple of seconds on a standard desktop computer.
More involved set-ups involving time-dependent plasma parameters or low
temperatures in combination with large runaway energies execute in minutes
or sometimes hours. Memory requirements range from a few hundred MB
(or less) to tens of GB, depending on resolution. CODE, which is written in
Matlab, has contributed to a number of studies and is used at several fusion
sites around the world.
The original version, described in Paper B, included the relativistic testparticle collision operator [80] and the Rosenbluth-Putvinski avalanche operator in Eq. (3.6), as well as the ability to find both time-dependent and
steady-state solutions for f . Subsequent extensions include: time-dependent
29
Chapter 3.
Simulation of runaway-electron momentum-space dynamics
plasma parameters, the field-particle term of the collision operator and an associated heat sink, the Chiu et al. avalanche operator (all described in Paper
C), and operators for synchrotron (Paper D, Sec. 4.2.3) and bremsstrahlung
(Paper G) radiation reaction. Various other technical features, such as efficient non-uniform automatically extending finite-difference grids, have also
been added [Conf. Contrib. V].
3.5 NORSE
Building on the experience gained from the work on CODE, NORSE (NOnlinear
Relativistic Solver for Electrons) was developed to extend the range of applicability of runaway-modeling tools to the regime where the runaway population
becomes substantial; i.e. the regime of greatest concern in practice. The fully
nonlinear treatment also makes it possible to consistently study phenomena
such as heating by the electric field, which is discussed in Paper F and Sec. 5.1.
The numerical implementation of NORSE – described in detail in Paper E –
differs substantially from that in CODE. In particular, a linearly implicit timeadvancement scheme is used to treat the nonlinear problem. The electronelectron collision operator implemented in NORSE is formulated in terms of
five potentials (analogous to the two so-called Rosenbluth potentials [94] in the
nonrelativistic case), which are functions of the distribution [96, 97]. By calculating the potentials explicitly in each time step from the known distribution,
the remainder of the kinetic equation can be formulated as a matrix equation,
which can be solved implicitly using standard techniques. Unlike in CODE, this
scheme requires the formation and inversion of a matrix in each timestep, and
is therefore in general more computationally demanding. If time-dependent
parameters are used, however; NORSE is marginally faster than CODE due to
an improved numerical implementation.
NORSE, also written in Matlab, uses a discretization scheme different to that in
CODE, as required by the nonlinear problem. Momentum space is discretized
on a two-dimensional non-uniform finite-difference grid, which allows for improved numerical efficiency as the grid points can be chosen in a way suited
to the problem at hand (i.e. smaller grid spacing close to ξ = 1 where the runaway tail forms, but larger grid spacing around ξ = 0 where the distribution
lacks fine-scale features). For the explicit calculation of the five potentials,
however; a mixed finite-difference–Legendre-mode representation is employed
(similar to CODE), as in this basis the potentials become one-dimensional in-
30
3.5 NORSE
tegrals over p. This representation is only used to calculate the potentials –
which are integral moments of f – and a small number of Legendre modes
are often sufficient for an accurate treatment. The mapping between the two
representations can be performed to machine precision at little computational
cost.
NORSE includes time-dependent plasma parameters, an operator for synchrotron
radiation reaction, a runaway region determined from the distribution in accordance with Eqs. (2.8) and (2.9), and elaborate heat and particle sinks.
31
4 Synchrotron radiation
Charged particles in accelerated motion emit radiation [100]. In the presence
of a magnetic field, the particles in a plasma follow helical orbits as a consequence of the Lorentz force [21, 22]; in other words, they are continuously
accelerated “inwards” – perpendicular to their velocities. The radiation emitted by electrons due to this motion is known as cyclotron radiation if the
particles are nonrelativistic (or mildly so) and synchrotron or betatron radiation if they are highly relativistic (the names come from the types of devices
where the radiation was first observed [101]). Synchrotron radiation has many
applications in the study of samples in condensed matter physics, materials
science, biology and medicine, where it is used in for instance scattering and
diffraction studies, and for spectroscopy and tomography [102]. The radiation
is usually produced using dedicated facilities (synchrotrons), but is also emitted in some natural processes, in particular in astrophysical contexts (where
it can be used as a source of information about the processes in question).
From the distinction between cyclotron and synchrotron radiation, it is evident that in a nonrelativistic plasma (with a temperature significantly below
511 keV), only the far tail of the electron distribution may emit synchrotron
radiation. The only plasma particles that reach highly relativistic energies are
the runaway electrons; the study of the synchrotron emission from a plasma
is thus a very important source of information about the runaways, and their
dynamics.
Synchrotron radiation plays an important role in several of the papers included
in this thesis. In Papers A and B, it is used as a source of information about
the runaway population – as a “passive” diagnostic not affecting the electron
distribution. This is discussed in Sec. 4.1. In Paper D (and also Papers H and
I, not included in this thesis), the impact of the emission on the distribution
is analyzed – i.e. the synchrotron radiation plays an “active” role through the
radiation reaction associated with its emission, as discussed in Sec. 4.2.
33
Chapter 4.
Synchrotron radiation
4.1 Emission and power spectra
The theory of synchrotron radiation was first derived by Schott in 1912 [103],
but was rediscovered, and to a large extent reworked, by Schwinger in the
40’s [104]. More recent, detailed discussions of the properties of synchrotron
radiation can be found in Refs. [100, 102, 105, 106]. In the rest frame of the
particle, the synchrotron radiation is emitted almost isotropically, however the
transformation to the lab frame introduces a strong forward-beaming effect.
Since the motion of the particle is predominantly parallel to the magnetic field
lines (the runaways are accelerated by Ek ), the synchrotron radiation will be
emitted in this direction as well, even though it is the perpendicular motion
of the particle that is the cause of the emission.
Since the synchrotron radiation is directed, its observation requires detectors in the right location and with the right field of view. In many tokamak
experiments this necessitates the use of dedicated cameras for the study of
synchrotron emission from runaways, and the number of such set-ups around
the world is limited but has seen an increase in recent years. Synchrotron
emission from runaways has now been observed in a number of tokamaks,
including TEXTOR [107, 108], DIII-D [41, 109], ASDEX Upgrade [39], Alcator C-Mod [Conf. Contrib. L], FTU [110], EAST [111], KSTAR [25], and
COMPASS [37].
Another important question is the emission spectrum, since it determines the
detector type to use. As part of the work on Paper A, the numerical tool
SYRUP (SYnchrotron emission from RUnaway Particles) was developed to
calculate the synchrotron spectra of both single electrons and runaway distributions. As we shall see, the emission has a distinct peak; for runaways
in tokamaks it is often located in the near infrared, at wavelengths of a few
µm. Electrons with energies above about 20–25 MeV do however also emit
a substantial amount of radiation in the visible range. The majority of the
observations above were done using fast visual cameras, due to the availability and good performance of the technology compared to infrared detectors
(although several IR systems are also in use). Visual cameras also provide a
more sensitive diagnostic of the highest-energy part of the runaway distribution, which is often considered to be of most interest.
In this section, we will first examine the synchrotron spectra emitted by single
electrons, followed by a generalization to a distribution of runaway electrons.
This topic is also discussed in more detail in Paper A.
34
4.1 Emission and power spectra
4.1.1 Single-particle spectrum in a straight magnetic field
The frequency of the cyclotron or synchrotron radiation emitted by a particle
is a multiple of the frequency with which it orbits the magnetic field line (the
gyro or cyclotron frequency ωc = eB/γme ). In the case of cyclotron emission,
the fundamental and the first few harmonics dominate completely, whereas
for the high-energy synchrotron emission, the high harmonics (up to some cut
off) are dominant. Since these are spaced very close together (compared to
the scale of the fundamental frequency), synchrotron radiation essentially has
a continuous frequency spectrum [105]. The emission can span a large part of
the electromagnetic spectrum, from microwaves to hard x rays, depending on
the frequency of the gyro motion and the electron energy.
In terms of quantities convenient for plasma physics [105], the synchrotron
power spectrum emitted by a fast electron can be expressed as
Z ∞
1
ce2
P(λ) = √
K5/3 (l) dl ,
(4.1)
3 ε0 λ3 γ 2 λc /λ
where Kν (x) is the modified Bessel function of the second kind (of order ν),
and λc is a critical wavelength given by
λc =
4π cme γk
4π c
=
,
3 ωc γ 3
3 eBγ 2
(4.2)
with γk = (1 − vk2 /c2 )−1/2 the relativistic γ factor due to the motion parallel to the magnetic field, and B the magnetic field strength. The spectrum,
which peaks at λ ' 0.42λc , is plotted in Fig. 4.1 for a few different parameter
sets. There is a sharp cutoff at short wavelengths, but a much slower decay
towards longer wavelengths. Note that both the peak wavelength and the
emitted power are sensitive to both the particle energy and pitch. In particular, it is possible to produce similar synchrotron spectra using significantly
different parameter sets, which makes it difficult to treat the inverse problem of finding the parameters of a particle (or distribution) that produced a
certain spectrum.
4.1.2 Single-particle spectrum in a toroidal magnetic field
Equation (4.1) considers the radiation emitted due to pure gyro motion around
a straight field line. In a tokamak, particle orbits are more complicated since
35
Chapter 4.
Synchrotron radiation
Figure 4.1: Synchrotron power spectrum for a single electron with kinetic
energy Ek and pitch described by tan θ = v⊥ /vk (with θ the pitch angle).
both the motion around the torus, that due to the helicity of the field lines,
and various drifts contribute. The synchrotron power spectrum for a particle
trajectory including the gyro motion, the motion along a toroidal magnetic
field, and vertical centrifugal drift was derived by I. M. Pankratov in 1999
[112], and is
Z ∞
ce2
P(λ) =
g(y) J0 aζy 3 sin (h(y)) dy
3
2
ε0 λ γ
0
Z ∞
π
,
(4.3)
−4a
y J00 aζy 3 cos (h(y)) dy −
2
0
where a = η/(1 + η 2 ), g(y) = y −1 + 2y, h(y) = 3ζ y + y 3 /3 /2,
ζ=
η=
4π
R
p
,
3
3 λγ
1 + η2
eBR v⊥
ωc R
'
tan θ,
2
γme vk
c
(4.4)
(4.5)
R is the tokamak major radius, Jν (x) is the Bessel function, Jν0 (x) its derivative with respect to the argument, and θ is the pitch angle. The parameter η
36
4.1 Emission and power spectra
is the ratio between the perpendicular and drift velocities of the particle and
determines how much the radius of curvature (and thus the synchrotron emission) varies along the particle orbit [112, 113]. The parameter ζ is proportional
to the ratio λmax /λ, where λmax is the wavelength where the spectrum peaks
(the exact expression for which depends on the parameter regime [113]). The
integrands in Eq. (4.3) are products of Bessel functions and trigonometric
functions and are highly oscillatory with respect to the variable of integration (y). Because of this, numerical integration – although possible – is not
straight-forward. For wavelengths shorter than the maximum, the oscillations
become particularly rapid, as ζ (which appears in the arguments of both the
Bessel and trigonometric functions) becomes large.
In Ref. [112], two asymptotic forms of Eq. (4.3) are also derived. These use
approximations for the integrals, meaning that they are more suited for numerical implementation. In Paper A, the three formulas of Ref. [112], together
with Eq. (4.1), are studied and compared for a variety of tokamak parameters
and it is concluded that the cylindrical limit (Eq. 4.1) is a good approximation to Eq. (4.3) in large devices, whereas in devices with small major radius,
one of the asymptotic expressions is more suitable in terms of approximating
Eq. (4.3). In general, however, the power spectra from the various expressions
are similar.
4.1.3 Spectrum from a runaway distribution
The total synchrotron power emitted by an electron in circular motion is [104]
Ptot =
e2 ω⊥ 3 4
β γ ,
6πε0 ρ ⊥
(4.6)
where ω⊥ is the angular velocity, ρ is the radius of curvature and β⊥ = v⊥ /c.
In a homogeneous magnetized plasma, the angular velocity and curvature
radius are the Larmor frequency and radius, respectively: ωc and rL = v⊥ /ωc .
This gives
Ptot =
e4
e2 ωc2 3 4
e4
2 2 2
β⊥ γ =
B
β
γ
=
B 2 p2⊥ .
⊥
6πε0 v⊥
6πε0 m2e c
6πε0 m2e c
(4.7)
The total emitted power thus scales as p2⊥ = γ 2 (v⊥ /c)2 = γ 2 sin2 θ (v/c)2 ≈
γ 2 θ2 , with θ the particle pitch angle, meaning that the most energetic particles with the largest pitch angles emit most strongly. It has therefore been
37
Chapter 4.
Synchrotron radiation
assumed that the emission from these particles completely dominates the spectrum, and when interpreting synchrotron spectra and emission patterns, the
simplification of considering a mono-energetic beam of electrons with a single
pitch – referred to here as the “single-particle approximation” – has frequently
been employed (see for instance Refs. [109, 114–116]).
Less approximate synchrotron spectra can be calculated by using the average
emission from the entire runaway distribution, according to
Z
Z
2π
f (p|| , p⊥ ) P p⊥ dp⊥ dp|| =
P̄ (p|| , p⊥, λ) dp⊥ dp|| ,
(4.8)
P (λ) =
nr Sr
Sr
where P = P(p|| , p⊥, λ) is one of the single particle emission formulas (i.e.
Eqs. 4.1 and 4.3), Sr is the runaway region in momentum space, and P̄ is the
integrand – the contribution to the emitted synchrotron power from a given
region of momentum space. Spectra calculated from runaway distributions
using the above equation are studied in detail in Papers A and B.
The validity of the single-particle approximation can be assessed by examining the contribution to the total synchrotron emission at a given wavelength
from different parts of momentum space. This is done in Fig. 4.2. The figure shows a runaway distribution and the corresponding contribution to the
emission (the quantity P̄ ) for two different wavelengths1 . In this case, the
spectrum (calculated using Eq. 4.1) peaks at around 15 µm. At short wavelengths (Fig. 4.2b), the emission is localized to the particles of largest energy
and pitch angle, as expected from the simple argument above. At somewhat
longer wavelengths (Fig. 4.2c), however; particles in a larger part of momentum
space contribute (and the emission is much stronger in general). In addition,
for this wavelength, the main contribution is not from the particles with the
largest pitch-angle, but P̄ instead peaks closer to the parallel axis. Both these
effects lead to the single-particle approximation becoming a poor estimate in
general (at the very least, different particle energies and pitches would have
to be used to approximate the distribution at different wavelengths).
This fact is also evident from the comparison of single-particle and runawaydistribution synchrotron spectra, where the spectra generally differ both qualitatively and quantitatively, as shown in Papers A and B. Including the full
runaway distribution in the calculation is thus absolutely necessary to obtain
1 The
parameters used were T = 300 eV, n = 5 · 1019 m−3 , E = 1 V/m, Zeff = 1 and
B = 5.3 T, and the distribution was obtained by running CODE for 17000 collision times,
including synchrotron-radiation back-reaction, but neglecting avalanche runaway generation.
38
4.1 Emission and power spectra
-10
9
-15
6
3
-20
0
-25
9
2
6
1
3
0
0
1
9
6
0.5
3
0
0
0
5
10
15
20
25
Figure 4.2: a) Electron distribution in 2D momentum space and b)–c) contribution to the corresponding synchrotron emission at wavelengths b)
λ = 700 nm and c) λ = 7 µm. The plotted quantity in b) and c)
is P̄norm = P̄ / max[P̄7µm ], i.e. P̄ normalized to its maximum value for
λ = 7 µm. Note the difference in emission amplitude for the two wavelengths.
39
Chapter 4.
Synchrotron radiation
accurate results, and in this sense SYRUP constitutes a significant improvement
over previous methods. The synchrotron spectrum is very sensitive to changes
in the plasma parameters and electric-field strength. This sensitivity is due
to the dependence on the exact shape of the runaway distribution, something
which cannot be captured by the mono-energetic approximation.
It is clear from the discussion above that particles of many different energies
and pitches contribute to the observed synchrotron spectrum. This is also true
for the spatial shape of the detected radiation spot, where particles at different
spatial locations and with different momenta contribute to overlapping regions
on the detector plane. This added complication makes the task of extracting
information about the runaway distribution from the synchrotron image highly
non-trivial.
In the analysis in Paper A, analytical avalanche distributions were used. The
analytical formula (Eq. 3.7) represents a steady-state limit, however, and is
not able to capture dynamical effects or describe the synchrotron emission in
the early stages of the runaway-population evolution. In addition, it does not
include the effect of radiation-reaction losses, which can have a significant effect when E/Ec is small. In Paper B, numerical distributions from CODE were
used to study both dynamic phenomena and distributions where the Dreicer
mechanism was dominant. Excellent agreement was also found between the
numerical distribution and Eq. (3.7) at sufficiently late times.
4.2 Radiation-reaction force
As an electron emits radiation, it receives an impulse in the opposite direction
due to the conservation of momentum. There is therefore a radiation-reaction
force Fsyn associated with the emission of synchrotron radiation, which acts to
slow the particle down (and reduce its pitch angle). Synchrotron emission only
becomes important at relativistic energies, however; contrary to collisional
friction, the radiation reaction force increases with particle speed (in accordance with the estimate in the previous section). This completely changes
the force balance for runaways at high momentum and non-vanishing pitch
angles2 . In the following discussion it is convenient to consider only the onedimensional single-particle force balance to qualitatively illustrate the impact
of the synchrotron radiation reaction on the dynamics. In practice, however;
2 The
force balance exactly on the parallel axis remains unchanged, as non-vanishing p⊥
is required for the emission of synchrotron radiation.
40
4.2 Radiation-reaction force
|F|
Ff
eE
Fsyn
eE>Ff
FC
eE*c
eEc
vth
v1
vmin
Sr
v2
c
v
Figure 4.3: Schematic representation of the forces associated with total friction (Ff ), collisional friction (FC ), synchrotron radiation reaction (Fsyn ),
the (classical) critical field (Ec ) and the critical field including synchrotron
radiation reaction (Ec∗ ). The two critical velocities v1 and v2 corresponding
to the electric field E are also shown, together with the runaway region
(Sr ), and the speed at which the total friction force is minimized (vmin ).
See Fig. 2.2 for comparison. Note that the velocity scale is chosen for clarity
– in practice both vmin and v2 lie close to c.
the problem involves transport processes in two-dimensional momentum space
and must in general be treated using numerical tools, see Chapter 3.
The force balance in the presence of synchrotron radiation reaction is depicted
in Fig. 4.3. Two effects are of particular interest in the figure: firstly, the
radiation-reaction force effectively prevents runaways from reaching arbitrary
energies, and secondly, it raises the critical field for runaway generation. We
will discuss these two effects in turn in the following sections.
4.2.1 A limit on the achievable runaway energy
As can be seen in Fig. 4.3; for a given electric field E, there are two speeds
(v1 and v2 ) larger than the thermal speed for which the total friction force
equals the accelerating force: Ff (v1,2 ) = |eE|. Runaway is only possible for
v1 < v < v2 , meaning that particles are unlikely to reach energies significantly
higher than that corresponding to v2 . This has been suggested as a possi-
41
Chapter 4.
Synchrotron radiation
ble mechanism limiting the achievable runaway energy, both in the case of
synchrotron and bremsstrahlung radiation reaction [60, 66, 117, 118].
An additional consequence of the existence of an upper bound to the runaway region is that particles tend to accumulate in the vicinity of the speed
v2 in velocity space. Provided certain conditions are satisfied, momentumspace transport mechanisms associated with this process can even lead to the
formation of a non-monotonic feature – a “bump” – on the runaway tail, as
discussed in Papers G, H and I (not included in the thesis).
4.2.2 Effect on the critical field
As indicated by Fig. 4.3, the minimum of the friction force for any nonvanishing ξ is no longer found at v = c (see Sec. 2.1.1), but at some intermediate speed vmin satisfying v1 < vmin < v2 . (Note that, since the synchrotron
emission is a relativistic effect, vmin will nevertheless be close to c.) Therefore, the minimum field necessary for runaway generation to occur is raised
accordingly:
|eEc∗ (ξ)| = Ff (vmin , ξ),
(4.9)
where Ec∗ (ξ) is the critical field at a given ξ in the presence of radiation reaction
forces and Ec∗ > Ec for all non-vanishing ξ, as depicted in Fig. 4.3.
Note that the force balance on the parallel axis (ξ = 1) remains unaffected by
the radiation reaction, and thus the critical field Ec is unchanged if collisional
diffusion is neglected. In practice, however; diffusive and dynamic processes
play an important role in determining the effective critical field. The full
problem is studied in Paper D using simulations in CODE, and it is found
that the synchrotron radiation reaction can reduce the Dreicer growth rate by
orders of magnitude for E/Ec close to unity, corresponding to an increase in
the effective critical field.
4.2.3 Operator for synchrotron radiation reaction
To understand the full role of radiation-reaction effects on plasma dynamics,
the single-particle treatment considered above is insufficient. Fully kinetic
simulations are necessary to capture the interplay between the various processes affecting the momentum-space transport. Such calculations can be
42
4.2 Radiation-reaction force
performed using CODE or NORSE, and for this purpose an operator describing
the radiation reaction is needed.
The radiation-reaction force can be calculated from the Abraham-LorentzDirac force affecting an electron [119],
e2 γ 2
3γ 2
γ2
3γ 2
2
Frad =
v̈ + 2 (v · v̇) v̇ + 2 v · v̈ + 2 (v · v̇) v , (4.10)
6πε0 c3
c
c
c
where v is the velocity of the particle. Assuming that the magnetic force
dominates, so that the particle is predominantly accelerated perpendicular to
its velocity (v · v̇ = 0), the expression can be simplified to
γp 1 − ξ 2
p
Fsyn = −
(4.11)
τ
p r
pξ 1 − ξ 2
ξ
Fsyn
=−
,
(4.12)
γτr
where τr = 6πε0 (me c)3 /e4 B 2 is the radiation-damping timescale. The radiationreaction force enters the kinetic equation (3.2) as an operator of the form
(∂/∂p) · (Fsyn f ), and using Eqs. (4.11) and (4.12), the explicit form is
∂
1 − ξ2
∂f
2
∂f
· (Fsyn f ) = −
−ξ
+ 4p2 +
f .
(4.13)
γ2p
∂p
γτr
∂p
∂ξ
1−ξ 2
The force acts to limit both the particle energy and pitch, which is to be
expected as the emitted synchrotron power is proportional to Ptot ∼ γ 2 θ2 (see
Sec. 4.1.3). Fig. 4.4 shows the effect on the distribution: its width in p⊥ and
extension in pk is reduced in the presence of a magnetic field. Equations (4.11)
and (4.12) were derived in Ref. [117], the first paper to properly consider the
role of radiation reaction in runaway dynamics, but in that paper a high-energy
limit was used which lead to an incomplete expression in place of Eq. (4.13),
as pointed out in [120].
43
Chapter 4.
Synchrotron radiation
Figure 4.4: Electron distribution functions in CODE after 700 thermal collision times, a) without and b) with synchrotron radiation reaction included.
The parameters were T = 3 keV, n = 1 · 1019 m−3 , E = 0.05 V/m, Zeff = 1
and B = 6 T.
44
5 Nonlinear effects and slide-away
When studying the behavior of runaway electrons, a very common approximation is to treat the runaways as a perturbation to an electron population
in local thermal equilibrium. This linearization of the distribution function
f around a Maxwellian significantly simplifies the collision operator Cee in
Eq. (3.2), as well as the numerical method needed to solve the kinetic equation. However, this approach is not valid when the runaway population becomes substantial, or the linearization becomes invalid for other reasons (such
as when a strong electric field E & Esa shifts the bulk of the electron population). There is also a risk that the linearized treatment fails to accurately
describe subtle phenomena such as feedback loops (which may for instance
cause rapid depletion of the thermal population). The tool NORSE was developed to be able to treat such situations, as discussed in Chapter 3. Here we
will introduce two particular nonlinear effects considered in Papers E and F:
Ohmic heating and slide-away.
5.1 Ohmic heating
The accelerating force of the parallel electric field affects all electrons, not only
runaways. However, for most particles, the energy gained from the electric
field is quickly redistributed into random motion through collisions with other
plasma particles. As a consequence, the electric field acts as a source of
heat, primarily affecting the thermal bulk of the electron population – this is
commonly referred to as Ohmic or Joule heating. The associated change in
the energy W of the electron population is given by the energy moment of the
electric-field term in the kinetic equation:
Z
dW
E ∂f
3
2
= d p (γ − 1)me c −
·
,
(5.1)
dτ
Ec ∂p
with τ = νrel t the time in units of the collision time for relativistic electrons.
If the electric field is sufficiently strong, the heating of the bulk can be sub-
45
Chapter 5.
Nonlinear effects and slide-away
stantial. This can have consequences for electron slide-away, as discussed in
the next section.
The Ohmic heating is automatically accounted for in a nonlinear treatment of
the kinetic equation, making NORSE able to consistently model this process.
This is less straight-forward to do in a linearized tool such as CODE, since
the heating effectively involves a change to the Maxwellian around which the
distribution is linearized; either the heat must be removed, or the properties
of the Maxwellian modified.
The heat supplied by the electric field does not always stay in the plasma,
however. Spatial temperature gradients lead to heat transport (a fundamental problem in fusion reactors), although this process may or may not be
relevant on the runaway acceleration timescale. More important in a cold
post-disruption fusion plasma may be the energy radiated away via atomic
transitions in partially ionized impurity ions (the so-called line radiation), or
spatial transport of particles and energy due to unstable modes (sometimes
driven by the runaways themselves – see for instance [90] and Paper K). As
a consequence, the plasma temperature may stay constant, or even decrease,
despite the existence of a strong electric field.
5.2 Electron slide-away
As discussed in Sec. 2.1.2: if the electric field is stronger than the collisional
friction in the entire momentum space, all electrons are accelerated, which is
known as the slide-away regime [1, 57]. Recalling that the slide-away field
Esa ∼ n/T (see Eq. 2.5), where these quantities are those of the bulk (since
the maximum of the friction force is located at around v = vth ), a transition
to the slide-away regime can be induced through either of the following four
mechanisms:
I: E > Esa – For electric fields stronger than the classical slide-away field
Esa , slide-away is immediate
II: E . Esa – Although the field is slightly weaker than Esa , it distorts the
distribution, which lowers the friction. This process is a positive feedback loop, since the reduced friction makes acceleration easier, leading
to further distortion of the distribution, reduced friction, etc. Eventually, the friction is sufficiently reduced that the electric-field acceleration
dominates everywhere, leading to slide-away. This process happens on
46
5.2 Electron slide-away
timescales comparable to the thermal collision time, and is thus very
quick.
III: E < Esa , no or inadequate heat sink – The field is significantly weaker
than Esa but supplies heat to the distribution. This heat is not efficiently
removed, which leads to Ohmic heating of the bulk electrons. The increase in temperature reduces the slide-away field, eventually leading
to slide-away as the electric field becomes comparable to Esa (t). The
process can be slow or quick, depending on the initial value of E/Esa
and the efficacy of the heat sink.
IV: E < Esa , adequate heat sink – The heat supplied by the electric field
is removed and the temperature is kept constant. The electric field
causes prolonged substantial runaway generation, which eventually leads
to depletion of the bulk population of electrons. This reduces the slideaway field, and eventually slide-away is reached. This process is generally
the slowest.
Mechanisms I–III are discussed in Paper E. In Paper F, the focus is the influence on the transition to slide-away of the properties of the heat sink. Both
mechanisms III and IV are observed in a tokamak-relevant scenario, depending on the details of the heat sink. The feedback loop related to mechanism II
is also discussed, as mechanisms III and IV exhibit the behavior of mechanism
II just before slide-away is reached.
In an idealized situation, slide-away should eventually occur in all systems
with a persistent electric field. In practice, however; this is not observed.
If E Esa , the timescale of the slide-away transition is slow compared to
most processes in the plasma. In this case, cold electrons are supplied to
the thermal population, compensating for the particles running away and
thus preventing a transition in accordance with mechanism IV. Also in the
absence of efficient cooling (mechanism III), slide-away may be prevented by
the feedback between the current and the electric field. The electric field in a
tokamak disruption is generated as a consequence of a reduction in the plasma
current (due to a quick cooling of the plasma), however if runaway generation
becomes strong enough to significantly affect the total current, the electric
field will be reduced. This may interrupt a transition through mechanisms
III or IV before slide-away is reached. Fields sufficiently strong to cause
slide-away through mechanisms I–II are uncommon in fusion experiments.
Nevertheless, as demonstrated in Paper F, non-linear simulations can be vital
for the understanding of some realistic tokamak scenarios where a linearized
approach quickly becomes invalid.
47
6 Concluding remarks
In order to reduce the threat posed by runaway electrons to future fusionenergy devices, progress on several fronts are necessary. The understanding of
runaway generation and loss mechanisms needs to be improved – both theoretically and experimentally – so that reliable operation scenarios and efficient
mitigation schemes can be developed. At the core of the runaway phenomenon
lies dynamics in momentum space, which is the topic of this thesis. The most
important mechanisms affecting these dynamics are summarized schematically
in Fig. 6.1. In this thesis, these mechanisms (apart from bremsstrahlung radiation reaction) have been investigated using purpose-built numerical tools,
leading to several new insights concerning runaway dynamics. The present
Chapter provides a summary of the papers that constitute this work, as well
as a short outlook.
6.1 Summary of the included papers
An important component in increasing the understanding of runaway dynamics is to improve the capability to analyze experimentally observed runaway
beams and to extract the information available in the few diagnostic measurements that are sensitive to the runaway parameters. To this end, Paper A
is focused on calculating the synchrotron spectrum emitted by a distribution
of runaway electrons: a significant improvement over previous methods which
typically interpret the observed spectra using a single-particle approximation
for the runaway population (as discussed in connection with Fig. 4.2). As
shown in the paper, this approximation fails to capture both qualitative and
quantitative features of the synchrotron spectrum, and the use of distributionintegrated spectra is essential to accurately infer the runaway parameters. The
paper analyses the spectra obtained using analytical avalanche distributions
(Eq. 3.7). Together with the possibility to easily obtain runaway-electron distribution functions numerically using CODE or NORSE, the work in this thesis
makes a significant contribution to the interpretation of synchrotron spectra,
and thereby to the experimental analysis of runaway-electron parameters.
49
Chapter 6.
Concluding remarks
Synchrotron
radiation reaction
Large-angle
collisions
Pitch-angle scattering
E-field
Collisional friction
Parallel collisional diffusion
Bulk heating
and cooling
Bremsstrahlung
radiation reaction
Figure 6.1: Schematic presentation of the various effects of importance for
runaway-electron momentum-space dynamics and their qualitative effect on
the distribution function. Bremsstrahlung radiation reaction is included for
completeness, even though it has not been discussed in detail in this thesis.
Paper B describes and validates the development of CODE, a tool for calculating the time-evolution of the electron distribution function (or its steady-state
shape) in the presence of electric-field acceleration and Coulomb collisions. In
the paper, the obtained distributions are used to calculate synchrotron spectra
(in accordance with Paper A), giving access to parameter regimes and evolution times not properly modeled by the analytical avalanche distribution.
Paper C expands the capability of CODE by introducing time-dependent plasma
parameters (enabling the modeling of dynamic scenarios such as disruptions);
a conservative collision operator essential for calculating e.g. the plasma conductivity; and the avalanche source term in Eq. (3.5). A scenario dominated
by hot-tail runaway generation is investigated, and the effect on the avalanche
50
6.1 Summary of the included papers
growth rate of the choice of avalanche operator is quantified. It is found that
the commonly used Rosenbluth-Putvinski avalanche operator can both overestimate and underestimate the avalanche growth rate significantly, depending
on the parameter range.
Apart from acting as an observable diagnostic for the runaway distribution,
the emission of synchrotron radiation also affects the distribution itself, as
discussed in Sec. 4.2. To model this behavior, Paper D introduces an operator
describing synchrotron radiation reaction into CODE and uses it to investigate
the effect of the radiation reaction on the critical field for runaway generation.
The paper also explains that a large part of the observed modification to the
critical field (initially attributed to so-called “anomalous losses”, which include
synchrotron radiation reaction) is in fact likely to be just a manifestation of the
temperature dependence of the Dreicer runaway-generation rate (Eq. 2.10),
so that the parameter determining the runaway growth is E/ED rather than
E/Ec . The paper also shows that redistribution of particles in velocity space
can give the impression (because of reduced synchrotron emission) that the
runaway population is decaying, when in fact both the number and total
kinetic energy of the runaways keep increasing. Again, these insights impact
the interpretation of experimental observations and thus contribute to the
understanding of runaway parameters in practice.
Paper E describes the development of NORSE. The motivation behind this
new Fokker-Planck tool was to investigate the impact of collisional nonlinearities on the runaway dynamics. Specifically, NORSE makes it possible to study
situations where the runaways constitute a substantial part of the electron
distribution, or the electric field is significant compared to ED . This had not
previously been done in a framework allowing for relativistic particle energies.
The paper highlights the fact that a transition to the slide-away regime can
be initiated for electric fields well below the traditional slide-away field Esa .
This line of investigation is then continued in Paper F, which uses an ITERdisruption scenario to explore the importance of Ohmic heating of the bulk
electron population. In addition, the paper studies the impact on the slideaway dynamics of the efficiency of the available heat-loss mechanisms. It is
found that in the absence of spatial-transport and current–electric-field feedback effects, the electron population in an ITER disruption should eventually
transition to a slide-away regime, however the time scale of this transition
depends strongly on the heat-loss rate. This could potentially have large
consequences for the understanding of runaway dynamics in ITER, although
further investigations are needed.
51
Chapter 6.
Concluding remarks
6.2 Outlook
In this thesis, the focus has been on modeling of the dynamics of runaway
electrons in momentum space. These dynamics have sometimes been misunderstood or misinterpreted, as in the case of observed modifications to the
critical electric field, which motivated the work on Paper D. The tools developed in this thesis can help avoid future such misunderstandings – as they
facilitate the investigation of runaway dynamics – and make more accurate
interpretation of experimental data possible.
All results presented in this thesis were obtained using predefined electricfield evolutions, i.e. the applied electric field was not affected by the electron
distribution in any way and was therefore not calculated self-consistently. As
long as the runaway population and current are small, this approximation is
adequate, however in cases where the runaways contribute a substantial part
of the plasma current (or otherwise significantly affect the plasma evolution),
a self-consistent treatment is essential. This point is of particular concern in
the scenario considered in Paper F. To build on the work presented here,
a logical step forward is to include a self-consistent electric field, taking the
evolution of the electron distribution into account.
Another area for further research is to extend the numerical treatment to include one spatial (radial) dimension. This would make it possible to capture
magnetic trapping effects, as well as collisional radial transport of the runaways. Naturally, such a development is not without complications, both analytically and numerically, and the increased dimensionality puts much higher
demands on the computational resources. Although Fokker-Planck tools (such
as LUKE [44, 45]) that include a radial coordinate do exist, they are not primarily focused on runaway research. Neither do they necessarily include all
the relevant effects (for instance LUKE includes trapping effects, but not consistent radial transport).
Some progress towards including the two effects mentioned above have been
made, and is described in Conf. Contrib. V. The adopted approach is to
couple a Fokker-Planck solver (in this case CODE) to the 1D fluid code GO
[72, 121–123], which evolves the plasma parameters and current, and handles
radial electric-field diffusion. This approach could serve as a first step towards
a self-consistent model, however further work is needed since CODE does not
include trapping effects. A complete tool able to consistently calculate the
electron distribution function as a function of momentum, radius and time
in a disruption scenario would contribute significantly to the understanding
52
6.2 Outlook
of runaway generation and loss, and would be of great value to the fusion
community.
The work presented in this thesis has brought new insights into the dynamics
of runaways and the analysis of the radiation they emit. Hopefully, the information and tools described herein will bring us one step closer to stable and
reliable operation of fusion devices.
53
Bibliography
[1] H. Dreicer, Electron and ion runaway in a fully ionized gas I, Physical
Review 115, 238 (1959), DOI: 10.1103/PhysRev.115.238.
[2] H. Dreicer, Electron and ion runaway in a fully ionized gas II, Physical
Review 117, 329 (1960), DOI: 10.1103/PhysRev.117.329.
[3] A. Einstein, Zur Elektrodynamik bewegter Körper, Annalen der Physik
322, 891 (1905), DOI: 10.1002/andp.19053221004.
[4] P. Helander, L.-G. Eriksson, and F. Andersson, Runaway acceleration
during magnetic reconnection in tokamaks, Plasma Physics and Controlled Fusion 44, B247 (2002), DOI: 10.1088/0741-3335/44/12B/318.
[5] W. J. Nuttall, Fusion as an energy source: Challenges and opportunities, Tech. Rep. (Institute of Physics Report, 2008) http://www.iop.org/
publications/iop/2008/file 38224.pdf.
[6] F. F. Chen, An Indispensable Truth - How Fusion Power Can Save the
Planet (Springer, 2011).
[7] J. P. Freidberg, Plasma Physics And Fusion Energy (Cambridge University Press, 2007).
[8] T. Hender, J. Wesley, J. Bialek, A. Bondeson, A. Boozer, R. Buttery, A. Garofalo, T. Goodman, R. Granetz, Y. Gribov, O. Gruber, M. Gryaznevich, G. Giruzzi, S. Günter, N. Hayashi, P. Helander, C. Hegna, D. Howell, D. Humphreys, G. Huysmans, A. Hyatt, A. Isayama, S. Jardin, Y. Kawano, A. Kellman, C. Kessel,
H. Koslowski, R. L. Haye, E. Lazzaro, Y. Liu, V. Lukash, J. Manickam,
S. Medvedev, V. Mertens, S. Mirnov, Y. Nakamura, G. Navratil, M. Okabayashi, T. Ozeki, R. Paccagnella, G. Pautasso, F. Porcelli, V. Pustovitov, V. Riccardo, M. Sato, O. Sauter, M. Schaffer, M. Shimada,
P. Sonato, E. Strait, M. Sugihara, M. Takechi, A. Turnbull, E. Westerhof, D. Whyte, R. Yoshino, H. Zohm, and the ITPA MHD, Disruption
55
Bibliography
and Magnetic Control Topical Group, Chapter 3: MHD stability, operational limits and disruptions, Nuclear Fusion 47, S128 (2007), DOI:
10.1088/0029-5515/47/6/S03.
[9] E. Hollmann, G. Arnoux, N. Commaux, N. Eidietis, T. Evans,
R. Granetz, A. Huber, D. Humphreys, V. Izzo, A. James, T. Jernigan, M. Lehnen, G. Maddaluno, R. Paccagnella, P. Parks, V. Philipps,
M. Reinke, D. Rudakov, F. Saint-Laurent, V. Sizyuk, E. Strait, J. Wesley, C. Wong, and J. Yu, Plasma-surface interactions during tokamak
disruptions and rapid shutdowns, Journal of Nuclear Materials 415, S27
(2011), DOI: 10.1016/j.jnucmat.2010.10.009, Proceedings of the 19th International Conference on Plasma-Surface Interactions in Controlled Fusion.
[10] A. H. Boozer, Theory of runaway electrons in ITER: Equations, important parameters, and implications for mitigation, Physics of Plasmas 22,
032504 (2015), DOI: 10.1063/1.4913582.
[11] A. Gurevich, G. Milikh, and R. Roussel-Dupre, Nonuniform runaway
air-breakdown, Physics Letters A 187, 197 (1994), DOI: 10.1016/03759601(94)90062-0.
[12] H. E. Tierney, R. A. Roussel-Dupré, E. M. D. Symbalisty,
W. H. Beasley, Radio frequency emissions from a runaway
tron avalanche model compared with intense, transient signals
thunderstorms, Journal of Geophysical Research 110 (2005),
10.1029/2004JD005381.
and
elecfrom
DOI:
[13] U. S. Inan and N. G. Lehtinen, Production of terrestrial gamma-ray
flashes by an electromagnetic pulse from a lightning return stroke, Geophysical Research Letters 32 (2005), DOI: 10.1029/2005GL023702.
[14] T. F. Bell, V. P. Pasko, and U. S. Inan, Runaway electrons as a source
of red sprites in the mesosphere, Geophysical Research Letters 22, 2127
(1995), DOI: 10.1029/95GL02239.
[15] G. D. Holman, Acceleration of runaway electrons and Joule heating in solar flares, Astrophysical Journal 293, 584 (1985), DOI:
10.1086/163263.
[16] H. Lesch and W. Reich, The origin of monoenergetic electrons in the arc
of the galactic center – particle acceleration by magnetic reconnection,
Astronomy and Astrophysics 264, 493 (1992).
56
Bibliography
[17] T. Fülöp and G. Papp, Runaway positrons in fusion plasmas,
Physical Review Letters 108, 225003 (2012), DOI: 10.1103/PhysRevLett.108.225003.
[18] T. Fülöp and M. Landreman, Ion runaway in lightning discharges,
Physical Review Letters 111, 015006 (2013), DOI: 10.1103/PhysRevLett.111.015006.
[19] T. Fülöp and S. Newton, Alfvénic instabilities driven by runaways
in fusion plasmas, Physics of Plasmas 21, 080702 (2014), DOI:
10.1063/1.4894098.
[20] J. Wesson, Tokamaks, 2nd ed. (Oxford Science Publications, 1997).
[21] O. Heaviside, On the electromagnetic effects due to the motion of electrification through a dielectric, Philosophical Magazine 27, 324 (1889),
DOI: 10.1080/14786448908628362.
[22] H. A. Lorentz, La Théorie electromagnétique de Maxwell et son application aux corps mouvants, Archives néerlandaises des sciences exactes et
naturelles, Vol. 25 (E. J. Brill, 1892) pp. 363–552.
[23] B. Esposito, J. R. Martı́n-Solı́s, F. M. Poli, J. A. Mier, R. Sánchez,
and L. Panaccione, Dynamics of high energy runaway electrons in the
Frascati Tokamak Upgrade, Physics of Plasmas 10, 2350 (2003), DOI:
10.1063/1.1574328.
[24] H. W. Lu, X. J. Zha, F. C. Zhong, L. Q. Hu, R. J. Zhou, and
EAST Team, Investigation of runaway electrons in the current rampup by a fully non-inductive lower hybrid current drive on the EAST
tokamak, Physica Scripta 87, 055504 (2013), DOI: 10.1088/00318949/87/05/055504.
[25] M. Cheon, J. Kim, Y. An, D. Seo, and H. Kim, Observation of
the loss of pre-disruptive runaway electrons in KSTAR ohmic plasma
disruptions, Nuclear Fusion 56, 126004 (2016), DOI: 10.1088/00295515/56/12/126004.
[26] J. Wesson, R. Gill, M. Hugon, F. Schüller, J. Snipes, D. Ward,
D. Bartlett, D. Campbell, P. Duperrex, A. Edwards, R. Granetz, N. Gottardi, T. Hender, E. Lazzaro, P. Lomas, N. L. Cardozo, K. Mast,
M. Nave, N. Salmon, P. Smeulders, P. Thomas, B. Tubbing, M. Turner,
and A. Weller, Disruptions in JET, Nuclear Fusion 29, 641 (1989), DOI:
10.1088/0029-5515/29/4/009.
57
Bibliography
[27] P. de Vries, M. Johnson, B. Alper, P. Buratti, T. Hender, H. Koslowski,
V. Riccardo, and JET-EFDA Contributors, Survey of disruption
causes at JET, Nuclear Fusion 51, 053018 (2011), DOI: 10.1088/00295515/51/5/053018.
[28] E. Lenz, Ueber die Bestimmung der Richtung der durch elektrodynamische Vertheilung erregten galvanischen Ströme, Annalen der Physik 107,
483 (1834), DOI: 10.1002/andp.18341073103.
[29] R. Gill, B. Alper, A. Edwards, L. Ingesson, M. Johnson, and D. Ward,
Direct observations of runaway electrons during disruptions in the
JET tokamak, Nuclear Fusion 40, 163 (2000), DOI: 10.1088/00295515/40/2/302.
[30] V. Plyusnin, V. Riccardo, R. Jaspers, B. Alper, V. Kiptily, J. Mlynar,
S. Popovichev, E. de La Luna, F. Andersson, and JET EFDA contributors, Study of runaway electron generation during major disruptions in
JET, Nuclear Fusion 46, 277 (2006), DOI: 10.1088/0029-5515/46/2/011.
[31] C. Reux, V. Plyusnin, B. Alper, D. Alves, B. Bazylev, E. Belonohy,
A. Boboc, S. Brezinsek, I. Coffey, J. Decker, P. Drewelow, S. Devaux,
P. de Vries, A. Fil, S. Gerasimov, L. Giacomelli, S. Jachmich, E. Khilkevitch, V. Kiptily, R. Koslowski, U. Kruezi, M. Lehnen, I. Lupelli, P. Lomas, A. Manzanares, A. M. D. Aguilera, G. Matthews, J. Mlynář,
E. Nardon, E. Nilsson, C. P. von Thun, V. Riccardo, F. Saint-Laurent,
A. Shevelev, G. Sips, C. Sozzi, and JET contributors, Runaway electron
beam generation and mitigation during disruptions at JET-ILW, Nuclear Fusion 55, 093013 (2015), DOI: 10.1088/0029-5515/55/9/093013.
[32] E. Hollmann, P. Parks, D. Humphreys, N. Brooks, N. Commaux,
N. Eidietis, T. Evans, R. Isler, A. James, T. Jernigan, J. Munoz,
E. Strait, C. Tsui, J. Wesley, and J. Yu, Effect of applied toroidal
electric field on the growth/decay of plateau-phase runaway electron currents in DIII-D, Nuclear Fusion 51, 103026 (2011), DOI: 10.1088/00295515/51/10/103026.
[33] E. Hollmann, M. Austin, J. Boedo, N. Brooks, N. Commaux, N. Eidietis,
D. Humphreys, V. Izzo, A. James, T. Jernigan, A. Loarte, J. MartinSolis, R. Moyer, J. Muñoz-Burgos, P. Parks, D. Rudakov, E. Strait,
C. Tsui, M. V. Zeeland, J. Wesley, and J. Yu, Control and dissipation of runaway electron beams created during rapid shutdown experiments in DIII-D, Nuclear Fusion 53, 083004 (2013), DOI: 10.1088/00295515/53/8/083004.
58
Bibliography
[34] E. Marmar, A. Bader, M. Bakhtiari, H. Barnard, W. Beck, I. Bespamyatnov, A. Binus, P. Bonoli, B. Bose, M. Bitter, I. Cziegler, G. Dekow,
A. Dominguez, B. Duval, E. Edlund, D. Ernst, M. Ferrara, C. Fiore,
T. Fredian, A. Graf, R. Granetz, M. Greenwald, O. Grulke, D. Gwinn,
S. Harrison, R. Harvey, T. Hender, J. Hosea, K. Hill, N. Howard,
D. Howell, A. Hubbard, J. Hughes, I. Hutchinson, A. Ince-Cushman,
J. Irby, V. Izzo, A. Kanojia, C. Kessel, J. Ko, P. Koert, B. LaBombard,
C. Lau, L. Lin, Y. Lin, B. Lipschultz, J. Liptac, Y. Ma, K. Marr, M. May,
R. McDermott, O. Meneghini, D. Mikkelsen, R. Ochoukov, R. Parker,
C. Phillips, P. Phillips, Y. Podpaly, M. Porkolab, M. Reinke, J. Rice,
W. Rowan, S. Scott, A. Schmidt, J. Sears, S. Shiraiwa, A. Sips, N. Smick,
J. Snipes, J. Stillerman, Y. Takase, D. Terry, J. Terry, N. Tsujii, E. Valeo, R. Vieira, G. Wallace, D. Whyte, J. Wilson, S. Wolfe, G. Wright,
J. Wright, S. Wukitch, G. Wurden, P. Xu, K. Zhurovich, J. Zaks, and
S. Zweben, Overview of the Alcator C-Mod research program, Nuclear
Fusion 49, 104014 (2009), DOI: 10.1088/0029-5515/49/10/104014.
[35] F. Saint-Laurent, G. Martin, T. Alarcon, A. L. Luyer, P. B. Parks,
P. Pastor, S. Putvinski, C. Reux, J. Bucalossi, S. Bremond, and
P. Moreau, Overview of runaway electron control and mitigation experiments on Tore Supra and lessons learned in view of ITER, Fusion
Science and Technology 64, 711 (2013).
[36] Z. Y. Chen, W. C. Kim, Y. W. Yu, A. C. England, J. W. Yoo,
S. H. Hahn, S. W. Yoon, K. D. Lee, Y. K. Oh, J. G. Kwak, and
M. Kwon, Study of runaway current generation following disruptions
in KSTAR, Plasma Physics and Controlled Fusion 55, 035007 (2013),
DOI: 10.1088/0741-3335/55/3/035007.
[37] M. Vlainic, J. Mlynar, J. Cavalier, V. Weinzettl, R. Paprok, M. Imrisek, O. Ficker, M. Varavin, P. Vondracek,
and J.-M. Noterdaeme, Post-disruptive runaway electron beams in the COMPASS
tokamak, Journal of Plasma Physics 81, 475810506 (2015), DOI:
10.1017/S0022377815000914.
[38] G. Pautasso, P. McCarthy, C. Fuchs, R. Dux, S. Potzel, G. Papp,
M. Sertoli, G. Tardini, K. Lackner, A. Scarabosio, A.Mlynek, L. Giannone, the ASDEX Upgrade team, and the EUROfusion MST team,
Generation and suppression of runaway electrons in ASDEX Upgrade
disruptions, Europhysics Conference Abstracts 39E, P1.134 (2015),
http://ocs.ciemat.es/EPS2015PAP/pdf/P1.134.pdf.
59
Bibliography
[39] G. Papp, G. Pautasso, J. Decker, M. Gobbin, P. McCarthy, P. Blanchard, D. Carnevale, D. Choi, S. Coda, B. Duval, R. Dux, B. Erdős,
B. Esposito, O. Ficker, R. Fischer, C. Fuchs, C. Galperti, L. Giannone, A. Gude, B. Labit, K. Lackner, T. Lunt, L. Marelli, P. Martin, A. Mlynek, E. Nardon, M. Maraschek, P. Marmillod, M. Nocente,
Y. Peysson, P. Piovesan, V. Plyusnin, G. Pokol, P. Poloskei, S. Potzel,
C. Reux, F. Saint-Laurent, O. Sauter, B. Sieglin, U. Sheikh, C. Sommariva, W. Suttrop, G. Tardini, M. Teschke, D. Testa, W. Treutterer,
M. Valisa, ASDEX Upgrade Team, TCV Team, and the EUROfusion
MST1 Team, Runaway electron generation and mitigation on the european medium sized tokamaks ASDEX Upgrade and TCV, Proceedings
of the IAEA Fusion Energy Conference, Kyoto, Japan, EX/9–4 (2016).
[40] C. Barnes and J. Strachan, Sawtooth oscillations in the flux of runaway electrons to the PLT limiter, Nuclear Fusion 22, 1090 (1982), DOI:
10.1088/0029-5515/22/8/010.
[41] C. Paz-Soldan, N. W. Eidietis, R. Granetz, E. M. Hollmann, R. A.
Moyer, J. C. Wesley, J. Zhang, M. E. Austin, N. A. Crocker, A. Wingen,
and Y. Zhu, Growth and decay of runaway electrons above the critical
electric field under quiescent conditions, Physics of Plasmas 21, 022514
(2014), DOI: 10.1063/1.4866912.
[42] R. S. Granetz, B. Esposito, J. H. Kim, R. Koslowski, M. Lehnen, J. R.
Martin-Solis, C. Paz-Soldan, T. Rhee, J. C. Wesley, L. Zeng, and ITPA
MHD Group, An ITPA joint experiment to study runaway electron generation and suppression, Physics of Plasmas 21, 072506 (2014), DOI:
10.1063/1.4886802.
[43] J. R. Martı́n-Solı́s, B. Esposito, R. Sánchez, F. M. Poli, and L. Panaccione, Enhanced production of runaway electrons during a disruptive
termination of discharges heated with lower hybrid power in the Frascati Tokamak Upgrade, Physical Review Letters 97, 165002 (2006), DOI:
10.1103/PhysRevLett.97.165002.
[44] Y. Peysson, J. Decker, and R. W. Harvey, Advanced 3-D electron fokkerplanck transport calculations, AIP Conference Proceedings 694, 495
(2003), DOI: 10.1063/1.1638086.
[45] J. Decker and Y. Peysson, DKE: A fast numerical solver for the 3D drift
kinetic equation, Tech. Rep. EUR-CEA-FC-1736 (Euratom-CEA, 2004).
[46] E. Nilsson, J. Decker, Y. Peysson, R. S. Granetz, F. Saint-Laurent, and
M. Vlainic, Kinetic modelling of runaway electron avalanches in tokamak
60
Bibliography
plasmas, Plasma Physics and Controlled Fusion 57, 095006 (2015), DOI:
10.1088/0741-3335/57/9/095006.
[47] S. Chiu, M. Rosenbluth, R. Harvey, and V. Chan, Fokker-Planck
simulations mylb of knock-on electron runaway avalanche and bursts
in tokamaks, Nuclear Fusion 38, 1711 (1998), DOI: 10.1088/00295515/38/11/309.
[48] R. W. Harvey, V. S. Chan, S. C. Chiu, T. E. Evans, M. N. Rosenbluth,
and D. G. Whyte, Runaway electron production in DIII-D killer pellet
experiments, calculated with the CQL3D/KPRAD model, Physics of
Plasmas 7, 4590 (2000), DOI: 10.1063/1.1312816.
[49] E. M. Hollmann, P. B. Aleynikov, T. Fülöp, D. A. Humphreys, V. A.
Izzo, M. Lehnen, V. E. Lukash, G. Papp, G. Pautasso, F. SaintLaurent, and J. A. Snipes, Status of research toward the ITER disruption mitigation system, Physics of Plasmas 22, 021802 (2015), DOI:
10.1063/1.4901251.
[50] M. Rosenbluth and S. Putvinski, Theory for avalanche of runaway electrons in tokamaks, Nuclear Fusion 37, 1355 (1997), DOI: 10.1088/00295515/37/10/I03.
[51] ITER physics basis, Nuclear Fusion 39, 2137 (1999).
[52] http://www.iter.org.
[53] M. Lehnen, K. Aleynikova, P. Aleynikov, D. Campbell, P. Drewelow,
N. Eidietis, Y. Gasparyan, R. Granetz, Y. Gribov, N. Hartmann,
E. Hollmann, V. Izzo, S. Jachmich, S.-H. Kim, M. Kočan, H. Koslowski,
D. Kovalenko, U. Kruezi, A. Loarte, S. Maruyama, G. Matthews,
P. Parks, G. Pautasso, R. Pitts, C. Reux, V. Riccardo, R. Roccella,
J. Snipes, A. Thornton, and P. de Vries, Disruptions in ITER and
strategies for their control and mitigation, Journal of Nuclear Materials 463, 39 (2015), DOI: 10.1016/j.jnucmat.2014.10.075, proceedings
of the 21st International Conference on Plasma-Surface Interactions in
Controlled Fusion Devices Kanazawa, Japan May 26-30, 2014.
[54] P. Helander and D. J. Sigmar, Collisional Transport in Magnetized Plasmas (Cambridge University Press, 2002).
[55] F. F. Chen, Introduction to plasma physics and controlled fusion, 2nd
ed., Vol. 1 (Plenum Press, 1984).
[56] J. Connor and R. Hastie, Relativistic limitations on runaway electrons,
Nuclear Fusion 15, 415 (1975), DOI: 10.1088/0029-5515/15/3/007.
61
Bibliography
[57] B. Coppi, F. Pegoraro, R. Pozzoli, and G. Rewoldt, Slide-away distributions and relevant collective modes in high-temperature plasmas,
Nuclear Fusion 16, 309 (1976).
[58] G. Fussmann, On the motion of runaway electrons in momentum space,
Nuclear Fusion 19, 327 (1979), DOI: 10.1088/0029-5515/19/3/005.
[59] H. Smith, P. Helander, L.-G. Eriksson, and T. Fülöp, Runaway electron
generation in a cooling plasma, Physics of Plasmas 12, 122505 (2005),
DOI: 10.1063/1.2148966.
[60] J. R. Martı́n-Solı́s, J. D. Alvarez, R. Sánchez, and B. Esposito,
Momentum-space structure of relativistic runaway electrons, Physics of
Plasmas 5, 2370 (1998), DOI: 10.1063/1.872911.
[61] P. Aleynikov and B. N. Breizman, Theory of two threshold fields for relativistic runaway electrons, Physical Review Letters 114, 155001 (2015),
DOI: 10.1103/PhysRevLett.114.155001.
[62] C. Liu, D. P. Brennan, A. Bhattacharjee, and A. H. Boozer, Adjoint
fokker-planck equation and runaway electron dynamics, Physics of Plasmas 23, 010702 (2016), DOI: 10.1063/1.4938510.
[63] J. R. Martı́n-Solı́s, R. Sánchez, and B. Esposito, On the effect of synchrotron radiation and magnetic fluctuations on the avalanche runaway
growth rate, Physics of Plasmas 7, 3814 (2000), DOI: 10.1063/1.1287215.
[64] Y. A. Sokolov, ”Multiplication” of accelerated electrons in a tokamak,
JETP Letters 29, 218 (1979).
[65] N. Besedin and I. Pankratov, Stability of a runaway electron beam,
Nuclear Fusion 26, 807 (1986), DOI: 10.1088/0029-5515/26/6/009.
[66] R. Jayakumar, H. Fleischmann, and S. Zweben, Collisional avalanche
exponentiation of runaway electrons in electrified plasmas, Physics Letters A 172, 447 (1993), DOI: 10.1016/0375-9601(93)90237-T.
[67] A. V. Gurevich, On the theory of runaway electrons, Sov. Phys. JETP
12, 904 (1961).
[68] M. D. Kruskal and I. B. Bernstein, PPPL report MATT-Q-20, p. 174,
(1962).
[69] R. H. Cohen, Runaway electrons in an impure plasma, Physics of Fluids
19, 239 (1976), DOI: 10.1063/1.861451.
62
Bibliography
[70] S. Putvinski, P. Barabaschi, N. Fujisawa, N. Putvinskaya, M. N. Rosenbluth, and J. Wesley, Halo current, runaway electrons and disruption
mitigation in ITER, Plasma Physics and Controlled Fusion 39, B157
(1997), DOI: http://dx.doi.org/10.1088/0741-3335/39/12B/013.
[71] P. Helander, H. Smith, T. Fülöp, and L.-G. Eriksson, Electron kinetics in a cooling plasma, Physics of Plasmas 11, 5704 (2004), DOI:
10.1063/1.1812759.
[72] H. Smith, P. Helander, L.-G. Eriksson, D. Anderson, M. Lisak, and
F. Andersson, Runaway electrons and the evolution of the plasma current in tokamak disruptions, Physics of Plasmas 13, 102502 (2006), DOI:
10.1063/1.2358110.
[73] H. M. Smith and E. Verwichte, Hot tail runaway electron generation
in tokamak disruptions, Physics of Plasmas 15, 072502 (2008), DOI:
http://dx.doi.org/10.1063/1.2949692.
[74] H. M. Smith, T. Fehér, T. Fülöp, K. Gál,
and E. Verwichte, Runaway electron generation in tokamak disruptions, Plasma
Physics and Controlled Fusion 51, 124008 (2009), DOI: 10.1088/07413335/51/12/124008.
[75] D. Mosher, Interactions of relativistic electron beams with high
atomic-number plasmas, Physics of Fluids 18, 846 (1975), DOI:
10.1063/1.861219.
[76] Y. L. Igitkhanov, The effect of non-Coulomb scattering of relativistic electrons on the generation of runaways in multicomponent plasma, Contributions to Plasma Physics 52, 460 (2012), DOI:
10.1002/ctpp.201210034.
[77] K. Aleynikova, P. Aleynikov, S. Konovalov, A. Teplukhina, and
V. Zhogolev, Interaction of runaway electrons with high-Z impurities, Europhysics Conference Abstracts 37D, O5.103 (2013),
http://ocs.ciemat.es/EPS2013PAP/pdf/O5.103.pdf.
[78] J. R. Martı́n-Solı́s, A. Loarte, and M. Lehnen, Runaway electron dynamics in tokamak plasmas with high impurity content, Physics of Plasmas
22, 092512 (2015), DOI: 10.1063/1.4931166.
[79] L. Hesslow, Effect of Screened Nuclei on Fast Electron Beam Dynamics, Master’s thesis, Chalmers University of Technology (2016),
http://publications.lib.chalmers.se/records/fulltext/236294/236294.pdf.
63
Bibliography
[80] G. Papp, M. Drevlak, T. Fülöp, and P. Helander, Runaway electron
drift orbits in magnetostatic perturbed fields, Nuclear Fusion 51, 043004
(2011), DOI: 10.1088/0029-5515/51/4/043004.
[81] G. Papp, M. Drevlak, T. Fülöp, and G. I. Pokol, The effect of resonant
magnetic perturbations on runaway electron transport in ITER, Plasma
Physics and Controlled Fusion 54, 125008 (2012), DOI: 10.1088/07413335/54/12/125008.
[82] G. Papp, M. Drevlak, G. Pokol, and T. Fülöp, Energetic electron
transport in the presence of magnetic perturbations in magnetically confined plasmas, Journal of Plasma Physics 81, 475810503 (2015), DOI:
10.1017/S0022377815000537.
[83] H. Knoepfel and D. Spong, Runaway electrons in toroidal discharges,
Nuclear Fusion 19, 785 (1979), DOI: 10.1088/0029-5515/19/6/008.
[84] X. Guan, H. Qin, and N. J. Fisch, Phase-space dynamics of runaway
electrons in tokamaks, Physics of Plasmas 17, 092502 (2010), DOI:
10.1063/1.3476268.
[85] L. Zeng, H. R. Koslowski, Y. Liang, A. Lvovskiy, M. Lehnen, D. Nicolai, J. Pearson, M. Rack, H. Jaegers, K. H. Finken, K. Wongrach,
Y. Xu, and the TEXTOR team, Experimental observation of a
magnetic-turbulence threshold for runaway-electron generation in the
TEXTOR tokamak, Physical Review Letters 110, 235003 (2013), DOI:
10.1103/PhysRevLett.110.235003.
[86] G. Papp, P. W. Lauber, M. Schneller, S. Braun, H. R. Koslowski, and
the TEXTOR Team, Interaction of runaway populations with fast particle driven modes, Europhysics Conference Abstracts 38F, P2.032 (2014),
http://ocs.ciemat.es/EPS2014PAP/pdf/P2.032.pdf.
[87] P. Aleynikov and B. Breizman, Stability analysis of runaway-driven
waves in a tokamak, Nuclear Fusion 55, 043014 (2015), DOI:
10.1088/0029-5515/55/4/043014.
[88] Y. Liu, Y. Dong, X. Peng, C. Chen, Y. Zhang, J. Gao, O. Pan, and
X. Duan, Mitigation of runaway current with supersonic molecular beam
injection on HL-2A tokamak, 26th IAEA Fusion Energy Conference,
Kyoto, Japan, EX/9–3 (2016).
[89] T. Fülöp, G. Pokol, P. Helander, and M. Lisak, Destabilization of
magnetosonic-whistler waves by a relativistic runaway beam, Physics
of Plasmas 13, 062506 (2006), DOI: 10.1063/1.2208327.
64
Bibliography
[90] G. Pokol, T. Fülöp, and M. Lisak, Quasi-linear analysis of whistler
waves driven by relativistic runaway beams in tokamaks, Plasma
Physics and Controlled Fusion 50, 045003 (2008), DOI: 10.1088/07413335/50/4/045003.
[91] A. Kómár, G. I. Pokol, and T. Fülöp, Electromagnetic waves destabilized by runaway electrons in near-critical electric fields, Physics of
Plasmas 20, 012117 (2013), DOI: 10.1063/1.4776666.
[92] A. Kómár, G. I. Pokol, and T. Fülöp, Interaction of electromagnetic waves and suprathermal electrons in the near-critical electric field
limit, Journal of Physics: Conference Series 401, 012012 (2012), DOI:
10.1088/1742-6596/401/1/012012.
[93] L. D. Landau, Kinetic equation for the Coulomb effect, Physikalische
Zeitschrift der Sowjetunion 10, 154 (1936).
[94] M. N. Rosenbluth, W. M. MacDonald, and D. L. Judd, Fokker-Planck
equation for an inverse-square force, Physical Review 107, 1 (1957),
DOI: 10.1103/PhysRev.107.1.
[95] S. T. Beliaev and G. I. Budker, The relativistic kinetic equation, Soviet
Physics-Doklady 1, 218 (1956).
[96] B. J. Braams and C. F. F. Karney, Differential form of the collision integral for a relativistic plasma, Physical Review Letters 59, 1817 (1987),
DOI: 10.1103/PhysRevLett.59.1817.
[97] B. J. Braams and C. F. F. Karney, Conductivity of a relativistic
plasma, Physics of Fluids B: Plasma Physics 1, 1355 (1989), DOI:
10.1063/1.858966.
[98] O. Embréus, Kinetic modelling of runaway in plasmas, Licentiate
thesis,
Chalmers
University
of
Technology
(2016),
http://publications.lib.chalmers.se/records/fulltext/236745/236745.pdf.
[99] C. Møller, Zur theorie des durchgangs schneller elektronen
durch materie, Annalen der Physik 406, 531 (1932), DOI:
10.1002/andp.19324060506.
[100] J. D. Jackson, Classical Electrodynamics, 3rd ed. (Wiley & Sons, 1999).
[101] F. R. Elder, A. M. Gurewitsch, R. V. Langmuir, and H. C. Pollock,
Radiation from electrons in a synchrotron, Physical Review 71, 829
(1947), DOI: 10.1103/PhysRev.71.829.5.
65
Bibliography
[102] E.-E. Koch, Handbook on Synchrotron Radiation, Vol. 1a (North Holland, 1983).
[103] G. A. Schott, Electromagnetic Radiation (Cambridge University Press,
1912).
[104] J. Schwinger, On the classical radiation of accelerated electrons, Physical
Review 75, 1912 (1949), DOI: 10.1103/PhysRev.75.1912.
[105] G. Bekefi, Radiation Processes in Plasmas (Wiley & Sons, 1966).
[106] L. D. Landau and E. M. Lifshitz, The Classical Theory of Fields, 4th ed.,
Course of Theoretical Physics, Vol. 2 (Butterworth-Heinemann, 1975).
[107] K. Finken, J. Watkins, D. Rusbuldt, W. Corbett, K. Dippel, D. Goebel,
and R. Moyer,
Observation of infrared synchrotron radiation from tokamak runaway
electrons in TEXTOR,
Nuclear Fusion 30, 859 (1990), DOI: 10.1088/0029-5515/30/5/005.
[108] K. Wongrach, K. Finken, S. Abdullaev, R. Koslowski, O. Willi, L. Zeng,
and the TEXTOR Team, Measurement of synchrotron radiation from
runaway electrons during the TEXTOR tokamak disruptions, Nuclear
Fusion 54, 043011 (2014), DOI: 10.1088/0029-5515/54/4/043011.
[109] J. H. Yu, E. M. Hollmann, N. Commaux, N. W. Eidietis, D. A.
Humphreys, A. N. James, T. C. Jernigan, and R. A. Moyer, Visible
imaging and spectroscopy of disruption runaway electrons in DIII-D,
Physics of Plasmas 20, 042113 (2013), DOI: 10.1063/1.4801738.
[110] B. Esposito, L. Boncagni, P. Buratti, D. Carnevale, F. Causa, M. Gospodarczyk, J. Martin-Solis, Z. Popovic, M. Agostini, G. Apruzzese, W. Bin,
C. Cianfarani, R. D. Angelis, G. Granucci, A. Grosso, G. Maddaluno,
D. Marocco, V. Piergotti, A. Pensa, S. Podda, G. Pucella, G. Ramogida,
G. Rocchi, M. Riva, A. Sibio, C. Sozzi, B. Tilia, O. Tudisco, M. Valisa, and FTU Team, Runaway electron generation and control, Plasma
Physics and Controlled Fusion 59, 014044 (2017), DOI: 10.1088/07413335/59/1/014044.
[111] R. J. Zhou, L. Q. Hu, E. Z. Li, M. Xu, G. Q. Zhong, L. Q. Xu, S. Y. Lin,
J. Z. Zhang, and the EAST Team, Investigation of ring-like runaway
electron beams in the EAST tokamak, Plasma Physics and Controlled
Fusion 55, 055006 (2013), DOI: 10.1088/0741-3335/55/5/055006.
[112] I. M. Pankratov, Analysis of the synchrotron radiation spectra of runaway electrons, Plasma Physics Reports 25, 145 (1999).
66
Bibliography
[113] I. M. Pankratov, I. V. Pavlenko, and O. A. Pomazan, Analysis of synchrotron radiation emitted by runaway electrons in tokamaks, Journal of
Kharkiv National University 1059, 39 (2013), series ”Nuclei, Particles,
Fields”.
[114] I. M. Pankratov, Analysis of the synchrotron radiation emitted by runaway electrons, Plasma Physics Reports 22, 535 (1996).
[115] R. Jaspers, N. J. L. Cardozo, A. J. H. Donné, H. L. M. Widdershoven,
and K. H. Finken, A synchrotron radiation diagnostic to observe relativistic runaway electrons in a tokamak plasma, Review of Scientific
Instruments 72, 466 (2001), DOI: 10.1063/1.1318245.
[116] R. J. Zhou, I. M. Pankratov, L. Q. Hu, M. Xu, and J. H. Yang, Synchrotron radiation spectra and synchrotron radiation spot shape of runaway electrons in Experimental Advanced Superconducting Tokamak,
Physics of Plasmas 21, 063302 (2014), DOI: 10.1063/1.4881469.
[117] F. Andersson, P. Helander, and L.-G. Eriksson, Damping of relativistic
electron beams by synchrotron radiation, Physics of Plasmas 8, 5221
(2001), DOI: 10.1063/1.1418242.
[118] M. Bakhtiari, G. J. Kramer, M. Takechi, H. Tamai, Y. Miura,
Y. Kusama, and Y. Kamada, Role of bremsstrahlung radiation in limiting the energy of runaway electrons in tokamaks, Physical Review Letters 94, 215003 (2005), DOI: 10.1103/PhysRevLett.94.215003.
[119] W. Pauli, Theory of Relativity (Dover, New York, 1981).
[120] R. Hazeltine and S. Mahajan, Radiation reaction in fusion plasmas, Physical Review E 70, 046407 (2004), DOI: 10.1103/PhysRevE.70.046407.
[121] L.-G. Eriksson, P. Helander, F. Andersson, D. Anderson, and M. Lisak,
Current dynamics during disruptions in large tokamaks, Physical Review
Letters 92, 205004 (2004), DOI: 10.1103/PhysRevLett.92.205004.
[122] K. Gál, T. Fehér, H. Smith, T. Fülöp, and P. Helander, Runaway
electron generation during plasma shutdown by killer pellet injection,
Plasma Physics and Controlled Fusion 50, 055006 (2008).
[123] T. Fehér, H. M. Smith, T. Fülöp, and K. Gál, Simulation of runaway
electron generation during plasma shutdown by impurity injection in
ITER, Plasma Physics and Controlled Fusion 53, 035014 (2011).
67
Included papers (A–F)
Paper A
A. Stahl, M. Landreman, G. Papp, E. Hollmann, and T. Fülöp,
Synchrotron radiation from a runaway electron distribution in tokamaks,
Physics of Plasmas 20, 093302 (2013).
http://dx.doi.org/10.1063/1.4821823
http://arxiv.org/abs/1308.2099
PHYSICS OF PLASMAS 20, 093302 (2013)
Synchrotron radiation from a runaway electron distribution in tokamaks
€ p1
€ lo
A. Stahl,1 M. Landreman,2 G. Papp,1,3 E. Hollmann,4 and T. Fu
1
Department of Applied Physics, Nuclear Engineering, Chalmers University of Technology and Euratom-VR
Association, SE-412 96 G€
oteborg, Sweden
2
Plasma Science and Fusion Center, MIT, Cambridge, Massachusetts 02139, USA
3
Department of Nuclear Techniques, Budapest University of Technology and Economics,
Association EURATOM, H-1111 Budapest, Hungary
4
Center for Energy Research, University of California, San Diego, La Jolla, California 92093-0417, USA
(Received 9 August 2013; accepted 16 August 2013; published online 19 September 2013)
The synchrotron radiation emitted by runaway electrons in a fusion plasma provides information
regarding the particle momenta and pitch-angles of the runaway electron population through the
strong dependence of the synchrotron spectrum on these parameters. Information about the
runaway density and its spatial distribution, as well as the time evolution of the above quantities,
can also be deduced. In this paper, we present the synchrotron radiation spectra for typical
avalanching runaway electron distributions. Spectra obtained for a distribution of electrons are
compared with the emission of mono-energetic electrons with a prescribed pitch-angle. We also
examine the effects of magnetic field curvature and analyse the sensitivity of the resulting
spectrum to perturbations to the runaway distribution. The implications for the deduced runaway
electron parameters are discussed. We compare our calculations to experimental data from DIII-D
and estimate the maximum observed runaway energy. [http://dx.doi.org/10.1063/1.4821823]
I. INTRODUCTION
Understanding the process of runaway beam formation
and loss in tokamaks is of great importance, due to the potentially severe damage these electrons may cause in
disruptions. In present tokamaks, runaway electrons have
energies between a few hundred keV and tens of MeV, and
in a next-step device like ITER, they are projected to reach a
maximum energy of up to 100 MeV.1 Runaway electrons
emit synchrotron radiation,2–5 the spectrum of which
depends on the velocity-space distribution of the radiating
particles. Therefore, the spectrum can be used to obtain
information about the departure of the velocity distribution
from isotropy and about the energy of the particles. The
emitted radiation can also be an energy loss mechanism,6
although in tokamaks this loss is not appreciable unless the
electrons have very large energies, above 70 MeV.3
Many theoretical studies of the synchrotron radiation of
the energetic population have been done before, either using
approximate electron distribution functions or assuming
straight magnetic field lines.7–9 In several studies, the synchrotron emission from a single particle is used as an approximation for the entire runaway distribution,4,5 using a
specific momentum and pitch-angle for the electrons, often
identified as the maximum momentum and pitch-angle of the
electrons in the runaway beam. In the present work, we use
an electron distribution function typical of avalanching runaway electron populations in tokamak disruptions. As we
will show, taking into account the whole distribution is
important, since synchrotron radiation diagnostics based on
single particle emission can give misleading results.
Furthermore, we will illustrate that synchrotron radiation can
be used to detect signs of modification of the electron distribution, which can occur due to for instance wave-particle
interaction.
1070-664X/2013/20(9)/093302/9/$30.00
The structure of the paper is as follows. In Sec. II, we
give several expressions for the radiated synchrotron power,
including the effect of field-curvature. We also discuss the
applicability of these expressions in different contexts.
Section III is devoted to the analysis of the synchrotron radiation spectrum from an avalanching runaway electron distribution. We will describe the parametric dependences on
magnetic field, density, temperature, effective charge, and
electric field. In Sec. IV, we discuss the potential use of synchrotron radiation as a diagnostic. We also present a comparison between the synchrotron spectrum calculated for the
avalanching runaway electron distribution and an experimentally measured synchrotron spectrum from DIII-D. Our
conclusions will be summarized in Sec. V.
II. SYNCHROTRON EMISSION FORMULAS
The power radiated by an electron with Lorentz factor
c 1 at wavelength k in the case of straight magnetic field
lines is10
ð
1 ce2 1
P cyl ðkÞ ¼ pffiffiffi 3
K5=3 ðlÞdl ;
(1)
3 0 k c2 kc =k
where e is the electron charge, c is the speed of light, 0 is
the
vacuum permittivity, kc ¼ ð4pcme ck Þ=ð3eBc2 Þ; cjj ¼ 1=
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 v2jj =c2 ; me is the electron rest mass, B is the magnetic field, k denotes the component along the magnetic
field, and K ðxÞ is the modified Bessel function of the second kind. The radiation is emitted in a narrow beam in
the parallel direction due to relativistic effects.10 In a
tokamak, the effects of magnetic field line curvature and
curvature drift have to be taken into account. This has
been done in Ref. 11, where the following expression was
obtained:
20, 093302-1
093302-2
P full ðkÞ ¼
Stahl et al.
(ð
Phys. Plasmas 20, 093302 (2013)
dy
3
1
ð1 þ 2y2 ÞJ0 ðay3 Þsin n y þ y3
2
3
0 y
ð
4g 1
3
1
p
dyyJ 0 0 ðay3 Þcos n y þ y3
;
2
1þg 0
2
3
2
(2)
ce2
e0 k3 c2
1
where a ¼ ng=ð1 þ g2 Þ,
4p
R
pffiffiffiffiffiffiffiffiffiffiffiffiffi ;
3 kc3 1 þ g2
(3)
eBR v? xc R v?
’
;
cme v2k
cc vk
(4)
n¼
g¼
R is the tokamak major radius, J ðxÞ is the Bessel function,
and J0 ðxÞ its derivative. The integrands in Eq. (2) are highly
oscillatory and the calculation of synchrotron spectra can
become computationally heavy. This motivates examining
more approximate formulas which are less complex, especially when considering possible diagnostic applications. In
Eqs. (21) and (26) of Ref. 11, two limits of Eq. (2) are given.
These two limits are obtained by first expanding in n 1,
which can be translated
pffiffiffiffiffiffiffiffiffiffiffiffiffito a condition for the wavelength
k ð4p=3ÞR=ðc3 1 þ g2 Þ. Then, to obtain the first of the
two expressions, Eq. (2) is also expanded in the smallness of
the argument of the Bessel functions, leading to the condition ng ⱗ 1 þ g2 . The resulting approximative formula is
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffi
ce2 2 1 þ g2 n
4g
I
ðaÞ
þ
I
ðaÞ
; (5)
P as1 ðkÞ e
0
1
40
1 þ g2
k5 Rc
where I ðxÞ is the modified Bessel function. P as1 was the
expression used to calculate the synchrotron radiation of an
avalanching population of positrons in Ref. 12 and in fitting
of the synchrotron spectrum in the optical range in DIII-D in
Ref. 5. The two conditions required for validity of Eq. (5)
can be summarized as g=ð1 þ g2 Þⱗ1=n 1, which leads to
a rather narrow validity range for P as1 . Figures 1(a) and 1(b)
show the range of wavelengths for which P as1 is valid
(kl ⱗ k ku ) for different runaway momenta in DIII-D-size
and ITER-size tokamaks, respectively. Note that the wavelength should be much smaller than the solid line(s) in the
figure for P as1 to be valid. It is clear that for wavelengths in
the 0.1–1 lm range (as in the measurements described in
Ref. 5), the approximative formula P as1 is only valid for
particles with large normalized momenta p ¼ cv=c, and not
necessarily for all values of v? =vk .
To obtain the second limit of Eq. (2) (Eq. (26) in Ref. 11),
k ð4p=3ÞRg=½c3 ð1 þ gÞ3 has to be fulfilled. Equation (2) then simplifies to
pffiffiffi 2
4p R 1
3 ce c ð1 þ gÞ2
:
exp
P as2 ðkÞ ¼
p
ffiffi
ffi
3 kc3 1 þ g
g
8p 0 k2 R
(6)
(7)
The condition in Eq. (6) is more strict than the one stemming
from n 1; it is only necessary to fulfill Eq. (6) for Eq. (7)
to be valid. Figures 2(a) and 2(b) show the upper bound for
the wavelength given by Eq. (6). We conclude that for the
visible part of the spectrum, P as2 could be a suitable approximative formula for runaway electron beams with p < 50 and
v? =vjj < 0:1. In the opposite case, when p and v? =vjj are
large then either the full expression P full , or in some cases
P as1 , can be used.
In general, the difference between the emitted power
given by P cyl (valid in the cylindrical limit) and P full (including field line curvature) is not very large if we consider only
emission by a single particle. Single particle synchrotron
spectra calculated by P cyl and P full , as well as the approximate formulas P as1 and P as2 are shown in Figs. 3(a) and 3(b)
for particles with normalized momentum p ¼ 50 (corresponding to a particle energy of roughly 25 MeV) and
v? =vk ¼ 0:1 in two different tokamaks. For such particles,
the peak emission is for wavelengths of a few lm (the near
infrared part of the spectrum). Figure 3(a) shows that for
medium-sized tokamaks (such as DIII-D), P full is closely
approximated by P as2 . This is not surprising, as P as2 is valid
in most of the wavelength range considered (especially for
shorter wavelengths), whereas P as1 is only valid for longer
wavelengths for these parameters. For large tokamaks (such
as ITER), P full is best approximated by P cyl , as the effects of
field curvature become small for such large major radii.
Figure 3(b) shows that P as2 is not a good approximation in
this case, which is expected, since P as2 is not valid in this
region.
Figures 3(c) and 3(d) investigate the energy dependence
of the above conclusions. The quantity plotted is
log10 ðP i ðkÞÞ. Figure 3(c) confirms that P as2 is a good
approximation to P full in DIII-D for a wide range of runaway
energies. For the highest energies, agreement is still very good
FIG. 1. Upper and lower bounds on the
wavelength k for which P as1 is valid.
Note the logarithmic scale on the vertical axis. The parameters used are
(a) B ¼ 2:1 T and R ¼ 1:67 m and (b)
B ¼ 5:3 T and R ¼ 6 m.
093302-3
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
FIG. 2. Upper bounds on the wavelength k for which P as2 is valid. Note
the logarithmic scale on the vertical axis.
The parameters used are (a) B ¼ 2:1 T
and R ¼ 1:67 m and (b) B ¼ 5:3 T and
R ¼ 6 m.
for short wavelengths, but less so for longer wavelengths.
This agrees with Fig. 2(a), which indicates that P as2 is no longer valid for high energies and long wavelengths. Figure 3(c)
also shows that for a tokamak this size, the difference between
P cyl and P full increases with p, and using P cyl is not recommended if quantitative agreement is sought. In an ITER-like
device, however, Fig. 3(d) indicates that P cyl approximates
P full very well over the whole energy range considered.
Formally, P full reduces to P cyl when R ! 1 and ck ’ c/v?
(where this latter relation is equivalent to cv? / c 1).
III. SPECTRUM FROM RUNAWAY ELECTRON
DISTRIBUTIONS
In Refs. 4 and 5, the synchrotron spectrum is calculated by multiplying the single particle spectrum by the
number of runaways with a specific pitch-angle and
momentum. In this section, we investigate how the synchrotron spectrum changes if we take into account the
whole runaway electron distribution instead of the single
particle approximation considered above. We calculate the
synchrotron emission integrated over a runaway electron
distribution using
PðkÞ ¼
2p
nr
ð
fRE ðp; vÞ P i ðp; v; kÞ p2 dp dv ;
(8)
Rr
where fRE is the runaway distribution function, P i is one
of the single particle emission formulas discussed in Sec. II,
v ¼ pk =p is the cosine of the pitch-angle, and nr is the runaway electron density. The runaway region of momentum
space Rr is defined by a separatrix ps ¼ ðE 1Þ1=2 such that
all particles with p > ps are considered runaways.13 Here,
E ¼ Ejj =Ec is the parallel electric field Ejj normalized to the
critical field Ec ¼ me c=ðesÞ, with s ¼ ð4pre2 ne c ln KÞ1 the
collision time for relativistic electrons, re the classical electron radius, ne the electron density, and lnK the Coulomb
logarithm. As we normalize to nr ; PðkÞ is the average emission per runaway. The alternative choice of normalizing by
the runaway current Ir was also considered, and it was found
that all results presented below are essentially unchanged
aside from an overall scale factor, since the speed of all runaways is nearly c.
In large tokamak disruptions, secondary runaway generation is expected to dominate over primary generation, in
which case the runaway distribution will grow approximately
FIG. 3. Single particle synchrotron
emission from different emission formulas. (a) and (b) show emitted spectra
for particles with v? =vk ¼ 0:1 and
p ¼ 50 and tokamak parameters corresponding to (a) DIII-D and (b) ITER.
The solid (blue) line corresponds to
the expression including the field-line
curvature, P full . The dotted (black) line
is the cylindrical limit, P cyl . The
dashed-dotted (red) and dashed (green)
lines correspond to the approximative
expressions P as1 and P as2 , respectively. (c) and (d) show contours of
log10 ðP i ðkÞÞ (with P i in units of
W=lm) for various particle momenta
and compares (c) P cyl ; P full , and P as2
and (b) P cyl and P full .
093302-4
Stahl et al.
exponentially in time: @fRE =@t / fRE . In this case of exponential growth, the electron distribution can be approximated
by14
!
^ 2
pk
nr E^
Ep
?
exp fRE ðpk ; p? Þ ¼
; (9)
cz lnK 2pk
2pcz pk lnK
where E^ ¼ ðE 1Þ=ð1 þ Zeff Þ; Zeff is the effective ion
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
charge and cz ¼ 3ðZeff þ 5Þ=p, and the momentum space
coordinates are related to p and v through pjj ¼ pv and
pffiffiffiffiffiffiffiffiffiffiffiffiffi
p? ¼ p 1 v2 . Derivation of Eq. (9) assumes strong anisotropy (p? pjj ) and high electric field (E 1). In addition to the lower boundary p ¼ ps of the runaway region, an
upper cut-off p ¼ pmax of the distribution will be introduced.
This cut-off is physically motivated by the finite life-time of
the accelerating electric field and the presence of loss mechanisms, such as radiation and radial transport.
As it was shown in Sec. II, the inclusion of field curvature effects via the use of P full rather than P cyl had little effect
on the synchrotron emission of a single particle in an ITERsized device. The effect is larger in smaller devices. When
the complete runaway distribution is taken into account, these
conclusions still hold. Figure 4 shows synchrotron spectra
calculated using Eq. (8) together with the distribution in
Eq. (9) and the emission formulas P cyl ; P full ; P as1 , and P as2 .
The calculation was performed for both a DIII-D-size and an
ITER-size device, as the field curvature is what separates the
different formulas. The parameters used in the calculation in
Fig. 4 are maximum normalized momentum pmax ¼ 100 (corresponding to a maximum runaway energy of roughly
50 MeV), parallel electric field Ek ¼ 2 V=m, effective charge
Zeff ¼ 1, background electron density ne ¼ 3 1020 m3 ,
and background plasma temperature T ¼ 10 eV. The relatively low temperature is what is expected after a thermal
quench in a disruption. In DIII-D, the post thermal-quench
temperature is estimated to be as low as T ¼ 2 eV.15
Figure 4(a) shows that in DIII-D, P full is well approximated by P as2 , especially in the short wavelength slope
region of the spectrum. In ITER, P cyl is a good approximation, as shown in Fig. 4(b). This is expected since the field
curvature is much smaller here. These results are consistent
with the conclusion regarding single particles in Fig. 3. For
simplicity, throughout the remainder of this paper, we will
use P cyl when calculating synchrotron spectra (except for the
comparison with DIII-D data in Sec. IV B). Synchrotron
spectra calculated by P cyl and P full are qualitatively similar
Phys. Plasmas 20, 093302 (2013)
for both small and large machines and are also often quantitatively similar for large machines.
The single particle synchrotron emission formulas are independent of the plasma temperature, effective charge, density, and the strength of the electric field. These quantities do
however affect the shape of the runaway distribution, which
in turn affects the synchrotron emission. Figure 5 shows scans
in these parameters, the magnetic field, and maximum momentum pmax of the distribution. The baseline scenario corresponds to the parameters used in Fig. 4 together with
B ¼ 3 T. Since P cyl is used, there is no dependence on R.
Figure 5 shows that the average synchrotron emission
increases with B; T; Zeff ; ne , and pmax , but decreases with
increasing electric field strength. The dependence on ne and
E is particularly strong, and we note that the average emission can vary over several orders of magnitude. This variation
is completely missing from the single particle approximation
used in Sec. II. If, as a disruption mitigation technique, a
large amount of material is injected into the plasma (for
instance in the form of a massive gas injection), the increase
in density would lead to increased synchrotron emission from
the runaways (if the mitigation is unsuccessful). This could
give the impression of an increase in the number of runaways
even though this is not necessarily the case. The figure also
shows that the wavelength of peak emission shifts appreciably with varying parameter values. Generally, an increased
average emission is accompanied by a shift of the peak emission towards shorter wavelengths. The total synchrotron
emission of a single particle scales roughly as ðcv? =vk Þ2 .4
Thus, the most strongly emitting particles are highly energetic with large pitch-angle. These particles emit at shorter
wavelengths, so the shift of the wavelength of peak emission
with increased total emission is expected.
In light of the particle energy dependence of the emitted
synchrotron power, the decrease in emission with increasing
electric field strength may seem a little surprising, as a stronger accelerating field leads to more highly energetic particles. The explanation can be found in the shape of the
runaway beam. Figure 6 shows the runaway distribution,
Eq. (9), in (pk ; p? )-space for three of the parameter sets in
the electric field scan in Fig. 5(b). The figure shows that the
distribution, in addition to being extended in pk , becomes
more narrow in p? as the electric field strength increases.
This leads to lower average-per-particle emission by virtue
of the pitch-angle dependence of P cyl , despite the presence
of a greater number of highly energetic particles.
FIG. 4. Comparison of the synchrotron
spectrum from a runaway distribution
(Eq. (9)), as calculated using P cyl ;
P full ; P as1 , or P as2 . Normalizing the
emitted power by the runaway current
Ir instead of by nr gives negligible difference in these figures or any figures
below (the curves are not even distinguishable), since most runaways move
at speed c.
093302-5
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
FIG. 5. Synchrotron spectra calculated using Eq. (8) together with P cyl and Eq. (9). Note that the spectra are normalized to the runaway density. Unless otherwise noted, the parameters are pmax ¼ 100; Ek ¼ 2 V=m; Zeff ¼ 1; ne ¼ 3 1020 m3 ; T ¼ 10 eV, and B ¼ 3 T. For this scenario, Ec ¼ 0:15 V=m.
FIG. 6. Shape of the analytical avalanche distribution (Eq. (9)) for three
of the parameter sets in Fig. 5(b). The
plot shows contours of the quantity
log10 jfRE =nr j.
093302-6
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
FIG. 7. Synchrotron spectra (average emission per particle) calculated using the runaway distribution in Eq. (9) and P cyl for DIII-D-like and ITER-like cases.
The synchrotron spectrum from a single particle with p ¼ 100 and v? =vjj ¼ 0:15 is also shown. Note that the single particle spectra have been multiplied by a
small factor to fit on the same scale. The parameters used for the distributions are pmax ¼ 100 and A: Ek ¼ 2 V=m; Zeff ¼ 1; ne ¼ 5 1019 m3 ; T ¼ 2 eV; B :
Ek ¼ 10 V=m; Zeff ¼ 1:5; ne ¼ 1 1020 m3 ; T ¼ 2 eV; C : Ek ¼ 2 V=m; Zeff ¼ 1; ne ¼ 5 1020 m3 ; T ¼ 10 eV; D : Ek ¼ 10 V=m; Zeff ¼ 2; ne ¼ 1 1021 m3 ;
T ¼ 10 eV.
Figure 7 shows a comparison of the average synchrotron
spectrum calculated for the runaway distribution Eq. (9) and
for a single particle. The figure clearly shows that using the
single-particle emission overestimates the synchrotron emission per particle by several orders of magnitude. (Note that
the values for the emitted power per particle were divided by
a large number to fit in the same scale.) The overestimation
is caused by the fact that the single-particle approximation
assumes that all particles emit as much synchrotron radiation
as the most strongly emitting particle in the actual distribution, as discussed in Sec. I. Furthermore, the wavelength of
peak emission is shifted towards shorter wavelengths when
using this approximation. Using the single-particle approximation can thus give misleading results regarding both the
spectrum shape and the total emission strength.
IV. SYNCHROTRON RADIATION AS A RUNAWAY
ELECTRON DIAGNOSTIC
The interest in the synchrotron emission of runaways is
primarily motivated by its potential as a runaway diagnostic.
In principle, the distribution can be determined by acquiring
an experimental synchrotron spectrum and comparing it to
calculations using Eq. (8) for a range of pmax , provided all
other relevant parameters are known. There are however several problems with this approach. First, the complete synchrotron spectrum is not known. Detectors are only sensitive
in a limited wavelength range, which is likely to also contain
contaminating radiation from other sources in the plasma.
Second, the relevant plasma parameters are not always well
known, especially during disruptions. This can lead to significant uncertainty in the computed synchrotron spectrum, as
the parameter scans in Fig. 5 indicated. Using a single particle approximation for the runaway distribution seemingly
avoids the second issue, but as we have seen, it also ignores
factors that can influence the emission by orders of
magnitude.
A. Spectrum slope and maximum runaway energy
Simple measurements of the synchrotron power for different wavelengths on the steep slope of the spectrum have
been used to estimate the runaway energy,4 using the single
particle emission formulas and assuming mono-energetic
runaways with well-defined pitch-angle. In this case, there is
a monotonic relationship between the slope and the particle
energy (as the wavelength of peak emission decreases
monotonically with increasing p). The slope can be obtained
through a relative measurement of the synchrotron power at
two wavelengths, S ¼ Pðk1 Þ=Pðk2 Þ. However, as the runaway distribution is sensitive to the plasma parameters,
when taking it into account there is in general no such simple
relationship between the slope of the spectrum and the maximum runaway energy in the distribution. If all other parameters are fixed the relation still holds, as is shown in Fig. 8(a).
This follows naturally from the relation for single particles,
as when pmax is increased, more particles that emit at short
wavelengths are included, and the average emission correspondingly shifts towards shorter wavelengths, affecting the
slope. But if the plasma parameters are uncertain, the slope
FIG. 8. Spectra calculated using the analytical avalanche distribution Eq. (9)
and P cyl . In (a), the parameters used are
the same as the baseline scenario in
Fig. 5, but with different maximum particle momenta. All the curves in panel
(b) have the same slope S, as calculated
with k1 ¼ 1:5 lm, k2 ¼ 2:8 lm. The
plasma parameters that differ between
the curves are indicated in the figure.
The remaining parameter values are
ne ¼ 3 1020 m3 and B ¼ 3 T.
093302-7
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
can be misleading. Figure 8(b) shows multiple spectra with
the same slope S for k1 ¼ 1:5 lm and k2 ¼ 2:8 lm. Using
only a measurement of S in the above range, they cannot be
distinguished, despite the appreciable difference in average
emission. This type of two-point slope measurement can be
performed using physical wavelength filters placed in front
of the detector,4 in which case measurements are constrained
to specific k1 and k2 that cannot be easily changed. The pmax
of the different spectra in Fig. 8(b) range from 50 to 90, with
only modest variation of the plasma parameters E; Zeff , and
T (all of which are hard to estimate during disruptions).
Thus, if the plasma properties are uncertain, there is no clear
correlation between S and pmax of the distribution. Another
weakness of using the slope is the difficulty in asserting that
both measurement points are actually located on the approximately linear part of the spectrum. As the plasma parameters
change, the peak of the spectrum may shift (as discussed in
connection with Fig. 5). Choosing k1 and k2 that are suitable
for a wide range of different conditions (as when using physical filters) is not easy. Instead of using the slope directly,
one should calculate the emission for an assumed beam-like
distribution function (e.g., similar to Eq. (9)), and iteratively
find the pmax, which fits the synchrotron spectrum best.
B. Synchrotron emission in DIII-D
It is interesting to investigate how a synchrotron spectrum calculated for an avalanching distribution compares
with an experimentally measured synchrotron spectrum from
DIII-D. In the specific experimental scenario we consider
(shot number 146 704 and time t ¼ 2290 ms16), the loop
voltage is 7 V, the density 3:9 1019 m3 , and the plasma
current Ip ¼ 0:15 MA, measured near the end of a runaway
plateau phase. The runaway density can be estimated from
the current using nr ¼ Ip =ðecAre Þ, where Are is the area of
the runaway beam. The runaway beam radius in this case
was around 20 cm. The temperature is assumed to be 1:5 eV
and Zeff ¼ 1. For synchrotron emission by mono-energetic
runaway electrons, the conversion to the measured brightness can be done using Eq. (2) in Ref. 5
Bðk; h; cÞ ¼ Pðk; h; cÞ
2R
nr ;
ph
(10)
where R is the major radius (of the runaway beam) and
h ¼ v? =vk is the tangent of the particle pitch-angle. Taking
into account the runaway distribution, we calculate the
brightness as
ð vmax ð pmax
1
BðkÞ ¼ 4R
Pðk;hðvÞ;cðpÞÞf ðp;vÞp2 dpdv; (11)
vmin pmin hðvÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffi
where hðvÞ ¼ tanðarccosðvÞÞ ¼ 1 v2 =v and cðpÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ p2 þ 1; pmin ¼ ðE 1Þ1=2 and the integration limits
for the pitch-angle are vmin ¼ 0, vmax ¼ 1. Since we consider
the visible part of the spectrum, all pmin below p ¼ 50 produce identical results, as only the highest energy particles
emit in this range. Equation (10) is strictly valid for
1=c h.5 As we are interested in the complete distribution
with both small c and small h, we use instead the effective
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
viewing aperture heff h2 þ c2 þ ðrlens =r0 Þ2 . Here,
rlens ¼ 2 cm is the lens aperture of the detector and r0 ’ 2 m
is the distance between the detector and the runaway beam.
Introducing heff into Eq. (11), we find
ð
1
Pðk; v; pÞ f ðp; vÞ p2 dpdv :
BðkÞ ¼ 4R
(12)
Rr heff ðp; vÞ
Figure 9 shows a comparison of spectra calculated using
Eq. (12) together with P as2 and Eq. (9), and the experimentally measured spectra for different runaway beam radii (the
beam is assumed to have circular cross-section), pmax, loop
voltages, and densities. The good agreement for rre ¼ 20 cm
and pmax ¼ 130 leads us to estimate the maximum runaway
electron energy to be around 65 MeV. This is much larger
than the mean energy of several MeV estimated from other
diagnostics.16
For comparison, we also fit the experimental data with
synchrotron spectra from a mono-energetic runaway population (using Eq. (10)), for different particle energies and
pitch-angles. As in Ref. 5, we assume that 1% of the runaway
population (calculated with rre ¼ 20 cm) has the specific
energy considered. The results are shown in Fig. 10. This fitting procedure gives a lower estimate for the maximum runaway energy, at about 40–50 MeV, depending on pitch-angle.
C. Effect of wave-particle interaction
Another instance where the synchrotron spectrum from
a complete runaway distribution is useful is in investigations
of mechanisms that affect the shape of the distribution itself.
One such mechanism is resonant wave-particle interactions,
and here we consider their effect on the synchrotron spectrum through a modification of part of the distribution given
in Eq. (9). A runaway distribution is normally strongly
peaked around the parallel direction (v ¼ 1), i.e., it has a
high degree of anisotropy in momentum space (see for
instance Fig. 6). Wave-particle interaction tends to drive the
distribution towards isotropy through pitch-angle scattering
of electrons with resonant momenta.17 A simple way to simulate the decrease in anisotropy is to introduce a flat profile
in part of momentum space, as indicated in Fig. 11.
The usual integral for the total emitted power, Eq. (8), is
split up into three regions in momentum-space. The first and
third parts remain unmodified, with the usual distribution
function fRE. In the second (middle) part, the distribution
function is assumed to be flat. We denote the lower and
upper boundaries of this region pL and pU, respectively. The
momentum space volume of the shaded block in the figure
should be the same as that of the part of the distribution it
replaces, which gives us a condition from which to calculate
the appropriate height of the block. The integration of the
normal distribution is taken over the entire v-range
(v 2 ½0; 1). As the distribution decreases exponentially with
decreasing v, the contribution from particles with low v is
very small. When the modifications are introduced, however,
the contribution could be substantial, and we need to restrict
the extent of the block for the modified part of the
093302-8
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
FIG. 9. Measured visible spectrum in
DIII-D during the runaway plateau at
t ¼ 2290 ms in shot 146 704. The data
are a superposition of synchrotron
radiation from runaways and line radiation from the background plasma.
Theoretical synchrotron spectra are
also shown for various (a) runaway
beam radii, (b) maximum normalized
momenta pmax , (c) loop-voltages Vloop ,
and (d) densities n. Unless otherwise
noted the parameters are pmax ¼ 130;
rre ¼ 0:2 m, n ¼ 3:9 1019 m3 , and
Vloop ¼ 7 V, which are indicated by the
red (dashed-dotted) lines.
distribution in v (v 2 ½vmin ; 1). The introduction of vmin can
be seen as a compensation for the fact that in reality the
pitch-angle scattered particles are not evenly distributed in v.
Letting fc ðp; vÞ ¼ h be a constant distribution where h represents the height of the block, and equating the momentum
space volume of the block with that of the part of the distribution it replaces, we have
ð 1 ð pU
ð 1 ð pU
f ðp;vÞp2 dpdv ¼ 2p
fc ðp;vÞp2 dpdv
V ¼ 2p
0 pL
vmin pL
2p
¼ h ð1 vmin Þðp3U p3L Þ:
3
(13)
We may solve this for h, and obtain
FIG. 10. Measured visible spectrum in DIII-D during the runaway plateau at
t ¼ 2290 ms in shot 146 704. Spectra from several mono-energetic populations calculated using P as2 are also shown. The number of runaways used to
obtain the spectra was 1% of nr calculated from the runaway current (assuming rre ¼ 20 cm).
ð 1 ð pU
3
f ðp; vÞp2 dpdv
h¼
0
pL
ð1 vmin Þðp3U p3L Þ
(14)
as the block height that conserves the total number of particles. We emphasize that the above modification represents
a “worst case scenario” in terms of the effect on the spectrum. In a more realistic case, the modifications would be
less severe.
The analytical avalanche distribution (Eq. (9)) was
modified according to the above, with pL ¼ 25 and pU ¼ 35
since this is a typical range where wave-particle interactions
manifest.17 The maximum pitch-angle in the modified region
was set to p? =pk ¼ 0:2 (vmin ¼ 0:98), which is qualitatively
consistent with experimental estimates of the maximum
runaway pitch-angle.4,5 In Fig. 12, modified distributionintegrated synchrotron spectra are shown and compared with
those of unmodified distributions. From the figure, it is clear
that there is an appreciable increase in the average emission
of the runaways as a result of the modifications to the distribution. Again, this increase is related to the pitch-angle
dependence of P cyl . The isotropization broadens the distribution in pitch-angle which leads to a higher average emission.
Due to the difference in the synchrotron spectrum, the onset
of a particle-wave resonance should be detectable. However,
as we have seen before, there are also other changes in
plasma parameters that could have a similar effect on the
synchrotron emission.
Our goal in this exercise is not to explore the parameter
space of artificially modified distributions—the modifications introduced above are too crude to lead to quantitative
conclusions—but rather to illustrate the sensitivity of the
synchrotron spectrum to the details of the runaway distribution. The analysis here shows that the spectrum from a
093302-9
Stahl et al.
Phys. Plasmas 20, 093302 (2013)
FIG. 11. Schematic runaway distribution with modifications emulating the
effects of wave-particle interaction.
also illustrated that using the slope of the spectrum for estimating the runaway energy can be misleading, and in general
one should calculate the emission from an assumed approximative distribution and iteratively find the maximum
runaway energy to fit the synchrotron spectrum. Finally,
through a comparison with an experimental synchrotron
spectrum from DIII-D, we have estimated the maximum runaway electron energy in that particular experimental scenario
to be around 65 MeV.
ACKNOWLEDGMENTS
FIG. 12. Synchrotron spectra from unmodified and modified runaway
distributions for different electric field strengths. The parameters used
are pmax ¼ 50; pL ¼ 25; pU ¼ 35; Zeff ¼ 1:6; ne ¼ 3 1020 m3 ; T ¼ 10 eV,
and B ¼ 3 T. For these parameters, the critical field is Ec ¼ 0:15 V=m. The
maximum pitch-angle for the particles in the modified region was set to
p? =pk ¼ 0:2.
distribution modified by particle-wave interaction can imply
runaway parameters distinctly different from those that are
actually present, especially if only a limited part of the spectrum is considered. Failure to include such effects can thus
lead to incorrect conclusions regarding the runaway beam
properties.
V. CONCLUSIONS
The synchrotron emission spectrum can be an important
diagnostic of the runaway electron population. In some previous work, synchrotron spectra have been interpreted under
the assumption that all runaways have the same energy and
pitch-angle. In practice, however, runaway electrons have a
wide distribution of energies and pitch-angles. When taking
into account the full distribution, the most suitable approximative emission formula may not be the one that has been
used in previous work (P as1 ). Instead, depending on the
major radius of the device and the actual runaway electron
distribution, either P cyl (for large devices) or P as2 (for
medium-sized devices) are more suitable. Although the single particle synchrotron emission formulas do not depend on
the plasma temperature, effective charge, density or electric
field strength, the total synchrotron emission is sensitive to
these parameters, as they determine the shape of the runaway
distribution.
We have shown that the single-particle emission overestimates the synchrotron emission per particle by orders of
magnitude, and the wavelength of the peak emission is
shifted to shorter wavelengths compared with the spectrum
from an avalanching runaway electron distribution. We have
The authors are grateful to Y. Kazakov for fruitful discussions. This work was funded by the European Communities
under Association Contract between EURATOM, HAS, and
Vetenskapsrådet. The views and opinions expressed herein do
not necessarily reflect those of the European Commission.
M.L. was supported by the United States Department of
Energy’s Fusion Energy Postdoctoral Research Program
administered by the Oak Ridge Institute for Science and
Education. E.H. was supported in part by the United States
Department of Energy under DE-FG02-07ER54917.
1
G. Papp, M. Drevlak, T. F€ul€op, and G. I. Pokol, Plasma Phys. Controlled
Fusion 53, 095004 (2011).
D. Winske, Th. Peter, and D. A. Boyd, Phys. Fluids 26, 3497 (1983).
3
K. H. Finken, J. G. Watkins, D. Rusbult, W. J. Corbett, K. H. Dippel,
D. M. Goebel, and R. A. Moyer, Nucl. Fusion 30, 859 (1990).
4
R. Jaspers, N. J. Lopes Cardozo, A. J. H. Donn, H. L. M. Widdershoven,
and K. H. Finken, Rev. Sci. Instrum. 72, 466 (2001).
5
J. H. Yu, E. M. Hollmann, N. Commaux, N. W. Eidietis, D. A.
Humphreys, A. N. James, T. C. Jernigan, and R. A. Moyer, Phys. Plasmas
20, 042113 (2013).
6
F. Andersson, P. Helander, and L.-G. Eriksson, Phys. Plasmas 8, 5221
(2001).
7
P. A. Robinson, Plasma Phys. Controlled Fusion 27, 1037 (1985).
8
E. Moghaddam-Taaheri and L. Vlahos, Phys. Fluids 31, 99 (1988).
9
K. Kato and I. H. Hutchinson, Phys. Fluids 30, 3809 (1987).
10
G. Bekefi, Radiation Processes in Plasmas (Wiley, 1966).
11
I. M. Pankratov, Plasma Phys. Rep. 25, 145 (1999).
12
T. F€ul€op and G. Papp, Phys. Rev. Lett. 108, 225003 (2012).
13
H. Smith, P. Helander, L.-G. Eriksson, and T. F€ul€op, Phys. Plasmas 12,
122505 (2005).
14
T. F€ul€op, G. Pokol, P. Helander, and M. Lisak, Phys. Plasmas 13, 062506
(2006).
15
E. M. Hollmann, P. B. Parks, D. A. Humphreys, N. H. Brooks, N.
Commaux, N. Eidietis, T. E. Evans, R. Isler, A. N. James, T. C. Jernigan,
J. Mu~noz, E. J. Strait, C. Tsui, J. Wesley, and J. H. Yu, Nucl. Fusion 51,
103026 (2011).
16
E. M. Hollmann, M. E. Austin, J. A. Boedo, N. H. Brooks, N. Commaux,
N. W. Eidietis, D. A. Humphreys, V. A. Izzo, A. N. James, T. C. Jernigan,
A. Loarte, J. Martin-Solis, R. A. Moyer, J. M. Mu~noz-Burgos, P. B. Parks,
D. L. Rudakov, E. J. Strait, C. Tsui, M. A. Van Zeeland, J. C. Wesley, and
J. H. Yu, Nucl. Fusion 53, 083004 (2013).
17
G. Pokol, T. F€ul€op, and M. Lisak, Plasma Phys. Controlled Fusion 50,
045003 (2008).
2
Paper B
M. Landreman, A. Stahl, and T. Fülöp,
Numerical calculation of the runaway electron distribution function and associated synchrotron emission,
Computer Physics Communications 185, 847-855 (2014).
http://dx.doi.org/10.1016/j.cpc.2013.12.004
http://arxiv.org/abs/1305.3518
Computer Physics Communications 185 (2014) 847–855
Contents lists available at ScienceDirect
Computer Physics Communications
journal homepage: www.elsevier.com/locate/cpc
Numerical calculation of the runaway electron distribution function
and associated synchrotron emission
Matt Landreman a,b,∗ , Adam Stahl c , Tünde Fülöp c
a
University of Maryland, College Park, MD, 20742, USA
b
Plasma Science and Fusion Center, MIT, Cambridge, MA, 02139, USA
c
Department of Applied Physics, Nuclear Engineering, Chalmers University of Technology and Euratom-VR Association, Göteborg, Sweden
article
info
Article history:
Received 15 May 2013
Received in revised form
11 September 2013
Accepted 3 December 2013
Available online 9 December 2013
Keywords:
Fokker–Planck
Runaway electrons
Relativistic
Plasma
Kinetic
Synchrotron emission
abstract
Synchrotron emission from runaway electrons may be used to diagnose plasma conditions during a
tokamak disruption, but solving this inverse problem requires rapid simulation of the electron distribution
function and associated synchrotron emission as a function of plasma parameters. Here we detail
a framework for this forward calculation, beginning with an efficient numerical method for solving
the Fokker–Planck equation in the presence of an electric field of arbitrary strength. The approach is
continuum (Eulerian), and we employ a relativistic collision operator, valid for arbitrary energies. Both
primary and secondary runaway electron generation are included. For cases in which primary generation
dominates, a time-independent formulation of the problem is described, requiring only the solution of a
single sparse linear system. In the limit of dominant secondary generation, we present the first numerical
verification of an analytic model for the distribution function. The numerical electron distribution function
in the presence of both primary and secondary generation is then used for calculating the synchrotron
emission spectrum of the runaways. It is found that the average synchrotron spectra emitted from realistic
distribution functions are not well approximated by the emission of a single electron at the maximum
energy.
© 2013 Elsevier B.V. All rights reserved.
1. Introduction
Due to the decrease in the Coulomb collision cross section
with velocity, charged particles in an electric field can ‘‘run away’’
to high energies. In tokamaks, the resulting energetic particles
can damage plasma-facing components and are expected to be a
significant danger in the upcoming ITER experiment. Electrons are
typically the species for which runaway is most significant [1,2],
but runaway ions [3] and positrons [4,5] can also be produced.
Relatively large electric fields are required for runaway production,
and in tokamaks these can arise during disruptions or in sawtooth
events. Understanding of runaway electrons and their generation
and mitigation is essential to planning future large experiments
such as ITER.
Runaway electrons emit measurable synchrotron radiation,
which can potentially be used to diagnose the distribution function, thereby constraining the physical parameters in the plasma.
The runaway distribution function and associated synchrotron
∗ Correspondence to: Institute for Research in Electronics and Applied Physics,
University of Maryland, College Park, MD 20742, USA. Tel.: +1 651 366 9306.
E-mail addresses: [email protected], [email protected]
(M. Landreman).
0010-4655/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cpc.2013.12.004
emission depend on the time histories of the local electric field E,
temperature T , average ion charge Z , and density n. To infer these
quantities (and the uncertainty in these quantities) inside a disrupting plasma using the synchrotron emission, it is necessary to
run many simulations of the runaway process, scanning the various
physical parameters. To make such a scan practical, computational
efficiency is important.
To this end, in this work we demonstrate a framework for rapid
computation of the runaway distribution function and associated
synchrotron emission for given plasma parameters. The distribution function is computed using a new numerical tool named CODE
(COllisional Distribution of Electrons). Physically, the distribution
function is determined by a balance between acceleration in the
electric field and collisions with both electrons and ions. The calculation in CODE is fully relativistic, using a collision operator
valid for both low and high velocities [6] and it includes both primary and secondary runaway electron generation. If primary runaway electron generation dominates, CODE can be used in both
time-dependent and time-independent modes. The latter mode of
operation, in which a long-time quasi-equilibrium distribution
function is calculated, is extremely fast in that it is necessary only
to solve a single sparse linear system. Due to its speed and simplicity, CODE is highly suitable for coupling within larger more
848
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
expensive calculations. Besides the inverse problem of determining plasma parameters from synchrotron emission, other such applications could include the study of instabilities driven by the
anisotropy of the electron distribution function, and comprehensive modeling of disruptions.
Other numerical methods for computing the distribution
function of runaways have been demonstrated previously, using
a range of algorithms. Particle methods follow the trajectories of
individual marker electrons. Deterministic particle calculations [7]
can give insight into the system behavior but cannot calculate
the distribution function, since diffusion is absent. Collisional
diffusion may be included by making random adjustments to
particles’ velocities, an approach which has been used in codes
such as ASCOT [8] and ARENA [9]. For a given level of numerical
uncertainty (noise or discretization error), we will demonstrate
that CODE is more than 6 orders of magnitude faster than a particle
code on the same computer. Other continuum codes developed to
model energetic electrons include BANDIT [10], CQL3D[11,12] and
LUKE [13,14]. These sophisticated codes were originally developed
to model RF heating and current drive, and contain many features
not required for the calculations we consider. For example, CQL3D
contains ∼90,000 lines of code and LUKE contains ∼118,000
lines, whereas CODE contains <1200 lines (including comments).
While future more elaborate modeling may require the additional
features of a code like CQL3D or LUKE, for the applications we
consider, we find it useful to have the nimble and dedicated tool
CODE. For calculations of non-Maxwellian distribution functions
in the context of RF heating, an adjoint method [15] can be a
useful technique for efficient solution of linear inhomogeneous
kinetic equations. However, the kinetic equation we will consider
is nonlinear (if avalanching is included) and homogeneous, so the
adjoint method is not applicable.
In several previous studies, a single particle with a representative momentum and pitch-angle is used as an approximation for
the entire runaway distribution [16,17] when computing the synchrotron emission. In this paper, we present a computation of the
synchrotron radiation spectrum of a runaway distribution in various cases. By showing the difference between these spectra and
those based on single particle emission we demonstrate the importance of taking into account the entire distribution.
The remainder of the paper is organized as follows. In
Section 2 we present the kinetic equation and the collision
operator used. Section 3 details the discretization scheme and
calculation of the primary runaway production rate, with typical
results shown in Section 4. The avalanche source term and its
implementation are described in Section 5. In this section we also
demonstrate agreement with an analytic model for the distribution
function [18]. Computation of the synchrotron emission spectrum
from the distribution function is detailed in Section 6, and
comparisons to single-particle emission are given. We conclude
in Section 7.
2. Kinetic equation and normalizations
We begin with the kinetic equation
∂f
− eEb · ∇p f = C {f } + S .
(1)
∂t
Here, −e is the electron charge, E is the component of the electric
field along the magnetic field, b = B/B is a unit vector along
the magnetic field, ∇p is the gradient in the space of relativistic

momentum p = γ mv, γ = 1/ 1 − v 2 /c 2 , v = |v | is the speed, m
is the electron rest mass, c is the speed of light, C is the electron
collision operator, and S represents any sources. All quantities
refer to electrons unless noted otherwise. Eq. (1) is the largeaspect-ratio limit of the bounce- and gyro-averaged Fokker–Planck
equation (Eq. (2) in [19]). Particle trapping effects are neglected,
which is reasonable since runaway beams are typically localized
close to the magnetic axis. We may write b · ∇p f in (1) in terms of
scalar variables using
1 − ξ 2 ∂f
∂f
+
(2)
∂p
p
∂ξ
where p = |p|, and ξ = p · b/p is the cosine of the pitch
angle relative to the magnetic field. The distribution
is
 function
defined such that the density n is given by n =
d3 p f , so f
has dimensions of (length × momentum)−3 , and we assume the
distribution function for small momentum to be approximately
√ the
Maxwellian fM = nπ −3/2 (mve )−3 exp(−y2 ) where ve = 2T /m
is the thermal speed, and y = p/(mve ) = γ v/ve is the normalized
b · ∇p f = ξ
momentum.
We use the collision operator from Appendix B of Ref. [6]. This
operator is constructed to match the usual nonrelativistic testparticle operator in the limit of v ≪ c, and in the relativistic limit it
reduces to the operator from Appendix A of Ref. [20]. The collision
operator is
C {f } =
1 ∂
p2 ∂ p

p2 CA

∂f
CB ∂
∂f
+ CF f + 2
(1 − ξ 2 )
∂p
p ∂ξ
∂ξ
(3)
where
Γ
Ψ (x),
v


Γ
δ 4 x2
CB =
Z + φ(x) − Ψ (x) +
,
2v
2
CA =
CF =
Γ
T
Ψ (x),
(4)
(5)
(6)

δ = ve /c, x = v/ve = y/ 1 + δ 2 y2 , Z is the effective ion charge,
√
Γ = 4π ne4 ln Λ = (3 π /4)νee ve3 m2
(7)
√
4
is identical
√ to the Γ defined in Refs. [6,20,21], νee = 4 2π e n
ln Λ/(3 mT 3/2 ) is the usual Braginskii electron collision frex
quency, φ(x) = 2π −1/2 0 exp(−s2 ) ds is the error function, and


1
dφ
Ψ (x) = 2 φ(x) − x
(8)
2x
dx
is the Chandrasekhar function. In the nonrelativistic limit δ → 0,
then y → x, and (3) reduces to the usual Fokker–Planck testparticle electron collision operator.
The collision operator (3) is approximate in several ways. First,
it originates from the Fokker–Planck approximation in which
small-angle collisions dominate, which is related to an expansion
in ln Λ ≫ 1. Consequently, the infrequent collisions with large
momentum exchange are ignored, so the secondary avalanche process is not included at this stage, but will be addressed later in
Section 5. Also, the modifications to the Rosenbluth potentials associated with the high-energy electrons are neglected, i.e. collisions with high-energy field particles are ignored.
The kinetic equation is normalized by multiplying through
with m3 ve3 π 3/2 /(νee n), and defining the normalized distribution
function
F = (π 3/2 m3 ve3 /n)f
(9)
so that F → 1 at p → 0. We also introduce a normalized electric
field
Ê = −eE /(mve νee )
(10)
which, up to a factor of order unity, is E normalized by the Dreicer
field. The normalized time is t̂ = νee t and the normalized source
is Ŝ = Sm3 ve3 π 3/2 /(νee n). We thereby obtain the dimensionless
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
equation
849
Then the operation
∂F
∂F
+ Ê ξ
∂y
∂ t̂
√
3 π
−
4
√
3 π
−
4
+ Ê
1 − ξ 2 ∂F
2L + 1
∂y


1 ∂ 2 Ψ (x) ∂ F
y
+
2
Ψ
(
x
)
F
y2 ∂ y
x ∂y


1
δ 4 x2 ∂
∂F
Z + φ(x) − Ψ (x) +
(1 − ξ 2 )
2
2xy
2
∂ξ
∂ξ
y
= Ŝ .
2
Notice that this equation has the form of a linear inhomogeneous
3D partial differential equation:
(12)
for a linear time-independent differential operator M. If a timeindependent equilibrium solution exists, it will be given by F =
M −1 Ŝ.
Since both the electric field acceleration term and the collision
operator in the kinetic equation (1) have the form of a divergence
of a flux in velocity space, the total number
 of particles
 is constant
in time in the absence of a source: (d/dt ) d3 p f = d3 p S. However, runaway electrons are constantly gaining energy, so without
a source at small p and a sink at large p, no time-independent distribution function will exist. From another perspective, a nonzero
source is necessary to find a nonzero equilibrium solution of (11),
because when Ŝ = 0, (11) with ∂/∂ t̂ = 0 is a homogeneous equation with homogeneous boundary conditions. (The boundary conditions are that F be regular at y = 0, ξ = −1, and ξ = 1, and that
f → 0 as y → ∞.) Thus, the solution of the time-independent
problem F = M −1 Ŝ for Ŝ = 0 would be F = 0.
To find a solution, we must either consider a time-dependent
problem or include a nonzero S. In reality, spatial transport can give
rise to both sources and sinks, and a sink exists at high energy due
to radiation. When included, secondary runaway generation (considered in Section 5) also introduces a source. To avoid the added
complexity of these sinks and sources and simultaneously avoid
the intricacies of time dependence, when restricting ourselves to
primary generation we may formulate a time-independent prob−y2
lem as follows. We take Ŝ = α e
for some constant α , representing a thermal source of particles. Eq. (11) for ∂/∂ t̂ = 0 may
be divided through by α and solved for the unknown F /α . Then α
may be determined by the requirement F (p = 0) = 1, and F is
then obtained by multiplying the solution F /α by this α .
The constant α represents the rate at which particles must be
replenished at low energy to balance their flux in velocity space to
high energy. Therefore, α is the rate of runaway production. As we
do not introduce a sink at high energies, F will have a divergent
integral over velocity space.
CODE can also be run in time-dependent mode. Once the
velocity space coordinates and the operator M are discretized, any
implicit or explicit scheme for advancing a system of ordinary
differential equations (forward or backward Euler, Runge–Kutta,
trapezoid rule, etc.) may be applied to the time coordinate. (Results
shown in this paper are computed using the trapezoid rule.) Due to
the diffusive nature of M, numerical stability favors implicit timeadvance schemes.
3. Discretization
We first expand F in Legendre polynomials PL (ξ ):
F (y, ξ ) =
∞

L=0
FL (y)PL (ξ ).
1
PL (ξ )(·)dξ
(14)
−1
is applied to the kinetic equation. Using the identities in the
Appendix, we obtain
(11)
∂F
+ MF = Ŝ
∂ t̂

(13)

∞  
L+1
∂ FL 
L
∂
+
Ê
δL+1,ℓ +
δL−1,ℓ
2L + 3
2L − 1
∂y
∂ t̂
ℓ=0


(L − 1)L
Ê (L + 1)(L + 2)
δL+1,ℓ −
δL−1,ℓ
+
y
2L + 3
2L − 1
√
√ 

3 π Ψ (x)
∂2
3 π 2Ψ (x)
dx dΨ
−
δL,ℓ 2 −
+
δL,ℓ
4
x
∂y
2
y
dy dx
√ 

3 π 1 dx dΨ
2Ψ (x)
Ψ (x) dx
∂
−
+
− 2
+ 2Ψ (x) δL,ℓ
4
x dy dx
xy
x dy
∂y
√ 


3 π
δ 4 x2
+
Z + φ(x) − Ψ (x) +
L(L + 1)δL,ℓ Fℓ = ŜL (15)
2
8xy
2
2 2 −3/2
where dx/dy = (1 + δ y )
and ŜL = (2L + 1) 2
−1
1
−1
, dΨ /dx = 2π −1/2 e−x − (2/x)Ψ (x),
2
Ŝ dξ is the appropriate Legendre mode
of Ŝ (y, ξ ) =
L=0 ŜL (y)PL (ξ ). Note that the collision operator is
diagonal in the L index, and the electric field acceleration term is
tridiagonal in L.
It is useful to examine the L = 0 case of (15), which corresponds
to (half) the integral of the kinetic equation over ξ :
∞

∂ F0
Ê
1 ∂
− 2
−y2 F1
y ∂y
3
∂ t̂
√


3 π 2 Ψ (x) ∂ F0
+
y
+ 2Ψ (x)F0
= Ŝ0 .
(16)
4
x ∂y
∞
Applying 4π −1/2 y dy y2 (·) for some boundary value yb , and
b
assuming the source is negligible in this region, we obtain
1 dnr
4

Ê
−y2 F1
νee n dt̂
3
π
√


3 π 2 Ψ (x) ∂ F0
y
+ 2Ψ (x)F0
+
4
x ∂y
= −√
(17)
y=yb
where nr is the number of runaways,
 meaning the number
 ∞ of electrons with y > yb , so that nr = y>y d3 p f = 2π mv y dp p2
1
−1
b
e b
dξ f . The runaway rate calculated from (17) should be inde-
pendent of yb in steady state (as
 ylong as yb is in a region of Ŝ0 = 0),
which can be seen by applying y b2 dy y2 (·) to (16). We find in pracb1
tice it is far better to compute the runaway production rate using
(17) than from the source magnitude α , since the latter is more
sensitive to the various numerical resolution parameters.
To discretize the equation in y, we can apply fourth-order finite
differences on a uniform grid. Alternatively, for greater numerical
efficiency, a coordinate transformation can be applied so grid
points are spaced further apart at high energies. The y coordinate
is cut off at some finite maximum value ymax . The appropriate
boundary conditions at y = 0 are dF0 /dy = 0 and FL = 0 for
L > 0. For the boundary at large y, we impose FL = 0 for all L. This
boundary condition creates some unphysical grid-scale oscillation
at large y, which may be eliminated by adding an artificial diffusion
c1 y−2 (∂/∂ y)y2 exp(−[y − ymax ]/c2 )∂/∂ y localized near ymax to the
linear operator. Suitable values for the constants are c1 = 0.01
and c2 = 0.1. This term effectively represents a sink for particles,
850
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
a
b
log10
Fig. 1. (Color online) Typical results of CODE, obtained for δ = 0.1, Ê = 0.1, and Z = 1. (a) Normalized distribution function F for p⊥ = 0. Results are plotted for two
different sets of numerical parameters ({Ny = 300, ymax = 20, Nξ = 20} and {Ny = 1200, ymax = 40, Nξ = 40}). The results overlap completely, demonstrating excellent
convergence. A Maxwellian is also plotted for comparison. (b) Contours of F at values 10z for integer z. Bold contours indicate F = 10−5 and 10−10 .
which must be included in the time-independent approach due to
the particle source at thermal energies. Since this diffusion term
is exponentially small away from ymax , the distribution function
is very insensitive to the details of the ad-hoc term except very
near ymax . All results shown hereafter are very well converged with
respect to doubling the domain size ymax , indicating the results are
insensitive to the details of the diffusion term.
4. Results for primary runaway electron generation
Fig. 1 shows typical results from a time-independent CODE
computation. To verify convergence, we may double Nξ (the
number of Legendre modes), double Ny (the number of grid points
in y), and double the maximum y (ymax ) at fixed y grid resolution
(which requires doubling Ny again). As shown by the overlap of the
solid red and dashed blue curves in Fig. 1a, excellent convergence
is achieved for the parameters used here. Increasing the ad-hoc
diffusion magnitude c1 by a factor of 10 for the parameters of the
red curve causes a relative change in the runaway rate (computed
using (17) for yb = 10) of |dnr /dt̂ (c1 = 0.1) − dnr /dt̂ (c1 =
0.01)|/[dnr /dt̂ (c1 = 0.01)] < 10−9 , demonstrating the results
are highly insensitive to this diffusion term. As expected, the
distribution function is increased in the direction opposite to the
electric field (p∥ > 0). While the distribution function is reduced
in the direction parallel to the electric field (p∥ < 0) for y <
5, F is actually slightly increased for y > 5 due to pitch-angle
scattering of the high-energy tail electrons, an effect also seen
in Fokker–Planck simulations of RF current drive [22]. The pitchangle scattering term can be artificially suppressed in CODE, in
which case F is reduced in the direction parallel to the electric field
for all y.
Fig. 2 compares the distribution functions obtained from the
time-independent and time-dependent approaches. At sufficiently
long times, the time-dependent version produces results that are
indistinguishable from the time-independent version.
For comparison with previously published results, we show in
Fig. 3 results by Kulsrud et al. [23], who considered only the nonrelativistic case δ → 0. The agreement with CODE is exceptional.
The runaway production rate in CODE is computed using (17) for
yb = 10. (Any value of yb > 5 gives indistinguishable results.)
Ref. [23] uses a different normalized electric field EK which is re√
lated to Ê by EK = 2(3 π )−1 Ê, and in Ref. [23] the runaway
√ rate is
also normalized by a different collision frequency νK = 3 π/2 νee .
It should also be noted that the Kulsrud computations are timedependent, with a simulation run until the flux in velocity space
reaches an approximate steady state. Each CODE point shown in
Fig. 3 took approximately 0.08 s on a single Dell Precision laptop
with Intel Core i7-2860 2.50 GHz CPU and 16 GB memory, running
in MATLAB. Faster results could surely be obtained using a lowerlevel language.
Fig. 2. (Color online) The distribution function from time-dependent CODE at
various times. At t = 1000/νee , the distribution function is indistinguishable
from the solution obtained using the time-independent scheme (t = ∞) over the
momentum range shown.
Fig. 3. (Color online) Benchmark of CODE in the nonrelativistic limit δ → 0 against
data in Table 1 of Ref. [23].
To emphasize the speed of CODE, we have directly compared
it to the ARENA code [9] for computing the runaway rate using
the parameters considered in [23]. ARENA is a Monte Carlo
code written in Fortran 90 and designed specifically to compute
the runaway distribution function and runaway rate. Detailed
description of the current version of ARENA is given in Refs. [24,25].
Both codes were run on a single thread on the same computer
with an Intel Xeon 2.0 GHz processor. ARENA required 49,550 s to
reproduce the left square point in Fig. 3, and 5942 s to reproduce
the top-right square point. 50,000 particles were required for
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
reasonable convergence. For comparison, at a similar level of
convergence, time-independent CODE required 0.00106 s and
0.000696 s for the two respective points, and time-dependent
CODE required 0.0307 s and 0.00082 s respectively. Thus, for these
parameters, both time-independent and time-dependent CODE
require less than 1.5 × 10−7 as many cpu-hours as ARENA for the
same hardware.
5. Secondary runaway electron generation
In the previous sections we used the Fokker–Planck collision
operator, which includes ‘‘distant’’ (large impact parameter)
collisions but not ‘‘close’’ (small impact parameter) collisions in
which a large fraction of energy and momentum are transferred
between the colliding particles. Close collisions are infrequent
compared to distant collisions, and are therefore neglected in the
Fokker–Planck operator. However, close collisions may still have
a significant effect on runaway generation, since the density of
runaways is typically much smaller than the density of thermal
electrons which may be accelerated in a close collision. The
production of runaways through close collisions is known as
secondary production, or as avalanche production since it may
occur with exponential growth. To simulate secondary generation
of energetic electrons, we use a source term derived in [19],
starting from the Møller scattering cross-section in the w ≫ 1
limit, with w = p/(mc ) = δ y a normalized momentum. In
this limit, the trajectories of the primary electrons are not much
deflected by the collisions. The source then takes the form
∂
δ(ξ − ξ2 ) 2
S=
4π τ ln Λ
w ∂w
nr
1


1
√
1−
1 + w2
,
(18)
where 1/τ = 4π ne4 ln Λ/(m2 c 3 ) is the collision frequency for
relativistic electrons,
nr is the density of the fast electrons and
√
ξ2 = w/(1 + 1 + w 2 ) is the cosine of the pitch angle at which the
runaway is born. (Our Eq. (18) differs by a factor m3 c 3 compared to
the source
our distribution function
 in Ref. [19] since we normalize

as n = d3 p f instead of n = d3 w f . There is also a factor of
2π difference due to the different normalization of the distribution
function.)
Due to the approximations used to derive S, care must be taken
in several regards. First, to define nr in (18), it is not clear where
to draw the dividing line in velocity space between runaways and
non-runaways. One possible strategy for defining nr is to compute
the separatrix in velocity space between trajectories that will have
bounded and unbounded energy in the absence of diffusion, and
to define the runaway density as the integral of f over the latter
region [26]. This approach may somewhat overestimate the true
avalanche rate, since it neglects the fact that some time must elapse
between an electron entering the runaway region and the electron
gaining sufficient energy to cause secondary generation. As most
runaways have ξ ≈ 1, we may approximate the separatrix by
setting dw/dt = 0 where dw/dt = eE /(mc ) − (1 + 1/w 2 )/τ
defines the trajectory of a particle with ξ = 1, neglecting diffusion
in momentum and pitch angle. The runaway region is therefore
w > wc where wc = [(E /Ec ) − 1]−1/2 and Ec = mc /(eτ ) is the
1
∞
critical field, and so we take nr = 2π m c −1 dξ w dw w f . (We
c
cannot define nr by the time integral of (17), since (17) is no longer
valid when S is nonzero away from p ≈ 0.) A second deficiency of
(18) is that S is singular at w → 0, so the source must be cut off
below some threshold momentum. Following Ref. [12], we choose
the cutoff to be wc . Neither of the cutoffs discussed here would be
necessary if a less approximate source term than (18) were used,
but derivation of such an operator is beyond the scope of this paper.
Normalizing and applying (14) as we did previously for the
other terms in the kinetic equation, the source included in CODE
3 3
2
851
becomes
ŜL =
nr 3πδ 5 2L + 1
n 16 ln Λ
2
PL (ξ2 )
(1 −
√
1
1 + w 2 )2
√
1 + w2 y
.
(19)
When secondary generation is included, CODE must be run in timedependent mode.
To benchmark the numerical solution of the kinetic equation
including the above source term by CODE, we use the approximate
analytical expression for the avalanche distribution function
derived in Section II of Ref. [18]:

k
faa (w∥ , w⊥ ) =
w∥
exp γ̃ t −

−
E /Ec − 1

Z +1
γ̃ τ
E /Ec − 1
2
w⊥
2w∥
w∥

(20)
where k is a constant. The quantity γ̃ is the growth rate γ̃ =
(1/f )∂ f /∂ t, which must be independent of both time and velocity
for (20) to be valid. Eq. (20) is also valid only where p∥ ≫ p⊥ and in
regions of momentum space where S is negligible. (This restriction
is not a major one since S = 0 everywhere except on the ξ = ξ2
curve.) If most of the runaway distribution function is
 accurately
described by (20), then we may approximate nr ≈
d3 p faa =


∞
∞
2π m3 c 3 −∞ dw∥ 0 dw⊥ w⊥ faa , giving γ̃ = (1/nr )dnr /dt and
k = nr e−γ̃ t
τ
(21)
2 π m3 c 3 ( 1 + Z )
where nr e−γ̃ t is constant. (Eq. (21) may be inaccurate in some
situations even if (20) is accurate in part of velocity-space, because
(21) requires (20) to apply in all of velocity-space.) Figs. 4
and 5 show comparisons between distributions from CODE and
(20)–(21) for two different sets of parameters. More precisely, the
quantity plotted in Fig. 4–5 is log10 (m3 c 3 f /nr ). To generate the
figures, CODE is run for a sufficiently long time that (1/f )∂ f /∂ t
becomes approximately constant. The resulting numerical value
of (1/nr )dnr /dt is then used as γ̃ when evaluating (20)–(21). For
a cleaner comparison between CODE and analytic theory in these
figures, we minimize primary generation in CODE in these runs
by initializing f to 0 instead of to a Maxwellian. For both sets
of physical parameters, the agreement between CODE and (20)
is excellent in the region where agreement is expected: where
p∥ ≫ p⊥ and away from the curve ξ = ξ2 .
6. Synchrotron emission
Using the distribution functions calculated with CODE, we now
proceed to compute the spectrum of emitted synchrotron radiation. Due to the energy dependence of the emitted synchrotron
power, the emission from runaways completely dominates that of
the thermal particles. The emission also depends strongly on the
pitch-angle of the particle. In a cylindrical plasma geometry, the
emitted synchrotron power per wavelength at wavelength λ from
a single highly energetic particle is given by [27]
4π
P (γ , γ∥ , λ) = √
ce2
3 λ γ
3

2
∞
K5/3 (l) dl,
(22)
λc /λ
where the two-dimensional
 momentum of the particle is determined by γ and γ∥ = 1/ 1 − v∥2 /c 2 , Kν (x) is a modified Bessel
function of the second kind, and
λc =
4π mc 2 γ∥
3 eBγ 2
,
(23)
where B is the magnetic field strength.
Using CODE we will demonstrate that the synchrotron radiation spectrum from the entire runaway distribution is substantially
852
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
(1 − w 2 ξ 2 /(1 + w2 ))−1 , and integrating (22) over the runaway region R in momentum space, we obtain the total synchrotron emission from the runaway distribution. Normalizing to nr , we find that
the average emitted power per runaway particle at a wavelength
λ is given by
P (λ) =
Fig. 4. (Color online) Contour plots of the long-time distribution function from
CODE (shown in two different coordinate systems), obtained for E /Ec = 40 (Ê =
0.532), Z = 3, δ = 0.1 and t = 5000/νee . Results are plotted for the numerical
parameters Ny = 1500, ymax = 1500 and Nξ = 100, with time step dt = 10/νee .
The analytical distribution in (20)–(21) for the same physical parameters is also
plotted for comparison, together with part of the curve where avalanche runaways
are created (ξ = ξ2 ).
Fig. 5. (Color online) Contour plots of the long-time distribution function from
CODE (shown in two different coordinate systems), obtained for E /Ec = 100 (Ê =
0.332), Z = 1, δ = 0.05 and t = 6000/νee . Results are plotted for the numerical
parameters Ny = 1500, ymax = 3000 and Nξ = 180, with time step dt = 25/νee .
The analytical distribution in (20)–(21) is also plotted for comparison, together with
part of the curve where avalanche runaways are created (ξ = ξ2 ).
different from the spectrum obtained from a single particle approximation. By transforming to the more suitable coordinates w
and ξ , related to γ and γ∥ through γ 2 = 1 + w 2 and γ∥2 =
2π
nr

f (w, ξ ) P (w, ξ , λ) w 2 dw dξ .
(24)
R
Up to a factor ecA, where A is the area of the runaway beam, normalization by nr is equivalent to normalization by the runaway
current, since the emitting particles all move with velocity ≈ c.
The per-particle synchrotron spectra generated by the CODE
distributions in Figs. 4 and 5 were calculated using this formula,
and are shown in Fig. 6, together with the spectra radiated by
electron distributions for other electric field strengths. For the
physical parameters used, we note that the peak emission occurs
between 7 and 25 µm. The synchrotron spectra show a decrease
in per-particle emission with increasing electric field strength.
Even though a stronger electric field leads to more particles with
high energy (and thus high average emission), it also leads to
a more narrow distribution in pitch-angle. This reduction in the
number of particles with large pitch-angle leads to a decrease in
average emission. Both figures confirm that the average emission
is reduced for higher electric fields, implying that the latter
mechanism has the largest impact on the spectrum.
In calculating the spectra, the runaway region of momentum
space, R, was defined such that the maximum particle momentum
was wmax = 50 (which translates to ymax = 500 and ymax = 1000
respectively for the cases shown in Figs. 4 and 5), corresponding
to a maximum particle energy of ≃25 MeV. Physically the cutoff
at large energy can be motivated by the finite life-time of the
accelerating electric field and the influence of loss mechanisms
such as radiation. Since the radiated synchrotron power increases
with both particle energy and pitch, this truncation of the
distribution is necessary to avoid infinite emission, although the
precise value for the cutoff depends on the tokamak and on
discharge-specific limitations to the maximum runaway energy.
For the low-energy boundary of R, wmin = wc = [(E /Ec ) − 1]−1/2
was used, and all particles with ξ ∈ [0, 1] were included. Although
no explicit cutoff was imposed in ξ , the distribution decreases
rapidly as this parameter decreases from 1 (as can be seen in Figs. 4
and 5) and there are essentially no particles below some effective
cutoff value.
Fig. 6 also shows the synchrotron spectrum from single particles with momentum corresponding to the maximum momentum
of the distributions (w = 50), and several values of pitch-angle
y⊥ /y∥ . This single-particle ‘‘approximation’’ is equivalent to using
a 2D δ -function model of the distribution, as was done in Refs.
[28,16] (and with some modification in [17]). The figure shows that
this approximation significantly overestimates the synchrotron
emission per particle. Note that in the figure, the values for the
emitted power per particle were divided by a large number to fit
in the same scale. The overestimation is not surprising, since the δ function approximation effectively assumes that all particles emit
as much synchrotron radiation as one of the most strongly emitting
particles in the actual distribution. The figure also shows that the
δ -function approximation leads to a different spectrum shape, with
the wavelength of peak emission usually shifted towards shorter
wavelengths. In order to obtain an accurate runaway synchrotron
spectrum, it is thus crucial to use the full runaway distribution in
the calculation.
In the cases shown in Fig. 6, the runaway electron distribution
is dominated by secondary generation. For comparison, in Fig. 7–8
we show a case where the distribution is dominated by primary
generation. Fig. 7a shows contours of a distribution from primaries
only, together with a distribution obtained with the avalanche
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
a
853
Emitted Power [10–15 W/µm]
Emitted Power [10–15 W/µm]
b
Wavelength [µm]
Wavelength [µm]
Fig. 6. (Color online) Synchrotron spectra (average emission per particle) for the runaway distributions in (a) Fig. 4 and (b) Fig. 5. Emission spectra from the CODE
distributions in Figs. 4 and 5 are shown in solid black, together with spectra from distributions with varying electric field strength but otherwise identical physical parameters.
A magnetic field of B = 3 T was used. The synchrotron spectra from single particles with w = 50 and various pitch angles are also shown. (These single-particle spectra are
the same in figures a and b, as the particle parameters are independent of simulation settings.)
Emitted Power
[10–16
W/µm]
a
b
Wavelength [µm]
Fig. 8. (Color online) Synchrotron spectra (average emission per particle) for the
runaway distributions shown in Fig. 7 for magnetic field B = 3 T.
Fig. 7. (Color online) (a) Contour plots of the distribution function from primaries
only (black solid line), together with a distribution obtained with the avalanche
source enabled (dashed red line), for E /Ec = 10 (Ê = 0.523), Z = 1 and δ = 0.2 at
t = 45/νee . The quantity plotted is log10 (F ). Results are plotted for the numerical
parameters Ny = 20, ymax = 100 and Nξ = 130, with time step dt = 0.02/νee .
(b) Contour plots of the distribution function at different times with the avalanche
source enabled, using the above parameters.
source enabled, and confirms that the distribution is dominated by
primaries, except for a small number of secondary runaways generated along the curve ξ = ξ2 . Fig. 7(b) shows contour plots with
the avalanche source enabled, for three different times. The physical parameters used in Fig. 7 are temperature T = 10 keV, density
n = 5 × 1019 m−3 , effective charge Z = 1, and electric field E =
0.45 V/m. The collision time in this case is 0.39 ms, so the times
shown in the figure correspond to 5.9, 11.8 and 17.6 ms, which
correlates well with the time-scale of the electric field spike for a
typical disruption in DIII-D (see e.g. Fig. 2 in [29]). Fig. 8 compares
the synchrotron spectra from the distributions shown in Fig. 7.
The main difference compared to the case dominated by secondary
generation (Fig. 6) is the generally longer wavelengths in the spectrum. The reason is the low runaway electron energy (w . 10) in
the runaway electron distribution in this case. The small peak at
short wavelengths in the spectra including the avalanche source
stems from the secondary runaways generated at ξ = ξ2 (visible
in Fig. 7(a)).
In principle, we may also use the synchrotron spectra from distributions calculated through CODE to estimate the maximum energy of the runaways in existing tokamaks. However, due to the
region of sensitivity of the available detectors, there is only a limited wavelength range in which calculated spectra can be fitted
to experimental data in order to determine the maximum runaway energy. The available range often corresponds to the short
wavelength slope of the spectrum, where the emitted power shows
an approximately linear dependence on wavelength. Indeed, the
short-wavelength spectrum slope has been used to estimate the
maximum runaway energy in experiments [16]. If the runaway
distribution function is approximated by a δ -function at the maximum available energy and pitch angle, there is a monotonic relationship between the short-wavelength spectrum slope and the
maximum particle energy (at fixed pitch angle). Such a relationship
holds because increasing the particle energy leads to more emission at shorter wavelengths, resulting in a shift of the wavelength
of peak emission towards shorter wavelengths, and a corresponding change in the spectrum slope.
854
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
Using an integrated synchrotron spectrum from a CODE distribution is much more accurate than the single particle approximation, but it also introduces additional parameters (Ê, δ , Z ). If the
physical parameters are well known, a unique relation still holds
between the spectrum slope and the maximum particle energy.
During disruptions however, many parameters (like the temperature and the effective charge) are hard to measure with accuracy.
As the shape of the underlying distribution depends on the values
of the parameters, the synchrotron spectrum will do so as well.
This complexity is apparent in Fig. 6, where the single particle approximation produces identical results in the two cases, whereas
the spectra from the complete distributions are widely different.
The dependence on distribution shape makes it possible in principle for two sets of parameters to produce the same spectrum
slope for different maximum energies. Given this insight, using
the complete runaway distribution when modeling experimentally
obtained spectra is necessary for an accurate analysis and reliable
fit of the maximum particle energy. In this context, CODE is a very
useful tool with the possibility to contribute to the understanding
of runaways and their properties.
7. Conclusions
In this work, we have computed the synchrotron emission
spectra from distribution functions of runaway electrons. The
distribution functions are computed efficiently using the CODE
code. Both primary (Dreicer) and secondary (avalanche) generation
are included. A Legendre spectral discretization is applied to the
pitch-angle coordinate, with high-order finite differences applied
to the speed coordinate. A nonuniform speed grid allows high
resolution of thermal particles at the same time as a high maximum
energy without a prohibitively large number of grid points. If
secondary generation is unimportant, the long-time distribution
function may be calculated by solving a single sparse linear
system. The speed of the code makes it feasible to couple to
other codes for integrated modeling of complex processes such
as tokamak disruptions. CODE has been benchmarked against
previous analytic and numerical results in appropriate limits,
showing excellent agreement. In the limit of strong avalanching,
CODE demonstrates agreement with the analytic distribution
function (20) from Ref. [18].
The synchrotron radiation spectra are computed by convolving
the distribution function with the single-particle emission. We
find that the radiation spectrum from a single electron at
the maximum energy can differ substantially from the overall
spectrum generated by a distribution of electrons. Therefore,
experimental estimates of maximum runaway energy based on the
single-particle synchrotron spectrum are likely to be inaccurate.
A detailed study of the distribution-integrated synchrotron
spectrum and its dependences on physical parameters can be
found in [30].
In providing the electron distribution functions (and thus
knowledge of a variety of quantities through its moments), the
applicability of CODE is wide, and the potential in coupling CODE
to other software, e.g. for modeling of runaway dynamics in
disruptions, is promising. For a proper description of the runaways
generated in disruptions it is important to take into account
the evolution of the radial profiles of the electric field and fast
electron current self-consistently. This can be done by codes such
as GO, initially described in Ref. [31] and developed further in
Refs. [32,33]. GO solves the equation describing the resistive
diffusion of the electric field in a cylindrical approximation coupled
to the runaway generation rates. In the present version of GO, the
runaway rate is computed by approximate analytical formulas for
the primary and secondary generation. Using CODE, the analytical
formulas can be replaced by a numerical solution for the runaway
rate which would have several advantages. One advantage would
be that Dreicer, hot-tail and secondary runaways could all be
calculated with the same tool, avoiding the possibilities for doublecounting and difficulties with interpretations of the results. Also,
in the present version of GO, it is assumed that all the runaway
electrons travel at the speed of light, an approximation that
can be easily relaxed using CODE, which calculates the electron
distribution in both energy and pitch-angle. Most importantly,
the validity region of the results would be expanded, as the
analytical formulas are derived using various assumptions which
are often violated in realistic situations. The output would be a selfconsistent time and space evolution of electric field and runaway
current, together with the electron distribution function. This
information can then be used for calculating quantities that depend
on the distribution function, such as the synchrotron emission or
the kinetic instabilities driven by the velocity anisotropy of the
runaways.
Acknowledgments
This work was supported by US Department of Energy grants
DE-FG02-91ER-54109 and DE-FG02-93ER-54197, by the Fusion
Energy Postdoctoral Research Program administered by the Oak
Ridge Institute for Science and Education, and by the European
Communities under Association Contract between EURATOM and
Vetenskapsrådet. The views and opinions expressed herein do not
necessarily reflect those of the European Commission. The authors
are grateful to J. Rydén, G. Csépány, G. Papp, P. Helander, E. Nilsson,
J. Decker, and Y. Peysson for fruitful discussions.
Appendix. Integrals of Legendre polynomials
Here we list several identities for Legendre polynomials which
are required for the spectral pitch-angle discretization. To evaluate
the ξ integral of the ξ ∂ F /∂ y term in (11), we use the recursion
relation
ξ PL (ξ ) =
L+1
2L + 1
PL+1 (ξ ) +
L
2L + 1
PL−1 (ξ )
(A.1)
where PL−1 is replaced by 0 when L = 0. Applied to the relevant
integral in (11), and noting the orthogonality relation (2L +
1
1)2−1 −1 PL (ξ )Pℓ (ξ )dξ = δL,ℓ , we find
2L + 1

2
1
dξ ξ PL (ξ )Pℓ (ξ ) =
−1
L+1
2L + 3
δℓ,L+1 +
L
2L − 1
δℓ,L−1 . (A.2)
Similarly, to evaluate the ξ integral of the ∂ F /∂ξ term in (11), we
use the recursion relation
(1 − ξ 2 )(dPL /dξ ) = LPL−1 (ξ ) − Lξ PL (ξ )
(A.3)
to obtain
2L + 1

2
1
dξ PL (ξ )(1 − ξ 2 )
−1
dPℓ
dξ
(L + 1)(L + 2)
(L − 1)L
=
δℓ,L+1 −
δℓ,L−1 .
2L + 3
2L − 1
(A.4)
Finally, the pitch-angle scattering collision term gives the integral
2L + 1
2

1
dξ PL (ξ )
−1
∂
∂
(1 − ξ 2 ) Pℓ (ξ ) = −(L + 1)LδL,ℓ .
∂ξ
∂ξ
References
[1] H. Dreicer, Phys. Rev. 117 (1960) 329.
[2] J.W. Connor, R.J. Hastie, Nucl. Fusion 15 (1975) 415.
[3] H.P. Furth, P.H. Rutherford, Phys. Rev. Lett. 28 (1972) 545.
(A.5)
M. Landreman et al. / Computer Physics Communications 185 (2014) 847–855
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
P. Helander, D. Ward, Phys. Rev. Lett. 90 (2003) 135004.
T. Fülöp, G. Papp, Phys. Rev. Lett. 51 (2012) 043004.
G. Papp, M. Drevlak, T. Fülöp, P. Helander, Nucl. Fusion 51 (2011) 043004.
J.R. Martín-Solís, J.D. Alvarez, R. Sánchez, B. Esposito, Phys. Plasmas 5 (1998)
2370.
J.A. Heikkinen, S.K. Sipilä, T.J.H. Pättikangas, Comput. Phys. Comm. 76 (1993)
215.
L.-G. Eriksson, P. Helander, Comput. Phys. Comm. 154 (2003) 175.
M.R. O’Brien, M. Cox, D.F.H. Start, Nucl. Fusion 26 (1986) 1625.
S.C. Chiu, M.N. Rosenbluth, R.W. Harvey, V.S. Chan, Nucl. Fusion 38 (1998)
1711.
R.W. Harvey, V.S. Chan, S.C. Chiu, T.E. Evans, M.N. Rosenbluth, Phys. Plasmas 7
(2000) 4590.
Y. Peysson, J. Decker, R.W. Harvey, AIP Conf. Proc. 694 (2003) 495.
Y. Peysson, J. Decker, AIP Conf. Proc. 1069 (2008) 176.
C. Karney, N. Fisch, Phys. Fluids 29 (1986) 180.
R. Jaspers, N.J.L. Cardozo, A.J.H. Donné, H.L.M. Widdershoven, K.H. Finken, Rev.
Sci. Instrum. 72 (2001) 466.
J.H. Yu, E.M. Hollmann, N. Commaux, N.W. Eidietis, D.A. Humphreys,
A.N. James, T.C. Jernigan, R.A. Moyer, Phys. Plasmas 20 (2013) 042113.
T. Fülöp, G. Pokol, P. Helander, M. Lisak, Phys. Plasmas 13 (2006) 062506.
M.N. Rosenbluth, S.V. Putvinski, Nucl. Fusion 37 (1997) 1355.
P. Sandquist, S.E. Sharapov, P. Helander, M. Lisak, Phys. Plasmas 13 (2006)
072108.
C. Karney, N. Fisch, Phys. Fluids 28 (1985) 116.
855
[22] C. Karney, N. Fisch, Phys. Fluids 22 (1979) 1817.
[23] R.M. Kulsrud, Y.-C. Sun, N.K. Winsor, H.A. Fallon, Phys. Rev. Lett. 31 (1973) 690.
[24] J. Rydén, Monte Carlo Simulation of Runaway Electrons, Master’s Thesis,
Chalmers University of Technology (2012), http://publications.lib.chalmers.
se/records/fulltext/179211/179211.pdf.
[25] G. Csépány, Kinetic Simulation of Runaway Electrons in Tokamaks, Master’s Thesis, Budapest University of Technology and Economics (2012),
http://deep.reak.bme.hu/~cheoppy/diploma-thesis-2012.pdf.
[26] H. Smith, P. Helander, L.-G. Eriksson, T. Fülöp, Phys. Plasmas 12 (2005) 122505.
[27] G. Bekefi, Radiation Processes in Plasmas, Wiley, 1966.
[28] K.H. Finken, J.G. Watkins, D. Rusbuldt, W.J. Corbett, K.H. Dippel, D.M. Goebel,
R.A. Moyer, Nucl. Fusion 30 (1990) 859.
[29] E.M. Hollmann, M.E. Austin, J.A. Boedo, N.H. Brooks, N. Commaux, N.W. Eidietis,
D.A. Humphreys, V.A. Izzo, A.N. James, T.C. Jernigan, A. Loarte, J. Martin-Solis,
R.A. Moyer, J.M. Munoz-Burgos, D.L. Rudakov, E.J. Strait, C. Tsui, M.A.V. Zeeland,
J.C. Wesley, J.H. Yu, Nucl. Fusion 53 (2013) 083004.
[30] Stahl A., Landreman M., G. Papp, E. Hollmann, T. Fülüp, Phys. Plasmas 20 (2013)
093302.
[31] H. Smith, P. Helander, L.-G. Eriksson, D. Anderson, M. Lisak, F. Andersson, Phys.
Plasmas 13 (2006) 102502.
[32] K. Gál, T. Fehér, H. Smith, T. Fülöp, P. Helander, Plasma Phys. Controll. Fusion
50 (2008) 055006.
[33] T. Fehér, H.M. Smith, T. Fülöp, K. Gál, Plasma Phys. Controll. Fusion 53 (2011)
035014.
Errata, Paper B
ˆ In Eq. (11), the partial derivative with respect to y in the third term
should be replaced by a partial derivative with respect to ξ, so that the
first row reads
∂F
∂F
1 − ξ ∂F
+ Ê ξ
+ Ê
∂y
y ∂ξ
∂ t̂
ˆ On the left-hand side of Eq. (17), the collision frequency should be removed, so that the expression starts with
1 dnr
= ...
n dt̂
Paper C
A. Stahl, O. Embréus, G. Papp, M. Landreman and T. Fülöp,
Kinetic modelling of runaway electrons in dynamic scenarios,
Nuclear Fusion 56, 112009 (2016).
http://dx.doi.org/10.1088/0029-5515/56/11/112009
http://arxiv.org/abs/1601.00898
Nuclear Fusion
International Atomic Energy Agency
Nucl. Fusion 56 (2016) 112009 (10pp)
doi:10.1088/0029-5515/56/11/112009
Kinetic modelling of runaway electrons
in dynamic scenarios
A. Stahl1, O. Embréus1, G. Papp2, M. Landreman3 and T. Fülöp1
1
Department of Physics, Chalmers University of Technology, Göteborg, Sweden
Max Planck Institute for Plasma Physics, Garching, Germany
3
University of Maryland, College Park, MD, USA
2
E-mail: [email protected]
Received 5 January 2016, revised 4 March 2016
Accepted for publication 29 March 2016
Published 22 July 2016
Abstract
Improved understanding of runaway-electron formation and decay processes are of prime
interest for the safe operation of large tokamaks, and the dynamics of the runaway electrons
during dynamical scenarios such as disruptions are of particular concern. In this paper, we
present kinetic modelling of scenarios with time-dependent plasma parameters; in particular,
we investigate hot-tail runaway generation during a rapid drop in plasma temperature. With the
goal of studying runaway-electron generation with a self-consistent electric-field evolution,
we also discuss the implementation of a collision operator that conserves momentum and
energy and demonstrate its properties. An operator for avalanche runaway-electron generation,
which takes the energy dependence of the scattering cross section and the runaway distribution
into account, is investigated. We show that the simplified avalanche model of Rosenbluth and
Putvinskii (1997 Nucl. Fusion 37 1355) can give inaccurate results for the avalanche growth
rate (either lower or higher) for many parameters, especially when the average runaway energy
is modest, such as during the initial phase of the avalanche multiplication. The developments
presented pave the way for improved modelling of runaway-electron dynamics during
disruptions or other dynamic events.
Keywords: runaway electrons, Fokker–Planck equation, avalanche generation, hot-tail
generation, linearized collision operator
(Some figures may appear in colour only in the online journal)
1. Introduction
Kinetic simulation is the most accurate and useful method
for investigating runaway-electron dynamics, and we recently
developed a new tool called code (collisional distribution of electrons [5]) for fast and detailed study of these
processes. code solves the spatially homogeneous kinetic
equation for electrons in 2D momentum space, including
electric-field acceleration, collisions, avalanche runaway generation and synchrotron-radiation-reaction losses [5–7]. In
code, momentum space is discretized using finite differences
in momentum and a Legendre-mode decomposition in pitchangle cosine. Often, the time evolution of the distribution
is the desired output, but a (quasi-)steady-state solution can
also be efficiently obtained through the inversion of a single
sparse system (in the absence of an avalanche source). code
has been used to study the spectrum of the synchrotron radiation emitted by runaways [5], the corresponding influence of
Runaway electrons, a phenomenon made possible by the
decrease of the collisional friction with particle energy [1,
2], are common in plasmas in the presence of strong external
electric fields or changing currents. The tightly focused beam
of highly relativistic particles can be a serious threat to the
first wall of a fusion reactor, due to the possibility of localized melting or halo-current generation [3]. In the quest for
avoidance or mitigation of the harmful effects of runawayelectron losses, a greater understanding of the runawayelectron phenom­enon is required [4]. Improved knowledge
of runaway-electron formation mechanisms, dynamics and
characteristics will benefit the fusion community and contribute to a stable and reliable operation of reactor-scale
tokamaks.
0029-5515/16/112009+10$33.00
1
© 2016 EURATOM Printed in the UK
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
κ = m3v˜e 3π 3 / 2 /n˜, we have also defined the distribution function
F = F ( y, ξ ) = κ f (normalized so that F( y = 0) = 1 for a
Maxwellian with T = T˜ and n = n˜), time tˆ = ν˜ee t, and electric field Eˆ = −eE /mv˜eν̃ee, as well as the normalized operators
˜ /3m2v˜e 3
Cˆ = C κ /ν˜ee and Sˆ = Sκ /ν˜ee, with ν˜ee = 16 π e4n˜ ln Λ
the reference electron thermal collision frequency,−e, m and
v the charge, rest mass and speed of the electron, and γ the
relativistic mass factor. Note that |Eˆ | = (3 π /2)E /ED, with ED
the Dreicer field [1]. C is the Fokker–Planck collision operator
and S an operator describing close (large-angle) Coulomb col­
lisions. These operators will be discussed more thoroughly in
sections 3 and 4, respectively; for now we just state the formulation of the collision operator employed in [5] using the
normalizations above:
the emission on the distribution function [6–8], and the factors influencing the critical electric field for runaway-electron
generation [6, 9].
In this paper we describe improvements to code which
enable us to investigate the effect of hot-tail runaway generation on the distribution (section 2). This process can be the
dominant mechanism in rapidly cooling plasmas. We also
discuss the implementation of a full linearized collision operator, and demonstrate its conservation properties (section 3).
The use of this operator is necessary in cases where the correct plasma conductivity is required, and our implementation indeed reproduces the Spitzer conductivity [10] for weak
electric fields. In addition, an improved model for the largeangle (knock-on) Coulomb collisions leading to avalanche
multiplication of the runaway population [11], is described in
section 4. This model takes the energy dependence of the runaway distribution into account, and uses the complete energydependent Møller scattering cross section [12]. We find that
its use can in some cases lead to significant modifications to
the avalanche growth rate, compared to the more simplified
model of Rosenbluth and Putvinskii [13].
The improvements described in this work enable the
detailed study of runaway processes in dynamic situations
such as disruptions, and the conservative collision operator
makes self-consistent calculations of the runaway population
and current evolution in such scenarios feasible [14].
⎛∂ ⎡
⎛1 ∂
2 ⎞ ⎤
tp
⎢ y2Ψ ⎜
+ 2 ⎟F ⎥
Cˆ = cC v¯e 3y−2 ⎜⎜
∂
∂
y
x
y
v
⎝
¯
⎣
e ⎠ ⎦
⎝
+
cξ ∂
∂F ⎞
⎟.
(1 − ξ 2 )
∂ξ ⎟⎠
2x ∂ξ
(2)
The superscript tp denotes that this is the test-particle part
of the linearized collision operator Cl discussed in section 3.
Here (and throughout the rest of this paper), a bar denotes
a quantity normalized to its reference value (i.e. v¯e = ve /v˜e),
x = y /γ = v /v˜e is the normalized speed, cC = 3 π ν¯ee /4,
cξ = Z eff + Φ − Ψ + v¯e 2δ 4x 2 /2, Z eff is the effective ion charge,
Φ = Φ(x /v¯e ) and Ψ = Ψ(x /v¯e ) = v¯e 2[Φ − v¯e−1x dΦ/d(x /v¯e )]/2x 2
are the error and Chandrasekhar functions, respectively,
and δ = v˜e /c (with c the speed of light) is assumed to be a
small parameter (i.e. the thermal population is assumed to be
non-relativistic).
Changes to the plasma temperature manifest as shifts in
the relative magnitude of the various terms in equation (2)
(through δ and the quantities with a bar), as well as a change
in the overall magnitude of the operator, whereas changes
in density only have the latter effect. In both cases, the distribution is effectively colliding with (and relaxing towards)
a Maxwellian different from the one native to the reference
momentum grid. Heat or particles are introduced to (or
removed from) the bulk of the distribution when using this
scheme, as all changes to plasma parameters are described
by changes to the Maxwellian. This provides a powerful way
of simulating rapid cooling, for instance associated with a
tokamak disruption.
2. Time-dependent plasma parameters
To be able to investigate the behavior of the electron population
in dynamic scenarios such as disruptions or sawtooth crashes,
it is necessary to follow the distribution function as the plasma
parameters change. To this end, code has been modified to
handle time-dependent background-plasma parameters. Since
the kinetic equation is treated in linearized form, the actual
temperature and density of the distribution are determined
by the background Maxwellian used in the form­ulation of
the collision operator. This allows for a scheme where the
kinetic equation is normalized to a reference temperature T̃
and number density ñ, so that the discretized equation can
be expressed on a fixed reference grid in momentum space
(throughout this paper, we will use a tilde to denote a reference quantity). By changing the properties of the Maxwellian
equilibrium around which the collision operator is linearized,
the evolution of the plasma parameters can be modelled on the
reference grid without the need for repeated interpolation of
the distribution function to new grids.
Analogously to [5], the kinetic equation in 2D momentum
space for the electron distribution function f experiencing an
electric field E (parallel to the magnetic field) and collisions,
can be expressed as
2.1. Hot-tail runaway-electron generation
If the time scale of the initial thermal quench in a disruption
event is short enough—comparable to the collision time—
the tail of the initial Maxwellian electron distribution will
not have time to equilibrate as the plasma cools. The particles in this supra-thermal tail may constitute a powerful
source of runaway electrons, should a sufficiently strong
electric field develop before they have time to reconnect
with the bulk electrons. This process is known as hot-tail
generation, and can be the dominant source of runaways
⎛ ∂F
1 − ξ 2 ∂F ⎞
∂F
+
+ Eˆ ⎜ξ
⎟ = Cˆ {F} + Sˆ{F}.
(1)
ˆ
y ∂ξ ⎠
∂t
⎝ ∂y
Here we have introduced a convenient normalized momentum
y = γv /v˜e, where v˜e = 2T˜ /m is the reference electron thermal
speed, and the cosine of the pitch angle ξ = y∥ / y. Using
2
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 1. (a) Temperature and electric-field evolution in equations (3) and (4). (b) Parallel (ξ = 1) electron distributions (solid) and
corresponding Maxwellians (dashed) at several times during the temperature drop in (a). A momentum grid with a fixed reference
temperature T˜ = 100 eV was used and the distributions are normalized to F(y = 0) in the final time step to facilitate a comparison.
For the temperature evolution in equation (3), analytical
results for the hot-tail runaway generation were obtained in
[18]. Assuming the background density to be constant, the
runaway fraction at time t can be written as
under certain conditions [15, 16]. It has previously been
investigated analytically or using Monte-Carlo simulations
[16, 17] or purpose-built finite-difference tools [17, 18].
Using code to model a temperature drop enables the efficient study of a wider range of scenarios, and allows full
use of other capabilities of code, such as avalanche generation or synchrotron radiation reaction. Here, we restrict
ourselves to a proof-of-principle demonstration, and leave a
more extensive investigation to future work.
To facilitate a comparison to the theoretical work by Smith
and Verwichte [18], we will model a rapid exponential temper­
ature drop, described by
∞⎡
n r,dir
(u3 − 3τ )2 / 3 ⎤ −u2 2
4
⎢1 − 3c
⎥ e u du,
=
(5)
n
(u − 3τ )2 / 3 ⎦
π uc ⎣
∫
where τ (t ) = (3 π /4)νee(t − t ) = (3 π /4)(tˆ − tˆ ) is a nor­
malized time, u(t ) = x[0] + 3τ (t ), x[0] is the speed nor­
malized to the initial thermal speed, and uc is related to the
critical speed for runaway generation: uc(t ) = x[c0] + 3τ (t ).
Equation (5), which corresponds to equation (18) in [18], is
only valid when a significant temperature drop has already
taken place (as manifested by the appearance of the cooling
time scale t as a ‘delay’ in the expression for τ, see [18]).
Equation (5) is derived in the absence of an electric field; only
an exponential drop in the bulk temperature is assumed. The
electric field shown in figure 1(a) is only used to define a runaway region, so that the runaway fraction can be calculated.
In other words, it is assumed that the electric field does not
have time to influence the distribution significantly during the
temper­ature drop.
The runaway fraction calculated using equation (5) includes
only the electrons in the actual runaway region, i.e. particles
whose trajectories (neglecting collisional momentum-space
diffusion) are not confined to a region close to the origin.
In this case, the lower boundary of the runaway region is
given in terms of the limiting (non-relativistic) momentum
y for a given ξ: ycξ = (δ 2[(ξ + 1)E /2Ec − 1])−1/2 [17], where
Ec = 4πe3n ln Λ/mc 2 is the critical electric field for runaway
generation [19]. The temperature drop does however lead to an
isotropic high-energy tail (in the absence of an electric field).
By defining the runaway region as y > yc = (δ 2[E /Ec − 1])−1 / 2,
thereby including all particles with v > vc, equation (5) can be
simplified to
T (t ) = Tf + (T0 − Tf )e−t / t,
(3)
with T0 = 3.1 keV the initial temperature, Tf = 31 eV the final
temperature, and t = 0.3 ms the cooling time scale. We also
include a time-dependent electric field described by
E (t ) ⎛ E ⎞
T0
=⎜ ⎟
,
(4)
⎝ ED ⎠0 T (t )
ED
with (E /ED )0 = 1/530 the initial normalized electric field.
The temperature and electric-field evolutions are shown in
figure 1(a) and are the same as those used in figure 5 of [18],
as are all other parameters in this section.
Figure 1(b), in which the additional parameters
n = 2.8 ⋅ 1019 m−3, and Z eff = 1 were used, illustrates the distribution-function evolution during the temperature drop. The
figure shows that as the temperature decreases, most of the
electrons quickly adapt. At any given time t, the bulk of the distribution remains close to a Maxwellian corresponding to the
current temperature T (t). The initially slightly more energetic
electrons, although part of the original bulk population, thermalize less efficiently. On the short cooling time-scale, they
remain as a distinct tail, and as the thermal speed decreases
they become progressively less collisional. This process is evident in the first three time steps shown (t = 0.025–0.83 ms). In
the final time step, the electric field has become strong enough
to start to affect the distribution, and a substantial part of the
high-energy tail is now in the runaway region. This can be
seen from the qualitative change in the tail of the distribution,
which now shows a positive slope associated with a strong
flow of particles to higher momenta.
2
2
nr
=
uc e−uc + erfc(uc ),
(6)
n
π
where erfc(x) is the complementary error function. By default,
code uses such an isotropic runaway region, which is a good
3
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 2. Hot-tail runaway density obtained using code—with (blue, dashed) and without (yellow, dash–dotted; red, dotted) an electric
field included during the temperature drop—and the analytical estimates equations (5) and (6) (black, solid), for the temperature and E-field
evolution in figure 1(a). An (a) ξ-dependent and (b) isotropic lower boundary of the runaway region was used. The collision operator in
equation (2) was used for the blue and yellow lines, whereas its non-relativistic limit was used for the red and black lines.
approximation in the case of only Dreicer and avalanche generation (especially once the runaway tail has become substantial); however, in the early stages of hot-tail-dominated
scenarios, the isotropic runaway region significantly overestimates the actual runaway fraction, and the lower boundary
ycξ must be used.
Figure 2 compares the runaway density evolution computed with code, using both ξ-dependent and isotropic
runaway regions, to equations (5) and (6), respectively. The
parameters of the hot-tail scenario shown in figure 1 were
used, and no avalanche source was included in the calcul­
ation. The collision operator used in [18] is the non-relativistic limit of equation (2), with cξ = 0 (since the distribution
is isotropic in the absence of an electric field). code results
using both this operator (red, dotted) and the full equation (2)
(yellow, dash–dotted) are plotted in figure 2, with the latter
producing ∼ 50% more runaways in total. This difference can likely be explained by the relatively high initial
temper­ature (3 keV) in the scenario considered, in which
case the non-relativistic operator is not strictly valid for the
highest-energy particles. Good agreement between code
results and equations (5) and (6) (black, solid) is seen for
the saturated values in the figure. A code calculation where
the electric-field evolution is properly included in the kinetic
equation (corresp­onding to the distribution evolution in
figure 1(b)) is also included (blue, dashed), showing increased
runaway production. With the isotropic runaway region (figure
2(b)), the increase is smaller than a factor of 2, and neglecting
the influence of the electric field can thus be considered reasonable for the parameters used, at least for the purpose of
gaining qualitative understanding. With the ξ-dependent runaway region (figure 2(a)), the change in runaway generation is
more pronounced, and the inclusion of the electric field leads
to an increase by almost an order of magnitude. Note that the
final runaway density with the electric field included is very
similar in figures 2(a) and (b), indicating that the details of the
lower boundary of the runaway region become unimportant
once the tail is sufficiently large. Throughout the remainder of
this paper we will make use of the isotropic runaway region.
We conclude that, in order to obtain quantitatively accurate results, the electric field should be properly included,
and a relativistic collision operator should be used. This is
especially true when modelling ITER scenarios, where the
initial temperature can be significantly higher than the 3 keV
used here.
3. Conservative linearized Fokker–Planck collision
operator
Treating the runaway electrons as a small perturbation to a
Maxwellian distribution function, the Fokker–Planck operator for electron–electron collisions [20, 21] can be linearized and written as C{ f } C l{ f } = C tp + C fp. The so-called
test-­particle term, C tp = C nl{ f1 , fM}, describes the perturbation colliding with the bulk of the plasma, whereas the
field-­particle term, C fp = C nl{ fM , f1}, describes the reaction
of the bulk to the perturbation. Here C nl is the non-linear
Fokker–Planck–Landau operator, fM denotes a Maxwellian,
and f1 = f − fM the perturbation to it ( f1 fM). Collisions
described by C{ f1 , f1 } are neglected since they are second
order in f1. The full linearized operator Cl conserves particles,
momentum and energy. Since it is proportional to a factor
exp(−y 2 ), the field-particle term mainly affects the bulk of the
plasma, and is therefore commonly neglected when studying
runaway-electron kinetics. The test-particle term in equation (2) only ensures the conservation of particles, however,
not momentum or energy.
Under certain circumstances, it is necessary to use a fully
conservative treatment also for the runaway problem, in particular when considering processes where the conductivity of
the plasma is important. In the study of runaway dynamics
during a tokamak disruption using a self-consistent treatment
of the electrical field, accurate plasma-current evolution is
essential, and the full linearized collision operator must be
used. A non-linear collision operator valid for arbitrary particle (and bulk) energy has been formulated [22, 23]. The col­
lision operator originally implemented in code is the result
of an asymptotic matching between the highly relativistic
limit of the test-particle term of the linearized version of that
operator, with the usual non-relativistic test-particle operator
[24], and is given in equation (2). The relativistic field-particle
term is significantly more complicated, however, and its use
would be computationally more expensive. Here we instead
4
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 3. (a) Parallel momentum and (b) energy moments of the distribution function in code, using different collision operators. Initially,
E = 50 V m−1 and Z eff = 1 were used, but for t > t0, the electric field was turned off and the ion charge set to Z eff = 0. Using two Legendre
modes for the field-particle term was sufficient to achieve good conservation of energy and parallel momentum.
implemented the non-relativistic field-particle term, as formulated in [25, 26]. As will be shown, this operator (together
with the non-relativistic limit of equation (2)) accurately
reproduces the Spitzer conductivity for sufficiently weak electric fields and temperatures where the bulk is non-relativistic.
Using the normalization in section 2, the field-particle term is
is modest (the operator ∇2v is proportional to l2, with l the
Legendre mode index, and G and H therefore decay rapidly
with increasing l).
The conservation properties of the full non-relativistic col­
lision operator (equations (2) and (7)), as well as the relativistic
test-particle operator in equation (2), are shown in figure 3.
As an electric field is applied to supply some momentum
and energy to the distribution, the parallel momentum (figure
3(a)) quickly reaches a steady-state value corresponding to
the plasma conductivity, which differs by about a factor of
two for the two operators (see below). The electric field is
turned off at t = t0 = 100 collision times (and Z eff = 0 is
imposed to isolate the behavior of the electron–electron col­
lision operator), at which point the parallel momentum for the
operator in equation (2) (blue, dashed) is lost on a short time
scale as the distribution relaxes back towards a Maxwellian. In
contrast, the full linearized operator (black, solid) conserves
parallel momentum in a pure electron plasma, as expected.
The electric field continuously does work on the distribution, a large part of which heats the bulk electron population, but the linearization of the collision operator breaks
down if the distribution deviates too far from the equilibrium
solution. As long as a non-vanishing electric field is used
together with an energy conserving collision operator, an
adaptive sink term removing excess heat from the bulk of
the distribution must be included in equation (1) to guarantee
a stable solution. Physically this accounts for loss processes
that are not included in the model, such as line radiation,
bremsstrahlung and radial heat transport. The magnitude of
the black line in figure 3(b) therefore reflects the energy content of the runaway population—not the total energy supplied by the electric field—since a constant bulk energy is
enforced. The energy sink is not included for t > t0 (since
E = 0), however, and the energy conservation observed is
due to the properties of the collision operator itself. Again,
the use of the collision operator in equation (2) is associated
with a quick loss of kinetic energy as soon as the electric
field is removed.
The electrical conductivity of a fully ionized plasma subject to an electric field well below the Dreicer value—the
Spitzer conductivity—can be expressed as
⎡ 2x 2 ∂ 2G
⎤
2
c
−2 2
fp
− 2 H + 4πF ⎥ ,
Cˆ = 3C/ 2 e−v¯e x ⎢ 4
(7)
⎣ v¯e ∂x 2
⎦
v¯e
π
where G and H are the Rosenbluth potentials, obtained from
the distribution using
v˜e 2∇2v H = −4πF ,
v˜e 2∇2v G = 2H .
(8)
The system of equations composed of equations (7) and (8),
together with the non-relativistic limits of equations (1) and
(2) (y → x and δ → 0), is discretized (see [5]) and solved using
an efficient method described in [27]. The equations are combined into one linear system of the form
⎛ M11 M12 M13 ⎞⎛ F ⎞ ⎛ Si ⎞
⎜M M
0 ⎟⎟⎜⎜ G ⎟⎟ = ⎜⎜ 0 ⎟⎟,
(9)
22
⎜ 21
⎝ 0 M31 M33 ⎠⎝ H ⎠ ⎝ 0 ⎠
where the first row describes the kinetic equation (1) (with Si
representing any sinks or sources), and the second and third
rows correspond to equation (8). This approach makes it possible to consistently solve for both the Rosenbluth potentials
and the distribution with a single matrix operation. Since
there is no explicit need for the Rosenbluth potentials, however, G and H can be eliminated by solving the block system
analytically:
1
−1
((10)
M11 − [M12 − M13 M−
33 M32] M 22 M21)F ≡ MF = Si.
If only the test-particle operator (equation (2)) is used, M
reduces to M11. Since the Rosenbluth potentials are defined
through integrals of the distribution, the field-particle term
introduces a full block for each Legendre mode into the
normally sparse matrix describing the system. However,
the integral dependence on F also implies that significantly
fewer modes are required to accurately describe the potentials (compared to F ), and the additional computational cost
5
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 4. (a) Conductivity (normalized to the Spitzer value) and (b) normalized runaway density, as functions of time for different collision
operators (non-relativistic full linearized: solid; relativistic test-particle: dashed) and E-field strengths (E /ED = 1%: yellow; E /ED = 5%:
red; E /ED = 6%: blue), considering only Dreicer runaway generation. The parameters T = 1 keV, n = 5 ⋅ 1019 m−3 and Z eff = 1 were used.
where n r is the number density of runaway electrons, n̄ is the
density normalized to its reference value, and δ D is the Dirac
δ-function. In the derivation, the momentum of the incoming
particle is assumed to be very large (simplifying the scattering
cross section) and its pitch-angle vanishing (ξ = 1). It is also
assumed that the incoming particle is unaffected by the interaction. These conditions imply that the generated secondary particles are all created on the curve ξ = ξ2 = δy /(1 + 1 + δ 2y 2 )
ne 2
σS = L (Z eff )
,
(11)
Z eff mνee
where L (Z eff ) is a transport coefficient which takes the value
L 2 in a pure hydrogen plasma [10]. Figure 4 demonstrates
that the conductivity calculated with code reproduces the
Spitzer value for moderate electric-field strengths, if the conservative collision operator is used, and the initial Maxwellian
adapts to the applied electric field on a time scale of roughly
10 collision times. For field strengths significantly larger than
Ec, the conductivity starts to deviate from σS, as a runaway tail
begins to form (figure 4(b)); in this regime, the calculation in
[10] is no longer valid. Using the collision operator in equation (2) consistently leads to a conductivity which is lower by
about a factor of 2, as expected (see for instance [28]). The
runaway growth is also affected, with the conserving operator
leading to a larger runaway growth rate.
(which is a parabola in [ y∥, y⊥] -space), and that all runaways
(from the point of view of the avalanche source) are assumed
to have momentum p = γv /c = δy 1 (since SˆRP ∝ n r). They
can therefore contribute equally strongly to the avalanche process. This has the peculiar and non-physical consequence that
particles can be created with an energy higher than that of any
of the existing runaways. The δ-function in ξ is numer­ically
ill-behaved, as it produces significant oscillations (Gibbs
phenom­enon) when discretized using the Legendre-mode
decomposition employed in code (see figure 5(a)).
An operator that relaxes the assumption of very large runaway momentum has been presented by Chiu et al [11]. It has
the form
4. Improved operator for knock-on collision
The Fokker–Planck collision operators discussed in section 3 accurately describe grazing collisions—small-angle
deflections which make up the absolute majority of particle
interactions in the plasmas under consideration. Large-angle
collisions are usually neglected as their cross section is significantly smaller, but in the presence of runaway electrons they
can play an important role in the momentum space dynamics,
as an existing runaway can transfer enough momentum to a
thermal electron in one collision to render it a runaway, while
still remaining in the runaway region itself. Such knock-on
collisions can therefore lead to an exponential growth of the
runaway density—an avalanche [13, 29].
In the absence of a complete solution to the Boltzmann
equation, we model avalanche runaway generation using
an additional source term in the kinetic equation (1), evaluated for y > yc. A commonly used operator was derived by
Rosenbluth and Putvinski [13] and takes the form
⎡
⎛
3πδ 3
n
1 ∂
1
SˆRP = r n¯ 2 ⎢
δ D(ξ − ξ2 ) 2 ⎜⎜
˜
⎢
n
y
y
∂
16
ln
Λ
−
+ δ 2y 2
1
1
⎝
⎣
ˆCh( y, ξ ) = n¯ 2πe n˜δ x ( y )4F ( y ) Σ(γ , γ ),
S(13)
in
in
in
m2c 3 ν˜ee y 2ξ
4
where
Σ(γ , γin ) =
−
3
⎡
γin 2
⎢(γin − 1)2
(γin 2 − 1)(γ − 1)2(γin − γ )2 ⎣
⎤
(γ − 1)(γin − γ )
(2γin 2 + 2γin − 1 − (γ − 1)(γin − γ ))⎥
⎦
γin 2
(14)
is the Møller scattering cross section [12] and F is the
pitch-angle-averaged distribution of incoming runaways
with properties yin and γin. All incoming particles are thus
still assumed to have zero pitch angle (ξ = 1), but their
energy distribution is properly taken into account. In
code, F is computed from the 0th Legendre mode of F;
F = 2F0.
From the conservation of 4-momentum in a collision, the
momentum-space coordinates are related through
⎞⎤
⎟⎥ ,
⎟⎥
⎠⎦
(12)
6
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 5. Contour plots of the magnitude of the source in (a) equation (12) and (b) equation (13) in (y∥,y⊥) momentum space, given the
same electron distribution. The plotted quantity is log10 Ŝ and yc defines the lower bound of the runaway region. The angle-averaged source
magnitudes are shown in (c). The parameters T = 1 keV, n = 5 ⋅ 1019 m−3, Z eff = 1 and E = 1 V m−1, with max( y ) = 70, were used to
obtain the distribution, and the simulation was run for 300 collision times with primary generation only.
(γ − 1)(γin + 1)
ξ=
,
(15)
(γ + 1)(γin − 1)
(as expected). This reduces the sensitivity of the avalanche
growth rate to the choice of momentum cut-off (as long as
ycut − off ⩽ yc), and reaffirms our choice ycut − off = yc.
Figure 5(c) shows the source terms integrated over pitchangle, and as expected, the source in equation (13) extends
only up to y ymax / 2, whereas the source in equation (12) is
non-vanishing also for larger momenta. The amount of secondary runaways generated by the two sources agrees well at
low energies, but less so further away from the bulk. In this
particular case, the total source magnitude ∫ Sˆ y 2 dy dξ agrees
to within 25%, as most of the secondaries are created close to
the boundary of the runaway region.
which restricts the region where the source is non-vanishing
(this relation is analogous to the parabola ξ2 in the case of
the operator in equation (12)). Since the electrons participating in a collision are indistinguishable, it is sufficient
to consider only the cases where the energy of the created
secondary runaway is less than half of the primary energy,
(γ − 1) ⩽ (γin − 1)/2, which with the above equation leads to
the condition ξ ⩽ ξmax = γ /(γ + 1) . By the same argument,
the maximum attainable runaway energy in the simulation
(the maximum of the momentum grid) leads to the condition
ξ ⩾ ξmin = (γ − 1)(γmax + 1)/(γ + 1)(γmax − 1) .
The magnitudes of the two sources (12) and (13) are computed from a given typical runaway distribution function,
and shown in figures 5(a) and (b). Curves corresponding to
the parabola ξ2, as well as the limits ξmin and ξmax are also
included. Note that the amount of numerical noise is significantly reduced for the source in equation (13). In order to
avoid double-counting the small-angle collisions described by
the Fokker–Planck–Landau collision operator C, the knockon source must be cut off at some value of momentum sufficiently far from the thermal bulk. As can be seen from the
figure, however, the magnitude of both sources increases
with decreasing momenta, and the avalanche growth rate is
therefore sensitive to the specific choice of momentum cutoff. Since our particular interest is the generation of runaway
electrons, we choose to place the cut-off at y = yc, so that
the sources are non-vanishing only in the runaway region
[5, 15]. Secondary particles deposited just below the
threshold—although not technically runaways—could eventually diffuse into the runaway region, thereby potentially
increasing the Dreicer growth rate. In [30], such effects
were however shown to be negligible for the operator in
equation (12), indicating that the vast majority of particles
deposited at y < yc are slowed down rather than accelerated
4.1. Avalanche growth rates for the different operators
In general, the avalanche growth rate produced by the two
sources can differ substantially. We will illustrate this point
by considering the Møller cross section in more detail. We
choose to quantify the source magnitude for an arbitrary distribution by computing the cross section, integrated over the
energy of the outgoing (secondary) particle and normalized
to r 20, with r0 the classical electron radius. In other words, we
look at the total normalized cross section for an incoming particle with γin to participate in a knock-on collision resulting in
avalanche [31]:
KCh(γin ) =
∫γ
(γin − 1) /2 + 1
c
Σ(γ , γin )dγ
⎡ γ 2
γin 2
= (γin 2 − 1)−1 ⎢ in +
⎣ γc − 1 γc − γin
+
⎤
2γin − 1 ⎛ γc − 1 ⎞ γin + 1
ln⎜
− γc⎥ ,
⎟+
⎥⎦
γin − 1 ⎝ γin − γc ⎠
2
(16)
where γc = (E /Ec )/(E /Ec − 1) corresponds to the critical
momentum for runaway generation and the upper integration
7
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
Figure 6. (a) Contours (black, white) of the ratio of total cross-sections (KCh /KRP) for an electron with pin to contribute to the avalanche
process, as a function of pin = γinvin /c = γin 2 − 1 and E /Ec. (b) Ratio of avalanche growth rates ( [ΓCh − ΓD]/[ΓRP − ΓD] ) in code
simulations. The parameters T ∈ [0.1 eV, 5 keV], E /Ec ∈ [1.1, 1000] , n = 5 ⋅ 1019 m−3 and Z eff = 1 were used.
was set to either tmax, the first time step for which n r > 5%, or
the first time step in which the growth rate started to become
affected by the proximity of the runaway tail to the end of the
simulation grid, whichever occurred first. The parameters of
the scan were chosen to focus on the most interesting region
of figure 6(a)—by performing longer simulations on larger
momentum grids, the upper part of the figure could also be
studied. Exact agreement between figures 6(a) and (b) can not
be expected, since the source, in addition to the cross section,
depends on the details of the runaway distribution. Figure 6(a)
should thus be viewed as a simplified analytical estimate for
figure 6(b). The different regions identified in figure 6(a)
are still apparent in figure 6(b), however they are somewhat
shifted in parameter space. In particular, the region where the
Rosenbluth–Putvinski operator produces a higher growth rate
is larger, whereas the opposite region—where the operator
in equation (13) dominates—is smaller, or at least shifted to
higher values of E /Ec.
Figure 7 shows all the data points in figure 6(b), as a function of temperature. The figure confirms that the region where
the more accurate operator produces a significantly higher
growth rate is only accessible at temperatures T < 100 eV
(in the domain of validity of a linearized treatment). As is evident in the figure, however, regions where the Rosenbluth–
Putvinski operator significantly overestimates the avalanche
growth rate (points below 1 on the vertical axis) are present
at all temperatures. The operator in equation (13) is thus of
general interest.
Since the electric field spike responsible for the acceleration of runaways during a tokamak disruption is induced by
the temperature drop, and therefore occurs slightly later than
the drop itself, the temperature is low during the majority
of the acceleration process. For significant runaway acceleration, E /Ec 1 is therefore required, and during the initial
part of the acceleration process, parameters are likely those
corresponding to the blue region of figure 6(b), where the
improved avalanche source produces a significantly higher
growth rate than the Rosenbluth–Putvinski operator. Postthermal-quench temperatures in ITER are expected be as low
as 10 eV and peak electric fields in disruptions can reach 80 V
m−1 or more [32]. Towards the end of the thermal quench, the
boundary stems from the condition leading to ξmax. This
expression is relevant to the source in equation (13), which
uses the complete cross section (14), whereas for the more
simple source in equation (12), only the leading-order term in
γin in the scattering cross section is taken into account. This
corresponds to taking the high-energy limit of the above equation, so that
1
KRP =
(17)
γc − 1
becomes a simple constant.
To systematically explore the relative magnitude of the
two sources, the ratio KCh /KRP is plotted in figure 6(a). As
expected, the two expressions agree very well at high primary momenta. At somewhat lower momenta, of the order
γ ≈ p 5, two distinct regions are discernible. For E /Ec 10
(the orange region), the simplified cross section is larger than
the full expression, and the Rosenbluth–Putvinski operator
(12) is likely to overestimate the avalanche generation. For
E /Ec 10, the opposite is true, and the operator in equation (13) has a significantly larger cross section for E /Ec 30
(the blue region). The more accurate operator (13) should thus
be expected to produce more runaways when the runaway
population is at predominantly low energies, and E /Ec is large.
For both of these conditions to be fulfilled simultaneously (and
at the same time avoid a slide-away scenario), the temperature
must be low so that E /ED 1 even for large E /Ec. The effect is
also likely to be most apparent at relatively early times, before
the runaway tail has extended to multi-MeV energies.
code simulations support the above conclusions and show
excellent qualitative agreement, as shown in figure 6(b).
The figure shows the ratio of final avalanche growth rates
(ΓCh − ΓD)/(ΓRP − ΓD), with Γi = n−r 1(dn r /dtˆ ) the growth rate
obtained in a code run using source i (here the subscript D
denotes pure Dreicer generation). Each marker in the figure is
thus computed from three separate code runs. As a proxy for
pin, the average runaway momentum pr,av in the final time step
tf of the simulation without a source was used, and for a given
E /Ec, different pr,av were obtained from simulations with
varying values of T (and corresponding values of E /ED). The
simulations were run for tmax = 5000 collision times, and tf
8
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
implementation described. The operator was found to reproduce the expected Spitzer conductivity in the relevant parameter regime and showed excellent conservation properties.
The use of such an operator is essential for the correct current
evolution in self-consistent modelling, and in particular when
studying the interplay between current and electric-field evolution and runaway-electron generation during a disruption.
The process of avalanche multiplication of the runaway
population via close Coulomb collisions was also considered,
and an improved operator, relaxing some of the approx­imations
of the commonly used Rosenbluth–Putvinski operator, was
discussed. It was found that the avalanche growth rate can be
significantly affected—increased for low temperatures and
high E /Ec and decreased for low E /Ec—by the use of the new
operator. The change to the growth rate can be especially large
during the early stages of the runaway acceleration process,
thus potentially affecting the likelihood of a given runaway
seed transforming into a serious runaway beam, and use of
the improved operator is of particular relevance in disruption
scenarios.
The work presented in this paper paves the way for a
better understanding of runaway-electron dynamics in rapidly
changing scenarios, for instance during tokamak disruptions.
It enables more accurate assessment of the risks posed by
runaway electrons in situations of experimental interest, par­
ticularly in view of future tokamaks such as ITER.
Figure 7. Ratio of avalanche growth rates ( [ΓCh − ΓD] / [ΓRP − ΓD] )
in code simulations, as a function of temperature. The same
parameters as in figure 6 were used.
normalized electric field is then E /Ec ≈ 1300 (with E = 80 V
m−1, T = 50 eV and n = 1 ⋅ 10 20 m−3). A typical ITER disruption would thus (at least initially) be firmly in the blue region
of figure 6(b), and the avalanche growth should be significantly higher than what the Rosenbluth–Putvinski source predicts. As the temperature is low, the runaways will also spend
a comparatively long time at low momenta ( p 1), where
the disagreement between the operators is most pronounced.
Note that, according to the figure, an average runaway
energy of several MeV (p > 5–10) is needed for the difference between the growth rates to become small for all E /Ec,
at which point the most energetic electrons will have reached
energies of several tens of MeV or more. However, since the
electric field changes rapidly, the runaways may experience
parameters corresponding to both the orange and blue regions
in figure 6(b) before reaching such energies. Further work is
therefore needed to assess the overall impact on the avalanche
growth of using the improved operator (13), although it is
clear that its use is essential for accurate analysis.
Acknowledgments
The authors are grateful to I. Pusztai and E. Hirvijoki for
fruitful discussions. This work has been carried out within the
framework of the EUROfusion Consortium and has received
funding from the Euratom research and training programme
2014–2018 under grant agreement No 633053. The views and
opinions expressed herein do not necessarily reflect those of
the European Commission. The authors also acknowledge
support from Vetenskapsrådet, the Knut and Alice Wallenberg
Foundation and the European Research Council (ERC2014-CoG grant 647121).
5. Summary
Runaway electrons are intimately linked to dynamic scenarios, as they predominantly occur during disruptions and
sawtooth events in tokamaks. An accurate description of their
dynamics in such scenarios requires kinetic modelling of rapidly changing plasma conditions, and mechanisms such as
hot-tail runaway generation add to the already interesting set
of phenomena of importance to the evolution of the runaway
population.
In this paper we have described the modelling of several
such processes, using the numerical tool code to calculate the
momentum-space distribution of runaway electrons. In particular, we have investigated rapid-cooling scenarios where
hot-tail runaway-electron generation is dominant. Good
agreement with previous theoretical work was observed, but
code simulations also allow for flexible study of a variety of
parameter regimes not readily accessible in analytical treatments, and involving other processes such as avalanche generation or synchrotron radiation.
Furthermore, the full linearized non-relativistic Fokker–
Planck–Landau collision operator was discussed, and its
References
[1] Dreicer H. 1959 Phys. Rev. 115 238
[2] Dreicer H. 1960 Phys. Rev. 117 329
[3] Hollmann E.M. et al 2015 Phys. Plasmas 22 021802
[4] Boozer A.H. 2015 Phys. Plasmas 22 032504
[5] Landreman M., Stahl A. and Fülöp T. 2014 Comput. Phys.
Commun. 185 847
[6] Stahl A., Hirvijoki E., Decker J., Embréus O. and Fülöp T.
2015 Phys. Rev. Lett. 114 115002
[7] Hirvijoki E., Pusztai I., Decker J., Embréus O., Stahl A. and
Fülöp T. 2015 J. Plasma Phys. 81 475810502
[8] Decker J., Hirvijoki E., Embréus O., Peysson Y., Stahl A.,
Pusztai I. and Fülöp T. 2015 Plasma Phys. Control. Fusion
58 025016
[9] Paz-Soldan C. et al 2014 Phys. Plasmas 21 022514
[10] Spitzer L. and Harm R. 1953 Phys. Rev. 89 977
[11] Chiu S.C., Rosenbluth M.N., Harvey R.W. and Chan V.S. 1998
Nucl. Fusion 38 1711
9
A. Stahl et al
Nucl. Fusion 56 (2016) 112009
[23] Braams B.J. and Karney C.F.F. 1989 Phys. Fluids B 1 1355
[24] Papp G., Drevlak M., Fülöp T. and Helander P. 2011 Nucl.
Fusion 51 043004
[25] Catto P.J. and Tsang K.T. 1977 Phys. Fluids 20 396
[26] Li B. and Ernst D.R. 2011 Phys. Rev. Lett. 106 195002
[27] Landreman M. and Ernst D.R. 2013 J. Comput. Phys. 243 130
[28] Helander P. and Sigmar D. 2002 Collisional Transport in
Magnetized Plasmas (Cambridge: Cambridge University
Press)
[29] Jayakumar R., Fleischmann H.H. and Zweben S. 1993 Phys.
Lett. A 172 447
[30] Nilsson E., Decker J., Peysson Y., Granetz R.S., SaintLaurent F. and Vlainic M. 2015 Plasma Phys. Control.
Fusion 57 095006
[31] Aleynikov P. and Breizman B.N. 2015 Phys. Rev. Lett.
114 155001
[32] Papp G., Drevlak M., Fülöp T., Helander P. and Pokol G.I.
2011 Plasma Phys. Control. Fusion 53 095004
[12] Møller C. 1932 Ann. Phys. 406 531
[13] Rosenbluth M.N. and Putvinski S.V. 1997 Nucl. Fusion
37 1355
[14] Papp G., Stahl A., Drevlak M., Fülöp T., Lauber Ph.W. and
Pokol G.I. 2015 Eur. Conf. Abstr. 39E P1.173
[15] Harvey R.W., Chan V.S., Chiu S.C., Evans T.E.,
Rosenbluth M.N. and Whyte D.G. 2000 Phys. Plasmas
7 4590
[16] Helander P., Smith H., Fülöp T. and Eriksson L.-G. 2004 Phys.
Plasmas 11 5704
[17] Smith H., Helander P., Eriksson L.-G. and Fülöp T. 2005 Phys.
Plasmas 12 122505
[18] Smith H.M. and Verwichte E. 2008 Phys. Plasmas 15 072502
[19] Connor J.W. and Hastie R.J. 1975 Nucl. Fusion 15 415
[20] Landau L.D. 1936 Phys. Z. Sowjetunion 10 154
[21] Rosenbluth M.N., MacDonald W.M. and Judd D.L. 1957
Phys. Rev. 107 1
[22] Beliaev S.T. and Budker G.I. 1956 Sov. Phys.—Dokl. 1 218
10
Paper D
A. Stahl, E. Hirvijoki, J. Decker, O. Embréus, and T. Fülöp,
Effective critical electric field for runaway electron generation,
Physical Review Letters 114, 115002 (2015).
http://dx.doi.org/10.1103/PhysRevLett.114.115002
http://arxiv.org/abs/1412.4608
week ending
20 MARCH 2015
PHYSICAL REVIEW LETTERS
PRL 114, 115002 (2015)
Effective Critical Electric Field for Runaway-Electron Generation
1
A. Stahl,1,* E. Hirvijoki,1 J. Decker,2,1 O. Embréus,1 and T. Fülöp1
Department of Applied Physics, Chalmers University of Technology, SE-412 96 Gothenburg, Sweden
2
École Polytechnique Fédérale de Lausanne (EPFL), Centre de Recherches en Physique des Plasmas (CRPP),
CH-1015 Lausanne, Switzerland
(Received 26 August 2014; published 17 March 2015)
In this Letter we investigate factors that influence the effective critical electric field for runaway-electron
generation in plasmas. We present numerical solutions of the kinetic equation and discuss the implications
for the threshold electric field. We show that the effective electric field necessary for significant runawayelectron formation often is higher than previously calculated due to both (1) extremely strong dependence
of primary generation on temperature and (2) synchrotron radiation losses. We also address the effective
critical field in the context of a transition from runaway growth to decay. We find agreement with recent
experiments, but show that the observation of an elevated effective critical field can mainly be attributed to
changes in the momentum-space distribution of runaways, and only to a lesser extent to a de facto change
in the critical field.
DOI: 10.1103/PhysRevLett.114.115002
PACS numbers: 52.25.Xz, 52.55.Fa, 52.55.Pi, 52.65.Ff
Introduction.—In a plasma, an electron beam accelerated
by an electric field is damped by collisional friction against
the bulk plasma and by emission of electromagnetic
radiation. Since the collisional friction decreases with
increasing velocity of the electrons, a large enough electric
field may overcome the collisional damping and accelerate
electrons to relativistic speeds, leading to the formation
of a runaway-electron (RE) beam. In laboratory plasmas,
much attention has been given to the potentially dangerous,
highly relativistic RE beams that can be generated in
tokamak disruptions [1]. Runaway acceleration can also
occur in nondisruptive plasmas due to the Ohmic electric
field, if the plasma density is low. In addition, runaway
electrons are ubiquitous in atmospheric and space plasmas,
e.g., as a source of red sprites in the mesosphere [2] and in
lightning discharges in thunderstorms [3], and their occurrence in solar flares has been suggested [4].
The critical (threshold) electric field for runaway
generation, Ec ¼ ne e3 ln Λ=4πϵ20 me c2 , is a classic result
in plasma physics [5]. Because of relativistic effects, it is
the weakest field at which electron runaway is possible.
Here, ne and me are the number density and rest mass of the
electrons, respectively, ln Λ is the Coulomb logarithm, c is
the speed of light, and e is the magnitude of the elementary
charge. Recent experimental evidence from several tokamaks indicates that the electric field strength necessary for
RE generation could in fact be several times larger than the
critical electric field [6–8]. As the classic expression for the
critical field only considers the balance between electric
field and Coulomb collisions, many potential mechanisms
affecting the RE generation are left out. In this Letter we
investigate the role of the background plasma temperature
and synchrotron radiation reaction as possible explanations
for these observations.
0031-9007=15=114(11)=115002(5)
An appreciation for the importance of temperature and
synchrotron effects can be gained by considering the
energy balance for an electron experiencing electric field
acceleration, collisional damping, and the AbrahamLorentz radiation reaction force [9]:
3γ 2
γ2
3γ 2
Frad ¼ k v̈ þ 2 ðv · v_ Þ_v þ 2 v · v̈ þ 2 ðv · v_ Þ2 v ;
c
c
c
ð1Þ
where k ¼ e2 γ 2 =ð6πε0 c3 Þ, v is the electron velocity, and
γ ¼ ½1 − ðv=cÞ2 −1=2 is the relativistic mass factor. At the
critical electric field, acceleration due to the electric field
balances the friction due to collisions and radiation, so the
particle energy is constant. This means that γ_ ¼ 0 and
v· v_ ¼0, implying that qE·vþFc ·v−ðe2 γ 4 =6πϵ0 c2 Þ_v · v_ ¼0.
At constant energy, v_ ¼ ðe=γme Þv × B, so that
v_ · v_ ¼ ð1 − ξ2 Þv2 ω2c . Here B is the magnetic field,
ξ ¼ p∥ =p is the cosine of the particle pitch angle, p ¼ jpj ¼
γv=c is the magnitude of the normalized momentum,
and ωc ¼ eB=γme is the Larmor frequency. This means
that for the electric field to accelerate the electron, it has to
be larger than
E 1 c2
2ε0 B2 1 − ξ2 v 2
γ ≡ hðv; ξÞ:
> 2þ
Ec ξ v
3ne me ln Λ ξ c
ð2Þ
Note that hðv; ξÞ > 1, always. In Fig. 1, h is illustrated for
typical ITER [10] plasma parameters as a function of ξ and
the electron kinetic energy Ekin. From this simple estimate
of the single electron energy balance, we see that the value
of E=Ec needed to accelerate electrons can be significantly
larger than unity and increases with Ekin in the MeV range.
This warrants a more thorough investigation of the RE
dynamics close to the critical electric field. The first term
115002-1
© 2015 American Physical Society
102
function describing the particle speeds is determined by the
temperature. This introduces a temperature dependence to
the effective critical field, since the number of particles with
speed above any threshold speed is temperature dependent.
Mathematically, this can be understood from the primary
runaway growth rate [5]:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dnr
∼ ne νee E −3ð1þZeff Þ=16 exp ½−1=ð4EÞ − ð1 þ Zeff Þ=E dt
101
0.6
E/Ec
ξ=v /v
0.8
0.4
100
0.2
−1
0
10
1
2
10
2
10
Ekin / (mec )
FIG. 1 (color online). Magnitude of the normalized electric
field (E=Ec ) necessary to compensate for collisional friction and
synchrotron damping for a single electron in an ITER-like plasma
with ne ¼ 1020 m−3 , T e ¼ 10 keV, and B ¼ 5 T.
10000
0
3000
E/E
D =1
−4
1000
−8
300
0.1
100
0.0
2
0.0
1
30
10
1.2 2 3
− 12
− 16
−1
log10 (n−1
e dnr /dt) [s ]
in Eq. (2) is related to the collisional damping, which is
temperature dependent, while the second term—which
vanishes for purely parallel motion (ξ ¼ 1)—is due to the
radiation reaction force. We will consider both these effects
in more detail.
The single electron estimate depends strongly on ξ, v,
and various plasma parameters, but neglects the collisional
Coulomb diffusion that spreads the electrons in velocity
space. An accurate estimate for the threshold electric field
can thus only be obtained using kinetic calculations that take
into account the details of the electron distribution function.
Here we make use of COllisional Distribution of Electrons
(CODE) [11], an efficient finite-difference–spectral-method
tool that solves the two-dimensional momentum space
kinetic equation in a homogeneous plasma. The Coulomb
collision operator in CODE is valid for arbitrary electron
energies [12]. Often secondary (or avalanche) generation of
REs, resulting from knockon collisions between REs and
thermal electrons, is the dominant RE generation mechanism, and CODE has also been equipped with an operator
describing this process [13]. For the present work, an
operator for synchrotron emission backreaction based on
the analysis in Refs. [14,15] was implemented. With the
numerical electron distribution function from CODE, we
may investigate the runaway generation dynamics for a
wide range of plasma parameters, to calculate, e.g., the
synchrotron radiation spectra of the REs [16], or to study
wave-particle interactions [17,18]. The parameters used
in this Letter reflect those common to magnetic fusion
experiments, but the arguments are generally applicable.
In particular, no effects specific to fusion plasmas (such as
a toroidal field configuration) have been assumed.
Temperature dependence of the critical electric field.—
At E ≳ Ec , only electrons already moving with approximately the speed of light may run away. Since the number of
plasma particles is finite (especially in a laboratory context),
the actual highest speed achieved by the background
electrons may be significantly less than c. Thus, if the
critical speed for RE generation at a given E field is larger
than this maximum speed, no electrons will be able to run
away. The width (in velocity space) of the distribution
(where νee ¼ ne e4 ln Λ=4πε20 m2e v3th is the collision frepffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
quency, vth ¼ 2T e =me is the electron thermal velocity,
and Zeff is the effective charge number of the plasma), which
is exponentially small in E ¼ E=ED ¼ ðT e =me c2 ÞðE=Ec Þ,
where ED is the Dreicer field [19]. There is thus an inherent
temperature dependence in the primary runaway growth
rate at a given value of E=Ec , and for significant RE
production on a short time scale it is not enough to only
require E > Ec [1,20].
We define a runaway electron as any electron with
p > pc ¼ ðE=Ec − 1Þ−1=2 , where pc is the critical momentum for electron runaway. In the absence of avalanche
generation, a quasisteady state for the RE distribution can
be calculated using CODE. In Fig. 2, the RE growth rate
for this primary distribution is displayed as a function of
the electron temperature and E=Ec . The figure indicates
that, for all temperatures T e ≲ 5 keV, the fraction of the
electron population that runs away in 1 s is less than 10−20
for all E=Ec < 1.5. In a plasma with ne ≲ 1020 m−3 and a
volume of a few tens of m3 (typical of fusion experiments),
essentially no runaway production (let alone detection) is
thus to be expected. It is also clear from the figure that,
for lower temperatures, a stronger normalized electric field
would be required for significant RE production (note that
Fig. 2 essentially covers the whole temperature range of
magnetic fusion plasma operation). We note that this
temperature dependence may increase the sensitivity of
future hotter tokamaks (like ITER) to the problem of
deleterious RE formation. The white and black contours
in Fig. 2 show the corresponding values of E=ED , and
we may conclude that E=ED must be at least larger than
1%–2% for significant runaway formation to occur, in
Te [eV]
1
10
week ending
20 MARCH 2015
PHYSICAL REVIEW LETTERS
PRL 114, 115002 (2015)
− 20
5
10
30
E/Ec
100
300
FIG. 2 (color online). Primary runaway growth rate (particle
fraction per second) as a function of temperature and electric field,
in the absence of synchrotron effects. White and black contours
refer to E=ED . The plasma parameters ne ¼ 5 × 1019 m−3 and
Zeff ¼ 1.5 were used.
115002-2
agreement with previous analytical findings [1,20]. In
practice, there are thus two conditions that must be
fulfilled: E=Ec > 1 and E=ED > k, for some k ∼ 1%–2%.
The second criterion is more restrictive for temperatures
below 5.1 and 10.2 keV (for k ¼ 1% and 2%), respectively.
Momentum loss due to synchrotron emission.—The
importance of synchrotron backreaction as a limiting factor
for the maximum energy achieved by REs has been
discussed before [21], and its importance in RE dynamics
has been investigated [14], also in the context of the critical
field for RE generation [8]. An accurate description of
the RE dynamics close to the critical field based on first
principles does, however, require kinetic modeling, and we
will investigate the effect of the synchrotron emission on
the effective critical field using CODE.
In a homogeneous plasma, the distribution function f
for electrons experiencing an electric field, Coulomb
collisions, and synchrotron radiation backreaction is determined by the gyro-averaged Fokker-Planck equation,
∂f eE∥
∂f 1 − ξ2 ∂f
∂
þ
þ · ðFrad fÞ ¼ Cffg þ Sava ;
ξ þ
∂t me c ∂p
∂p
p ∂ξ
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}
radiation
where E∥ is the parallel (to −B) electric field, Cf·g is
the collision operator, Sava is the source of secondary
(avalanche) runaways, and Frad is given by Eq. (1). Note
that the operator ð∂=∂pÞ · ðFrad fÞ conserves the number
of particles, unlike the corresponding operator used in
Ref. [14]. (The simplified operator used in Ref. [14] was
justified due to the focus on the high-energy tail of the
distribution function.) The magnetic force jFm j ≃ ωc me v,
characterized by the Larmor frequency ωc ¼ eB=γme,
typically dominates both the electric force and the radiation
reaction force. This implies that v · v_ ≃ 0, and that in the
coordinates p and ξ, the term accounting for the effects of
synchrotron radiation backreaction can be written as
∂
1 ∂ γp3 ð1 − ξ2 Þ
· ðFrad fÞ ¼ − 2
f
∂p
τr
p ∂p
2
∂ ξð1 − ξ Þ
f ;
þ
∂ξ
γτr
0.7
1000
0.5
300
0.3
100
0.1
30
2
3
5
10
30
0.9
m−3]
(a)
3000
19
10000
where τr ¼ 6πε0 ðme cÞ3 =ðe4 B2 Þ [14] is the radiation damping time scale. What ultimately determines the relative
importance of the synchrotron effects is the ratio of the
collision time 1=νee to the radiation time scale τr . For a
given magnetic field, we therefore expect the largest effect
on the distribution for high temperatures and low densities,
2
since 1=ðνee τr Þ ∼ T 3=2
e B =ne .
The radiation reaction force acts as an additional drag,
which increases with particle momentum. Therefore, it
ultimately prevents the REs from reaching arbitrary energies [21], and given enough time, the system will reach a
steady state where the RE growth rate vanishes. This occurs
only once the REs have reached very high energies; in the
initial phase (which is our interest in this section), the RE
growth rate is well defined. The rate is calculated as the flux
through a sphere of constant p, located well inside the RE
region. The change in this RE growth rate in CODE as a
result of the synchrotron radiation reaction is presented in
Fig. 3. From the figure, we conclude that the synchrotron
losses can reduce the RE rate substantially for weak E
fields—by several orders of magnitude at high temperatures and low densities—and it is therefore essential to
include these effects when considering near-critical RE
dynamics. The sharp cutoff for weak fields is in line with
the change in effective critical field associated with the
inclusion of the synchrotron drag [see Eq. (2)]. The full
kinetic simulation thus agrees qualitatively with the singleparticle estimate in the Introduction. For stronger electric
fields, the effects are less pronounced.
We note that in the post-thermal-quench conditions
associated with disruptions in tokamaks (low T e, high
ne ), the effects of synchrotron radiation reaction on the RE
growth rate are likely to be negligible, whereas in the case
of RE generation during the plasma current ramp-up phase
(high T e , low ne ), they can be substantial. There is thus a
qualitative difference in the momentum space dynamics in
these two cases—at least for near-critical electric fields—
and conclusions from ramp-up (or flattop) scenarios do not
necessarily apply under postdisruption conditions.
Critical field under experimental conditions.—Until now
we have only considered primary (Dreicer [19]) generation,
i.e., when the electrons gradually diffuse through velocity
space due to small-angle collisions and run away as they
ne [10
electric field
Te [eV]
week ending
20 MARCH 2015
PHYSICAL REVIEW LETTERS
PRL 114, 115002 (2015)
0.7
3
0.5
1
0.3
0.3
0.1
30
0.9
(b)
10
0.1
2
3
4
5
6
7
8
9
10
E/Ec
E/Ec
FIG. 3 (color online). Contour plots of the (a) temperature and (b) density dependence of the ratio between the primary RE growth rate in
CODE with and without synchrotron effects included. The parameters B ¼ 4 T, Zeff ¼ 1.5, and (a) ne ¼ 1 × 1019 m−3 , (b) T e ¼ 2 keV
were used. To ensure reliable results in (a), the parameter region has been restricted as the growth rates are negligible for low T e and
E fields, and E=ED approaches unity for high T e and E fields (cf. Fig. 2).
115002-3
PRL 114, 115002 (2015)
week ending
20 MARCH 2015
PHYSICAL REVIEW LETTERS
reach the critical velocity. An electron can enter the
runaway region also through a sudden collision at close
range, which throws it above the critical speed in a single
event. This leads to avalanche multiplication of the REs,
which is the dominant mechanism in many cases. The
avalanche growth rate is γ ava ∝ nRE ðE=Ec − 1Þ [13] (where
nRE is the runaway density), and the total runaway growth
rate is then the sum of primary and avalanche processes.
Above the critical field, the RE population will be
growing and below it will be decaying; according to the
estimate in Ref. [13] (taking only losses due to Coulomb
collisions into account), the transition should occur at
E=Ec ¼ 1. However, in an experiment designed to test
this at the DIII-D tokamak [7], the measured transition from
RE growth to decay occurred at E=Ec ¼ 3–5. In the
experiments, after first generating a substantial RE population, the plasma density was rapidly increased so as to
raise Ec . The value of E=Ec for the transition was
determined from the change from growth to decay in the
hard-x-ray (HXR) signal, as well as from the synchrotron
emission (in the visual range). These signals were thus
treated as a straightforward representation of the number of
REs. This assumption is not necessarily valid in the present
context, as it neglects the influence of the distribution of
RE energies on the emitted radiation. We illustrate this by
calculating the synchrotron emission from RE distributions
obtained with CODE.
The synchrotron radiation emitted by an electron is
highly dependent on its energy and the curvature of its
trajectory, with highly energetic particles with large pitch
angles emitting most strongly and at the shortest wavelengths. The total emission is therefore very sensitive to the
shape of the electron distribution function [16]. A change
in the distribution shape may in fact lead to a reduction in
the emitted synchrotron power, even though the size of the
RE population is constant, or even increasing. Here we use
the code SYRUP [16] to calculate the synchrotron spectrum
emitted by CODE distributions, in order to investigate the
role of this effect in the experiment in Ref. [7].
Accessing the physics underlying the evolution of the
RE distribution in the experiment may by done in CODE
by ramping down the electric field strength in the presence
of a significant RE tail, dominated by the avalanche
mechanism. The purpose here is to give a qualitative
explanation for the observed discrepancy. The CODE
calculation was started with a constant E=Ec ¼ 12, to
produce a significant RE tail, and after 1.2 s the electric
field was gradually ramped down during 1 s—a time scale
consistent with the experiments described in Ref. [7]. The
synchrotron spectrum at each time step was calculated
using SYRUP. In the calculations, the maximum particle
energy was restricted to 22 MeV, in agreement with
experimental observations [22]. (This maximum energy
limit cannot be attributed to the effects of the synchrotron
losses; other mechanisms, such as radial transport, must
(a)
(b)
FIG. 4 (color online). (a) Emitted synchrotron power in the
visual range and (b) kinetic energy contained in the RE
population during electric field ramp-down. In (a), each curve
is normalized to its peak value. The parameters T e ¼ 1.1 keV,
ne ¼ 2.5 × 1019 m−3 , and Zeff ¼ 1.2 were used. The black dotted
lines denote the beginning of the E-field ramp-down phase.
be invoked to explain it, but this is outside the scope of
this Letter.)
Figure 4(a) shows the total emitted power in the visual
spectral range (400–700 nm) during the simulation, for
various B-field strengths. The figure shows that as E=Ec
decreases, the emitted power transitions from growth to
decay in all cases, even though E=Ec is still well above
unity. There is a clear dependence on the magnetic field
strength, as the transition occurs at E=Ec ≃ 7.4 and 10.6 for
B ¼ 2.5 and 3.5 T, respectively. This suggests that the
origin of the effect is indeed the influence of the synchrotron reaction force on the distribution. In the experiments in
Ref. [7], the field was 1.5 T; in this case, we observe an
apparent effective critical field E=Ec ≃ 4.2, in agreement
with the experimental value of 3–5.
Although the RE growth rate decreases with E=Ec , in the
calculations it remains positive for all E=Ec > 1.1, 1.3, and
1.4 for B ¼ 1.5, 2.5, and 3.5 T, respectively. The effective
critical field is thus close to unity in all cases considered,
and is B dependent, as expected from the previous section.
The kinetic energy content of the RE population also
continues to increase well after the emitted synchrotron
power has started to decrease, as shown in Fig. 4(b). We
therefore conclude that what is observed in Fig. 4(a) is not
a fundamental change in the critical electric field but a
reduction in synchrotron emission. The changing electric
field leads to a reduced accelerating force, which modifies
the force balance, causing a redistribution of electrons in
velocity space towards lower energies. The density of
highly energetic particles with large pitch angles thereby
decreases, leading to a substantially reduced synchrotron
emission in the visual range.
Trends similar to those shown in Fig. 4(a) are seen
also in the infrared spectral range. Since several tokamaks
are equipped with fast visual or IR cameras dedicated
to observing the synchrotron emission from REs,
115002-4
PRL 114, 115002 (2015)
PHYSICAL REVIEW LETTERS
experimental study of the effect we describe should be
within reach. In particular, confirming the B dependence of
the apparent elevated critical field would be of interest.
In the experiments, E=Ec was decreased via a density
ramp-up. Increasing the density also indirectly modifies T e ,
Zeff , and the loop voltage [7] and reduces the magnitude of
the synchrotron effects. In the calculation in Fig. 4 these
changes were not taken into account; however, the plasma
parameters where chosen to reflect those observed at the
time of the transition from growth to decay in DIII-D shot
153545, considered in Ref. [7]. Although the trend shown
in Fig. 4 appears consistently and the results are largely
independent of the details of the ramp in electric field
strength and the energy cutoff, the specific value of E=Ec
for which the transition from growth to decay occurs is
sensitive to the plasma parameters. Similarly, the details of
the avalanche source used [13] are expected to affect only
the specific value of the transition, not the qualitative
behavior in Fig. 4 (indeed, the same trend is seen even
when the avalanche process is not included at all).
It was suggested in Ref. [7] that a significant part of the
detected HXR signal was due to RE bremsstrahlung
emission. Like the synchrotron emission, bremsstrahlung
is sensitive to the RE distribution function—the general
argument made above may be invoked to explain the
elevated growth-to-decay transition also in the case of
the HXR signal. The effects considered in this section thus
offer a plausible explanation for the mechanism behind the
experimentally detected elevated electric field for transition
from RE growth to decay. Our simulations show that the
observed increase is mainly an artifact of the methods used
to determine it, and only to a lesser extent the result of a
fundamental change to the critical field itself. The impact
on runaway mitigation schemes in future tokamaks is likely
to be negligible, especially considering that in a postdisruption scenario the impact on the distribution function
from synchrotron backreaction is small (due to the low T e
and high ne ).
Conclusions.—We have shown that several factors can
influence the effective critical electric field for both generation and decay of runaway electrons. The temperature
dependence of the RE growth rate means that in practice
E=ED > 1%–2% is required for substantial RE generation.
In addition, the drag due to synchrotron emission backreaction increases the critical field; for weak E fields, the
runaway growth rate can be reduced by orders of magnitude. The synchrotron effects on the distribution are most
prominent at high temperature and low density, however,
and their practical impact is likely negligible in postdisruption tokamak plasmas. By the same token, the effects
can be substantial during ramp-up and flattop.
Deducing changes to the size of the runaway population
using radiation can be misleading, as the emission is very
sensitive to the momentum-space distribution of the
week ending
20 MARCH 2015
runaways. This can lead to a perceived elevated critical
field in electric field ramp-down scenarios, despite a
continuing increase in the energy carried by the runaways
and their density. Our results are consistent with recent
experimental observations, giving a possible explanation
for the observed elevated critical field.
The authors are grateful to M. Landreman, G. Papp,
G. Pokol, and I. Pusztai for fruitful discussions. This
work has been carried out within the framework of the
EUROfusion Consortium and has received funding from
the Euratom research and training program 2014-2018
under Grant Agreement No. 633053. The views and
opinions expressed herein do not necessarily reflect those
of the European Commission.
*
[email protected]
[1] P. Helander, L.-G. Eriksson, and F. Andersson, Plasma
Phys. Controlled Fusion 44, B247 (2002).
[2] T. F. Bell, V. P. Pasko, and U. S. Inan, Geophys. Res. Lett.
22, 2127 (1995).
[3] A. V. Gurevich, G. M. Milikh, and R. A. Roussel-Dupre,
Phys. Lett. A 187, 197 (1994).
[4] G. D. Holman, Astrophys. J. 293, 584 (1985).
[5] J. W. Connor and R. J. Hastie, Nucl. Fusion 15, 415
(1975).
[6] R. S. Granetz, B. Esposito, J. H. Kim, R. Koslowski, M.
Lehnen, J. R. Martin-Solis, C. Paz-Soldan, T. Rhee, J. C.
Wesley, and L. Zeng, Phys. Plasmas 21, 072506 (2014).
[7] C. Paz-Soldan et al., Phys. Plasmas 21, 022514 (2014).
[8] J. R. Martin-Solis, R. Sanchez, and B. Esposito, Phys. Rev.
Lett. 105, 185002 (2010).
[9] W. Pauli, Theory of Relativity (Dover, New York, 1981).
[10] ITER, http://www.iter.org/
[11] M. Landreman, A. Stahl, and T. Fülöp, Comput. Phys.
Commun. 185, 847 (2014).
[12] G. Papp, M. Drevlak, T. Fülöp, and P. Helander, Nucl.
Fusion 51, 043004 (2011).
[13] M. N. Rosenbluth and S. V. Putvinski, Nucl. Fusion 37,
1355 (1997).
[14] F. Andersson, P. Helander, and L.-G. Eriksson, Phys.
Plasmas 8, 5221 (2001).
[15] R. D. Hazeltine and S. M. Mahajan, Phys. Rev. E 70,
046407 (2004).
[16] A. Stahl, M. Landreman, G. Papp, E. Hollmann, and T.
Fülöp, Phys. Plasmas 20, 093302 (2013).
[17] G. Pokol, T. Fülöp, and M. Lisak, Plasma Phys. Controlled
Fusion 50, 045003 (2008).
[18] A. Kómár, G. I. Pokol, and T. Fülöp, Phys. Plasmas 20,
012117 (2013).
[19] H. Dreicer, Phys. Rev. 117, 329 (1960).
[20] R. Jayakumar, H. H. Fleischmann, and S. J. Zweben, Phys.
Lett. A 172, 447 (1993).
[21] J. R. Martin-Solis, J. D. Alvarez, R. Sánchez, and B.
Esposito, Phys. Plasmas 5, 2370 (1998).
[22] E. M. Hollmann et al., Nucl. Fusion 53, 083004 (2013).
115002-5
Errata, Paper D
The color bar in Fig. 1 should extend between 1 and 7.4 (not 1 and 100), so
that the figure becomes:
Paper E
A. Stahl, M. Landreman, O. Embréus and T. Fülöp,
NORSE: A solver for the relativistic non-linear Fokker-Planck equation for electrons
in a homogeneous plasma,
Computer Physics Communications 212, 269-279 (2017).
http://dx.doi.org/10.1016/j.cpc.2016.10.024
http://arxiv.org/abs/1608.02742
Computer Physics Communications 212 (2017) 269–279
Contents lists available at ScienceDirect
Computer Physics Communications
journal homepage: www.elsevier.com/locate/cpc
NORSE: A solver for the relativistic non-linear Fokker–Planck
equation for electrons in a homogeneous plasma✩
A. Stahl a, *, M. Landreman b , O. Embréus a , T. Fülöp a
a
b
Chalmers University of Technology, Göteborg, Sweden
University of Maryland, College Park, MD, USA
article
info
Article history:
Received 9 August 2016
Received in revised form 20 October 2016
Accepted 31 October 2016
Available online 18 November 2016
Keywords:
Non-linear relativistic Fokker–Planck
equation
Kinetic plasma theory
Energetic electrons
Runaway electrons
a b s t r a c t
Energetic electrons are of interest in many types of plasmas, however previous modeling of their
properties has been restricted to the use of linear Fokker–Planck collision operators or non-relativistic formulations. Here, we describe a fully non-linear kinetic-equation solver, capable of handling large electricfield strengths (compared to the Dreicer field) and relativistic temperatures. This tool allows modeling
of the momentum-space dynamics of the electrons in cases where strong departures from Maxwellian
distributions may arise. As an example, we consider electron runaway in magnetic-confinement fusion
plasmas and describe a transition to electron slide-away at field strengths significantly lower than
previously predicted.
Program summary
Program title: NORSE
Program Files doi: http://dx.doi.org/10.17632/86wmgj758w.1
Licensing provisions: GPLv3
Programming language: Matlab
Nature of problem: Solves the Fokker–Planck equation for electrons in 2D momentum space in a homogeneous plasma (allowing for magnetization), using a relativistic non-linear electron–electron collision
operator. Electric-field acceleration, synchrotron-radiation-reaction losses, as well as heat and particle
sources are included. Scenarios with time-dependent plasma parameters can be studied.
Solution method: The kinetic equation is represented on a non-uniform 2D finite-difference grid and is
evolved using a linearly implicit time-advancement scheme. A mixed finite-difference-Legendre-mode
representation is used to obtain the relativistic potentials (analogous to the non-relativistic Rosenbluth
potentials) from the distribution.
© 2016 Elsevier B.V. All rights reserved.
1. Introduction
Energetic electrons, having speeds significantly larger than the
average speed of the thermal population, are ubiquitous in plasmas. Examples are found for instance in the solar corona [1] and
wind [2], and in solar flares [3,4]; in the ionosphere of the Earth [5]
and lightning discharges [6]; as well as in laboratory laser-plasma
accelerators [7] and inertial [8] and magnetic-confinement [9] fusion plasmas. In the latter case, understanding the dynamics of the
energetic electrons is of particular concern, as so-called runaway
electrons [10,11] generated during disruptions – events where the
plasma rapidly cools and strong electric fields are induced – have
✩ This paper and its associated computer program are available via the Computer
Physics Communication homepage on ScienceDirect (http://www.sciencedirect.
com/science/journal/00104655).
Corresponding author.
E-mail address: [email protected] (A. Stahl).
*
http://dx.doi.org/10.1016/j.cpc.2016.10.024
0010-4655/© 2016 Elsevier B.V. All rights reserved.
the potential to cause severe damage to a tokamak fusion reactor.
This problem is only expected to become more severe in future
devices since the runaway generation is exponentially sensitive to
the available plasma current [12,13].
In a spatially homogeneous plasma, the main processes influencing energetic-electron dynamics are: the presence of an
accelerating electric field; magnetization (causing directed motion); Coulomb collisions; dynamic changes in plasma parameters
such as the temperature; radiative losses (associated with synchrotron and fast-electron bremsstrahlung emission); and waveparticle interaction. The combined influence of these processes
has been shown to lead to phenomena such as bump-on-tail formation [14,15] and local isotropization [16,17] in the high-energy
tail of strongly anisotropic electron populations. Since analytic
treatment is possible only in special cases, the evolution of the
electron distribution function f must in general be studied using
kinetic simulations.
270
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
Many numerical tools solve the kinetic equation for f , taking
some subset of the processes mentioned above into account. In
collisional plasmas, the Fokker–Planck operator describing the
Coulomb collisions is the main source of complexity in the problem, which in general is described by a stiff integro-differential
diffusion equation, and numerical treatments can be broadly categorized based on the level of sophistication of the collision operator
employed. A number of continuum tools have been developed
that use linearized (around a Maxwellian) collision operators, especially in the case of fully non-relativistic problems, but also
in scenarios where the electrons are allowed to reach relativistic
energies [18–20]. In addition, several fully non-linear tools are
available for non-relativistic scenarios [21–28]; however, to our
knowledge, no tool treats the relativistic non-linear collision operator in its entirety. Both the integrated-tokamak-modeling tool
TASK and CQL3D (which is focused on heating and current drive in
tokamaks) successfully implement the first few Legendre modes
of the relativistic non-linear collision operator [29,30]. While this
approach guarantees the conservation of density, momentum and
energy, it cannot resolve fine structures in the momentum-space
distribution, making it unsuitable for accurate study of the fastelectron dynamics. In CQL3D, the implementation is general and
therefore in principle supports the use of any number of modes,
however in practice, the maximum number of modes cannot exceed 3 to 5 because of numerical problems [30].
In the magnetic-fusion community in particular, there is a
pressing need for a tool with the ability to handle situations where
relativistic particles comprise a significant part of the overall electron distribution, as these are the situations of greatest danger to
the integrity of the fusion device [13]. Such scenarios arise primarily when the electric field magnitude is (at least) a significant
fraction of the so-called Dreicer field [10], ED = ne3 ln Λ/4πϵ02 T ;
where n, T and −e are the electron number density, temperature
and charge, ln Λ is the Coulomb logarithm, and ϵ0 is the vacuum
permittivity. For such field strengths, the electric field overcomes
the maximum collisional friction force affecting the electrons.
However, also for E < ED , the distortion of the distribution can
become substantial, leading to the break-down of linearized codes.
In contrast, the so-called critical field Ec = Θ ED , with Θ = T /me c 2
the bulk temperature normalized to the electron rest mass, is
equivalent to the minimum collisional friction force experienced
by highly relativistic electrons. It therefore describes the weakest
field at which runaway-electron generation can occur [31], since
the accelerating force must overcome the friction (and the latter
decreases with increasing particle energy). For non-relativistic
bulk-electron temperatures, Ec is much less than ED and runawayelectron generation can in general be studied using a linearized
treatment since the electric field can fulfill both E > Ec and E ≪
ED simultaneously. A fully non-linear relativistic tool is however
needed in the scenarios of highest importance, where the runaway
population becomes comparable to the thermal population or the
electric field is of order ED .
In this paper, we describe such a tool: the new finite-difference
code NORSE (NOn-linear Relativistic Solver for Electrons), which
efficiently solves the kinetic equation in 2D momentum space.
NORSE includes a fully relativistic non-linear Fokker–Planck
operator for electron–electron collisions [32–34] and synchrotronradiation-reaction effects [14,35]. Time-dependent plasma parameters make the investigation of dynamic scenarios possible.
With a non-linear treatment, the distribution is not restricted to
being approximately Maxwellian, and strong electric fields (compared to ED ) can therefore be applied. However, if the distribution
departs strongly from a Maxwellian, concepts such as temperature
and the thermal collision time are not well defined. In many scenarios of practical interest, the distribution will nevertheless stay
close to collisional equilibrium, and we will make use of familiar
concepts where appropriate.
The kinetic equation and the operators for the various mechanisms mentioned above are discussed in Section 2. The numerical
implementation is then outlined in Section 3 and validated in
Section 4 through a comparison to previous work in several limits.
Finally, in Section 5 we use NORSE to investigate the properties
of strongly distorted electron distributions and the condition for
electron slide-away.
2. Kinetic equation
To study the momentum-space dynamics of energetic electrons, we will solve the kinetic equation neglecting any spatial
dependence. A point in the 3D momentum space is represented in
spherical coordinates by p = (p, ξ , ϕ ), with p = γ v/c the magnitude of the (normalized) momentum, ξ = p∥ /p the cosine of the
polar angle, and ϕ the azimuthal angle. Here v is the speed of the
particle, c is the speed of light, and γ is the relativistic mass factor.
The spherical symmetry of our problem is broken by the presence
of the electric field E, and we therefore let the electric field define
the parallel direction. If the plasma is magnetized, only the electricfield component parallel to B contributes to the acceleration (i.e.
B∥E), in which case ξ is the cosine of the pitch–angle and ϕ is
the gyro angle. We will assume the electron distribution function
f = f (t ; p, ξ ) to be independent of ϕ , reducing the problem to a
two-dimensional one.
The kinetic equation describing the evolution of f can be
written as
∂f
eE
∂f
∂
−
·
+
· (Fs f ) = C {f } + S ,
(1)
∂t
me c ∂ p
∂p
where Fs is the synchrotron-radiation-reaction force (in the presence of a magnetic field), C {f } is the Fokker–Planck collision operator describing microscopic Coulomb interactions between the
plasma particles, and S denotes sources and sinks (of
∫ for instance
heat or particles). The distribution f satisfies n = d3 p f , with n
the number density of electrons.
The parallel component of the momentum-space gradient, appearing in the term describing the Hamiltonian motion of the
electrons due to the electric field, becomes
E
E
·
(
)
∂f
∂f
1 − ξ 2 ∂f
= ξ
+
.
∂p
∂p
p
∂ξ
(2)
In what follows, we will detail the synchrotron-radiation-reaction
and collision terms of Eq. (1), as well as the various source terms.
2.1. Synchrotron-radiation reaction
The reaction force experienced by electrons emitting synchrotron radiation can be derived from the Lorentz–AbrahamDirac force. In a homogeneous plasma, it can be written as (see for
instance [14] and references therein)
( 3
)
(
)
1 ∂
γ p (1 − ξ 2 )
∂ ξ (1 − ξ 2 )
∂
· (FS f ) = − 2
f +
f
∂p
p ∂p
τr
∂ξ
γ τr
[
∂
f
∂
f
1 1 − ξ2
γ 2p
=−
−ξ
τ
γ
∂p
∂ξ
) ]
(r
2
2
+ 4p +
f ,
1 − ξ2
(3)
where
τr =
6π ϵ0 (me c)3
(4)
e4 B2
is the radiation time-scale. Here B is the magnetic field strength.
The total synchrotron power emitted by a relativistic particle
is proportional to p2⊥ = p2 (1 − ξ 2 ), and the back-reaction experienced by the electrons therefore increases with perpendicular
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
271
momentum. The efficacy of the synchrotron-radiation reaction is
thus closely linked to collisional pitch–angle scattering, which can
redistribute parallel momentum gained from the accelerating field.
through
2.2. Electron-ion collision operator
The five potentials are analogous to the two Rosenbluth potentials
g and h in the non-relativistic case [37], and reduce to these in
the appropriate limit. The notation (which differs from that in
Ref. [34], see Appendix B) has been chosen to highlight the existence of two ‘‘branches’’ of potentials, distinguished by the application of different La operators. Crucially, in the non-relativistic
limit (as La reduces to the Laplace operator), Υ0 = Π0 → h and
Υ1 = Π1 → g.
A sketch of the derivation of the explicit expressions obtained
in our coordinate system is given in Appendix A; here we list only
the final result, with the terms grouped according to the derivative
of f . The collision operator can be written as
In a fully ionized plasma, the collision operator C contains
contributions from collisions with electrons (Cee ) and ions (Cei ):
C {f } = Cee {f } + Cei {f }.
(5)
The electron–electron collision operator is the main source of
complexity in our problem, and will be discussed in Section 2.3.
In contrast, we will assume a stationary, Maxwellian ion population. This, together with the mass difference between the species
involved in the collision, significantly simplifies the operator for
electron-ion collisions (unless the ratio between ion and electron
temperatures is comparable to their mass ratio). In the ion restframe, the operator is [36]:
Cei {f } ≃ Zeff
νγ
p3
L{f } = Zeff
νγ
[
p3
]
∂f
(1 − ξ )
,
2 ∂ξ
∂ξ
1 ∂
2
2
where Zeff = n
j nj Zj is the effective charge (with the sum taken
over all ion species j), L is the Lorentz scattering operator, and
ν=
L2 Υ 1 = Υ 0 ,
L1 Π0 = f ,
L1 Π1 = Π0 .
Cee {f }
α
(6)
∑
−1
ne4 ln Λ
L0 Υ0 = f ,
C
(7)
4π ϵ02 m2e c 3
is the collision frequency for relativistic electrons. The operator Cei
describes pitch–angle scattering, but no change to the magnitude
of the electron momentum. This is because the ions are assumed
to be much heavier than the electrons (i.e. mi ≫ γ me ), so that the
energy lost by the electrons through collisions can be neglected.
C (p)
2.3. Electron–electron collision operator
To describe electron–electron collisions, we will use the fully
relativistic non-linear collision operator of Beliaev and Budker [32],
in the form developed by Braams and Karney [33,34]. The operator
is valid for collisions between arbitrary species of arbitrary energy
(i.e. the bulk population is not required to be non-relativistic). For
electron–electron collisions, it takes the form [34]
Cee {f } = α
(
)
∂f
∂
· D·
− Ff
∂p
∂p
C (ξ
2)
(8)
where α = 4πν/n, D is the diffusion tensor and F is the friction
vector. These are given by
D = γ −1 [LΥ− − (I + pp)Υ+ ] ,
C (ξ )
(9)
F = γ −1 KΠ ,
(10)
where I is the unit tensor and L and K are defined by
(
)
∂ 2 Υ−
∂ Υ−
,
· (I + pp) + (I + pp) p ·
∂ p∂ p
∂p
∂Π
KΠ = (I + pp) ·
.
∂p
LΥ− = (I + pp) ·
(11)
C (pξ )
(12)
Here, Υ− , Υ+ and Π are linear combinations of potential functions,
given by
Υ− = 4Υ2 − Υ1 ,
Υ+ = 4Υ2 + Υ1 ,
Π = 2Π1 − Π0 ,
(13)
where we denote the five potentials introduced by Braams and
Karney as Υ0 , Υ1 , Υ2 , Π0 and Π1 . These are defined using the
differential operator
La Ψ = (I + pp) :
(
)
∂ 2Ψ
∂Ψ
+ 3p ·
+ 1 − a2 Ψ ,
∂ p∂ p
∂p
(14)
2
∂ 2f
∂f
2 ∂ f
+ C (p)
+ C (ξ ) 2
2
∂p
∂p
∂ξ
2
(ξ ) ∂ f
(pξ ) ∂ f
+C
+C
+ C (f ) f ,
∂ξ
∂ p∂ξ
C (f )
(15)
2)
= C (p
with pre-factors C
(p2 )
L2 Υ2 = Υ1 ,
(i)
(16)
given by
γ 3 ∂ Υ−
p ∂p
γ (1 − ξ 2 ) ∂ 2 Υ−
γ ξ ∂ Υ−
−
+2 2
,
p2
∂ξ 2
p ∂ξ
∂
Υ2
1
(2 + 3p2 )(8Υ2 − Υ0 ) − 16γ
=
γp
∂p
(
)
∂ Υ1
∂ Υ0
γ 3 ∂ 2 Υ−
1 ∂ Υ−
+ 6γ
−γ
−2
+
∂p
∂p
p
∂ p2
p ∂p
(
)(
)
2
1
1
∂ Υ−
2 ∂ Υ−
+
2+ 2
2ξ
− (1 − ξ )
2
γp
p
∂ξ
∂ξ
∂Π
−γ
,
∂p
(
1 − ξ 2 γ 2 ∂ Υ−
=
2
γp
p ∂p
[
]
)
∂ 2 Υ−
∂ Υ−
1
+ 2 (1 − ξ 2 )
−
ξ
−
Υ
,
+
p
∂ξ 2
∂ξ
2
2
2
2
ξ (1 − ξ ) ∂ Υ−
γ (1 − ξ ) ∂ Υ−
γ ξ ∂ Υ−
=−
−2
−2 3
γ p4
∂ξ 2
p3
∂ p∂ξ
p ∂p
(
)
2
1 − ξ 2 ∂ Υ−
+
+3
γ p4
γ p2
∂ξ
(
)
∂ Υ2
∂ Υ1
∂ Υ0
∂Π
1 − ξ2
4
−3
+
+
−
2
γp
∂ξ
∂ξ
∂ξ
∂ξ
ξ
+ 2 2 Υ+ ,
γp
[
]
γ (1 − ξ 2 ) ∂ 2 Υ−
∂ Υ−
=2
,
p
−
p3
∂ p∂ξ
∂ξ
2
) ∂Π
∂ Π
1 (
= −γ
−
2 + 3p2
∂ p2
γp
∂p
1 − ξ 2 ∂ 2Π
ξ ∂Π
−
+2 2
.
γ p2 ∂ξ 2
γ p ∂ξ
= γ (8Υ2 − Υ0 ) − 2
(17)
(18)
(19)
(20)
(21)
(22)
2.4. Heat and particle sources
A strong electric field is a source of energy that quickly heats
the distribution function. In contrast to a linearized treatment
(where this heat must be removed to ensure the validity of the
272
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
linearization), this energy source is automatically accounted for
in the non-linear solution. Sometimes, it is however of interest to
remove the excess heat from the bulk as it is applied. In reality,
the bulk temperature is not always increasing during fast-particle
generation, for instance because of energy loss due to radiation
emission or heat conduction. A heat sink also serves as a way to
vary the temperature of the thermal population, which makes it
possible to model dynamic scenarios where the plasma parameters
change on a time scale similar to that of the acceleration dynamics.
To be able to model density changes, a particle source must also be
included.
An advantageous way to formulate a heat sink is to write it in
divergence form
)
∂ (
· kh Sh f = kh
∂p
(
2
p
∂ Sh (p)
∂
+ Sh (p)
f,
∂p
∂p
dW
dt
= me c 2
Sh (p) +
(
d3 p (γ − 1) −
Ω
+
eE
me c
∂f
∂p
·
)
∂
∂
· (Fs f ) − C {f } + kh
· (Sh f ) ,
∂p
∂p
(24)
(
)
γ (p)
) exp −
,
Θ
4π Θ K2 1/Θ
n
(25)
(
where Kν (x) is the modified Bessel function of the second kind (and
order ν ), has the energy moment
W (Θ ) =
me c 2 n
Θ K2 (1/Θ )
pmax,Ω
∫
( γ)
.
Θ
dp p2 (γ − 1) exp −
0
(26)
The magnitude kh can be determined from the requirement that
the energy supplied by the heat sink should equal W (Θ2 ) − W (Θ1 )
for two temperatures Θ1 and Θ2 at subsequent time steps. Here
pmax,Ω denotes the upper boundary of Ω in p; if pmax,Ω → ∞, the
above integral can be evaluated analytically, yielding
W (Θ ) = m e c 2 n
(
K3 (1/Θ )
K2 (1/Θ )
−1−Θ
)
≡ me c 2 nW (Θ ).
(27)
]
γ −1
+ ap ( Θ ) f M ;
Θ
(28)
a linear combination of the energy and density moments of a
Maxwellian (with an overall scaling factor kp analogous to kh ). The
quantity ap can be determined from the constraint that the energy
moment of Sp should vanish (so that the source supplies particles,
but no heat), giving
ap ( Θ ) = −
=
(23)
from which kh can be determined in each time step by demanding
that dW /dt = 0 (for an ideal heat sink). (Note that this approach does not automatically enforce energy conservation after
discretization; only physical sources of heat are taken into account.
If desirable, numerical heating caused by the discretization can be
eliminated by instead requiring the numerically calculated energy
moment of f to be constant.)
The same heat source can be used to induce changes to the
bulk temperature, but in this case the magnitude kh is calculated
differently. We note that the relativistic equilibrium Maxwell–
Jüttner distribution
fM (p) =
[
Sp = kp
)
since it will then automatically conserve particles (before discretization). Here Sh (p) is an isotropic function of momentum, with
Sh its p-component. kh is the magnitude, to be determined. In
practice, the exact momentum-space shape of the sink depends on
the processes responsible for the heat loss. A detailed investigation
of this is left for future work; here we let Sh have the shape of a
Maxwellian for simplicity.
Apart from the electric-field term, synchrotron-radiation reaction also changes the total heat content of the distribution by
removing energy, primarily at large particle momenta. However,
the momentum-space region Ω of interest need not necessarily
encompass the entire computational domain. For instance, it is
sometimes desirable to maintain a fixed energy content in the
thermal population, while simultaneously allowing the energetic
particles to gain energy. Physically, this corresponds to heat sinks
that only affect slow particles. In such cases, collisions may also
transfer energy into or out of Ω. The total energy change in Ω can
thus be written as
∫
Changes to the density can be introduced using a particle source
of the form
∫∞
1
Θ
2
Θ
−
∫0 ∞
0
dp p2 (γ − 1)2 exp(−γ /Θ )
dp p2 (γ − 1) exp(−γ /Θ )
3(1 + Θ )
W (Θ )
− 3.
(29)
As Θ → 0, the non-relativistic limit ap = −5/2 is recovered,
whereas in the ultra-relativistic case (Θ ≫ 1), ap → −4. The
density moment np of the source is given by
np
n
[
= kp
Θ
[
= kp
W (Θ )
]
+ a(p)
W (Θ ) + 2
Θ
−3
W (Θ ) + 1 + Θ
W (Θ )
]
,
(30)
from which the magnitude kp that gives a desired density change
can be determined. The bracket takes the asymptotic value −1 at
both Θ → 0 and Θ ≫ 1, but reaches a minimum of −1.18 for
intermediate temperatures.
3. Numerical method
3.1. Discretization
We choose to represent the distribution f on a two-dimensional
finite-difference grid in p and ξ , and use a 5-point stencil to
discretize the momentum-space derivatives. Moments of the distribution and other integrals are calculated using a composite
Simpson’s rule. The grid points can be chosen non-uniformly in
both p and ξ , making it possible to efficiently resolve both a
Maxwellian bulk (assuming there is one) and a high-energy tail.
Specifically, the p grid should preferably be densely spaced for
small p to resolve the bulk, but since the tail generally varies over
larger momentum scales, coarser spacing can be used at larger
momenta to reduce the computational expense. Similarly, the ξ
grid should be densely spaced close to ξ = 1 (the parallel direction)
to resolve the tail drawn out by the electric field. Alternatively,
in scenarios without a preferred direction of acceleration, a grid
which gives a uniform spacing in the polar angle (arccos ξ ) is often
appropriate. Due to the polar nature of the coordinate system, the
point at p = 0 is special; the value of the distribution at p = 0
should be independent of ξ . The total number of grid points is thus
Nξ × (Np − 1) + 1, with Np and Nξ the number of grid points in
the respective coordinate, and a single (rather than Nξ ) grid point
appropriately describes the system at p = 0.
For the calculation of the potentials Υi and Πi (here collectively
denoted by Ψ ), it is advantageous to decompose the ξ coordinate
in Legendre modes (rather than use a finite difference grid), since
these are eigenfunctions of the collision operator. The distribution
and potentials are then written as
f (p, ξ ) =
Nl
∑
l=0
fl (p)Pl (ξ ),
Ψ (p, ξ ) =
Nl
∑
Ψl (p)Pl (ξ ),
(31)
l=0
where Pl is the lth Legendre polynomial. The potentials are integral
moments of the distribution function, and their calculation is a
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
smoothing operation. Therefore, a small number Nl of Legendre
modes typically suffices to accurately describe the potentials, unless the bulk of the distribution deviates significantly from the
origin of the coordinate system. Thus, it is usually reasonable to
choose Nl to be much smaller than Nξ , the number of points in the
ξ grid.
The mapping between the 2D-finite-difference-grid and finitedifference-Legendre-mode representations can be formulated as
a single matrix operation, where a mapping matrix ML can be
constructed to represent the summation in Eq. (31). In general,
ML is not square, but the inverse mapping can be performed
by taking the Moore–Penrose pseudo-inverse of ML to find the
inverse in a least-squares sense. (This only needs to be done
once in each NORSE run.) The solution is exact in the sense that
norm[ML fl (p) − f (p, ξ )] is of order the round-off error, and the
mapping between the two representations can thus be performed
to machine precision at very small computational cost.
The parallel axis is the symmetry axis of the problem. Therefore,
we require that the derivative of the distribution with respect to
p⊥ at a fixed p∥ must vanish as p⊥ → 0. This condition must be
imposed as a boundary condition at p∥ = 0, but is automatically
satisfied for all non-vanishing p∥ . At p = pmax , we impose the
Dirichlet condition f (pmax ) = 0 for all ξ .
3.2. Calculation of potentials
The Legendre modes of the potentials Ψ can be calculated from
the distribution using Eq. (15), which becomes
L0,l Υ0,l = fl ,
L2,l Υ1,l = Υ0,l ,
L1,l Π0,l = fl ,
L1,l Π1,l = Π0,l ,
L2,l Υ2,l = Υ1,l ,
(32)
where
∂ 2Ψ
La,l Ψ = γ 2
+
∂ p2
(
)
(
∂Ψ
l(l + 1)
Ψ (33)
+ 3p
+ 1 − a2 −
2
p
∂p
p
2
)
is obtained by decomposing the differential operator La (described
in our coordinate system by Eq. (A.8)) into Legendre modes. Inverting Eqs. (32) results in operators which can determine the
potentials from an arbitrary f , and a block-diagonal sparse matrix
for each potential Ψ can be constructed to efficiently calculate Ψl
from fl for all l using a single matrix multiplication, in accordance
with the discussion in Section 3.4.
The above calculation requires that boundary conditions for
the potentials be specified. In general, for a function φ (p, ξ ) to
be continuous at p = 0, its Legendre modes φl (p) must satisfy
∂φl (0)/∂ p = 0 for l = 0 and φl (0) = 0 for l > 0. Boundary
conditions at p = pmax can be determined from Eq. (31) in Ref. [34],
which gives explicit expressions for the potentials Ψl in terms
of weighted integrals over fl . The calculation of these boundary
conditions is discussed in Appendix B.
3.3. Time advance
To advance the system in time, we employ a linearly implicit
time-advancement scheme based on the first-order backwardEuler method. The scheme avoids the restriction on the time
step imposed on explicit methods by the CFL condition, and is
straight-forward to implement, as it only requires building and
inverting a single matrix in each time step. Compared to fully
implicit methods, however; the time step has to be kept relatively
short, and the overall computation time can still be considerable
when simulating a long time span. As long as the time step is
short enough, good accuracy is achieved, and this simple scheme
is sufficient for our purposes.
The method is formulated as follows. The entire kinetic equation, excluding the time derivative, can in general be written as
273
an operator O{Ψ {f }, f }, where Ψ represents the five potentials
Υi and Πi , which depend on the distribution f . In a fully implicit
time-advancement scheme, this operator should be evaluated at
the next time step (k + 1): O{Ψ {f k+1 }, f k+1 }. If the potentials are instead evaluated based on the distribution at the current time step,
f k , O can be written as a regular matrix operation O{Ψ {f k }, f k+1 } =
k
k
Mmn
f k+1 , where the matrix Mmn
= M k (pm , ξn ) describes a set of
linear equations. This makes the time-advancement scheme linearly implicit, and M k can be explicitly evaluated in each time step
and the system solved using standard matrix-inversion techniques.
The Backward-Euler method for our problem can then be written
as
f k+1 = f k + ∆tM k f k+1 ,
(34)
where ∆t is the time step.
3.4. Performance
NORSE is written in Matlab, using an object-oriented structure.
To make efficient use of the Matlab language, care has been taken
to formulate the problem in terms of matrix multiplications and
avoid loops where they are detrimental to performance. To this
end, many parts of the operators of the kinetic equation are precalculated to speed up the matrix building in each time step. As an
example, the first term of the electron–electron collision operator
Eq. (16) at time step k (together with Eq. (17)) can be written as
2)
C (p
(
)
∂ 2 f k+1
= C0 Υ0k + C2 Υ2k + C− Υ−k D2pp f k+1 ,
∂ p2
(35)
where the various operators, defined as
C0 = −γ ,
C− = −2
D2pp =
γ3
p
∂2
,
∂ p2
C2 = 8γ ,
γ (1 − ξ 2 )
Dp −
Dp =
p2
∂
,
∂p
D2ξ ξ + 2
γξ
D2ξ ξ =
p2
(36)
Dξ
∂2
,
∂ξ 2
Dξ =
∂
,
∂ξ
(37)
are all independent of f , and can thus be pre-calculated. Constructing this part of the linear system in each time step is thus reduced
2
to determining the potentials Υik from f k and constructing C (p ) D2pp
in accordance with Eq. (35), using just a few matrix operations.
The above algorithm is efficient, making the matrix inversion
associated with the solution of the resulting linear system the most
costly part of each time step. The overall computational cost can be
reduced by approximately a factor of 2 by employing an iterative
scheme using the generalized minimal residual method (gmres
[38]), which is available in Matlab as a standard subroutine. By
periodically (every nLU time steps) solving the system exactly using
LU-factorization, and supplying the L and U factors as preconditioners for the next nLU − 1 steps, gmres converges in just a few
iterations if nLU is sufficiently small.
In certain scenarios – such as where an initial transient requires
high temporal resolution, but the subsequent relaxation happens
on a significantly longer time scale – adaptive time-step schemes
can be very effective in reducing the computational expense. Such
a scenario is for example considered in Section 4.1. Here, we use
a simple adaptive-time-step scheme based on information about
the number of iterations needed for convergence of the gmres
algorithm. If few gmres iterations are needed for convergence
(ngmres < nopt , where nopt is some desired optimal number), the
change in the distribution in each time step is small, indicating that
the step length can be increased. Conversely, the step length should
be reduced if ngmres > nopt .
Employing the techniques discussed above makes the implementation efficient, and moderately sized test cases usually run on
a standard laptop in less than a minute.
274
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
Fig. 1. Collisional relaxation from a starting distribution consisting of two shifted Maxwellians. Panels (a)–(d) show the 2D distribution at various times, and panel (g)
shows corresponding cuts along the positive parallel axis. Panel (e) shows the conservation of density and energy and panel (f) the time step used, as functions of time. The
numerical parameters Np = 250, Nξ = 65, Nl = 25, pmax = 10, and the initial time step dτ0 = 0.001 were used. A uniform grid was used in p, whereas a non-uniform grid
giving uniform spacing in the polar angle (arccos ξ ) was used for the ξ coordinate.
4. Tests and benchmarks
The kinetic equation solved by NORSE is valid for strongly
non-Maxwellian electron distributions, as well as relativistic temperatures and particle energies. In this section, we validate the
implementation by comparing to the two limits of arbitrary temperature but weakly distorted distribution (Section 4.2), and nonrelativistic but fully non-linear distribution (Section 4.3). However,
let us first look at a proof-of-principle scenario, demonstrating
both the non-linearity as well as the high-temperature validity of
NORSE . In this section, we will repeatedly make use of the normalized time τ = ν t (i.e. the time in units of relativistic-electron
collision times) and the normalized distribution F = f /fM (p = 0)
(so that if the initial distribution is a Maxwellian, F initially takes
the value unity at p = 0). We will also use the normalized electricfield magnitude Ê = eE /me c ν = E /Ec . Throughout the rest of the
paper, we will always apply fields with an implicit minus sign, so
that electrons will be accelerated towards positive p∥ .
4.1. Proof-of-principle non-linear scenario: collisional relaxation of a
two-maxwellian initial state
In this section we demonstrate the validity of the NORSE implementation by considering a basic non-linear test case: the collisional relaxation of two initially shifted Maxwellians. A shifted
Maxwell–Jüttner distribution (e.g. the equilibrium distribution
with temperature Θb in a frame boosted by pb in the parallel
direction, as seen from the stationary frame) takes the form
)
γb γ − pb p∥
(
) exp −
fM,b (Θb , pb ) =
,
(38)
Θb
4π Θb K2 1/Θb
√
with γb =
1 + p2b and p∥ = pξ . We consider two initial
Maxwellians, each with a temperature of 10 keV (Θb = 0.0196),
and each shifted the equivalent of three thermal speeds (pb =
0.59) along the symmetry axis (in opposite directions). The initial
n
(
state is depicted in Fig. 1(a), with panels (b)–(d) showing the
subsequent evolution of the distribution function. Panel (g) shows
a cut of the distribution along the positive parallel axis at the same
time steps. The parameter values E = 0 and B = 0 were used,
and to isolate the behavior of the non-linear electron self-collision
operator, a pure electron plasma was assumed (Zeff = 0). The
number density of each Maxwellian in its rest frame was set to
n = 1019 m−3 , resulting in a total initial number density of ntot =
2γb n = 2.326 · 1019 m−3 . The expected final-state Maxwellian
(cyan, thin dashed line in panel (g)) has a temperature of 61.3 keV,
which can be calculated by equating Eq. (27) (with n → ntot ) with
the combined energy content in the two shifted Maxwellians:
Wtot = 2Wb ,
Wb
[
]
= γb2 W + me c 2 n γb (γb − 1) + (γb2 − 1)Θb
(39)
(with W given by (27)), and solving for Θ . The final equilibrium
state shows excellent agreement with the theoretical prediction.
The relative error (compared to the initial value) in the density
and energy contents of the NORSE solution are shown in Fig. 1(e) as
functions of time. For the numerical parameters used, the density
is conserved to within 0.05%, whereas the relative error in energy
saturates at the 0.5% level. Fig. 1(f) shows the time step used by
the adaptive-time-step scheme, normalized to the initial time step.
In this particular case, the scheme is very effective since the time
evolution involves an initial transient followed by a comparatively
slow asymptotic relaxation. The final time step was approximately
104 times longer than the initial time step, and a total of 312 time
steps were used (as opposed to ∼ 4 · 105 had the initial time step
been used throughout the entire calculation).
4.2. Weak-electric-field limit: conductivity for relativistic temperatures
In Ref. [34], Braams and Karney use the relativistic electron–
electron collision operator to calculate the plasma conductivity for
a wide range of temperatures. The operator is linearized around
a stationary Maxwellian, and the zeroth and first Legendre modes
are calculated numerically as an initial-value problem. The results
are compiled in their Table 1, which contains normalized conductivities for Θ ∈ [0, 100] (recall that Θ = 1 corresponds to
T = me c 2 ≃ 511 keV) and Zeff ∈ [0, ∞]. The unit used is
√
σ̄ =
eme ln ΛZeff j
4πϵ02 T 3/2
E
,
(40)
where j is the current density.
To demonstrate that our implementation reproduces the above
results, we similarly calculate the conductivity of a quasi-steadystate distribution found by evolving the system from a Maxwellian
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
275
eventual loss of fast-electron confinement constitute a serious
threat to the plasma-facing components of fusion reactors. The
energy and current carried by the runaway electrons, and thus
the potential for damage, increase the larger the distortion of the
electron distribution. Unlike existing tools such as LUKE [19] and
CODE [20,41], NORSE can be used to study the cases of highest
runaway-electron growth rate.
5.1. Runaway region of momentum space
Fig. 2. Normalized conductivity in NORSE (lines) for various temperatures and
plasma compositions. Data points from Table 1 of Ref. [34] are also shown (squares).
The electric field corresponded to E = 10−3 ED , and n = 5·1019 m−3 and B = 0 were
used.
initial state using a constant electric field corresponding to E =
10−3 ED . However, we make no simplification to the collision operator and retain adequate resolution in ξ to accurately resolve
the distribution function in 2D momentum space. Fig. 2 overlays
NORSE results with the data in Ref. [34] for Zeff = 1, 2, 5 and 10,
and all tabulated temperatures. Excellent agreement is seen for all
parameters. In the figure, the data points at Θ = 0 in Ref. [34]
are compared to NORSE runs with Θ = 10−5 , however all temperatures Θ < 10−3 give good agreement as the obtained values of σ̄
are essentially independent of the temperature in this range.
4.3. Non-relativistic limit: highly anisotropic distributions
Several codes exist that solve the non-relativistic kinetic equation using non-linear collision operators. To validate the non-linear
aspect of NORSE, we will compare to conductivities reported by
Weng et al. in Ref. [39]. In their Fig. 3, conductivities as functions of
time are presented for electric fields as strong as the Dreicer field
ED , leading to highly distorted distributions. Results are shown for
E /ED = 0.01, 0.1 and 1, with Zeff = 1 and B = 0.
Fig. 3(a) reproduces the results in Fig. 3 of Ref. [39] using NORSE.
The units used are those of the original figure: conductivities are
given as j̄/Ê, with j̄ = jZeff /nec√
Θ 3/2 a normalized current density,
and the time unit used is Ê τ / Θ . The parameters Θ = 1 · 10−4
(corresponding to a temperature of T = 51eV) and n = 5·1019 m−3
were used in NORSE. Data points extracted from the figure in Ref.
[39] are included for comparison. The agreement is very good in
general, demonstrating that NORSE behaves as expected also for
highly non-linear distributions. The error is somewhat larger (and
systematic) for the weakest electric field, however in this case,
numerical heating in the results of Ref. [39] cannot be ruled out
[40]. The final distributions for the three field strengths are shown
in 3(b). As can be seen, the distributions deviate strongly from the
initial Maxwellian for the two higher field strengths, and even at
the weakest field, a substantial tail of runaway electrons is produced. For the strongest field, a small ‘‘bump’’ in the distribution
is seen at p∥ = 0, which indicates that electron-ion collisions
are strong enough to ‘‘capture’’ a sub-population of the electrons,
despite the strong accelerating electric field.
5. Application: Runaway electrons in fusion plasmas
As an example of how NORSE can be used to provide new physical insight, we consider the case of runaway-electron generation
in magnetically confined fusion plasmas. Under the influence of
strong electric fields, electrons quickly accelerate to relativistic
speeds since the friction force they experience decreases with
increasing velocity. The localized heat loads associated with the
What constitutes a runaway particle can be defined in several
ways. The definitions usually employed in theoretical works, as
well as many numerical tools, assume the distribution to be close to
Maxwellian and are therefore not directly applicable in our context
[10,42]. In addition, it has been pointed out that synchrotron radiation reaction may have a significant impact on the runaway region
[43–46], however the commonly used definitions only account
for collisional friction. We will define a runaway region based on
particle trajectories in momentum space [42], neglecting the effect
of diffusion but allowing for arbitrary electron distributions as well
as synchrotron radiation reaction.
For an arbitrary electron distribution, the lower boundary of the
runaway region (the separatrix) can be obtained by considering the
forces that affect a test particle:
∂Π
γ p(1 − ξ 2 )
+
,
(41)
∂p
τr
2
2
2
eE 1 − ξ
1 − ξ ∂Π
ξ (1 − ξ )
dξ
ξ
ξ
ξ
= FE − FC − FS =
+α
−
,
dt
me c
p
γ p2 ∂ξ
γ τr
dp
dt
= FEp − FCp − FSp =
eE
me c
ξ + αγ
(42)
where the expressions for the force associated with the electric
field, FEi , the collisional electron–electron friction FCi and the synchrotron radiation-reaction force FSi are taken from Eqs. (1), (A.3)
and (3), respectively. Asymptotically, particles on the separatrix
neither end up in the bulk population nor reach arbitrarily high
energies, but instead settle at the point pc of parallel force balance
at ξ = 1 (in the absence of diffusion). pc can be determined
from dp/dt = 0 at ξ = 1, since the separatrix becomes purely
perpendicular to the parallel axis as ξ → 1. The separatrix is then
traced out by numerically integrating the above equations from
ξ = 1 to ξ = −1. In the limit of non-relativistic temperature
(Θ ≪ 1), small departure from a Maxwellian, and B = 0, the result
agrees with the standard expression [42]. To ensure consistency
with the distribution, the separatrix in NORSE is calculated in each
time step.
5.2. Distortion-induced transition to electron slide-away
For electric fields stronger than approximately Esa = 0.215ED ,
all electrons in a Maxwellian distribution experience net acceleration, since the field overcomes the maximum of the collisional
friction force. This is known as electron slide-away [10,47]. However, in a non-linear treatment, the condition for slide-away can in
principle be modified since the collisional friction depends on the
shape of the electron distribution. The distortion of the distribution
associated with a moderately strong electric field turns out to
have a large effect on the effective Dreicer field at which the
transition to the slide-away regime occurs. This is illustrated in
Fig. 4, which shows the distribution at several time steps, as well
as the separatrix and the force balance (neglecting diffusion) as a
function of p at ξ = 1. In the figure, the distribution is evolved
under a constant electric field which initially corresponds to 5% of
the Dreicer field (E = 0.05ED,0 ≈ 0.23Esa,0 , with ED,0 and Esa,0
the Dreicer and slide-away fields at the initial temperature). The
distribution quickly becomes distorted, and soon after t = 0.15τ
276
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
Fig. 3. (a) Normalized conductivity in NORSE (lines) as a function of time for various E-field strengths. Data points extracted from Fig. 3 of Ref. [39] are also shown (squares).
(b) Cuts in the parallel direction through the final distributions in (a). The numerical parameters were Np = 300, Nξ = 55, Nl = 15, pmax = 0.3, using 2000, 400 and 300
time steps for E /ED = 0.01, 0.1 and 1, respectively.
Fig. 4. (a)–(d) Contour plots and (e) cuts along the parallel axis of the distribution at different times during a NORSE run. In (a)–(c), the white dashed lines are the separatrices
defining the lower boundary of the runaway region. (f) Sum of forces (neglecting diffusion) on the parallel axis. The physical parameters were Θ = 0.01 (T = 5.11 keV),
n = 5 · 1019 m−3 , Zeff = 1, Ê = 5 (corresponding to E /ED,0 = 0.05 and E = 0.22 V/m), B = 0, and Np = 100, Nξ = 35, Nl = 5, pmax = 3.54, and dτ = 2.8 · 10−4 were used.
the slide-away regime is reached. This can be seen in Fig. 4(f),
where the sum of forces (Eq. (41)) becomes positive everywhere
on the parallel axis, indicating that the electric field at that time
corresponds to the instantaneous slide-away field, E = Esa (t). No
separatrix therefore exists for later times (see Fig. 4(d)).
An important effect of the electric field is to quickly heat the
bulk of the distribution, and this turns out to be the main cause of
the induced transition to slide-away. Since (neglecting the weak
dependence on ln Λ) the Dreicer field ED ∼ 1/T , an increase in
temperature lowers the effective Dreicer field and thus the threshold for electron slide-away. An approximate effective temperature
Teff can be estimated from the energy moment W of the NORSE
distribution by solving Eq. (27) for Θ . The effective Dreicer field
can then easily be evaluated.
Fig. 5 highlights the importance of the heating effect by showing
the time to a transition to slide-away under constant electric fields
of various strengths (starting from an equilibrium distribution).
Lines denote when the effective temperature becomes such that
E > Esa (Teff ), whereas squares denote the actual transition in
NORSE, calculated from the force balance. The agreement between
these two values is very good in the entire range of electric-field
values, demonstrating that the bulk heating is the dominant effect
in the modification of the slide-away threshold. Only at fields
very close to Esa,0 does the values obtained using the effective
temperature noticeably overestimate the time to transition, indicating that here, other effects start to become important as well.
The figure also shows that the process leading to a transition to
slide-away is quick, also for relatively weak fields. At E /Esa,0 =
0.3 (E /ED,0 ≈ 0.065), the transition happens around 30 thermal
collision times for Zeff = 1, and at E /Esa,0 = 0.1 (E /ED,0 ≈ 0.022),
the corresponding figure is 500.
In practice, various processes can lead to heat losses, as previously noted. This can partially or entirely offset the heating caused
by the electric field, and in many situations the modification to
the slide-away threshold may not be as dramatic as demonstrated
here. In addition, a feedback mechanism commonly exists between the accelerating electric field and the distribution (through
changes in the plasma current). In such scenarios, a reduction
in the electric field may be induced due to the changes in the
distribution before these have become too extensive, thus limiting
the distortion and potentially avoiding a transition to slide-away
altogether.
6. Conclusions
The study of energetic-electron populations in plasmas has long
been of interest, but when considering relativistic particles in kinetic simulations, the work has so far been restricted to linearized
treatments of the Fokker–Planck collision operator. In this paper,
we remove that limitation by introducing a new efficient computational tool (NORSE) which includes the fully non-linear relativistic
collision operator in the differential form developed by Braams
and Karney, as well as electric-field acceleration and synchrotronradiation reaction. A 2D non-uniform finite-difference grid is used
to represent momentum space, however when evaluating the five
relativistic potentials (analogous to the two Rosenbluth potentials
in the non-relativistic case), a mixed finite-difference–Legendremode representation is used since the potentials are given by simple 1D integrals in a Legendre-mode decomposition. The system is
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
[
∂f
+ γ Dpξ (Υ− )
∂p
(
) ]
1 − ξ2
∂f
∂ Υ−
g
D
(
Υ
)
+
p
+
−
Υ
eξ
ξ
ξ
ξ
ξ
−
+
γ p2
∂p
∂ξ
277
(A.2)
and
Ff = γ
1 − ξ 2 ∂Π
∂Π
f ep +
f eξ ,
∂p
γ p2 ∂ξ
(A.3)
where
]
∂ 2 Υ−
∂ Υ−
∂ Υ−
−ξ
+p
,
2
∂ξ
∂ξ
∂p
[
]
1
∂ Υ−
∂ Υ−
Dϕϕ (Υ− ) = 4
p
−ξ
,
p (1 − ξ 2 )
∂p
∂ξ
[
]
2
2
1−ξ
∂ Υ−
2 ∂ Υ−
Dpξ (Υ− ) = Dξ p (Υ− ) =
p
−
p
p4
∂ p∂ξ
∂ξ
Dξ ξ (Υ− ) =
Fig. 5. Time to transition to slide-away as a function of electric-field strength for
various values of Zeff . Times calculated from the effective temperature of NORSE
distributions are indicated by lines, while the actual times obtained in NORSE (based
on force balance) are indicated by squares. Here, a subscript 0 indicates an initial
value, and τth,0 = τ /(2Θ0 )3/2 denotes the initial thermal-electron collision time.
The physical parameters were Θ0 = 1 · 10−4 , n = 1019 m−3 , B = 0.
evolved using a linearly implicit time-advancement scheme, and a
simple method for adapting the time step during runtime has been
implemented. NORSE has been successfully benchmarked in both
the relativistic-weak-field and non-relativistic–non-linear limits.
As an application, we have used NORSE to investigate scenarios relevant to the study of runaway electrons in magneticconfinement-fusion plasmas. We find that the quick heating of the
bulk associated with the application of medium to high-strength
electric fields (compared to the Dreicer field) leads to a transition
to the electron slide-away regime, despite the E field being weaker
than the threshold value Esa = 0.215ED for the initial distribution.
The time scale for this transition is relatively short, ranging from
a few to a few hundred thermal collision times for E fields in the
range E /Esa > 0.1. These effects cannot be consistently captured
in a linearized treatment, and this example thus illustrates that
NORSE opens new avenues of investigation into the dynamics of
relativistic electrons in plasmas.
Acknowledgments
The authors would like to thank E. Hirvijoki for his initial work
on the Braams and Karney collision operator, S.-M. Weng for his
helpful assistance, and I. Pusztai, S. Newton, T. DuBois and G. Wilkie
for constructive discussions. This work was supported by the
Swedish Research Council (Dnr. 2014-5510), the European Research Council (ERC-2014-CoG grant 647121), and the Knut and
Alice Wallenberg Foundation (Dnr. KAW 2013.0078) . A.S. would
also like to acknowledge travel support from Adlerbertska Forskningsstiftelsen.
Appendix A. Derivation of electron–electron collision operator
in (p, ξ ) coordinates
In our coordinate system, the non-zero components of the
metric are
gpp = 1,
gξ ξ =
p2
1 − ξ2
,
gϕϕ = p (1 − ξ ),
2
2
(A.1)
with g = 1/gii . Note also that the position vector is just p = p ep
(with ep the unit vector along p), since the coordinate system is
spherical. Using this and some algebra, we can write the terms in
the parenthesis of Eq. (8) as
ii
D·
∂f
=
∂p
[(
)
∂ 2 Υ−
∂ Υ−
∂f
γ3
+
γ
p
−
γ
Υ
+
∂ p2
∂p
∂p
]
γ (1 − ξ 2 )
∂f
+
g
D
(
Υ
)
ep
ξ
ξ
p
ξ
−
p2
∂ξ
1 − ξ2
p4
[
(1 − ξ 2 )
(A.4)
come from the expression for ∂ 2 Υ− /∂ p∂ p in the operator L.
Writing out Eq. (8) in components, we get
]p )
∂f
− Ff
∂p
∂p
([
]ξ )]
(
)
∂
∂f
+
D·
− Ff
≡ α Ap + Aξ ,
∂ξ
∂p
Cee {f } = α
[
1 ∂
p2
( [
p2 D ·
(A.5)
where the superscripts p and ξ denote the corresponding vector
components. Carrying out the differentiations, we find the p-term
to be
p2 A p =
∂ Υ−
∂p
)
2
∂
Υ
∂
Υ− ∂ 2 f
−
− γ (1 − ξ 2 )
+ 2γ ξ
2
∂ξ
∂ξ
∂ p2
([ 3
]
p
+
+ 2γ p (8Υ2 − Υ0 )
γ
[
]
∂ Υ2
∂ Υ1
∂ Υ0
∂ Υ−
+ γ p2 −16
+6
−
− 2γ 3
∂p
∂p
∂p
∂p
2
p(1 − ξ 2 ) ∂ 2 Υ−
3 ∂ Υ−
− 2γ p
−
∂ p2
γ
∂ξ 2
)
3
∂
Υ
p
ξ
∂
Υ
∂ 2 Υ− ∂ f
−
−
− γ (1 − ξ 2 )
+2
+ 2γ ξ
2
∂ p∂ξ
γ ∂ξ
∂ p∂ξ ∂ p
∂ 2f
2
+ γ p Dpξ (Υ− )
∂ p∂ξ
[( 3
)
]
p
∂ Dpξ (Υ− ) ∂ f
+
+ 2γ p Dpξ (Υ− ) + γ p2
γ
∂p
∂ξ
[( 3
)
]
2
∂
Π
∂
f
p
∂
Π
∂
Π
− γ p2
−
+ 2γ p
+ γ p2 2 f , (A.6)
∂p ∂p
γ
∂p
∂p
(
γ p2 (8Υ2 − Υ0 ) − 2γ 3 p
whereas the ξ -term becomes
∂ 2f
∂ Dpξ (Υ− ) ∂ f
+γ
∂ p∂ξ
∂ξ
∂p
(
) 2
1 − ξ2
∂ Υ−
∂ f
+
g
D
(
Υ
)
+
p
−
Υ
ξ
ξ
ξ
ξ
−
+
γ p2
∂p
∂ξ 2
[
(
)
ξ
∂ Υ−
+ − 2 2 g ξ ξ D ξ ξ (Υ − ) + p
− Υ+
γp
∂p
( (
)]
)
∂ 2 Υ−
∂ Υ+
∂f
1 − ξ 2 ∂ gξ ξ Dξ ξ (Υ− )
+
+
p
−
γ p2
∂ξ
∂ p∂ξ
∂ξ
∂ξ
[
]
1 − ξ 2 ∂Π ∂f
ξ ∂Π
1 − ξ 2 ∂ 2Π
−
+
2
−
f.
γ p2 ∂ξ ∂ξ
γ p2 ∂ξ
γ p2 ∂ξ 2
Aξ = γ Dpξ (Υ− )
(A.7)
278
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
Table B.1
Some analytical expressions for yl,a useful for validating the recursive calculation.
l\a
0
1
2
2
−(3 + 2p2 )/p3
−(15γ /p4 + 6γ /p2 )
−105γ 2 /p5 − 42γ 2 /p3 + 27/p3 + 18/p
−945γ 3 /p6 − 378γ 3 /p4 + 483γ /p4 + 258γ /p2
−3γ /p3
3/p2 − 15γ 2 /p4
−5(12γ /p3 + 21γ /p5 )
−45(7γ 2 /p4 + 21γ 2 /p6 + 1/p2 )
−3/p3
−(5/2)γ /p2 × (15 − 15γ 2 /p2 + 21/p2 )
−(35/2)γ 2 /p3 × (15 − 15γ 2 /p2 + 21/p2 ) + 15/p3
−(315/2)γ 3 /p4 × (15 − 15γ 2 /p2 + 21/p2 ) + 135γ /p4
+ 30γ /p2 × (15 − 15γ 2 /p2 + 21/p2 )
3
4
5
Combining and re-grouping the terms according to the derivative
of f , we arrive at the final expressions (17)–(22). In obtaining some
of these results, we have used the differential operator La , which
can be written as
(
)
∂ 2Ψ
2
∂Ψ
+
+ 3p
2
∂p
p
∂p
(
)
2ξ ∂ Ψ
(1 − ξ 2 ) ∂ 2 Ψ
− 2
+ 1 − a2 Ψ ,
(A.8)
+
p2
∂ξ 2
p ∂ξ
together with L2 Υ± = 4Υ1 ± Υ0 , to remove third order derivatives.
La Ψ = γ 2
Appendix B. Boundary conditions for the potentials Υi and Πi
at p = pmax
To calculate the potentials Υi,l and Πi,l (or collectively Ψl ) from
fl , boundary conditions at p = pmax must be specified. These can be
determined from Eq. (31) in Ref. [34], which for the final grid point
becomes
Ψl (pmax ) =
pmax
∫
Nl,∗ (pmax , p′ )
0
p ′2
γ′
fl (p′ )dp′ ,
(B.1)
where the place holder ∗ denotes a set of indices, distinct for each
potential. These indices – which in the notation of Ref. [34] specify
the order of differential operators La to apply to obtain a given
potential from f (cf. Eq. (15)) – are
Υ0 : 0,
Υ1 : 02,
Υ2 : 022,
Π0 : 1,
Π1 : 11.
(B.2)
The quantity Nl,∗ is defined as
Nl,∗ (p, p′ )
⎧
y (p)j (p′ ),
⎪
⎨ l,a l,a ′
if ∗ = a,
yl,a (p)jl,aa′ (p ) + yl,aa′ (p)jl,a′ (p′ ), if ∗ = aa′ ,
=
(B.3)
′
′
′
⎪
⎩yl,a (p)jl,aa′ a′′′ (p′′ ) + yl,aa′ (p)jl,a′ a′′ (p ) + yl,aa′ a′′ (p)jl,a′′ (p ),
if ∗ = aa a ,
where yl,a (p) and jl,a (p) are two independent solutions to the homogeneous equation La,l Ψl,a = 0 (here Ψl,a represents one of the
one-index potentials; either Υ0 and Π0 , depending on the value
of a), and the other yl,∗ and jl,∗ can be calculated from these using
relations given in [34]. The problem of finding Nl,∗ can be reduced
to recursively calculating jl,a for a = 0, 1, 2, and all l of interest,
however the recursive calculation is numerically non-trivial. A
method for achieving accurate results is outlined in Appendix 7 of
[34]. Validation of the obtained jl,a can be done using the equation
after Eq. (A4) in [34] for any l ≥ 0, but the calculation of yl,a from
jl,a requires that the recursion be performed also for l < 0. For
completeness, Table B.1 lists some analytic expressions for yl,a (not
contained in Eqs. (A26–A27) in [34]) which are useful in validating
the recursive algorithm. Note also that Eq. (A28c) in [34] has a typo,
and should read y1[1]2 = −(1 + 2z 2 )γ /z 2 .
References
[1] D.A. Roberts, J.A. Miller, Geophys. Res. Lett. 25 (1998) 607–610. http://dx.doi.
org/10.1029/98GL00328.
[2] W.G. Pilipp, H. Miggenrieder, M.D. Montgomery, K.H. Mühlhäuser,
H. Rosenbauer, R. Schwenn, J. Geophys. Res.: Space Phys. 92 (1987) 1075–1092.
http://dx.doi.org/10.1029/JA092iA02p01075.
[3] J.A. Van Allen, S.M. Krimigis, J. Geophys. Res. 70 (1965) 5737–5751. http:
//dx.doi.org/10.1029/JZ070i023p05737.
[4] G.D. Holman, Astrophys. J. 293 (1985) 584–594. http://dx.doi.org/10.1086/
163263.
[5] T.F. Bell, V.P. Pasko, U.S. Inan, Geophys. Res. Lett. 22 (1995) 2127–2130. http:
//dx.doi.org/10.1029/95GL02239.
[6] A. Gurevich, G. Milikh, R. Roussel-Dupre, Phys. Lett. A 187 (1994) 197–203.
http://dx.doi.org/10.1016/0375-9601(94)90062-0.
[7] V. Malka, J. Faure, Y. A. Gauduel, E. Lefebvre, A. Rousse, K.T. Phuoc, Nature Phys.
4 (2008) 447–453. http://dx.doi.org/10.1038/nphys966.
[8] M. Tabak, D.S. Clark, S.P. Hatchett, M.H. Key, B.F. Lasinski, R.A. Snavely, S.C.
Wilks, R.P.J. Town, R. Stephens, E.M. Campbell, R. Kodama, K. Mima, K.A.
Tanaka, S. Atzeni, R. Freeman, Phys. Plasmas 12 (2005) 057305 . http://dx.doi.
org/10.1063/1.1871246.
[9] P. Helander, L.-G. Eriksson, F. Andersson, Plasma Phys. Control. Fusion 44
(2002) B247–62. http://dx.doi.org/10.1088/0741-3335/44/12B/318.
[10] H. Dreicer, Phys. Rev. 115 (1959) 238–249. http://dx.doi.org/10.1103/PhysRev.
115.238.
[11] H. Dreicer, Phys. Rev. 117 (1960) 329–342. http://dx.doi.org/10.1103/PhysRev.
117.329.
[12] M. Rosenbluth, S. Putvinski, Nucl. Fusion 37 (1997) 1355–1362. http://dx.doi.
org/10.1088/0029-5515/37/10/I03.
[13] A.H. Boozer, Phys. Plasmas 22 (2015) 032504 . http://dx.doi.org/10.1063/1.
4913582.
[14] E. Hirvijoki, I. Pusztai, J. Decker, O. Embréus, A. Stahl, T. Fülöp, J. Plasma Phys.
81 (2015) 475810502. http://dx.doi.org/10.1017/S0022377815000513/.
[15] J. Decker, E. Hirvijoki, O. Embreus, Y. Peysson, A. Stahl, I. Pusztai, T. Fülöp,
Plasma Phys. Control. Fusion 58 (2016) 025016 . http://dx.doi.org/10.1088/
0741-3335/58/2/025016.
[16] G. Pokol, T. Fülöp, M. Lisak, Plasma Phys. Control. Fusion 50 (2008) 045003 .
http://dx.doi.org/10.1088/0741-3335/50/4/045003.
[17] G.I. Pokol, A. Kómár, A. Budai, A. Stahl, T. Fülöp, Phys. Plasmas 21 (2014)
102503. http://dx.doi.org/10.1063/1.4895513.
[18] S. Chiu, M. Rosenbluth, R. Harvey, V. Chan, Nucl. Fusion 38 (1998) 1711–1721.
http://dx.doi.org/10.1088/0029-5515/38/11/309.
[19] J. Decker, Y. Peysson, DKE: A fast numerical solver for the 3D drift kinetic
equation, Technical Report EUR-CEA-FC-1736, Euratom-CEA, 2004.
[20] M. Landreman, A. Stahl, T. Fülöp, Comput. Phys. Commun. 185 (2014) 847–855.
http://dx.doi.org/10.1016/j.cpc.2013.12.004.
[21] M. McCoy, A. Mirin, J. Killeen, Comput. Phys. Commun. 24 (1981) 37–61. http:
//dx.doi.org/10.1016/0010-4655(81)90105-3.
[22] L. Chacón, D. Barnes, D. Knoll, G. Miley, J. Comput. Phys. 157 (2000) 654–682.
http://dx.doi.org/10.1006/jcph.1999.6395.
[23] L. Pareschi, G. Russo, G. Toscani, J. Comput. Phys. 165 (2000) 216–236. http:
//dx.doi.org/10.1006/jcph.2000.6612.
[24] S.-M. Weng, Z.-M. Sheng, M.-Q. He, H.-C. Wu, Q.-L. Dong, J. Zhang, Phys.
Plasmas 13 (2006) 113302 . http://dx.doi.org/10.1063/1.2370725.
[25] A. Pataki, L. Greengard, J. Comput. Phys. 230 (2011) 7840–7852. http://dx.doi.
org/10.1016/j.jcp.2011.07.005.
[26] E.S. Yoon, C.S. Chang, Phys. Plasmas 21 (2014) 032503 . http://dx.doi.org/10.
1063/1.4867359.
[27] W. Taitano, L. Chacón, A. Simakov, K. Molvig, J. Comput. Phys. 297 (2015)
357–380. http://dx.doi.org/10.1016/j.jcp.2015.05.025.
[28] E. Hirvijoki, J. Candy, E. Belli, O. Embreus, Phys. Lett. A 379 (2015) 2735–2739.
http://dx.doi.org/10.1016/j.physleta.2015.08.010.
[29] H. Nuga, A. Fukuyama, Prog. Nucl. Sci. Technol. 2 (2011) 78–84. http://dx.doi.
org/10.15669/pnst.2.78.
[30] Y. Petrov, R.W. Harvey, Benchmarking the fully relativistic collision operator in
CQL3D, 2009. URL: http://www.compxco.com/CompX-2009-1_Fully-Rel.pdf,
CompX report CompX-2009-1.
[31] J. Connor, R. Hastie, Nucl. Fusion 15 (1975) 415–424. http://dx.doi.org/10.
1088/0029-5515/15/3/007.
[32] S.T. Beliaev, G.I. Budker, Sov. Phys.-Doklady 1 (1956) 218.
[33] B.J. Braams, C.F.F. Karney, Phys. Rev. Lett. 59 (1987) 1817–1820. http://dx.doi.
org/10.1103/PhysRevLett.59.1817.
[34] B.J. Braams, C.F.F. Karney, Phys. Fluids B: Plasma Phys. 1 (1989) 1355–1368.
http://dx.doi.org/10.1063/1.858966.
[35] F. Andersson, P. Helander, L.-G. Eriksson, Phys. Plasmas 8 (2001) 5221–5229.
http://dx.doi.org/10.1063/1.1418242.
A. Stahl et al. / Computer Physics Communications 212 (2017) 269–279
[36] P. Helander, D.J. Sigmar, Collisional Transport in Magnetized Plasmas,
Cambridge University Press, 2002.
[37] M.N. Rosenbluth, W.M. MacDonald, D.L. Judd, Phys. Rev. 107 (1957) 1–6. http:
//dx.doi.org/10.1103/PhysRev.107.1.
[38] Y. Saad, M.H. Schultz, SIAM J. Sci. Stat. Comput. 7 (1986) 856–869. http://dx.
doi.org/10.1137/0907058.
[39] S.M. Weng, Z.M. Sheng, M.Q. He, J. Zhang, P.A. Norreys, M. Sherlock, A.P.L.
Robinson, Phys. Rev. Lett. 100 (2008) 185001 . http://dx.doi.org/10.1103/
PhysRevLett.100.185001.
[40] S.M. Weng, private communication, 2016.
[41] A. Stahl, O. Embréus, G. Papp, M. Landreman, T. Fülöp, Nucl. Fusion 56 (2016)
112009 . http://dx.doi.org/10.1088/0029-5515/56/11/112009.
279
[42] H. Smith, P. Helander, L.-G. Eriksson, T. Fülöp, Phys. Plasmas 12 (2005)
122505. http://dx.doi.org/10.1063/1.2148966.
[43] J.R. Martín-Solís, J.D. Alvarez, R. Sánchez, B. Esposito, Phys. Plasmas 5 (1998)
2370–2377. http://dx.doi.org/10.1063/1.872911.
[44] A. Stahl, E. Hirvijoki, J. Decker, O. Embréus, T. Fülöp, Phys. Rev. Lett. 114 (2015)
115002 . http://dx.doi.org/10.1103/PhysRevLett.114.115002.
[45] P. Aleynikov, B.N. Breizman, Phys. Rev. Lett. 114 (2015) 155001 . http://dx.doi.
org/10.1103/PhysRevLett.114.155001.
[46] C. Liu, D.P. Brennan, A. Bhattacharjee, A.H. Boozer, Phys. Plasmas 23 (2016)
010702 . http://dx.doi.org/10.1063/1.4938510.
[47] B. Coppi, F. Pegoraro, R. Pozzoli, G. Rewoldt, Nucl. Fusion 16 (1976) 309. URL:
http://stacks.iop.org/0029-5515/16/i=2/a=014.
Paper F
A. Stahl, O. Embréus, M. Landreman, G. Papp and T. Fülöp,
Runaway-electron formation and electron slide-away in an ITER post-disruption scenario.
Journal of Physics: Conference Series 775, 012013 (2016).
http://dx.doi.org/10.1088/1742-6596/775/1/012013
http://arxiv.org/abs/1610.03249
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
Runaway-electron formation and electron slide-away
in an ITER post-disruption scenario
A Stahl1 , O Embréus1 , M Landreman2 , G Papp3 and T Fülöp1
1
Department of Physics, Chalmers University of Technology, Göteborg, Sweden
Institute for Research in Electronics and Applied Physics, University of Maryland, College
Park, Maryland 20742, USA
3
Max Planck Institute for Plasma Physics, Garching, Germany
2
E-mail: [email protected]
Abstract. Mitigation of runaway electrons is one of the outstanding issues for the reliable
operation of ITER and other large tokamaks, and accurate estimates for the expected runawayelectron energies and current are needed. Previously, linearized tools (which assume the runaway
population to be small) have been used to study the runaway dynamics, but these tools are not
valid in the cases of most interest, i.e. when the runaway population becomes substantial. We
study runaway-electron formation in a post-disruption ITER plasma using the newly developed
non-linear code NORSE, and describe a feedback mechanism by which a transition to electron
slide-away can be induced at field strengths significantly lower than previously expected. If
the electric field is actively imposed using the control system, the entire electron population is
quickly converted to runaways in the scenario considered. We find the time until the feedback
mechanism sets in to be highly dependent on the details of the mechanisms removing heat from
the thermal electron population.
1. Introduction
Runaway electrons pose a severe threat to the safety and reliability of ITER and other highplasma-current fusion devices [1]. The larger the runaway-electron population, the larger the
threat to the integrity of the device. However, if the electron momentum-space distribution
function becomes highly non-Maxwellian due to the presence of a high-energy tail of runaways,
existing numerical tools employing linearized collision operators are no longer valid. The same
is true if the electric field (even momentarily) becomes comparable to the Dreicer field [2].
We recently presented NORSE [3] – an efficient solver of the kinetic equation in a homogeneous
plasma – which includes the full relativistic non-linear collision operator of Braams & Karney
[4, 5]. NORSE – which will be discussed in Section 2 – is able to model Dreicer and hot-tail
runaway generation in the presence of electric fields of arbitrary strength and synchrotronradiation reaction: one of the most important energy-loss channels for runaways [6].
Since NORSE is able to treat highly distorted distributions, a range of new questions may be
addressed. One issue of particular interest is: will non-linear phenomena accelerate or dampen
the growth of runaways? Naturally, this is of great importance in view of ITER and other large
tokamaks, as it potentially impacts the requirements on the disruption mitigation system (the
design of which is currently being finalized) [1, 7]. In addition, the electric field is expected
to reach values as high as 80-100 V/m during the current quench in ITER [8], and runaway
Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution
of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.
Published under licence by IOP Publishing Ltd
1
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
generation is likely to be strong enough for 60% or more of the plasma current to be converted
to runaway current. In Section 3, we use NORSE to model the evolution of the electron population
in a typical ITER post-disruption scenario.
If the electric field is strong enough, the net parallel force experienced by electrons due to
the electric field and collisions becomes positive in the entire momentum space, leading to a
phenomenon known as electron slide-away. This is expected to happen when E > 0.215ED ≡
ESA , where ED = ne3 ln Λ/4π20 T is the so-called Dreicer field [2]; n, T and −e are the electron
number density, temperature and charge; ln Λ is the Coulomb logarithm; and 0 is the vacuum
permittivity. The associated surge in runaway current can have a large impact on the potential
for material damage, as well as the subsequent evolution of the parallel electric field. The slideaway process cannot be consistently modelled using linear tools such as CODE [9, 10] or LUKE
[11], which assume a Maxwellian background plasma and therefore require E ESA , as well as
that the runaway fraction is small nr /n 1.
A strong electric field represents a source of energy that quickly heats up the electron
distribution. This heating can induce a transition to the slide-away regime – even under a fixed
applied electric field which is initially below the threshold E < ESA – since the collisional friction
is lower in a hotter distribution. As a consequence, the Dreicer field is also lower, making the
effective normalized field E/ED,eff higher for a given field strength. If the temperature increase
is large enough, the slide-away regime is reached, which happens at a field of E/ED,eff = 0.215
in the case of a constant applied electric field, i.e. it coincides with the standard slide-away field
at the effective temperature Teff [3].
In practice, many processes act to remove heat from the plasma. In a cold post-disruption
plasma, line radiation and bremsstrahlung from interactions with partially ionized impurities
are important loss channels, as is radial heat transport. Including a heat sink in numerical
simulation of such scenarios is therefore desirable, and the sink effectively acts to delay or
prevent the transition to slide-away. In this paper, we demonstrate that the evolution of the
runaway electron population – including the time to reach slide-away – is highly sensitive to the
properties of the applied heat sink, making a detailed investigation of the various loss channels
an area of interest for future work.
2. NORSE
We will use the newly developed fully relativistic non-linear tool NORSE [3] to study the dynamics
of the electron population. NORSE, which is valid in spatially uniform plasmas, solves the kinetic
equation
eE ∂f
∂
∂f
−
·
+
· (Fs f ) = Cee {f } + Cei {f } + S,
(1)
∂t
me c ∂p ∂p
where f is the electron distribution function, t is the time, me and p are the electron rest
mass and momentum, E is the electric field, c is the speed of light, Fs is the synchrotronradiation-reaction force, Cee is the relativistic non-linear electron-electron collision operator, Cei
is the electron-ion collision operator, and S represents heat and particle sources or sinks. For a
detailed description of the various terms and operators, see Ref. [3]. For the remainder of this
paper, we define the electric field such that electrons are accelerated in the positive pk direction.
In NORSE, the particle momentum p is represented in terms of the magnitude of the normalized
momentum p = γv/c (where v is the velocity of the particle and γ is the relativistic mass factor)
and the cosine of the pitch angle ξ = pk /p. The kinetic equation is discretized using finite
differences in both p and ξ. A linearly implicit time-advancement scheme is used, where the
five relativistic Braams-Karney potentials [4] – analogous to the Rosenbluth potentials in the
non-relativistic case – are calculated explicitly from the known distribution. These are then
used to construct the electron-electron collision operator Cee , and the remainder of the kinetic
equation is solved implicitly.
2
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
For the results presented in this paper, a non-uniform finite-difference grid was used to
improve the computational efficiency. In the pitch-angle coordinate, the grid points were chosen
with a dense spacing close to ξ = ±1 – in particular at ξ = 1 where the runaway tail forms –
but with a sparser grid at intermediate ξ. In the p-direction, a grid mapping with a tanh step in
spacing was used in order to produce a grid with dense spacing at low momenta (to accurately
resolve the bulk dynamics), and larger spacing in the high-energy tail, where the scale length of
variations in f is larger.
The runaway region in NORSE was determined by studying particle trajectories in phase
space, neglecting momentum-space diffusion but including self-consistent collisional friction and
synchrotron-radiation reaction [3]. The trajectory that terminates at ξ = 1 and p = pc marks the
lower boundary of the runaway region, since particles that follow it neither end up in the bulk
nor reach arbitrarily high energies. The critical momentum pc is the momentum at which the
balance of forces in the parallel direction (ξ = 1) becomes positive (i.e. the lowest momentum
at which the accelerating force of the electric field overcomes the collisional and synchrotronradiation-reaction drag). If the balance of forces is positive for all p at ξ = 1, however, all
electrons experience a net acceleration, and the population is in the slide-away regime.
2.1. Heat sink
Including a heat sink (HS) in the numerical simulations is of great importance for accurate
modelling of the distribution evolution during a disruption. The heat sink
used in NORSE to
remove heat from the thermal population has the form S = ∂/∂p · kh Sh f , where Sh (p) is an
isotropic function of momentum (i.e. Sh k p) and kh (t) is the magnitude of the source. The terms
in the kinetic equation that affect the total energy content are the electric-field and synchrotronradiation-reaction terms, however; when considering a subset Ω of momentum space, collisions
can also transfer energy in or out of Ω, and a corresponding term must be included. The total
energy change dW/dt in Ω can thus be written as
Z
eE ∂f
∂
∂
dW
= me c2
d3 p (γ − 1) −
·
+
· (Fs f ) − C{f } + kh
· (Sh f ) .
(2)
dt
me c ∂p ∂p
∂p
Ω
The magnitude kh of the sink in each time step can be determined by requiring dW/dt = 0. In
this work, we take Ω to represent the
pthermal bulk of the distribution, which we define as all
particles with v < 4vth,0 , with vth,0 = 2T0 /me the thermal speed at the initial temperature T0 .
In this study, the p-component of Sh was chosen to have the shape of a Maxwellian at
the desired temperature T . In practice, the momentum dependence of the sink will be more
complicated and subject to the details of the particular physical processes at work. It will
also likely have a limited energy-removal rate, dictated by for instance spatial gradients or
impurity content, which could limit its efficiency in maintaining a given temperature. A detailed
investigation of the characteristics of the sink is left for future work; the aim of this paper is to
highlight the sensitivity of the runaway-electron evolution to the particulars of the sink, and for
that purpose we will impose a limit on the energy-removal rate, as will be discussed in the next
section.
3. Runaway generation in an ITER disruption
3.1. Post-disruption scenario
In this section, we use NORSE to model the evolution of the electron distribution during a typical
ITER disruption. The electric field evolution (which is shown in Fig. 1a) and other parameters
are taken from the ITER inductive scenario no. 2 in Ref. [8], but the temperature evolution
has been simplified to facilitate the numerical calculation. We assume the electron population
to be completely thermalized and use the final temperature T = T0 = 10 eV throughout our
3
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
Figure 1. a) Electric field in V/m (left vertical axis) and normalized to the Dreicer field ED at
the temperature T0 and density n0 (right vertical axis), as a function of time after the thermal
quench. b) Tail of the parallel electron distribution. Thin lines show f at tN (no HS), tW (weak
HS) and tS (strong HS), and thick lines show f immediately before the transition to slide-away.
simulation, together with the density n = n0 = 7.1·1019 m−3 . This is likely to underestimate
the runaway-electron generation in the early phase, in which the temperature is still dropping,
however the chosen set-up is sufficient for our purposes. We use the magnetic field on axis
(B = 5.3 T) and Zeff = 1. The initial current density in the scenario is j0 = 0.62 MA/m2 ,
however our calculations start from a Maxwellian distribution and make no attempt to maintain
the experimental current evolution explicitly.
To highlight the importance of the temperature evolution of the bulk, in Section 3.2 we will
consider three scenarios: no heat sink (subscript N), weak heat sink (W) and strong heat sink
(S). In the no-heat-sink scenario, all the energy supplied by the electric field will remain in
the simulation, leading to rapid bulk heating; with the strong heat sink, a bulk temperature of
T0 = 10 eV will be enforced in accordance with Eq. 2, i.e. any excess heat in the bulk region will
be removed using a heat sink. In the intermediate case of a weak heat sink, the energy-removal
rate of the heat sink will be restricted to 0.5 MW/m3 . This particlar value has been chosen at
will, but is meant to represent some inherent limitation in the physical processes responsible for
the energy loss. In both the weak and strong cases, the heat sink will affect only the thermal
population, allowing the supra-thermal tail to gain energy from the electric field. Physically, this
corresponds to processes not included in the simulation (such as radial transport or radiative
losses) which primarily affect the thermal population.
NORSE simulations of the evolution of the electron distribution function in the presence of
the electric field in Fig. 1a were performed for the three different scenarios. The simulations
were aborted when the runaway population reached nr /n = 1; i.e. a transition to the slide-away
regime was observed. The simulation results can however only be considered characteristic of
a natural ITER disruption for current densities comparable to, or somewhat larger than, the
initial value j0 , since after that point the strong response of the inductive electric field to the
increased local current would invalidate the E-field evolution used. We will therefore mark the
time where the current density reaches j/j0 > 5 in all plots, and denote it with tN , tW and tS ,
respectively, for the no-sink, weak-sink and strong-sink scenarios. The distribution evolution
at later times can only be considered accessible in scenarios where the loop voltage is actively
sustained using the control system. Nevertheless, this regime will turn out to be of interest,
since a non-linear feedback mechanism leading to a rapid transition to slide-away is observed.
4
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
Figure 2. a) Runaway fraction and b) current density normalized to its initial value j0 , as a
function of time after the thermal quench in the different heat-sink scenarios. The times tN , tW
and tS (vertical thin dashed lines), mark the time where the current density reaches j/j0 > 5
for the no-heat-sink, weak-heat-sink and strong-heat-sink scenarios, respectively.
3.2. Evolution of the runaway-electron population
In Fig. 1b, the tails of the distributions in the parallel direction are shown at tN , tW and
tS (thin lines), as well as at the final times (thick lines) in each scenario (just before the
transition to slide-away is reached). In the figure, the distribution is normalized such that
F = f /fM (t = 0, p = 0), where fM is a Maxwell-Jüttner distribution, so that F initially takes the
value unity at p = 0. The maximum achieved particle energies are highly dependent on whether
a heat sink was applied or not; in the no-heat-sink and weak-heat-sink scenarios, the particles
did not have time to reach relativistic energies, whereas in the strong heat sink case, p ≈ 19
and p ≈ 44 (corresponding to energies of roughly 9 and 22 MeV), were obtained at tS and just
before reaching slide-away, respectively. The reason for this is that, as we shall see, the current
density growth and subsequent transition to slide-away in the latter case occur at much later
times, and the runaways have time to gain more energy.
Figure 2a shows the evolution of the runaway fraction during the course of the simulation.
In the no-heat-sink case, the runaway fraction increases sharply, but as shown in Fig. 1b, the
runaways are all at low energy. The transition to slide-away happens already at t= 5.7 ms;
early on in the electric-field evolution (cf. Fig. 1). In the two scenarios employing a heat sink,
the growth in runaway fraction occurs later, but in the weak-heat-sink case the growth rate
is comparable to the case when a sink is absent once the process is initiated. In this case,
the transition to slide-away happens at t = 6.7 ms. With the strong heat sink, the runaway
population grows steadily, eventually dominating the entire distribution at t = 8.4 ms, however
in this case the transition is gradual, rather than rapid. In all three scenarios, including the one
with an ideal strong heat sink, the slide-away regime is thus reached even before the electric
field (calculated assuming a linear treatment) has reached its peak.
Figure 2b shows the evolution of the current density. It indicates that the rapid increase in
the runaway fraction is correlated with a similar increase in the current density, although in the
no-sink case, the growth rate is somewhat smaller. Again, the growth in the strong-heat-sink
case is gradual, rather than explosive. Note that in the no and weak heat-sink cases, the runaway
fraction is still negligibly small at the start of the rapid transition to slide-away. The transition
is thus not a non-linear phenomenon triggered by the size of the runaway population; it starts
in a regime where linearized tools are normally expected to be valid, and before the current
density becomes significantly larger than its initial value.
5
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
Figure 3. a) Effective temperature of the bulk population and b) corresponding effective
normalized electric-field strength (also taking changes to the bulk density into account) as
functions of time, in the different scenarios. The black solid line in panel b) corresponds to
the normalized field with T = T0 and n = n0 (cf. Fig. 1).
The explanation can be found by examining the thermal population. By comparing the energy
moment of the bulk of the distribution (WΩ ) with that of a relativistic Maxwellian (WM (T )),
WΩ = me c2
Z
Ω
d3 p (γ − 1)f = WM =
me c2 n
Θeff K2 (1/Θeff )
Z
0
γ
dp p2 (γ − 1) exp −
,
Θeff
pmax,Ω
(3)
an effective temperature Teff for a given distribution f can be determined by solving for
Θeff = Teff /me c2 . In the above equation, pmax,Ω is the upper boundary in p of the bulk region
in momentum space, and K2 (x) is the modified Bessel function of the second kind (and order
two). The effective temperature Teff is plotted in Fig. 3a as a function of time for all three
scenarios. In the no-heat sink case, it increases by roughly two orders of magnitude during
the simulation, i.e. when energy is not actively removed from the system, the Ohmic heating is
sufficient to heat the plasma to a temperature of about 700 eV before the onset of slide-away,
or 55 eV if the electric field is not artificially sustained. A similar (albeit weaker) tendency
is seen in the weak-heat-sink case, where the temperature increases to Teff ≈ 210 eV in the
phase leading up to the slide-away transition, or 25 eV in a non-driven case. This heating is a
consequence of the imposed limited maximum energy-removal rate of the heat sink in the weak
case, since the temperature is efficiently kept constant in the beginning of the simulation, where
(dW/dt)HS < 0.5 MW/m3 . The strong heat sink manages to keep Teff − T0 to within a few
tenths of an eV during the entire simulation, corresponding to a source with unlimited (or at
least higher than required) maximal energy-removal rate.
The significance of the observed bulk heating is its influence on the Dreicer field ED , which
(apart from the weak dependence on ln Λ) is inversely proportional to T . For a given electricfield strength, the normalized field E/ED thus increases as the bulk heats up. The effective
normalized electric field is shown in Fig. 3b, indicating that the rapid growth in the runaway
fraction in Fig. 2a is correlated with a sudden increase in the normalized electric field in the noheat-sink and weak-heat-sink cases. In the strong-heat-sink case, the increase in normalized field
is not caused by the temperature, which is kept constant during the entire simulation, but by
the decrease in the bulk density as the runaway population becomes substantial. As can be seen
from the yellow dash-dotted line in the figure, this has a similar effect as a temperature increase,
since ED ∼ nbulk . The effective E/ED starts to deviate from the baseline value (black solid line)
6
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
Figure 4. a) Evolution of the bulk of the electron distribution and b) balance of forces, in the
direction parallel to the electric field, just before the transition to slide-away in the weak-heatsink scenario. t̂ is the time relative to the transition to slide-away in µs.
already when nbulk /n0 ≈ 0.97. The feedback process is thus initiated when the runaway fraction
is just 3%; a regime where linear tools are expected to be valid.
The two effects of increasing Teff and decreasing nbulk lead to a positive feedback mechanism
which is responsible for the rapid growth in runaway fraction and current density seen in the
no-heat-sink and weak-heat-sink cases. Once the bulk temperature has increased enough (or
a high enough runaway tail is produced, as seen in the strong-heat-sink case), the normalized
electric field becomes strong enough to cause a depletion of the bulk through primary runaway
generation. The reduced bulk density in turn leads to a more efficient heating of the remaining
bulk particles. Both of these effects contribute to a reduction in the Drecier field ED ,
and a corresponding reduced collisional friction on the bulk electrons, which makes runaway
acceleration easier. This further increases the rate of bulk depletion, and so on. Eventually the
friction becomes low enough that the parallel balance of forces becomes positive everywhere,
marking the transition to the slide-away regime. At this point, the bulk of the distribution can
no longer be well described by a Maxwellian and the positive feedback mechanism makes the
transition possible even though E/ED,eff < 0.215 (in this case at around E/ED,eff ≈ 0.15).
The bulk depletion and associated change in the parallel force balance is shown in Fig. 4a
and Fig. 4b, respectively, in the phase leading up to the transition to slide-away in the weakheat-sink case. Initially, the force balance is positive in the tail – i.e. particles there experience
a net acceleration – while it is negative in the bulk, meaning particles are slowed down by
collisions. As the feedback process starts, the minimum in the sum of forces becomes gradually
less pronounced, in tandem with the depletion of the bulk population, up until the point where
the sum of forces becomes positive everywhere, and the slide-away regime is reached. This highly
non-linear process cannot be accurately captured by a linear model.
4. Discussion and conclusions
In this paper we have examined the evolution of the electron distribution and runaway generation
in an ITER-like post-disruption scenario where the electric fields reach values as high as 90 V/m.
With the help of the newly developed tool NORSE, which is a relativistic non-linear solver for
the electron momentum-space distribution function, we have shown that the slide-away regime,
i.e. a net parallel acceleration of electrons in all of momentum space, is reached in this scenario,
provided the electric field evolution used is artificially enforced by the control system. In the
stage leading up to the transition, a positive feedback mechanism sets in by means of which
the bulk quickly gets depleted by primary runaway generation which reduces the friction on the
7
Joint Varenna-Lausanne International Workshop on the Theory of Fusion Plasmas 2016
IOP Publishing
Journal of Physics: Conference Series 775 (2016) 012013
doi:10.1088/1742-6596/775/1/012013
thermal population, leading to further bulk depletion, until the point where the slide-away is
reached. This process can be initiated at significantly weaker fields than the slide-away field
ESA expected from linear theory.
The time to a transition to slide-away is highly dependent on the ability of loss processes
to remove heat from the thermal electron population, but even with an ideal sink (the strongheat-sink case in Section 3), complete runaway generation was seen 8.4 ms after the thermal
quench. These results were obtained without taking avalanche or hot-tail runaway generation
into account, which would only lead to more prominent runaway growth.
Also in the case of a disruption where the electric field is not artificially sustained, strong
bulk heating leads to a rapid growth in the runaway fraction because of the increase in the
normalized electric field E/ED , but the current density becomes large enough to significantly
affect the electric field evolution (supressing the growth of E) before the slide-away regime is
reached. This is observed in the absence of a heat sink, as well as with a heat sink with a limited
maximum energy-removal rate (in this case 0.5 MW/m3 ). If the efficiency of the heat sink
is not limited, the runaway fraction grows more slowly and the runaways have time to reach
significantly higher energies before the electric field becomes affected by the growing current
density. The severity of disruptions in ITER could thus be greatly affected by the properties of
the heat sinks present in the plasma.
The feedback mechanism described in this paper has important consequences for the
understanding of runaway-electron dynamics. With the entire electron population experiencing
a net accelerating force at much weaker electric fields than previously expected, very large
runaway-electron current generation is likely. This would impact the subsequent electricfield evolution, leading to a reduction in field strength and duration which could occur at
realatively early times if the heat sink has a limited energy removal rate. Therefore, it is
difficult to determine the magnitude of the effect on the current evolution and post-quench
dynamics without a self-consistent calculation of the electron distribution and the electric field.
Nevertheless, this paper shows that feedback effects play an important role in post-disruption
runaway dynamics, and that the details of the heat-loss channels may have a big impact on
what strength and duration of electric field can be tolerated before the positive feedback, and
possible subsequent transition to slide-away, is induced.
Acknowledgments
This work was supported by the Swedish Research Council (Dnr. 2014-5510), the European
Research Council (ERC-2014-CoG grant 647121), and the Knut and Alice Wallenberg
Foundation. M.L. was supported by the U.S. Department of Energy, Office of Science, Office of
Fusion Energy Science, under Award Numbers DE-FG02-93ER54197 and DE-FC02-08ER54964.
References
[1] Hollmann E M, Aleynikov P B, Fülöp T, Humphreys D A, Izzo V A, Lehnen M, Lukash V E, Papp G,
Pautasso G, Saint-Laurent F and Snipes J A 2015 Physics of Plasmas 22 021802
[2] Dreicer H 1959 Physical Review 115(2) 238–249
[3] Stahl A, Landreman M, Embréus O and Fülöp T 2016 Submitted to Computer Physics Communications
URL http://arxiv.org/abs/1608.02742
[4] Braams B J and Karney C F F 1987 Physical Review Letters 59 1817–1820
[5] Braams B J and Karney C F F 1989 Physics of Fluids B: Plasma Physics 1 1355–1368
[6] Stahl A, Hirvijoki E, Decker J, Embréus O and Fülöp T 2015 Physical Review Letters 114 115002
[7] Boozer A H 2015 Physics of Plasmas 22 032504
[8] Smith H, Helander P, Eriksson L G, Anderson D, Lisak M and Andersson F 2006 Physics of Plasmas 13
102502
[9] Landreman M, Stahl A and Fülöp T 2014 Computer Physics Communications 185 847 – 855 ISSN 0010-4655
[10] Stahl A, Embréus O, Papp G, Landreman M and Fülöp T 2016 Nuclear Fusion 56 112009
[11] Decker J and Peysson Y 2004 DKE: A fast numerical solver for the 3D drift kinetic equation Tech. Rep.
EUR-CEA-FC-1736 Euratom-CEA
8