ArticlePDF Available

Macroscopical model of streamer coronas around a spherical electrode

IOP Publishing
Plasma Sources Science and Technology
Authors:

Abstract and Figures

We present a model for streamer coronas emerging from a spherical electrode at high electrostatic potential. By means of a macroscopic streamer model and approximating the corona as a set of identical streamers with a prescribed spatial distribution around the electrode, we establish that coronas more densely packed with streamers are slower and more efficient at screening the electric field inside the streamers. We also apply our model to investigate the electrostatic potential at the boundary of the corona sheath that surrounds a leader and we underline the relevance of the rise-time of the leader potential during a leader step.
This content is subject to copyright. Terms and conditions apply.
Macroscopical model of streamer coronas
around a spherical electrode
M González
1,2
, F J Gordillo-Vázquez
1
and A Luque
1,3
1
Instituto de Astrofísica de Andalucía, IAA-CSIC Glorieta de la astronomía s/n, E-18008, Granada, Spain
2
Institut de Planétologie et dAstrophysique de Grenoble (IPAG), Université Grenoble Alpes, F-38058,
Grenoble Cédex 9, France
E-mail: marta.gonzalez@univ-grenoble-alpes.fr and aluque@iaa.es
Received 21 May 2019, revised 11 September 2019
Accepted for publication 16 October 2019
Published 12 November 2019
Abstract
We present a model for streamer coronas emerging from a spherical electrode at high
electrostatic potential. By means of a macroscopic streamer model and approximating the corona
as a set of identical streamers with a prescribed spatial distribution around the electrode, we
establish that coronas more densely packed with streamers are slower and more efcient at
screening the electric eld inside the streamers. We also apply our model to investigate the
electrostatic potential at the boundary of the corona sheath that surrounds a leader and we
underline the relevance of the rise-time of the leader potential during a leader step.
Keywords: streamers, corona, electric discharge
1. Introduction
When a sharp electrode such as a needle or a wire is subjected
to a high electric potential it ignites a type of electrical dis-
charge called corona. Depending on conditions such as
electrode geometry and rise time of the potential, the dis-
charge may exhibit a variety of forms [1,2]. A sufciently
sharp electrode on which a sufciently impulsive potential is
applied launches a corona composed of many thin laments
called streamers that propagate due to electron impact ioniz-
ation in a high-eld volume around their tips [35].
Streamer coronas exist on Earth in a variety of natural
and articial conditions. They appear as sprites [6,7]or blue
jets [811]on the upper atmosphere and they form a sheath
around the hot, highly conducting leader core of a long spark
[12]or an advancing lightning channel [13]. In articial
electrical systems, the undesired emergence of coronas can
cause severe negative effects like power loss or high-fre-
quency electromagnetic interferences. On the other hand,
streamer coronas have a variety of benecial applications
such as water and gas cleaning, electrostatic precipitation,
ozone creation, or even inactivation of bacteria (see
e.g. [14,15]).
Single streamers have been widely studied from a
microscopical standpoint with Monte Carlo [1618],uid
[1925], and hybrid [26]models. Other studies on specic
phenomena like branching [27,28], head-on collisions
between streamers [18,23,29,30], the effect of inhomo-
geneities of the medium [31,32]or the formation of luminous
structures inside the channels [33,34]are also available. The
interaction between streamers, however, is a very complex
matter and most of the works about it involve only two
streamers [3539]. Up to our knowledge, only the works by
Naidis [35], Akyuz et al [40]and Luque and Ebert [41]have
simulated a system with more than two streamers, and in all
cases the underlying system was heavily simplied.
Recently we presented a macroscopical model for strea-
mers treating them as advancing imperfect conductors [42].
This model allowed us to explore long term channel proper-
ties like channel inhomogeneities and their role in the for-
mation of streamer glows. Here we apply our model to
investigate the properties of streamer coronas composed of
many interacting streamers emerging from a spherical elec-
trode. Our aim is to understand how properties of the corona
depend on the density of streamers that compose it. As we see
below, the corona model presented here is considerably
Plasma Sources Science and Technology
Plasma Sources Sci. Technol. 28 (2019)115007 (13pp)https://doi.org/10.1088/1361-6595/ab4e7a
3
Author to whom any correspondence should be addressed.
Original 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.
0963-0252/19/115007+13$33.00 © 2019 IOP Publishing Ltd1
simplied but precisely for that reason it serves to build an
intuition about the coronas complex dynamics.
This paper is structured in six sections, the rst being this
introduction. In section 2we describe both the single streamer
and the streamer interaction model, explaining the underlying
assumptions and approaches. In section 3we describe the
numerical approximations and the implementation of the model.
In section 4we describe the results obtained in our simulations
for one conguration of experimental interest whereas in
section 5we investigate the potential drop within a corona.
Finally, in section 6we summarize the conclusions of our work.
2. Model description
In this section we start by briey summarizing the streamer
model used as base for the streamer corona. Then we explain
how to add the interaction between several streamers and the
numerical implementation used in this work.
2.1. General description of the single streamer model
We start with a brief description of the main equations and
assumptions of the single streamer model, as thoroughly
described in [42]. The streamer is modeled as a symmetric
one-dimensional imperfectly conducting charged channel
attached to an spherical electrode of radius aand propagating
in the zdirection (see gure 1(a)). This system evolves
according to electrodynamic and chemical models that are
coupled through the electric eld inside the streamer channel.
The transport of charge within the channel is determined
by charge conservation
t
q=−∇ ·j, where qis the charge
density and jthe current density. Integrating this equation
across a plane Γperpendicular to the streamer channel we
obtain
()
l
=-
t
I
z,1
where λis the linear charge density, and Iis the current
intensity:
()
òò
l==
GG
qSI j Sd, d. 2
z
Under the assumptions of negligible electron diffusion,
independence of mobilities from the electric eld, and uni-
form charge-carrier densities across the channel, the current
intensity is proportional to the integral of the electric eld on
the section of the streamer and (1)becomes a linear differ-
ential equation. The intensity Iis composed of three terms,
I=I
self
+I
bg
+I
surf
, that model the electric self-interaction
of the streamer, the inuence of an electrode and background
eld, and the effect of the surface current around the streamer
head as it propagates. This expression assumes that the
streamer radius is sufciently small that charge transport in
the transversal direction is instantaneous and we can describe
it using a linear charge density λ.
The terms I
α
, where αä{self, bg, surf}, can be
expressed in an integral form for each zin the streamer:
() () () ( ) ()
ò
s
pl¢¢
aa
Iz zzG zz z
4,d, 3
a
z
0
tip
where σ(z)is the conductivity of the streamer, ò
0
the vacuum
permitivity, and ()¢
a
G
zz,the kernel describing the inuence
on zof the charges in
¢z
.
The chemical model contains 17 species coupled by 78
reactions, as described in [42]. It boils down to a system of
ordinary differential equations describing the evolution of the
Figure 1. Model geometry of the simulations with (a)a single streamer and (b)two identical interacting streamers, S
0
and S
1
.
2
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
densities
n
siof each charge carrier s
i
,iä{1, K,n
spec
}
according to a set of reactions r
j
,jä{1, K,n
reac
}. The
equation governing the evolution of each species s
i
is given
by:
()
()
å
=
==
n
tAk n ,4
s
j
n
ij j
l
n
Ijl
1
,
1
,
i
j
reac in,
where A
i,j
is the net number of molecules of species s
i
created
each time reaction r
j
occurs (given by a difference on the
stochiometric coefcients in r
j
),k
j
is the rate coefcient of
reaction r
j
, and I(j,l)is the index of the reactant lof reaction
r
j
. The rate coefcients k
j
,jä{1, K,n
reac
}are functions of
the local electric eld, which in [42]is approximated by the
eld at the axis for each z. As we explain in section 3, in this
work we approximate instead the local eld at each zby the
average of the eld on the disk perpendicular to the streamer.
The propagation velocity of the streamer depends on the
streamer radius R
0
and the eld at the tip (estimated imme-
diately outside the streamer, since it is discontinuous at the
tip)following the expression derived by Naidis [43]. This
expression is a relationship between the streamer radius, its
velocity and the background electric eld. A nonlinear root
nder can be used to solve for the velocity given the other two
quantities. The main assumption behind Naidisformula is
that the electron density undergoes multiplication by a xed
factor within the area around the streamer head where the
electric eld is above breakdown. For positive streamers, the
generation of a sufcient quantity of electrons outside this
active area due to photo-ionization is implicitly assumed in
the derivation.
The model includes the effect of the eld ahead of the
streamer tip by imposing densities
n
s
0
iat the streamer tip, as
explained in [42]. Since streamers are initiated with a nite
length, the species densities are given initial conditions inside
the streamer body, which we set also as
n
s
0
i.
2.2. Additional streamers
Let us now consider the effect of additional streamers on the
model. Neighboring streamers inuence each other directly
through their electrostatic interaction; to incorporate this in
our model we introduce two assumptions, in addition to those
in [42]and described in section 2.1:
There is a perfectly symmetrical conguration of
streamers around a spherical electrode of radius a,so
we can consider that all the streamers are equal, i.e. that
the length, density charge, and other characteristics of all
the streamers S
i
,i>0 are the same as those of S
0
. This
allows us to concentrate on only one streamer, S
0
.
The streamer radius R
0
is small compared to the
minumum distance between streamers. This assumption
allows us to approximate S
i
,i>0asaninnitely thin
line of charge when we calculate its effect on S
0
.
Furthermore, we also neglect the variation within the
cross-section of S
0
of the electric eld created by S
i
.To
summarize: we consider the nite width of the streamers
only for the interactions between separate points of the
same streamer but not for the interaction between
different streamers.
The geometry of the effect exterted on S
0
by a neigh-
boring streamer S
1
is sketched in gure 1(b). We calculate the
electric eld at a point p
0
on streamer S
0
which, without loss
of generality, we consider to be on the zaxis. Thus, p
0
=(x
0
,
y
0
,z
0
)=(0, 0, z
0
)äS
0
.Asin[42], we only need the z
component of the electric eld induced in p
0
by a point charge
q
1
located at p
1
=(x
1
,y
1
,z
1
)äS
1
. This is:
() [()]
()
p
=-
++-
Ep qzz
xy zz4.5
pz,1
0
01
121201
23 2
1
Now, given that ()( )==pxyzluuu,, , ,
xyz
11111
111, where
a
u1is the αcomponent of the unit vector in the direction
generating S
1
, and l
1
is the distance from p
1
to the origin, (5),
can be rewritten as:
() []
()
p
=-
+-
Ep qzlu
lz zlu42
.6
pz
z
z
,1
0
01
1
120201 132
1
Then the contribution of the whole streamer S
1
is:
() ()( )
[]
()
ò
p
l
=-
+-
Ep lz lu
lz zlu l
1
42
d, 7
zS
z
z
1, 0
0
11 0 11
120201 132 1
1
where λ
1
is the linear charge density in S
1
.
Now, the electric current generated in the disk of S
0
at z
0
is:
() () () () ()sp=Iz z RzEz,8
zz1, 0 0 0 0 21, 0
where, as in [42],() ( (( ) ))=- -
R
zR zzR1exp
0 tip 0 12 is a
smooth function that models the variation of the radius along
the streamer channel.
Since all the streamers are identical we drop the sub-
scripts in l,λ, and σand, for each streamer S
i
,i>0, dene a
kernel
() ()( )
[]
()
p
=-
+-
Kz l Rz z lu
lz zlu
,2.9
ii
z
i
z
0
020
202032
Following the method of images, we add a virtual charge
¢
q(called image or mirror charge)to satisfy the boundary
conditions in the spherical electrode boundary. We need,
then, to add the effect of
¢
qlocated at a distance ¢<la
. Since
k¢=q
q
2, and
k¢=ll
2
with κ=a/lbeing a squeeze factor,
we derive the mirror kernel for each streamer S
i
with i>0:
() ( ) ()kk¢=-Kzl Kz l,,.10
mi i,0 0 2
From K
i
and K
m,i
we can dene a kernel G
i
so the
contribution of streamer S
i
to the current intensity in z
0
äS
0
has the form of (3). Adding up all kernels we obtain the
following expression for the total current intensity Iat each
point z
0
äS
0
:
() () () ( ) ( )
ò
s
pl¢¢
Iz zzG z z z
4,d, 11
a
z
00
0
tot 0
tip
where G
tot
is a kernel that now includes the self interactions of
streamer S
0
, the background and surface current, and the
contributions of all the surrounding streamers S
i
,i=1... n,
that is, =+++
å=
GGG G
i
n
tot self bg surf 1.
3
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
3. Numerical implementation
We use nite differences to solve the partial differential
equations from section 2. We apply a leap frog scheme to
couple the electrodynamical and the chemical parts of the
model, and a CrankNicolson discretization for both
equations (1)and (4). Figure 2shows the owchart used to
implement the calculations. At each time tthe streamer length
is discretized in a set of cells C
i
,i=1... Lwith boundaries
z
i±1/2
, where the right boundary of cell Lis z
tip
. In the elec-
trodynamical step z
tip
moves forward with a velocity obtained
from the expression derived by [43]and we solve the charge
transport system. Some variables describing the system, like
conductivity, densities and electric eld are dened at cell
boundaries (e.g. σ
i±1/2
), while the charge density λ
i
is con-
sidered constant within each cell i:
()
=-
-+
q
tII,12
iii12 12
where q
i
is the total charge in the cell and I
i±1/2
are the
current intensities at the boundaries. The current intensity at
the boundary of cell jis given by (11), which we integrate
numerically in each cell. This way, we can dene an inter-
action matrix M={m
ij
}, and the background component
vector b={b
i
}so at each time tthe current intensity at the
rightmost boundary of cell iis
() () () ( )
å
=+
+
=
Imtqtbt.13
i
j
L
ij ji12
1
For the single streamer model, each element m
ij
from the
interaction matrix is the cross-sectional current induced at the
boundary of cell iby an unitary charge in cell j, and each
element b
i
is the current induced at the boundary of cell iby
the background and electrode eld. In this corona model, the
interaction matrix includes also the effect of the jth cell from
all the other streamers. With this expression, we cast (12)as a
subtraction of interaction matrices and background vectors
forming a linear system.
The electrodynamical system is particularly apt for par-
allelization, since each element of the interaction matrix can
be calculated independently. We have, thus, accelerated its
performance using CUDA (integrated with python using
pyCUDA). As for the chemical processes, the discrete system
to solve is sparse and nonlinear, and we solve it iteratively,
using NewtonRaphson. The solution of this system is
accelerated using Fortran and OpenMP.
Our previous work [42]showed that, in the conditions of
the model, the electric eld at each point zof the streamer,
used to couple the electrodynamic and chemical parts of the
model, could be approximated by the eld at the axis. We
have found that the goodness of this approach decays when
we increase the number of streamers. For this reason in this
work we instead calculated an average electric eld across the
channel.
One point where we still need to calculate the eld at the
axis is at the tip. The eld at the axis in the single streamer
case is calculated using an expression analogous to (3)with a
different kernel G
ax
:
() ( ) ( ) ( )
ò
pl¢¢
Ez zGzzz
1
4,d, 14
a
z
ax
0
ax
tip
where
() ()( )
[() ( )] ()
p
¢=
¢+ -¢
Gzz Rz z z
Rz z z
,.15
ax
2
2232
To calculate the inuence of a streamer S
i
on the eld at
the axis in z
0
äS
0
, we derive a kernel G
ax,i
from (7),
() ( )() ( )
ò
pl=
Ez Gzlll
1
4,d. 16
iSiax, 0
0
ax, 0
1
This expression can be combined with (14)to obtain the nal
expression for the eld at the axis for z
0
äS
0
:
()
() () ( )
()() ()
ò
òå
pl
pl
¢¢
¢+¢¢
=
17
Ez zG zzz
zGzz G zz z
1
4,d
1
4,,d.
a
z
a
z
i
n
i
ax 0
0
ax,tot 0
0
ax 0
1
ax, 0
tip
tip
When z
0
=z
tip
, the integral of the eld at the axis (17)is
bounded and, thus, convergent, but for zz
0
the conv-
ergence order is increasingly large as z
0
approaches z
tip
.We
evaluate the eld at a point slightly beyond the actual tip
z
0
=z
tip
+ε
tip
, where
e
-
R10 n
tip tip with n
tip
an order
factor (typically 3). This ensures that we use an appropriate
ε
tip
, close enough to represent the eld at the tip, but far
enough that the resolution needed for the eld estimation is
reduced. Even after this, we had to use resolutions of the
order of 10
5
to obtain results with enough precision, so
additional optimization was needed. We added a variable
change to expand the integration domain in the tip cell, so
with our new variable ξ, the eld slope is not too steep. The
usual ξ=z
n
variable change helped but not enough so we
opted for a variable change capturing the decay of the radius
close to the tip. We used the following expression:
() ( ) ( )
/
x=-
--
-
ze1, 18
12
zz
R
tip
0
which is the inverse of the smoothing decay factor
applied to the radius near the streamer tip,
=
fR
(
(( ) ))---zz R1exp tip 0 12
. The variable change in (18)
allowed us to use the same resolution as in the rest of the
streamer, heavily accelerating the code while ensuring that a
good and robust approximation of the eld at the tip is
being used.
One nal element in our model is the distribution of n
streamers in the spherical surface of the electrode. One pos-
sibility is to distribute them randomly with a constant density
per unit surface. However this often leads to pairs of strea-
mers that are too close and that in reality would merge into a
single one. Therefore we look for a distribution of inception
points that in some sense maximizes the minimum distance
between them, accounting for the repulsion of streamers as
well as for the possibility of merging.
4
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
The distribution of points on the surface of a sphere that
maximizes the minimum distance between points is a classical
open problem (see e.g. [44]), described as one of the problems of
the century by Smale [45].Thecongurations of points solving
this problem are called spherical codes and have different
applications depending on the metric considered. Given the
electrostatic nature of our problem we opted for the so called
Thomson problem (as described in e.g. [46]), which aims to
determine the equilibrium positions of classical electrons on the
surface of a sphere, subject to their electrical repulsion. This
problem has been solved exactly for some numbers of particles,
where the equilibrium positions are the vertices of classic pla-
tonic solids and the distance between particles is constant.
However, these perfectly symmetrical congurations limit sig-
nicantly the number of streamers we can use, so we use quasi
symmetrical approximate solutions found numerically. In this
work, we use the database of solutions computed by [47],which
we apply to the distribution of streamers on the electrode sur-
face. We will use congurations with n=50, 100, 200, 400
streamer inception points, shown in gure 3.
Figure 2. Flowchart of the corona model. Calculations in the electrodynamic step accelerated with CUDA are marked in green, while Fortran
accelerated calculations in the chemical system are marked in pink.
5
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
4. Results
In this section we describe the results obtained in a cong-
uration which has similarities to laboratory conditions,
although the need of symmetry of our model makes it
impossible to recreate exact experimental congurations. All
the initial conditions and simulation parameters are shown in
table 1: we consider a positive streamer corona emerging
from a spherical electrode that is live with a potential of
50 kV. The electrode radius is 2 cm, resulting in mean
streamer separations at the electrodes surface varying from
about 3.5 mm to about one centimeter, as reected in table 2.
Since our model does not account for streamer inception, we
have to initiate the streamers already with a nite length,
which we chose as 1 cm.
4.1. Propagation of a single streamer compared to a streamer
inside a corona
Before describing the outcome from the many-streamers
model let us briey summarize the results obtained with our
model for single streamers [42].Ingure 4we plot the
electric eld created by a single streamer advancing from an
electrode with the initial conditions from table 1. The plot
shows the electric eld against the streamer length for all
recorded times, where time is indicated by a color scale. The
initial eld of the streamer is that of the electrode, and it is
screened in the interior of the streamer channel. The eld at
the tip increases with time, and when it is sufciently high,
the tip of the streamer starts advancing. The propagation,
in turn, softens both the screening in the streamer channel and
the increase of the eld at the tip. Increasing the potential of
the electrode, the leading ionization, or the radius of the
streamer channel will lead to faster propagation.
The right panel of gure 4shows the evolution of a
symmetrical corona with 200 streamers with the same initial
conditions (see table 1)as the single streamer model. Quali-
tatively, the general evolution of the streamer is similar
regardless of the number of streamers: screening of the eld
in the streamer channel and increase of the eld at the tip,
which triggers propagation. Quantitatively, however, the
effect of the additional streamers in the corona stands out
when we compare both panels from gure 4: the additional
streamers contribute to a better screening of the electric eld
and lead to a lower electric eld at the tip. The corona with
200 streamers, in gure 4, shows a eld inside the channel
approximately an order of magnitude lower than in the single
streamer case. The eld at the tip is also lower for 200
streamers although the differences are less than a factor 2.
These differences are enough, however, for the total propa-
gation distance to be substantially smaller, around a factor 6.
4.2. Dependence on the streamer density
Now that the general evolution of the streamer corona has
been described let us generalize our results to a varying
density of streamers. Figure 5shows the status of simulations
with different amounts of streamers at nal time, =t10 ns.
The top left panel of gure 5shows the electric eld in
the streamer channel for coronas with 1, 50, 100, 200, and
400 streamers, each shown in a different color. In all the
simulations, the eld shows a similar prole to that of a single
streamer, which is to be expected given that due to the
symmetry in our corona, all the streamers in the simulation
Figure 3. Congurations of n=50, 100, 200 and 400 streamers in a
spherical electrode following a Thompson distribution.
Table 1. Simulation parameters for the positive corona emerging
from a spherical electrode.
Variable Value Units
Electrode radius, a2cm
Electrode potential, V50 kV
Electron density at the tip, ne
010
20
electrons m
3
Streamer radius, R
0
1mm
Streamer initial length 1 cm
Background chemistry humid air
Pressure atmospheric
Final time 10 ns
Table 2. Density of streamers at the electrode and approximate mean
distance between them when considering a spherical electrode of
radius =a2c
m
. The density is d=n/4πa
2
and the mean distance
is d
1/2
.
Number of streamers, nSurface density Average separation
50 -
10 m
4
2
10 mm
100 ´-
2
10 m
42 7mm
200 ´-
4
10 m
42 5mm
400 ´-
810m
42 3.5 mm
6
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
behave as streamer S
0
(as explained in section 2). Even with
these assumptions, the effect of additional streamers in the
corona is clear: the electric eld decreases as the number of
streamer increases, both the peak eld and the eld along the
streamer channel. Since the streamer velocity increases with a
higher eld at the tip [43], as the number of streamer
increases, the propagation velocity decreases. This is sum-
marized in gure 6, where we plot the mean velocity of
coronas with varying number of streamers.
A special case is the simulation with n=400 streamers.
In this case the streamers barely propagate at all within 10 ns
because the electric eld at the tip is too low for sustained
streamer advance. What this case tells us is that there is limit
for the density of streamers in a corona imposed by electro-
static considerations. In our case the maximum number of
streamers was about 400 but this quantity depends on the
electrode potential and radius. Likely, the state in which a
perfectly symmetrical distribution of many streamers stops is
physically unstable and, due to small differences in the length
of the streamers, an actual corona would continue its propa-
gation with a lower density as a result of leaving behind the
shortest streamers.
Turning back to gure 5, its upper right panel shows the
charge density in the streamer channel against streamer length
at nal time for the same simulations with the same code
color as the upper left panel. For simulations with n<400,
the charge density increases with z, but the peak is reached
within the streamer body, slighly before the streamer tip,
where it decreases as it is invested in streamer propagation.
The charge density in the simulation with 400 streamers
decreases near the electrode and then steadily increases with
z, reaching the peak at z
tip
. In general, the charge density is
lower as we increase the number of streamers, both the peak
and near the electrode. However, the positive slope of the
charge density increases with n.
The lower left panel from gure 5shows the electron
density in the streamer channel against z. For the simulations
with signicant propagation, n<400, the electron density
increases with z, and the peak is reached at the tip, where the
leading electron density is =-
n
10 m
020 3 as per our bound-
ary condition. Both the electron density near the electrode and
its slope decrease with increasing n.
The observations detailed in this section lead us to the
rst and main conclusion of this work: the higher the density
of streamers in a corona, the slower is their propagation and
the lower is their internal eld. To understand this principle
let us consider two congurations with different streamer
densities d
1
and d
2
with d
1
<d
2
. Suppose that at a given time
the single-streamer charge density is the same in both con-
gurations. In the conguration with d
2
the electric eld is
more screened due to the neighboring streamers. Since this
implies a lower current, after some propagation it leads to a
lower charge density. Therefore in general the charge density
per streamer is lower in dense coronas. Since the electric eld
at the tip of each streamer results mostly from this charge
Figure 4. Time evolution of the electric eld in the streamer channel against z. The left panel shows the results for the simulation with a single
streamer and the right panel shows the results for a simulation of a corona with 200 streamers distributed as explained in the text. Different
colors show the electric eld at different times, according to the colorbar.
7
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
density, the propagation of dense coronas is also slower. This
process is combined with another principle described in [42]:
slower streamers have lower internal electric elds. These
lower electric elds in our case explain the depletion of
electrons observed in the lower panel of gure 5: this is due to
three-body attachment, which is more efcient for lower
elds.
4.3. Electric field screening
Let us now focus on the evolution at a xed point in the
streamer channel. Figure 7shows the electric eld (top
panels), and the electron density (bottom panels)for two xed
points in the streamer. We chose =
z
2c
m
(left panels)and
=
z
3c
m
(right panels), which are the two extremes of the
streamer at initial time. From a qualitative standpoint, all the
points in the streamer channel have an evolution similar to
that of the electrode boundary, and all the points beyond the
initial length of the streamer that are reached by it will behave
in a similar fashion to the initial tip.
The top panels of gure 7show the electric eld at the
electrode boundary =
z
2c
m
(top left panel), and at the initial
tip =
z
3c
m
(top right panel). At the electrode boundary, in
coronas with any number of streamers, the electric eld
initially decays at a rate that is faster for higher streamer
densities. For n<100 the eld remains quasi stationary at the
minimum value, but for denser coronas the eld slightly
increases with time, with a steeper slope for larger amounts of
streamers. At =
z
3c
m
, which initially is the tip of the
streamer, the eld increases until a value close to -
10 V m
71
is
reached, and then the streamer expansion starts and the eld
decreases. The evolution of the channel eld, then, becomes
similar to that of an inner point. The curves for the corona
with 400 streamers are jagged due to its marginal
Figure 5. Electric eld (top left panel), charge density (top right panel), and electron density (bottom left panel)against length at nal time
(=
t
10 ns). Each color in each plot represents the simulation of a corona with a different number of streamers, as shown in the legend in the
bottom right panel. Note that the maximum electron density (indicated here by a dashed line)is an input parameter of our model.
Figure 6. Streamer mean propagation velocity against number of
streamers.
8
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
propagation; as explained above this evolution is unstable and
not to be expected in actual coronas.
In the upper-left panel of gure 7we appreciate two
clearly differentiated regimes for the streamer base: (a)a fast,
roughly exponential, decay lasting for about 1 ns and (b)an
approximately constant eld afterwards.
The fast exponential decay (a)can be characterized by
tting the electric eld decay to () ( ) ( )t=-
E
tE t0exp
zz ,
where the parameter τrepresents the characteristic screening
time. We applied this t to times <<t
0
0.6 ns and the
resulting parameters τare plotted in gure 8. For comparison
that gure also shows the dielectric relaxation time (some-
times also called Maxwell relaxation time)computed as
τ
M
=ò
0
/σ, where ò
0
is the vacuum permitivity and σthe
electrical conductivity. Figure 8shows the Maxwell relaxa-
tion time using two approaches: with the inner channel con-
ductivity as a red line, and with the average corona
conductivity as a green line. For the average corona con-
ductivity we have weighted the channel conductivity with the
fraction of the surface covered by streamers,
n
Ra4
0
22, where
=
R
1m
m
0is maximum the radius of each streamer and
=
a
2c
m
is the radius of the electrode. Figure 8shows that
the dielectric relaxation time is a reasonably good approx-
imation for this transitory regime, although it shows sig-
nicant deviations for large streamer densities.
Regime (b), which is more representative for the typical
propagation of a streamer corona, sets in once the interior
eld is mostly screened. In that case there is a strong anti-
correlation between high elds and high conductivities in the
corona volume and therefore the dielectric relaxation
Figure 7. Electric eld (top panels), and electron density (bottom panels)against time at xed z. The left plots show the variables at
=
z
2c
m
, the electrode boundary, while the right plots show the results for the initial tip of the streamer, =
z
3c
m
. Each color in each plot
represents a simulation with a different number of streamers, while the rest of the initial parameters are those in table 1.
Figure 8. Characteristic decay times at the electrode boundary of the
electric eld (dots), and Maxwell relaxation time calculated from the
conductivity, for the channel value in each streamer (red), and for the
corona average (green)against number of streamers.
9
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
calculated with a mean conductivity is a poor estimate for the
evolution of the electric eld. A careful investigation of this
regime is beyond the scope of this paper and we leave it for a
future work.
Turning back to gure 7, its lower two panels show the
temporal evolution of the electron density at the two points
that we considered. In both cases, the electron density
decreases with time, although the decrease is larger when we
approach the electrode. At any point in the streamer the
depletion of electrons is faster as we increase the number of
streamers in the corona. As explained above, this is due to the
faster rate of three-body electron attachment at lower electric
elds.
5. Potential drop within a streamer corona
As an impulsive corona propagates away from an electrode, it
carries away part of the electrode potential. How efciently
this is performed depends on the strength of eld-screening
inside each of the streamers and, as we discussed above, this
is heavily inuenced by the streamer density in the corona. In
this section we investigate this process.
The main motivation for this part of our study is the
acceleration of electrons ahead of a lightning leader. This has
been proposed as a mechanism for the generation of
Terrestrial Gamma-ray Flashes (TGFs)[48,49], which are
intense bursts of energetic radiation connected to intra-cloud
lightning processes [50,51]. According to the thermal-run-
away model of TGFs [52,53], the high electric eld at the tip
of streamers pushes electrons from the bulk of the energy
distribution into a runaway regime where they accelerate
further and create relativistic-runaway electron avalanches
(RREA)[54]in the electric eld generated by a leader. A key
magnitude in this process is the total potential available for
the acceleration of electrons from the streamer tips to a
position far-away from the leader [55].
Let us rst consider the value of this potential in the
simulations described above, where the electrode is at a
potential =
V
50 kV
0. The potential at the streamer tips is
V
1
=V
0
ΔV, where the potential drop ΔVis
() ( )
ò
D=VEzzd, 19
a
ztip
with abeing the electrode radius, and E(z)the electric eld at
a distance zfrom the center of the electrode.
The left panel from gure 9shows the evolution of the
potential drop at the corona boundary, calculated using (19).
The right panel shows the drop at z
tip
at =t10 ns against the
number of streamers. As in section 4.3, we appreciate two
regimes in the evolution of the electric screening: a fast decay
from the background eld where the corona acts as an aver-
age air conductivity followed by a proper corona where the
Figure 9. Potential drop (V)between the electrode and the streamer tips. The left panel shows the evolution with time of the potential drop at
the corona boundary, where colors indicate simulations with different n. The right panel shows the potential drop at the corona boundary and
at nal time =
t
10 ns against the number of streamers.
10
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
potential drop increases with time. The key result here is that
even with 50 streamers (an average separation between
streamers of 1 cm at the electrode)the potential drop is less
than 10% of the total potential. A higher streamer density
implies even smaller potential drops. This means that almost
the full potential of the electrode is transferred to the
boundary of the corona.
To extend to our results to a situation closer to a lightning
leader we run a simulation with the parameters listed in table 3.
As our purpose is only to obtain a qualitative understanding of
the corona dynamics around the leader, we mimicked the leader
tip as a spherical electrode with radius 2 cm on which we apply a
potential () ( ( ))=--
V
tV tt1exp
0rise
with =
V
1M
V
0and
=t100 ns
rise . This dependence is an approximation to the rise
of the potential at the leader tip after a leader step. Given the
higher potential, we consider that the radius of the streamer is in
this case 5 mm. We consider here n=45 streamers, resulting in
a density »´ -
d
910m
32
at the leader surface. These para-
meters approximate the characteristics of a streamer burst around
a laboratory leader (see e.g. [56]).
Figure 10 shows the results of the simulation. The left
panel shows the electric eld against zwhereas the right panel
shows the potential drop against time. When we compare the
electric eld with the conguration shown in gure 4, dif-
ferences are obvious due mainly to the higher potential and
the nite rise time. Initially, the channel eld in the streamer
from gure 10 is low, the initial potential being too low. Soon
afterwards propagation starts and, due to the continuing leader
potential increase, despite the screening inside the channel,
the eld at the leader boundary is signicative. The eld at the
streamer tip remains below ´-
2
10 V m
71
due to the velocity
of the streamer, which averages -
7.6 10 m s
61
.
The right panel of gure 10 shows the potential drop at
the corona boundary against time. Due to the increase in the
leader potential, the streamers do not screen the eld ef-
ciently and at the end of the simulation, =t100 ns, when the
leader potential is »
V
0.6 M
V
, around half of that potential
is spent within the streamer corona. An electron starting at the
Table 3. Initial conditions for the leader conguration.
Variable Value Units
Number of streamers 45
Leader tip radius 2 cm
Potential rise time, t
rise
100 ns
Electrode (leader)potential, V
0
1mV
Electron density at the tip, ne
010
20
electrons m
3
Streamer radius, R
0
5mm
Streamer initial length 1 cm
Background chemistry humid air
Pressure Atmospheric
Final time 50 ns
Figure 10. Left panel: Electric eld against zcolor-coded for all times. Right panel: Potential drop (ΔV)at the corona boundary against time.
This simulation was done with the initial conditions listed in table 3, simulating a corona around a leader tip.
11
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
leader tip posseses around half of the leader potential avail-
able for acceleration.
With a nite potential rise time the electric eld inside
the streamers is established by the competion between
screening and the increasing electrode potential. On the right
panel of gure 10 we also show the potential drop of
a simulation with a faster rise time =t50 ns
rise , which is
signicantly lower than that for =t100 ns
rise . A proper
comparison, however, has to take into account the different
values of the streamer length and applied potential at a given
time. In our simulations we found that the average inner eld
for =t100 ns
rise stabilizes around ´
4
.5 10
5
Vm
1
whereas
for =t50 ns
rise it reaches ´
3
10
5
Vm
1
. This shows the
relevance of the potential rise-time in the dynamics of strea-
mers emerging from an electrode or a leader tip.
6. Conclusions
In this work we developed a macroscopical model for strea-
mer coronas where, to allow for the simulation of many
streamers, we simplied the corona as a symmetric cong-
uration of straight streamer channels. Nevertheless, the model
provides intuition about the effect of additional streamers on
the overall evolution of the corona. It also provides semi-
quantitative estimates of macroscopical corona properties. By
means of this model we reached the following conclusions:
As we increase the density of streamers in the corona, the
screening of the eld also increases. The electric eld at
the tip of the streamers decreases, which leads to slower
streamer propagation for large numbers of streamers.
Some macroscopical corona properties such as the
timescale of electric eld screening exhibit a clear
collective behavior. These properties cannot be under-
stood from the dynamics of a single streamer but rather
derive from the interactions between many of them.
The potential drop in a streamer corona is smaller for
streamer coronas with larger amounts of streamers. To
estimate the available potential for acceleration of
electrons ahead of a leader corona we need to consider
the rise time of the leader potential.
Further work on the modeling of streamer coronas should aim
at removing some of the strongest simplications in the
present model. One particularly desirable objective would be
to remove the unrealistical symmetry between all streamers
and allowing different propagation speeds.
Acknowledgments
This work was supported by the European Research Council
(ERC)under the European Union H2020 program/ERC grant
agreement 681257 and by the Spanish Ministry of Science
and Innovation under projects FIS2014-61774-EXP and
ESP2017-86263-C4-4-R. This project has also received
funding from the European Unionʼs Horizon 2020 research
and innovation program under the Marie Skłodowska-Curie
grant agreement SAINT 722337. The authors acknowledge
nancial support from the State Agency for Research of the
Spanish MCIU through the Center of Excellence Severo
Ochoaaward for the Instituto de Astrofísica de Andalucía
(SEV-2017-0709).
ORCID iDs
A Luque https://orcid.org/0000-0002-7922-8627
References
[1]Goldman M and Goldman A 1978 Corona Discharges
Gaseous Electronics ed M N Hirsh and H J Oskam (New
York: Academic)pp 21990
[2]Chang J S, Lawless P A and Yamamoto T 1991 IEEE Trans.
Plasma Sci. 19 1152
[3]van Veldhuizen E M and Rutgers W R 2002 J. Phys. D: Appl.
Phys. 35 2169
[4]Ebert U, Montijn C, Briels T M P, Hundsdorfer W,
Meulenbroek B, Rocco A and van Veldhuizen E M 2006
Plasma Sour. Sci. Technol. 15 S118
[5]Kochkin P, Lehtinen N, van Deursen A P J and Østgaard N
2016 J. Phys. D: Appl. Phys. 49 425203
[6]Franz R C, Nemzek R J and Winckler J R 1990 Science 249 48
[7]Ebert U, Nijdam S, Li C, Luque A, Briels T and
van Veldhuizen E 2010 J. Geophys. Res. (Space Phys.)115
A00E43
[8]Wescott E M, Sentman D, Osborne D, Hampton D and
Heavner M 1995 Geophys. Res. Lett. 22 1209
[9]Gordillo-Vázquez F J and Donkó Z 2009 Plasma Sour. Sci.
Technol. 18 034021
[10]Raizer Y P, Milikh G M and Shneider M N 2010 J. Geophys.
Res. (Space Phys.)115 A00E42
[11]Popov N A, Shneider M N and Milikh G M 2016 J. Atm. Sol.-
Terr. Phys. 147 121
[12]Les Renardières group 1978 Elektra 74 67216
[13]Dwyer J R and Uman M A 2014 Phys. Rep. 534 147
[14]van Veldhuizen E M 2000 Electrical Discharges for
Environmental Purposes: Fundamentals and Applications
(New York: Nova Science)
[15]Lee C, Kim J and Yoon J 2011 Chemosphere 82 1135
[16]Chanrion O and Neubert T 2010 J. Geophys. Res. (Space
Phys.)115 A00E32
[17]Sun A, Teunissen J and Ebert U 2014 J. Phys. D: Appl. Phys.
47 445205
[18]Köhn C, Chanrion O and Neubert T 2017 Geophys. Res. Lett.
44 2604
[19]Dhali S K and Williams P F 1985 Phys. Rev. A31 1219
[20]Kulikovsky A A 1995 J. Phys. D: Appl. Phys. 28 2483
[21]Pancheshnyi S, Ségur P, Capeillère J and Bourdon A 2008
J. Comput. Phys. 227 6574
[22]Luque A and Ebert U 2012 J. Comput. Phys. 231 904
[23]Ihaddadene M A and Celestin S 2015 Geophys. Res. Lett.
42 5644
[24]Lehtinen N G and Østgaard N 2018 J. Geophys. Res. (Atmos.)
123 6935
[25]Marskar R 2019 Plasma Res. Exp. 1015011
[26]Li C, Ebert U and Hundsdorfer W 2012 J. Comput. Phys.
231 1020
[27]Eichwald O, Bensaad H, Ducasse O and YousM 2012
J. Phys. D: Appl. Phys. 45 385203
[28]Luque A and Ebert U 2011 Phys. Rev. E84 046411
12
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
[29]Babich L and Bochkov E 2017 J. Phys. D: Appl. Phys. 50
455202
[30]Luque A 2017 J. Geophys. Res. (Atmos.)122 10
[31]Babaeva N Y and Kushner M J 2009 Plasma Sour. Sci.
Technol. 18 035009
[32]Luque A and Gordillo-Vázquez F J 2011 Geophys. Res. Lett.
38 L04808
[33]Luque A, Stenbaek-Nielsen H C, McHarg M G and
Haaland R K 2016 J. Geophys. Res. (Space Phys.)121 2431
[34]Malagón-Romero A and Luque A 2019 Geophys. Res. Lett.
46 4029
[35]Naidis G V 1996 J. Phys. D: Appl. Phys. 29 779
[36]Luque A, Ebert U and Hundsdorfer W 2008 Phys. Rev. Lett.
101 075005
[37]Bonaventura Z, Duarte M, Bourdon A and Massot M 2012
Plasma Sour. Sci. Technol. 21 052001
[38]Shi F, Liu N and Dwyer J R 2017 J. Geophys. Res. (Atmos.)
122 10
[39]Teunissen J and Ebert U 2017 J. Phys. D: Appl. Phys. 50
474001
[40]Akyuz M, Larsson A, Cooray V and Strandberg G 2003
J. Electrost. 59 11541
[41]Luque A and Ebert U 2014 New J. Phys. 16 013039
[42]Luque A, González M and Gordillo-Vázquez F J 2017 Plasma
Sour. Sci. Technol. 26 125006
[43]Naidis G V 2009 Phys. Rev. E79 057401
[44]Whyte L L 1952 Am. Math. Mon. 59 60611
[45]Smale S 1998 Math. Intell. 20 715
[46]Ashby N and Brittin W E 1986 Am. J. Phys. 54 776
[47]Wales D J and Ulker S 2006 Phys. Rev. B74 212101
[48]Fishman G J et al 1994 Science 264 1313
[49]Dwyer J R, Smith D M and Cummer S A 2012 Space Sci. Rev.
173 133
[50]Lu G, Blakeslee R J, Li J, Smith D M, Shao X M,
McCaul E W, Buechler D E, Christian H J, Hall J M and
Cummer S A 2010 Geophys. Res. Lett. 37 L11806
[51]Cummer S A, Lyu F, Briggs M S, Fitzpatrick G,
Roberts O J and Dwyer J R 2015 Geophys. Res. Lett.
42 7792
[52]Moss G D, Pasko V P, Liu N and Veronis G 2006 J. Geophys.
Res. (Space Phys.)111 A02307
[53]Celestin S, Xu W and Pasko V P 2012 J. Geophys. Res. (Space
Phys.)117 A05315
[54]Gurevich A V, Milikh G M and Roussel-Dupre R 1992 Phys.
Lett. A165 463
[55]Skeltved A B, Østgaard N, Mezentsev A, Lehtinen N and
Carlson B 2017 J. Geophys. Res. (Atmos.)122 8120
[56]Kostinskiy A Y, Syssoev V S, Bogatov N A, Mareev E A,
Andreev M G, Bulatov M U, Sukharevsky D I and
Rakov V A 2018 J. Geophys. Res. (Atmos.)123 536075
13
Plasma Sources Sci. Technol. 28 (2019)115007 M González et al
... One natural question to ask about these phenomena is whether streamer interaction has to be considered and, in that case, what is its influence on the global dynamics of the streamer system. The collective dynamics of streamer fronts have been investigated previously with models that treated streamers as one-dimensional advancing conductors [18,19] but a detailed, microscopical model of a streamer front requires expensive, three-dimensional computations that only recently have become within reach [20][21][22]. ...
... After 30 ns the simulations with five seeds have propagated about 8 mm less than the single-streamer simulations, which amounts to about 15% lower average speed. That a higher density of streamers leads to stronger screening and thus slows down the streamer propagation was already observed in a simplified model of a streamer corona around a spherical electrode [19]. Note however that, as we discussed above, the electric field in the longest streamer is enhanced by the presence of surrounding streamers. ...
... These points are consistent with previous results of macroscopic models such as [19]. ...
Article
Full-text available
We explore the dynamics of dense streamer channel fronts. We introduce a novel, fully three-dimensional, adaptive mesh refinement (AMR) streamer simulation code, which leverages the power of general-purpose graphical processing units (GPUs) to accelerate computations. Our code enables the simulation of systems comprising several parallel-propagating streamers, using appropriate boundary conditions to emulate an infinitely extended front of positive streamers in ambient air. Our findings reveal that denser streamer packings result in slower front propagation and increased electric field screening within the streamers. To interpret these results and progress towards developing a coarse-grained corona model, we present a streamlined model that effectively approximates the behavior of the comprehensive microscopic system.
... This model too was forced to neglect solutions of the full particle fluid equations, but did incorporate a finite streamer conductivity to create more realistic electric fields. An interesting model of many streamers simultaneously interacting was created by González et al [42]. In that work, the authors model hundreds of streamers symmetrically emerging from a spherical anode and interacting electrostatically using a simplified model. ...
... Such a model should be as simple as possible, to allow extension to many streamer ensembles, while also being accurate enough to give practically useful results. Indeed, the authors of that paper have pursued the same goal in recent work [32,42], but the problem is far from solved. The solution explored in this work is a compromise between the high fidelity models compared in [39] and the overly simplified single streamer constructions of [40,41]. ...
... One particularly interesting aspect of both the macroscopic parameter model and the 1.5D quasi-steady model is that the streamer velocity generally decreases as the background field decreases. This makes sense from the current conservation argument, and also agrees with the conclusion found by González et al [42] who noticed a correlation between the channel field and propagation velocity. ...
Article
Full-text available
Single streamers are currently well simulated using detailed computational models. Most of these models are inhibitively complex to use for modelling many-streamer interactions in a streamer corona. This work develops reduced order models of single positive streamers in atmospheric pressure air that replicate the core macroscopic behaviour of detailed models while using a simpler physics representation. Models are developed using the 1.5D framework, with emphasis placed on solving the equations of motion in the streamer reference frame. The solution in this quasi-steady frame is shown to be a good representation of the instantaneous state of the streamer. Finally, a method of uniquely characterizing the instantaneous state of a streamer using its macroscopic parameters (velocity, radius, tip electric field and channel electric field) is developed. This characterization is interpreted graphically, with streamers treated as quasi-steady structures which evolve in time at a rate much slower than the time scale of electron transport. Previous work in the literature is shown to be well captured by this interpretation.
... Another type of corona discharge extensively studied experimentally and numerically by researchers is streamers, which form on a sufficiently sharp electrode when an appropriate impulsive potential is applied [28]. In a study by Cruccolini et al. [29], a sinusoidal radio-frequency corona discharge in dry air was simulated using a multidimensional fluid model that considered 18 species and 64 chemical reactions in the first 10 pulses. ...
Article
Full-text available
The reactive oxygen and nitrogen species generated by plasma have demonstrated consequential effects on diverse commercial applications. Hence, studying the chemistry and spatial distribution of reactive species in plasma is imperative for understanding the influence of plasma in various applications. This study aims to systematically explore the plasma chemistry of a pin-to-plate negative direct current (DC) corona discharge in dry air, using simulations based on a two dimensional (2D) axisymmetric fluid model. The model encompasses a comprehensive set of chemical reactions involving 33 biomedically active species (ROS and RNS). This study entails a rigorous evaluation of the 2D spatial distribution of all chemical species, detailing their minimum and maximum values, at a needle voltage of −10 kV. To enhance visualization and enable comparisons, we integrate contour lines into the density distributions to indicate the average density of each species. N2(A3){\text{N}}_{2}\left({\text{A}}^{3}\sum\right) among nitrogen species, O3 and O2(a1Δ){\text{O}}_{2}\left({\text{a}}^{1}\Delta\right) among oxygen species, and N2O among NOx species exhibit the highest average density in the simulation domain. Furthermore, key reactions involved in the production and consumption of each species are thoroughly discussed. Additionally, the research examines the influence of needle voltage, ranging from −5 to −12.5 kV, on the peak and average densities of all species investigated. Lastly, to validate the simulation model, an experimental study of the pin-to-plate negative DC corona discharge is conducted, during which the voltage-current characteristics and optical emission spectrometry (OES) profiles are measured. The simulation results are in good agreement with the experimental data.
... Given the difficulties in simulating even just an isolated streamer filament [38], it is easy to understand why such claims are made. In a recent review Nijdam et al. [1] remarked that although fluid and particle simulations of streamers with tens or hundreds of branches are computationally unfeasible, reduced-order models of single filaments [36,39] are candidates for improved tree and fractal based models [40][41][42][43]. Although these models have been heralded for quite some time and could be used for simulating discharge trees, they are still in their infancy and they are unfortunately also excessively simplified. ...
Preprint
Full-text available
This paper contains the foundation for a new Particle-In-Cell model for gas discharges, based on Ito diffusion and Kinetic Monte Carlo (KMC). In the new model the electrons are described with a microscopic drift-diffusion model rather than a macroscopic one. We discuss the connection of the Ito-KMC model to the equations of fluctuating hydrodynamics and the advection-diffusion-reaction equation which is conventionally used for simulating streamer discharges. The new model is coupled to a particle description of photoionization, providing a non-kinetic all-particle method with several attractive properties, such as: 1) Taking the same input as a fluid model, e.g. mobility coefficients, diffusion coefficients, and reaction rates. 2) Guaranteed non-negative densities. 3) Intrinsic support for reactive and diffusive fluctuations. 4) Exceptional stability properties. The model is implemented as a particle-mesh model on cut-cell grids with Cartesian adaptive mesh refinement. Positive streamer discharges in atmospheric air are considered as the primary application example, and we demonstrate that we can self-consistently simulate large discharge trees.
Article
The most straightforward way to describe the gas electrical discharge phenomenon is a set of partial differential equations coupled with the transport equation and Poisson equation, but the analytical solution cannot be obtained because of the problem’s multivariable coupling and nonlinearity nature. In this article, this complex fluid model is transformed into a streamer head development model in a plasma background with an infinite volume. Based on the “coefficients decomposition and processes reorganization method”, a spatial evolution equation of the electron density of the streamer head is established. The other physical quantities of the streamer head, including average positive ion density, average net charge density, space charge electric field strength, and streamer velocity, also satisfy the same equation and maintain a consistent variation pattern. More importantly, this equation also has a closed-form analytical solution, even if it contains a variety of nonlinear processes such as collision ionization, photoionization, recombination, and diffusion, which can not only be used to study the type and dynamic characteristics of streamer discharge, but also provide theoretical guidance for numerical simulation and engineering practice.
Article
The interaction of lightning and electrostatic discharge with aircraft has been investigated for almost a century, and even though associated risks are not a concern in aviation safety, thanks to strict guidelines and measures, protection and mitigation are as much a science as an art. The state of the practice heavily relies on historical information, experience, and testing: an approach that has worked extremely well while aircraft variations have been incremental. These methods are becoming questionable when looking at the landscape of novel vehicles designed to achieve net zero carbon emissions, and for urban transport conquering the sky. In this article, it is argued that addressing the needs of these vehicles will require revisiting some of the fundamentals for leader inception, arc reattachment, and electrostatic discharge, and that an approach informed by physics can lead to models for design, as well as new solutions for protection and risk reduction. The article starts with an introduction to the field of lightning protection of aircraft, followed by an overview of the work performed by the author and collaborators. The emphasis is placed on building bridges between the practical application, at the global aircraft scale, and the fundamental physics of electrical breakdown, at the local discharge-point scale.
Article
Full-text available
We investigate the emergence of space stems ahead of negative leaders. These are luminous spots that appear ahead of an advancing leader mediating the leader's stepped propagation. We show that space stems start as regions of locally depleted conductivity that form in the streamers of the corona around the leader. An attachment instability enhances the electric field leading to strongly inhomogeneous, bright and locally warmer regions ahead of the leader that explain the existing observations. Since the attachment instability is only triggered by fields above 10kV/cm and internal electric fields are lower in positive than in negative streamers, our results explain why, although common in negative leaders, space stems and stepping are hardly observed if not absent in positive leaders. Further work is required to fully explain the streamer to leader transition, which requires an electric current persisting for timescales longer than the typical attachment time of electrons, around 100 ns.
Article
Full-text available
We use a three-specie fluid model of electric discharge in air to simulate streamer evolution from the avalanche-to-streamer transition to the collision of opposite-polarity streamers. We estimate the upper limit on the production of thermal runaway electrons, which is dominant during the second of these processes. More thermal runaways are produced if the ionization due to natural background and photoionization is reduced, due to possibility of creation of higher electric fields at streamer tips. The test-particle simulation shows, however, that these thermal runaway electrons have insufficient energies to become relativistic runaways. The simulations are done in constant uniform background fields of E0\mathit{E}_{0} = 4 and 6 MV/m. A simulation was also performed in E0\mathit{E}_{0} = 2 MV/m after formation of streamers in 4-MV/m field, in order to approximate the average background field created by \sim1-MV voltage over a \sim1-m electrode gap used in laboratory spark experiments. We conclude that the used fluid model is insufficient to explain X-ray observations during such experiments. We discuss the possible role of mechanisms which were not included in this or previous modeling but may play the deciding role in the electron acceleration and X-ray production during a streamer collision.
Article
Full-text available
A major obstacle for the understanding of long electrical discharges is the complex dynamics of streamer coronas, formed by many thin conducting filaments. Building macroscopic models for these filaments is one approach to attain a deeper knowledge of the discharge corona. Here, we present a one-dimensional, macroscopic model of a propagating streamer channel with a finite and evolving internal conductivity. We represent the streamer as an advancing finite-conductivity channel with a surface charge density at its boundary. This charge evolves self-consistently due to the electric current that flows through the streamer body and within a thin layer at its surface. We couple this electrodynamic evolution with a field-dependent set of chemical reactions that determine the internal channel conductivity. With this one-dimensional model, we investigate the formation of persisting structures in the wake of a streamer head. In accordance with experimental observations, our model shows that a within a streamer channel some regions are driven towards high fields that can be maintaned for tens of nanoseconds.
Article
Full-text available
We present a full electromagnetic model of streamer propagation where the Maxwell equations are solved self-consistently together with electron transport and reactions including photoionization. We apply this model to the collision of counter-propagating streamers in gaps tens of centimeters wide and with large potential differences of hundreds of kilovolts. Our results show that streamer collisions emit electromagnetic pulses that, at atmospheric pressure, dominate the radio frequency spectrum of an extended corona in the range from about 100 MHz to a few gigahertz. We also investigate the fast penetration, after a collision, of electromagnetic fields into the streamer heads and show that these fields are capable of accelerating electrons up to about 100 keV. By substantiating the link between X-rays and high-frequency radio emissions and by describing a mechanism for the early acceleration of runaway electrons, our results support the hypothesis that streamer collisions are essential precursors of high-energy processes in electric discharges.
Article
Full-text available
The hypothetical mechanism of electric field amplification at contact of positive and negative streamers in a streamer corona up to magnitudes required for the generation of runaway electrons and secondary Bremsstrahlung in the x-ray range, observed in long spark discharges in the open atmosphere, is analyzed. The development of two streamers, moving towards each other in interelectrode gaps of the centimetre range, is numerically simulated at applied voltages from 73 to 250 kV. It is shown that the size of the domain with strong electric field, with intensity sufficient for the thermal electron runaway, is of 1-2 mm. The mean field intensity in this domain increases up to magnitudes of ≈250-280 kV cm⁻¹. The maximum energy, to which electrons are capable of energizing in such field, is in the range of 20-70 keV. However, the electron energy is limited by an extremely small life-time of the strong field domain (less than 20 ps).
Article
Full-text available
We present an open-source plasma fluid code for 2D, cylindrical and 3D simulations of streamer discharges, based on the Afivo framework that features adaptive mesh refinement, geometric multigrid methods for Poisson's equation, and OpenMP parallelism. We describe the numerical implementation of a fluid model of the drift-diffusion-reaction type, combined with the local field approximation. Then we demonstrate its functionality with 3D simulations of long positive streamers in nitrogen in undervolted gaps, using three examples. The first example shows how a stochastic background density affects streamer propagation and branching. The second one focuses on the interaction of a streamer with preionized regions, and the third one investigates the interaction between two streamers. The simulations run on up to 10810^8 grid cells within less than a day. Without mesh refinement, they would require 410124\cdot 10^{12} grid cells.
Article
Full-text available
Several computer models exist to explain the observation of Terrestrial Gamma-ray Flashes (TGFs). Some of these models estimate the electric field ahead of lightning leaders and its effects on electron acceleration and multiplication. In this paper, we derive a new set of constraints to do more realistic modeling. We determine initial conditions based on in-situ measurements of electric field and vertical separation between the main charge layers of thunderclouds. A maximum electric field strength of 50 kV/cm at sea-level is introduced as the upper constraint for the leader electric field. The threshold for electron avalanches to develop of 2.86 kV/cm at sea-level is introduced as the lower value. With these constraints, we determine a region where acceleration and multiplication of electrons occur. The maximum potential difference in this region is found to be ∼52 MV and the corresponding number of avalanche multiplication lengths is ∼3.5. We then quantify the effect of the ambient electric field compared to the leader field at the upper altitude of the negative tip . Finally we argue that only leaders with the highest potential difference between its tips (∼600MV), can be candidates for the production of TGFs. However, with the assumptions we have used, these can not explain the observed maximum energies of at least 40 MeV. Open questions with regard to the temporal development of the streamer zone and its effect on the shape of the electric field remains.
Article
We discuss spatially and temporally adaptive implicit-explicit (IMEX) methods for parallel simulations of three-dimensional fluid streamer discharges in atmospheric air. We examine strategies for advancing the fluid equations and elliptic transport equations (e.g. Poisson) with different time steps, synchronizing them on a global physical time scale which is taken to be proportional to the dielectric relaxation time. The use of a longer time step for the electric field leads to numerical errors that can be diagnosed, and we quantify the conditions where this simplification is valid. Likewise, using a three-term Helmholtz model for radiative transport, the same error diagnostics show that the radiative transport equations do not need to be resolved on time scales finer than the dielectric relaxation time. Elliptic equations are bottlenecks for most streamer simulation codes, and the results presented here potentially provide computational savings. Finally, a computational example of 3D branching streamers in a needle-plane geometry that uses up to 700 million grid cells is presented.
Article
Properties of positive and negative leaders developing in air gaps ranging from 4 to 10 m that were subjected to 100/7,500-μs voltage impulses were examined using a two-frame, high-speed video camera with image enhancement. Abrupt extension (stepping) that culminated in a bright and structured corona streamer burst was observed for both negative (expected for the "classical" stepping process) and positive (expected for the so-called restrike process) leaders. Selected high-quality images of five negative and four positive leaders with pronounced corona streamer bursts are presented here. The morphology of corona streamer bursts was essentially independent of polarity. Streamer bursts exhibiting nearly spherical symmetry were observed. For the four positive leaders, the newly added channel sections (steps) were almost straight and had lengths ranging from about 50 to over 120 cm. For the five negative leaders, most of the steps were curved and their 2-D lengths were some tens of centimeters. It is generally thought that positive leaders in both long sparks and lightning extend continuously or exhibit optically unresolvable steps whose length is comparable to the leader tip size (1 cm or less) and that for sparks only when the absolute humidity is relatively high (>10 g/m³ or so) or voltage rise time is relatively long (around 1 ms or more) can larger steps occur. In this study, both modes of propagation for different branches of the same positive leader were observed.