HAL Id: hal-04212188
https://hal.science/hal-04212188
Submitted on 20 Sep 2023
HAL is a multi-disciplinary open access
archive for the deposit and dissemination of sci-
entic research documents, whether they are pub-
lished or not. The documents may come from
teaching and research institutions in France or
abroad, or from public or private research centers.
L’archive ouverte pluridisciplinaire HAL, est
destinée au dépôt et à la diusion de documents
scientiques de niveau recherche, publiés ou non,
émanant des établissements d’enseignement et de
recherche français ou étrangers, des laboratoires
publics ou privés.
Wave-number Space Networks in Plasma Turbulence
Ö. Gürcan
To cite this version:
Ö. Gürcan. Wave-number Space Networks in Plasma Turbulence. Reviews of Modern Plasma Physics,
2023, 7 (1), pp.20. �10.1007/s41614-023-00122-7�. �hal-04212188�
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma
Turbulence
Ö. D. Gürcan
Laboratoire de Physique des Plasmas, CNRS, Ecole
Polytechnique, Sorbonne Université, Université Paris-Saclay,
Observatoire de Paris, F-91120 Palaiseau, France.
Contributing authors: [email protected]olytechnique.fr;
Abstract
Turbulence commonly described in Fourier space due to its multi-scale
nature can be formulated using wave number space networks where each
node represents a wave-vector on a discretized wave-number space grid
that are connected to one another through triadic interactions denoted
as three body connections. This description that we call wave-number
space network formulation, while being very inefficient for numerical
implementation as compared for example to a pseudo-spectral formu-
lation of the same equations on a regular grid, provides an alternative
perspective and has conceptual advantages, such as the separation of
the equations and the nonlinear interactions. The network represents,
through its connections, the nonlinear interactions, and can be trun-
cated by dropping nodes, or connections corresponding to considering
only certain kinds of wave-numbers or certain kinds of interactions,
without modifying the equations themselves. This guarantees that the
underlying Hamiltonian structure of the equations remains unchanged,
and therefore one has the same conservation laws as the original system.
Wave-number space networks can also be reduced by lumping nodes that
have some similar characteristics together, in which case a reduction of
the equations through some sort of closure becomes necessary, for which
some possibilities are discussed. The network formulation can also be
used for analysing direct numerical simulations, and may be used for
discovering key nodes as well as training models for constructing reduced
systems. The goal of this review is to stimulate interest in thinking in
terms of networks, while dealing with problems in plasma turbulence
1
Springer Nature 2021 L
A
T
E
X template
2 Wave-number Space Networks in Plasma Turbulence
through a survey of what has been done in this subfield and what is pos-
sible for future studies, especially in the context of plasma turbulence.
1 Introduction
Transport of heat, particles and momentum in tokamak plasmas can be caused
by micro-turbulence, and regulated by meso-scale coherent structures such as
zonal flows[1, 2], geodesic acoustic modes (GAMs)[3, 4] or Alfvén eigenmodes
[5, 6] that the turbulence in these devices naturally generates (or regulates),
and their interactions. This makes the detailed understanding of plasma tur-
bulence and its self-regulation through these flow patterns one of the key
academic challenges that face the fusion community. Since the goal of the mag-
netized fusion program is to heat hydrogen ions to high enough temperatures
in order to overcome the Coulomb barrier and instigate fusion reactions, the
core of a magnetized fusion device, is to be extremely hot. On the other hand,
the region where the plasma touches the wall of the device, or the edge, should
be kept at lower temperatures in order to avoid melting the wall or other
plasma facing components. Details of this engineering problem, confounded by
the complexities of heating, operation, and large scale magnetohydrodynamic
(MHD) stability, results in a narrow range of available temperature gradients
between the edge and the core regions, driving a multitude of small scale in-
stabilities, such as the ion and electron temperature gradient driven modes
(ITG [7, 8]and ETG[9, 10]), where the source of the free energy is this back-
ground temperature gradient, trapped electron modes or dissipative drift waves
[11, 12]where the instability is a result of the non-adiabatic electron response
or interchange [13, 14] or resistive balooning modes[15, 16], where the insta-
bility source is the combination of magnetic curvature and pressure gradient
forces.
There is actually a whole zoology of similar modes where the free en-
ergy sources may be the gradients of more exotic quantities (such as parallel
velocity[17], current or resistivity[18]), or any number of combinations of those.
Given their potential importance, both linear and nonlinear physics of these
different instability mechanisms and the resulting “micro-turbulence” have
been studied thoroughly in the past[19, 20], using various approaches, from
linear to quasi-linear theory[21, 22], using analytical methods as well as direct
numerical simulations from simpler fluid systems[23], to gyro-kinetics[24, 25].
Albeit this multiplicity of physical mechanisms for small-scale instabilities
in tokamaks, the variety of parameters that control their behavior, and the
number of fluctuating fields that are involved in each instability, the generally
agreed upon view seems to be that the “plasma turbulence” is nevertheless a
generic notion that is somehow common in all these particular examples[26].
That is, while the instability mechanism that drives the system unstable, and
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 3
the waves that it generates are very different, the nonlinear “mode coupling”
mechanism that is triggered as a result of the interactions of these unstable
waves has a universal aspect[27]. However, it is also clear that plasma turbu-
lence, especially the kind we find in tokamaks, is not exactly the “universal”
in the same sense as the neutral fluid turbulence, which is usually described
using the Kraichnan-Kolmogorov phenomenology of the turbulent cascade[28].
Plasma turbulence is both similar to and different from neutral fluid turbu-
lence and is peculiar in various respects such as the importance of waves and
instabilities and therefore various resonant mechanisms[29, 30] or the fact that
the kinetic system provides a multitude of damping mechanisms [31, 32] and
hence the coexistence of unstable and damped modes[33], which results in a
turbulent “cascade”, without the presence of a clear inertial range.
In its most general formulation, the turbulent “cascade” in a bounded sys-
tem can be thought of as percolation of a triad interaction network with a
conserved quantity, like energy, where the network consists of discretized wave-
numbers[34]. In such a network, where each wave-number is a distinct node,
each node interacts with a set of pairs, with which it satisfies the triadic in-
teraction condition (i.e. k + p + q = 0 where k, p and q are the interacting
wave-numbers). For wave turbulence[35], additional constraints such as reso-
nance (or near resonance) among the frequencies of these wave-number nodes
[e.g . ω (k) + ω (p) ω (q) 0 ] can be invoked36. Due to their triadic na-
ture, the interactions in such a network are three body interactions, and the
resulting network is a three body interaction network. We call such networks,
wave-number space networks.
Consider a standard spectral formulation in a bounded system where the
nonlinear term is computed through convolution sums. One can view the con-
volution, as a sum over the underlying three body network consisting of all
possible combinations of triadic interactions. The introduction of the concept
of the “network” in this case is an equivalent but trivial reformulation of the
convolution sum. The network in such an example can be a regular grid, that
does not change and is made up of a huge number of elements. Treating a
convolution sum on a regular k-space grid as over an extended (but still some-
what regular) three body network where each triadic interaction is handled as
a separate connection, increases the computational complexity of the problem
considerably while introducing no apparent advantage. However when we want
to reduce the system, either dropping inactive nodes, or by lumping together
nodes that play similar roles, the network approach, provides some advantages
as well as an interesting perspective.
Networks appear in many problems in nature, and the discipline that is
devoted to their study is called the network science[37]. As neural-networks
have become extremely popular tools[38] for multivariate multiple regression
in science, policy and technology in recent years, and the study of topol-
ogy, structure and dynamics[39, 40] of biological[41], ecological[42], social [43]
and computer networks[44] has shown regular features in their complex self-
adaptation[45, 46], network science has become a central player in our quest to
Springer Nature 2021 L
A
T
E
X template
4 Wave-number Space Networks in Plasma Turbulence
understanding complex aspects of natural systems[47]. Network science gives
us tools that may provide insight into self organizing principles of these sys-
tems, such as the remarkable self-similarity that turbulent systems commonly
demonstrate. It has been argued recently that turbulence can be formulated
as a percolation on an evolving complex network, and some aspects of its be-
havior including intermittency can be related to generalities shared by other
complex networks, such as food webs, or the internet.
Use of the network abstraction to study nonlinear dynamics of turbulence
also provides interesting prospects[48, 49] especially in the context of plasma
turbulence. It may sometimes be possible to reduce complex networks by lump-
ing together certain similar elements. It is common, for example, to describe
food webs with species that play similar roles lumped together instead of la-
beling each distinct species separately (e.g. “whales” as a single node instead of
every single species of whale as separate nodes). In the context of networks of
wave-numbers, some regions in the wave-number domain may play similar roles
and can be lumped together. For example a description in terms of scales, is
a conceptual example of lumping together the wave-numbers that play similar
roles in the turbulent cascade, relevant to homogeneous, isotropic turbulence.
In the case of strong anisotropy, modes with wave-numbers, which has a
vanishing component (e.g. zonal flows as k
x
= 0 modes in geophysical fluid dy-
namics or k
y
= 0 modes in fusion plasmas) can be considered as an important
conceptual element of the anisotropic energy transfer in k-space[50, 51].
We know that in the study of plasma turbulence, particular meso-scale
structures play special roles in the turbulent self-organization. For example
the interactions between zonal flows and drift-wave turbulence, are commonly
referred to in the fusion community as predator-prey interactions[52], where
the zonal flows play the role of the predator and the underlying drift-wave
turbulence that drives them, play the role of the prey. One can even use the
Lotka-Volterra equation to model this state, and it is actually not unique to
fusion plasmas, or zonal flows, but is a feature of turbulent systems that are
not very far from marginal stability conditions, such as the conditions one
finds in transition to turbulence where only a finite number of modes would be
initially excited[53]. In fact, also in fusion plasmas, one observes these predator-
prey oscillations most clearly in near marginal stability conditions[54], or in
transitions, such as the Low to High confinement transition[55, 56]. The ex-
istence of this well established analogy between a particular state of plasma
turbulence -dominated by zonal flows- and an ecological system, nicely paves
the way towards the extended analogy of plasma turbulence as a network of
predator-prey relations, or a food web as it is called in ecology, and the con-
secutive natural step of abstraction of plasma turbulence in terms of “complex
dynamical networks” of which food web is just an example.
The remainder of the paper is organized as follows. The introduction con-
tinues with a simple, concrete example commonly used in plasma physics to
provide the context for the discussion on more abstract concepts into network
formulation. Section 2 is devoted to primitive wave-number space networks
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 5
(i.e. on regular rectangular grids), where the basic formulation is given in
Section 2.1, and the network concepts of energy transfer among nodes is dis-
cussed through conservation laws in Section 2.2. Section 2.3 discusses the
triadic instability assumption and phase dynamics, and section 2.4 considers
wave turbulence, and section 2.5 examines how this basic formulation can be
extended to multiple fields or kinetic systems, with section 2.6 focusing on
the effects of magnetic geometry, and localization of modes around rational
surfaces. Section 3 gives some examples of truncated network models, with
nested polyhedra models for fluids and MHD discussed in Section 3.1 and spi-
ral chain models for two dimensional turbulence discussed in Section 3.2. In
section 3.3, self-consistent quasi-linear models are discussed from the point of
view of truncated network models. In Section 4, reduction of triadic networks
are considered, with Section 4.1 describing how to deal with energy transfer
in reduced networks and Section 4.2 illustrating closure on reduced networks
through an eddy damped quasi-normal Markovian approximation. Section 5
provides some examples of ad-hoc models, which starts by reinterpreting shell
models as network models in Section 5.1 and then discussing small-world net-
work versions of those and their dynamics in Section 5.2. Section 6 is dedicated
to the use of wave-number space networks in analysis, with a short discussion
of model extraction in Section 6.1. Section 7 is summary and discussions.
1.1 Elementary Example: Modulational Instability
One of the earlier examples of reduced network models in fusion plasmas
were based on what is sometimes called the i-delta equations, as the weakly
non-adiabatic version of the Charney-Hasegawa-Mima system, using “a low
order k-space”, consisting of a basic wave-number space network of 10 or so
modes[57]. An even simpler example is the so-called modulational instability
calculation, which requires at the minimum the most unstable mode, a zonal
mode, and two sidebands, which means a minimum of 4 modes. The usual ex-
ample of a modulational instability calculation involves considering the most
unstable mode as the pump, and looking at the coupled evolution of the zonal
flow and sidebands assuming the energy in these are initially much smaller com-
pared to the pump mode, so that a linear stability analysis can be performed.
Such a linear stability analysis of the coupled zonal-flow/sidebands system in
the presence of the pump can be used to obtain the growth rate of the modu-
lational instability as well as the most unstable k
x
. Of course, one can instead
solve this low order system numerically, however in this case if want the system
to saturate, we need to add a second k
x
mode together with its sidebands (so
a minimum of 7 modes), which will act as the sink. A similar system of inter-
actions between zonal flows and drift waves were also considered in the past,
using balooning formalism and the gyrokinetic equation[58]. More recently, a
general network version similar to these systems was studied in detail for the
Hasegawa-Wakatani model[59]. Here we use the Hasegawa-Wakatani case in
the adiabatic (e.g. in the i-delta limit) as a simple elementary example in order
to provide a “plasma physics” introduction to the topic. Consider:
Springer Nature 2021 L
A
T
E
X template
6 Wave-number Space Networks in Plasma Turbulence
Figure 1 The dynamics and the network structure for the modulational instability exam-
ple. The most unstable mode (i.e. Φ
k0
) acts as a pump, exciting the zonal mode (i.e. Φ
1
)
and the sidebands (i.e. Φ
k1±
). The nodes are shown in a reduced k
x
-k
y
grid on the top
right plot, which also acts as a legend for the main plot. The outer modes that are shown
in grey circles are artificially damped, hence play the role of sink. In the final steady state,
without any large scale friction, the energy in Φ
1
, causes the interaction to effectively turn
off. Introduction of large scale friction would result in predator-prey like oscillations.
t
(1 + χ
k
) Φ
k
= iκk
y
Φ
k
+
1
2
X
M
kpq
Φ
p
Φ
q
(1)
where Φ
k
is the normalized electrostatic potential, χ
k
k
2
k
, where δ
k
is defined through the relationship between the electron density and the elec-
trostatic potential n
k
= (1
k
) Φ
k
, κ is the normalized diamagnetic velocity
and M
kpq
is the nonlinear interaction coefficient whose form is not important
for this example (see Section 6 for more details on definitions and normaliza-
tions). In order to study modulational instability it is common to consider a
subset of modes, such as:
Φ
k
= Φ
k0
+
X
Φ
+ Φ
kℓ+
+ Φ
kℓ
(2)
where Φ
k0
is the most unstable mode (withk
y
= k
y0
and k
x
= 0), Φ
are
the zonal modes (with k
x
= k
xℓ
and k
y
= 0) and Φ
kℓ±
are the sidebands
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 7
(with k
y
= k
y0
and k
x
= ±k
xℓ
) labeled by considering multiple radial wave-
numbers. In general we have 3n + 1 modes where n is the number of k
x
modes
considered. We can of course also add other k
y
modes, but having only a single
k
y
mode is sufficient for studying modulational instability. Note that for each
mode k = k
, its Hermitian conjugate k = k
should also be considered.
In order to have a stationary solution, we at least need to go to n = 2,
and choose the k
x0
and k
y0
to correspond to the most unstable modes of the
modulational instability and the linear instability respectively, and introduce
artificial damping on the equations for Φ
k±2
and Φ
2
. Such a system in the
absence of zonal flow damping (that is damping of the Φ
1
, which is the primary
modulationally unstable zonal component), evolves to a state of finite zonal
flows and stays at that final state with stationary zonal flows. Figure 1 shows
the results of the numerical integration of the low order dynamical system that
one gets from such a system (actually we have used the full Hasegawa-Wakatani
system with C = 10, κ = 1, see Section 6 for details).
2 Primitive Wave-number Space Networks
2.1 Basic Formulation
Consider a complex field ψ
k
, which represents the Fourier transform of a real
one through ψ
k
1
L
R
L
ψ (x) e
ik·x
d
n
x in a bounded domain L, for which we
can write:
t
ψ
k
+
k
ψ
k
=
1
2
X
M
kpq
ψ
p
ψ
q
(3)
where ω
k
is the complex frequency, whose imaginary part may represent the
instability -or dissipation in small scales- and M
kp q
is the symmetrized non-
linear interaction coefficient. This template form may represent a number of
different single field systems, such as for instance the Charney-Hasegawa-Mima
system with ω
k
=
v
k
y
1+k
2
and M
kpq
ˆ
z×p·q
(
q
2
p
2
)
(1+k
2
)
where v
is the normalized
background density gradient, L
x
and L
y
are the box dimensions.
The wave-number on a two dimensional regular square grid can be de-
fined using two integer indices
x
and
y
as k
x
= 2π (
x
N
x
/2) /L
x
and
k
y
= 2π (
y
N
y
/2) /L
y
, where L
x,y
is the size of the domain and N
x,y
are
the number of grid elements in each direction. We can flatten the two index
variables into a single integer using for example the column major order lin-
ear storage formula =
x
+
y
N
x
and use this as the node index. Note that
in the general case of n dimensions we can define the flattened node index via
the usual =
1
+
P
n
i=2
i
Q
i1
j=1
N
j
. This allows us to define the node wave-
number k
ˆ
xk
x
(
x
) +
ˆ
yk
y
(
y
) as a vector and node variable ψ
ψ
k
as a
complex number denoted by the node label . This way, the whole regular grid
of the n-dimensional wave-number domain can be thought of as a collection of
N =
Q
n
j=1
N
j
wave-number nodes.
Springer Nature 2021 L
A
T
E
X template
8 Wave-number Space Networks in Plasma Turbulence
We can thus write (3) as a dynamical equation for ψ
on a three body
network:
t
+
ψ
=
1
2
X
,ℓ
′′
i
M
ℓℓ
′′
ψ
ψ
′′
(4)
where i
denote the list of pairs that interact with the node . The list of pairs
i
needs to be computed by going over all the nodes that satisfy the triad
interaction condition k
+ k
+ k
′′
= 0. Written in this way, the equation (4)
is exactly the same as the Eqn. (3) on a regular k-space grid. However, the
former can be extended to irregular grids, or to a decimated Fourier space
and can be more easily modified to incorporate the dynamics of an equivalent
variable on a reduced network. Note also that the formulation of Eqn. (4) can
also be extended to windowed Fourier transforms, or wavelet coefficients, or
any other similar decomposition such as the Galerkin decomposition, except
that in that case ω
would be replaced by a non-diagonal matrix, representing
an operator and constructing the interaction network (that is i
as connections
and M
ℓℓ
′′
as weights) may be nontrivial.
In practice, the formulation in (4) using a regular grid can be implemented
in two steps. First, for each we find all possible
,
′′
pairs that satisfy the
triad interaction condition k
+ k
+ k
′′
= 0 and compute and record the
interaction coefficients M
ℓℓ
′′
. This constitutes our “network”, in that it is a
list of all three body connections between all the nodes of the network with
particular weights for each connection in the form of interaction coefficients.
Once the network is constructed, it can be stored and the complex dynamical
variables ψ
, corresponding to Fourier coefficients can be advanced on this
network using (4).
Searching for all possible pairs that satisfy the triadic interaction condition
is time consuming, but it can be ameliorated by scanning and
while using
k
± k
± k
′′
= 0 to solve for
′′
, keeping only k
< k
′′
part of the k-space
consisting of only k
y
0 modes (since the initial data is real). Notice that
one can also impose additional resonance conditions in this step in order to
implement weak wave-turbulence as a sparse network on a discrete regular
grid[60, 61].
On a properly constructed regular grid, (4) and (3) are mathematically
equivalent. This means that a numerical implementation of (4) as described
above, and say a pseudo-spectral implementation of (3) with the same forcing,
dissipation and initial and boundary conditions, should give exactly the same
evolution up to numerical precision. This can be verified, for example on a reg-
ular grid of resolution N
x
×N
y
= 256×256, beyond which the network method
starts to be impractical. This implies N
= (N
x
/2) × (N
y
/2) + (N
x
/2 1) ×
(N
y
/2 1) independent nodes, since in a real-to-complex transform, we have
N
y
/2 + 1 independent wave-numbers in the y direction, with the last one be-
ing the Nyquist wave-number and Φ
k
x
,0
= Φ
k
x
,0
on the k
y
= 0 axis due to
Hermitian symmetry. In such a network, the node that has the most connec-
tions is the smallest wave-number node that has N
t
= (N
x
2) × (N
y
2)/2
connections (i.e. triads that are connected to that node). Note also that for
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 9
standard 2D turbulence = 0 is unconnected since |k
| = |k
′′
| makes the in-
teraction coefficient vanish. Since M
ℓℓ
′′
= M
ℓℓ
′′
, by choosing
>
′′
, and
dropping the 1/2 in (4), we can reduce the maximum number of triads to
N
t
= (N
x
2) ×(N
y
2) /4.
Network formulation on a regular rectangular grid is extremely impractical
for any kind of meaningful resolution since its computational cost for a causal
formulation scales with N
3
, which would scale with N
9
(i.e. N
x
= N
y
= N
z
=
N ) for three dimensions, and there exists many efficient techniques for dealing
with turbulence on a regular grid. However since the same approach can be
used on a sparse network obtained from reduction such as the nested polyhedra
models that we will see in Section 3.1, which can describe a very large range
of scales using a relatively small number of nodes even in three dimensions,
they can be extremely powerful for computations as well. We argue that the
effort of writing down the network formulation on a regular rectangular grid is
nonetheless useful for establishing the connection to standard techniques, and
to provide a basis on which we can apply network reduction. For example, if we
know how to go from a regular rectangular grid to a particular reduced network
form, we can apply the same reduction to the data from direct numerical
simulations (DNS) that are usually on a rectangular grid (see section 6 for
some examples).
2.2 Conservation laws
If the nonlinear interaction coefficients in Eqn (4) have the symmetry
σ
M
ℓℓ
′′
+ σ
M
′′
+ σ
′′
M
′′
ℓℓ
= 0 (5)
where σ
is a coefficient that is a function of the node label , (i.e. a function
of the wave-number), the quadratic quantity defined as:
E
σ
total
X
σ
|ψ
|
2
can be shown to be conserved by the nonlinear dynamics, since
t
E
σ
=
X
,ℓ
′′
i
T
σ
ℓℓ
′′
+ P
σ
D
σ
(6)
where P
σ
is the production and D
σ
is the dissipation of the conserved quantity
labeled by σ at the site of node . If we use Eqn. (4), we get:
P
σ
D
σ
= 2γ
E
σ
since both energy injection at instability scales and the dissipation at small
scales come from the form of the “linear growth rate” [i.e. γ
= Im (ω
)] as a
function of k
which actually becomes negative as we go to small scales due to
Springer Nature 2021 L
A
T
E
X template
10 Wave-number Space Networks in Plasma Turbulence
1
2
3
4
123
134
Figure 2 The energy transfers in a network of triadic interactions between nodes 1 4,
connected by two triads 123 and 134. The triad interactions are show in the form of little
triangles, the transfer terms such as t
123
12
denote the energy transfer from node 1 to node 2
through the triad 123, which can also be denoted by t
3
12
. The total energy transfer between
two nodes is the sum of the transfers through each triad, as shown for the case of two triads
here with t
13
t
134
13
t
123
31
.
dissipation. The transfer rate T
ℓℓ
′′
in Eqn. (6) represents the energy transfer
from the nodes
and
′′
to the node , which can be defined explicitly as
T
σ
ℓℓ
′′
Re [σ
M
ℓℓ
′′
Φ
Φ
Φ
′′
] .
Since the interactions always appear as three body interactions, we have
T
σ
ℓℓ
′′
+ T
σ
′′
+ T
σ
′′
ℓℓ
= 0 as implied by Eqn. (5). When we sum the Eqn. (6)
over all nodes, we find that the total amount of conserved quantity (e.g. en-
ergy) increases or decreases only as a result of the difference between its total
injection and its total dissipation.
Dropping the label σ for convenience (e.g. considering energy), we can also
write
t
E
=
X
,ℓ
′′
t
′′
ℓℓ
+ P
D
(7)
where t
′′
ℓℓ
=
1
3
(T
ℓℓ
′′
T
′′
) represents the energy transfer from
to
mediated by
′′
. Note that for a given triad, with the node labels 1 , 2 and 3,
we have T
123
= t
3
12
+t
2
13
= t
3
12
t
2
31
, meaning that the energy transferred from
the nodes 3 and 2 to node 1 is the difference of the energy transferred from 2
to 1 mediated by 3 and the energy transferred from 1 to 3 mediated by 2. (see
figure 2) This allows us to transform the three body interaction network to a
simple network with edges that are weighted by the other components of the
triad, with multiple channels between the nodes.
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 11
We may further reduce the many connections between two nodes mediated
by different third nodes, by summing over the third node as:
t
E
=
X
t
ℓℓ
+ P
D
(8)
where t
ℓℓ
P
′′
t
′′
ℓℓ
. In figure 2 this corresponds to writing t
13
= t
4
13
+ t
2
13
=
t
4
13
t
2
31
= t
134
13
t
123
31
. The two connections in this example, t
134
13
and t
123
31
, are
part of two separate triad interactions, and therefore has two different phases
(see below). Summing over all these different connections belonging to different
triad interactions, having different phases, allows us to transform the system
from a multigraph (a graph which has multiple edges between two nodes) to a
simple graph (a graph with only one edge between two nodes) is an important
reduction of the network topology. However it results in loss of information
since once we sum multiple edges between two nodes into a single edge, there
is no way to get back the different edges that make up that single combined
edge. We also loose detailed information about the three-way relative phases
that determine the direction of the flux through a given triad, since we sum
over many triads in order to obtain the transfer between two nodes mediated
by all possible third nodes.
2.3 Triadic Instability Assumption and Phase dynamics
While the necessary condition for the existence of a link between the nodes in a
three body spectral network, representing turbulent mode coupling, is the triad
interaction condition between the wave vectors, this link accommodating an
actual transfer of energy between the nodes requires additional circumstances.
Given three nodes and a triad, one may usually estimate the direction of
energy transfer, through an analysis called the “instability assumption”[62, 63].
In the fusion context it would probably make more sense to call this “triadic
instability assumption”, since instability in that context rather refers to the
linear instability of the underlying system. In any case, the triadic instability
condition suggests that, if we start with an initial state such that ψ
O (1)
but ψ
ψ
′′
O (ϵ), a linear stability analysis for the perturbations ψ
and
ψ
′′
gives an instability condition:
M
′′
M
′′
ℓℓ
> 0 (9)
which results in the growth of ψ
and ψ
′′
resulting in a transfer of energy.
The growth can then be argued to continue until some kind of equipartition
between the three nodes ,
and
′′
. For example for incompressible two
dimensional turbulence the interactions coefficients are M
kpq
=
ˆ
z×p·q
(
q
2
p
2
)
k
2
[i.e. with ℓ,
,
′′
k, p, q] so that the instability condition M
pqk
M
qkp
> 0
implies
k
2
q
2
p
2
k
2
> 0, which is satisfied only if k is the middle wave-
number. This is a consequence of the intermediate axis theorem for rigid body
rotation and the equivalence of these two systems. More generally, the triadic
Springer Nature 2021 L
A
T
E
X template
12 Wave-number Space Networks in Plasma Turbulence
instability assumption means that the transfer is from the node which has
the interaction coefficient M
ℓℓ
′′
that has the opposite sign to the other two
M
′′
and M
′′
ℓℓ
, which has the same sign because of Eqn. (9) resulting in
T
ℓℓ
′′
< 0 while T
′′
> 0 and T
′′
ℓℓ
> 0. This works even in the presence
of linear growth and damping as long as the pump mode keeps increasing in
amplitude, at some point the nonlinear transfer mechanism will kick in. Note
that while the actual three wave system without any linear instability can
be solved exactly using Jacobi elliptic functions[64], the implications of these
solutions to the triadic instability assumption, where we only consider the
initial trends, which gives us an idea about the direction of the trasfers, until
we reach a stationary state either through statistical equipartition, or through
the non-equilibrium steady state between production and dissipation through
the nonlinear cascade processes.
However as it invokes statistical steady states such as the equipartition,
the analysis is usually based on the assumption that the phases are random
in a turbulent field. This is reasonable as long as the phases of the legs of the
triad that we are considering are not involved in some complicated conspiracy,
like for example all three nodes of the triad staying in a phase-locked state for
an extended period of time. For a system with internal free energy sources as
it is usually the case for plasma turbulence, linear frequencies can provide the
dominant term in the phase evolution of a given wave-number node. It may
be that these frequencies, possibly modified by nonlinear effects such as the
Doppler shift from large scale flows etc., reorganize themselves locally in order
to induce these phase-coherent states, which is in stark contrast to the case of
random phase.
Substituting ψ
= A
e
in (4) and assuming M
ℓℓ
′′
R for simplicity, we
get:
t
A
= γ
A
+
1
2
X
,ℓ
′′
i
M
ℓℓ
′′
A
A
′′
cos (ϕ
+ ϕ
+ ϕ
′′
) (10)
t
ϕ
= ω
rℓ
1
2
X
,ℓ
′′
i
M
ℓℓ
′′
A
A
′′
A
sin (ϕ
+ ϕ
+ ϕ
′′
) (11)
Note that for the more general case of complex M
ℓℓ
′′
the argument of the
sine and cosine in Eqn.(10-11) would be replaced by
ϕ
M
ℓℓ
′′
+ ϕ
+ ϕ
+ ϕ
′′
where ϕ
M
ℓℓ
′′
= arg (M
ℓℓ
′′
). The complete problem of plasma turbulence,
that we call the “primitive network”, that we usually solve in direct numerical
simulations involves Eqn. (4) or equivalently the Eqns. (11) and (10) or (6) on
a network constructed from a regular rectangular grid of wave-number nodes
in Fourier space.
However the point of the network formulation is reduction, and there are
many different ways one can reduce such a system depending on what the dom-
inant processes are and what one wants to describe. For example a blunt way
to do reduction is to directly truncate the Fourier space, so that we have a re-
duced system, that is somehow supposed to represent the full system. If such a
reduction is done using the original equations (i.e. Eqn. 4), it can typically give
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 13
us a sense of what the coupled system does qualitatively, but unless the trun-
cation is done respecting the statistical characteristics of the initial network, it
would modify things like energy equipartition solutions etc. In the same vein,
we can keep the nodes but reduce the links, which corresponds to keeping the
full regular rectangular grid in Fourier space, but only considering a certain
class of interactions (e.g. certain kinds of triads). Self-consistent quasi-linear
theory, commonly used in fusion and geophysical fluid dynamics applications
where one keeps interactions with large scales (zonal flows, and profiles) but
drops the interactions among small-scale fluctuations can be considered as an
example of this.
A different way to reduce the initial primitive network may be to use a
closure scheme, which would allow us to lump different nodes and links to-
gether in groups instead of using truncation or dropping links. Such a lumping
together of the nodes requires a closure that can represent multiple triadic in-
teractions as a single triadic (or two-body) interaction, which in turn requires
handling the statistics of phase relations. We can do this, for example by in-
voking the random phase approximation, which would allow us to use direct
interaction approximation (DIA)[65, 66] or the eddy damped quasi-normal
Markovian approximation (EDQNM)[67, 30], thus resulting in a reduced sys-
tem (say EDQNM equations) on a reduced network when we sum over groups
of nodes.
Another interesting case arises, when the dynamics is dominated by inter-
acting linear waves. In this limit, the resonant interactions between the linear
frequencies ω
result in a phase-locked state, which provides a natural closure
for the system of equations. It would also provide a natural reduction of the
network, since only those modes that also satisfy the resonant interaction con-
dition, i.e. ω
± ω
± ω
′′
= 0 are necessary and we can drop the others. It
can also be argued, without actually invoking wave-turbulence closure, that
the justification for using quasi-linear models in fusion plasmas is in fact the
idea that the system lacks obvious three-wave-resonances without large scale
flows, which can be zonal flows, Geodesic acoustic modes (GAMs), or other
large scale structures, so that the resonant interactions always have to involve
one of these flow structures[68].
2.4 Wave turbulence
Practicality of the network formulation rely on a suitable reduction of either
the nodes or the interactions of its underlying wave-number space network.
In this sense wave-turbulence provides a compelling scenario, since it allows
one to consider only a very small subset of all possible interactions due to the
resonance condition. While, wave turbulence, which describes the evolution of
an ensemble of weakly interacting waves[35, 69], obeying a linear dispersion
relation, is strictly applicable only in the asymptotic limit with linearly stable
waves, at the limit of infinite box size, it still presents a very powerful tool
for understanding the role the resonant or quasi-resonant interactions play in
the turbulent cascade. It can describe multiple statistical quantities using only
Springer Nature 2021 L
A
T
E
X template
14 Wave-number Space Networks in Plasma Turbulence
a conserved quantity called the wave-quanta, commonly denoted by N
k
E
k
k
where E
k
is the energy, which is equivalent to potential enstrophy
in the Charney-Hasegawa-Mima case with proper zonal flow response [apart
from a factor of 1/|k
y
|, which depends on the zonal flow response]. Network
formulation in the case of wave-turbulence is of great interest, and has been
studied in some detail in the past, as one can obviously decouple the geometric
study of the resonant manifold and the evolution of the wave-quanta on the
said manifold resulting in a major conceptual simplification.
Being an asymptotic theory, it is common to make the assumption of in-
finite size in the study of wave-turbulence, which also sidesteps the issue of
whether or not the discrete modes that are available in a finite system actu-
ally satisfy the resonance condition, since in that limit one has a continuous
k-space, hence an infinite number of nearly-resonant modes[69]. The usual
wave-kinetic equation for the Charney-Hasegawa-Mima system can be written
following Connaughton [70] as:
n
k
t
= S [n
k
] + f
k
γ
k
n
k
(12)
where the collision integral has the form:
S [n
k
] = 4π
Z
W
2
kpq
n
p
n
q
σ
ω
k
ω
q
n
p
n
k
σ
ω
k
ω
p
n
q
n
k
× δ
2
(k p q) δ (ω
k
ω
p
ω
q
) d
2
pd
2
q (13)
and σ denote the sign of its subscript, and the W
kpq
denote the nonlinear
interaction coefficient in the wave interaction representation:
W
kpq
=
ˆ
z × p ·q
2
r
|
p
y
q
y
k
y
|
p
2
q
2
(1 + p
2
) (1 + q
2
)
Some interesting observations for the Charney-Hasegawa-Mima system came
out of the study of the wave-kinetic equation, such as the identification of an
additional conservation law dubbed zonostrophy[71, 70]:
Z
k
arctan
k
x
+
3k
y
k
2
!
arctan
k
x
3k
y
k
2
!
2
3
(1 + k
2
)
k
y
In the study of fusion plasmas, on the other hand, it is more common to use
what is sometimes called weak-turbulence-theory (WTT), which is essentially
the same thing as it involves the same underlying assumptions, except that
one keeps the linear instability term, and the resulting kinetic equation is seen
as a Markovian statistical closure in the same vein as the EDQNM. We can
write the general form of the WTT equations somewhat symbolically as [30]:
t
C
k
2γ
lin
k
C
k
+ 2Reη
nl
k
C
k
= 2F
nl
k
(14)
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 15
where C
k
|Φ
k
|
2
,
η
nl
k
X
M
kpq
M
pkq
θ
kpq
(t) C
q
, (15)
and
F
nl
k
1
2
X
|M
kpq
|
2
Re [θ
kpq
(t)] C
p
C
q
(16)
with
θ
kpq
(t) πδ (ω
k
ω
p
ω
q
) .
Note that the first term in (13) corresponds to the incoherent term, given in
WTT by (16), and the last two term in (13) can be combined, by exchanging
p and q in one of the terms into the coherent term given in (15) [i.e. when
multiplied by n
k
as it appears in (14)]. Multiplying C
k
by σ
k
we can write
the equation for a conserved quantity E
σ
k
C
k
σ
k
. The advantage of this
formulation is that we can use it on a reduced network, exactly same way as
we would use any other closure. An example for the EDQNM closure on a
reduced network can be found in Section 4.2.
Wave turbulence on a discrete wave-number space network is also some-
times studied using the original equations (for example the Charney-Hasegawa-
Mima system) directly on a wave-number space network consisting of clusters
of triads[36, 70]. In this case, using the wave interaction representation ap-
pears to be a mere convenience since the equations are the same. However since
the resonance conditions, even including resonance broadening[72], makes the
topology of the k-space network, very sparse since the large majority of in-
teractions are effaced as a result of the resonance condition. This leads to the
creation of clusters of connected triads, that may be isolated or weakly con-
nected to one another, resulting in the blocking of the k-space cascade. Unlike
strong turbulence case, each wavenumber node in a wave-number space net-
work of wave-turbulence is involved in only a few (if any) triads. This is a
manifestation of the fact that the resonance condition, for example written for
the Charney-Hasegawa-Mima case as:
p
y
+ q
y
1 + p
2
+ q
2
+ 2p ·q
p
y
(1 + p
2
)
q
y
(1 + q
2
)
= 0
defines a curve for a given q = k
, and only the points p = k
that lie both
on the discrete k-space grid and on the resonance manifold gets connected
to this node. This makes it possible and somewhat practical to consider the
wave-turbulence as a network of “triads” that are connected by nodes, in an
inverted perspective to the point of view generally advocated in this review.
A key observation in this case is that since unconnected clusters will inde-
pendently conserve the quadratically conserved quantities, one has as many
conserved quantities as the number of clusters × the number of nonlinearly
conserved quantities if the system was fully connected[36]. This makes the
Springer Nature 2021 L
A
T
E
X template
16 Wave-number Space Networks in Plasma Turbulence
wave-turbulence cascade dependent on the topology of the network, getting
blocked if the clusters remain unconnected, and with an explicit percolation
phenomenon as the number of triads are increased[73].
The case of inhomogeneous wave-kinetics [74, 50] is also of particular inter-
est, especially in the context of self-consistent drift-wave/zonal flow evolution,
with radial propagation as well as scattering in wave-number due to the effects
of zonal flows. The wave-kinetic system that results is isomorphic to the Vlasov
Equation, with wave quanta playing the role of the distribution function:
t
N
k
+
ω
k
N
k
x
ω
x
N
k
k
x
= C (N, N) + F
k
where ω = ω
k
u
y
(x) k
y
is the basic drift wave frequency Doppler shifted
by the zonal flow u
y
(x), C (N, N) represents a collision integral describing
mode coupling similar to (13) and F
k
is the forcing and dissipation which can
be provided by a linear instability as well as external forcing and small scale
dissipation. The inhomogeneous wave-kinetic equation has proved extremely
useful in the study of transport and turbulence in fusion plasmas, being ap-
plied in various problems ranging from momentum transport[75] to turbulence
spreading[76]. Network formulation of the general class of kinetic systems, of
which the wave-kinetics is a member is discussed in the next section.
2.5 Formulation of Kinetic Theory
Going back to regular plasma or fluid turbulence, the form of Eqn. (4) given in
Section 2.1 is really strictly valid only for a single field system. Many plasma
problems even when reduced fluid equations are used, involve multiple fields.
In this more general case, the different fields at a given node are likely to be
coupled linearly, allowing also for the possibility of linear interactions between
nodes (e.g. toroidal mode coupling in tokamak plasmas), we can write the more
general network equation as:
t
ξ
α
+
X
L
αβ
ℓℓ
ξ
β
+
X
,ℓ
′′
i
M
αβγ
ℓℓ
′′
ξ
β
ξ
γ
′′
= 0 (17)
where L
αβ
ℓℓ
is an arbitrary linear matrix, i
is the interaction network and M
αβγ
ℓℓ
′′
are the nonlinear interaction coefficients. In the usual spectral formulation
L
αβ
ℓℓ
= L
αβ
δ
ℓℓ
, and one can in general define a set of alternative variables say
χ
α
that diagonalizes the Greek indices in order to write the problem in terms
its eigenmodes, in that particular case we can write
(
t
+
α
) χ
α
+
X
,ℓ
′′
i
M
αβγ
ℓℓ
′′
χ
β
χ
γ
′′
= 0
of course with a different M that describes the interactions between the
eigenmodes, which can be computed using standard rules of linear algebra.
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 17
A fluid system with a finite number of moments is in fact a closure of
the full kinetic system, so in general a distribution function, can be written
as a combination of a number of suitable functional forms. For instance it is
common to describe the distribution function of the Vlasov equation using
Hermite polynomials[77, 78] as in
f (x, v, t) =
1
N
v
X
ℓ,α
ξ
α
(t) H
α
(v) e
ik
xv
2
(18)
where N
v
is a normalization factor, H
α
is the Hermite polynomial of (in-
teger) order α, with hopefully only a finite number of Hermite polynomials
being sufficient to describe its evolution. When used in conjunction with Eqn.
(18) the network equation of (17) describes the interactions between wave-
number nodes that satisfy the triadic interaction conditions, where each node
has a number of complex variables (indicated by the Greek indices) represent-
ing the coefficients of Hermite polynomials, which correspond to consecutive
derivatives of Maxwellians.
In the same spirit, one can use a combination of Fourier-Bessel-Hermite
[79, 80] (or Fourier-Laguerre-Hermite [81] ) expansion which handles, spatial,
perpendicular and parallel velocity directions respectively:
f
x, v
, v
, t
=
X
ℓ,m,α
F
α
ℓm
(t) J
0
(κ
m
v
) H
α
v
e
ik
·xv
2
(19)
At this point we can either flatten the indices α and m as before so that we
are left with a single index which represents the generalized wave-number
in x, v space so that some version of Eqn. (4) can be used, or keep the form
of Eqn. (19) in order to write the wave-number space network equation in its
general form as:
t
f
α
ℓm
+
X
L
αβ
ℓℓ
f
β
m
+
X
m
;
′′
m
′′
i
M
αβγ
ℓℓ
′′
; mm
m
′′
f
β
m
f
γ
′′
m
′′
= 0
note that in gyrokinetics v
appears as a label in the linear term (so no
coupling between m’s), and the coupling condition for the triadic interactions
can be written as k
m
+ k
m
+ k
m
′′
= 0 where κ
m
|k
m
| for v
space as well.
Even though the condition for interaction in m is actually the same for that
in , because of the details of the way the system may be discretized in these
different variables makes the actual computation of the interaction network
topology rather complicated. There are actually many different alternatives
to the above approach and the usual velocity space formulation with finite
difference discretization can also be formulated as a complicated interaction
Springer Nature 2021 L
A
T
E
X template
18 Wave-number Space Networks in Plasma Turbulence
matrix, however since our focus is wave-number space networks, we pick a
spectral formulation also in the velocity space variable as the natural choice.
Note finally, that energy conservation for Eqn. (17) can still be written
using Eqn. (6), with:
E
σ
ξ
α
σ
αβ
ξ
β
T
σ
ℓℓ
′′
2Re
h
M
βλµ
ℓℓ
′′
σ
αβ
ξ
α
ξ
λ
ξ
µ
′′
i
F
σ
X
2Re
L
αλ
ℓℓ
σ
αβ
ξ
λ
ξ
β
so that we can write:
P
σ
=
(
F
σ
if F
σ
> 0
0 if F
σ
> 0
and D
σ
=
(
0 if F
σ
> 0
F
σ
if F
σ
> 0
where σ
αβ
= σ
βα
defines the conserved quantity.
2.6 Magnetic Shear and Rational Surfaces
Strictly speaking, the examples that are given up to this point were written
in Cartesian coordinates, and are therefore valid only in slab geometry. While
one can easily transform everything to arbitrary curvilinear geometry, for ex-
ample where one of the directions is aligned with the magnetic field as the
natural geometry in magnetic fusion devices[82], the actual details of using
such coordinates is in fact non-trivial. This is partly due to the way some lin-
ear effects work in fusion plasmas, naturally allows an important reduction of
the number of degrees of freedom. Among these effects, that of the magnetic
shear stands out.
In order to understand the effect of magnetic shear, first consider the case of
a sheared magnetic field in slab geometry with B = B
0
h
ˆ
z +
x
L
s
ˆ
y
i
. The parallel
wave-number is then defined as k
= k
y
x/L
s
, and is therefore a function
of the spatial variable x. We know that the Landau damping kills off any
fluctuation with k
v
th
> ω, resulting in a reduction of the amplitude for any
Fourier mode with k
> ω/v
th
. This results in the amplitude of the fluctuations
being localized to the region between the two Landau turning points x
±
=
±|
ωL
s
k
y
v
th
|. Of course a proper eigenmode analysis may incorporate various other
effects, including the effect of an external, or self-generated flow shear, and thus
give a more complete picture, but the basic concept of the localization of the
drift-wave eigenmode due to Landau damping of higher k’s is rather generic.
A sheared slab, can model the geometry of the magnetic field as a local
approximation. For example, choosing the poloidal flux ψ as a radial variable,
θ as a poloidal variable with a period 2π, and ζ as the toroidal variable, we
can construct a generic toroidal coordinate system. Perturbations in such a
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 19
system can be written in the general form as:
Φ (ψ, θ, ζ) =
X
n,m
Φ
nm
(ψ) e
i()
. (20)
Since the magnetic field is a function of ψ and has the form of a helix wrapped
around a torus, it is customary to define the toroidal winding number q (ψ)
, which is the ratio of the number of times the magnetic field turns around
the toroidal direction to the number of turns it makes in the poloidal direction,
as “the safety factor” because of its importance in magnetic stability. For an
axisymmetric tokamak with circular flux surfaces, this takes the familiar form
q =
rB
ϕ
RB
θ
, where r and R are the minor and major radius variables. One
can also define the effect of magnetic shear using the dimensionless parameter
ˆs rq
/q. More generally, the rate of change of the safety factor as a function
of the poloidal flux ψ determines the strength of the shear in the magnetic
field in the ψ direction. If we use a coordinates system that aligns itself to
the magnetic field locally (like the so-called Clebsch coordinates described in
some detail in Ref. 82), the effect enters through the non-diagonal terms in
the metric tensor. However, at least in tokamaks, it is more customary to use
toroidal coordinate ϕ as the direction of axisymmetry.
Since the magnetic field is sheared, each flux surface ψ has a different pitch
angle. The perturbations tend to be aligned to the field line (i.e. have k
0)
as we discussed above. However since they are also periodic in θ and ζ, this
can happen exactly, only when the perturbation is centered at what is called a
“rational surface”, where q (ψ) = m/n. This allows a perturbation of the form
(20) to align itself to the magnetic field:
Φ (ψ, θ, ζ) =
X
n
Φ
n
(ψ) e
in(ζ)
(21)
and because of Landau damping, the perturbation with n and m such that
q (ψ) = m/n will be localized to its rational surface ψ = ψ
nm
defined by this
relation.
The fact that plasma “turbulence” in tokamaks consists of modes localized
to their rational surfaces, has important implications for their network for-
mulation. In this picture, the usual mode coupling through triads is largely
restricted by the additional constraint that the interacting modes must have
spatial overlap. So while the standard two dimensional turbulence presents a
very densely coupled network of interactions, the quasi-two-dimensional tur-
bulence (since k
0) of tokamaks have a sparser interaction topology. This is
true even when one includes a large number of toroidal and poloidal modes, so
that rational surfaces are densely packed (i.e. each rational surface has many
nearby neighbors).
Note that the nonlinear term of the underlying fluid equations due to ad-
vection by the E × B velocity, represented by the Poisson bracket, can be
written in the coordinate system consisting of ψ and α [so that B = ψ ×α
Springer Nature 2021 L
A
T
E
X template
20 Wave-number Space Networks in Plasma Turbulence
say with α = ζ q (ψ) θ] as:
, Ω] =
ˆ
b × Φ ·Ω = B
Φ
ψ
α
Φ
α
ψ
.
2.6.1 Ballooning representation
When a perturbation of the form (20) is considered for a given n but for
different values of m in toroidal geometry with a standard (e.g. increasing)
profile of q, the e
imθ
factors from consecutive rational surfaces superpose in
such a way that while at θ = 0 they add up, at θ = π they cancel. This
causes an envelope-like dependence in θ direction with a maximum at θ = 0
direction, or the low field side of the tokamak (also called the bad curvature
side). This envelope structure, which makes the modes expanded towards the
low field side is called the “ballooning” structure. Details of the functional form
of ballooning depends on the functional form of the localization of the Fourier
modes around their rational surfaces. Since this ballooning structure suggests
a slow variation of the envelope of the amplitude in θ variable [i.e. f (θ) e
inqθ
where the dependence of f (θ) on θ is “slow”], we are tempted to use an eikonal
approximation. However the fact that the θ variable is periodic, complicates
the issue.
In order to see this consider a Gaussian centered at θ = 0 as the ballooning
function f (θ). For a periodic θ, we can not write this simply as f (θ) = e
θ
2
/2σ
2
as it would have a discontinuity at θ = 0. Instead a basic first order form
f (θ) = e
θ
2
2σ
2
+ e
(θ2π)
2
2σ
2
could be used (makes sense especially for σ 2π), in order to remove the
jump at θ = 0. However going back to (21), in order to avoid a jump in Φ, the
phase should also be continuous across the cut at θ = 0 (or θ = 2π), and thus
the actual form can be obtained by replacing θ θ 2π in f (θ) e
inqθ
and
adding this to itself. In other words the ballooning function that removes the
discontinuity at θ = 0 can be written as:
f (θ) = e
θ
2
2σ
2
+ e
(θ2π)
2
2σ
2
e
i2πnq(r)
Note that this particular example extends the range of θ from [0, 2π) to
[0, 4π) or equivalently (2π, 2π]. If the ballooning function has a larger sup-
port, we need to extend the range of θ until the support is fully contained in
the extended range. This leads us to introduce, what is called the ballooning
representation, or the ballooning transform [83, 84] as:
Φ = e
in(ϕqθ)
N
X
p=N
ˆ
Φ (θ 2πp) e
i2πnq(r)p
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 21
where the function
ˆ
Φ (θ 2πj)
ˆ
Φ (η) is defined as a function of the extended
ballooning angle η, such that the functional dependence on η is simpler (e.g.
the basic Gaussian form of f (η) = e
η
2
/2σ
2
in the example above). Note that
here we have used a discrete version of the more common, continuum version
of the ballooning transform for consistency with the network picture as well
as the conventions used in the rest of the paper.
The use of ballooning representation, as well as localization of the drift in-
stabilities to rational surfaces, seems better adapted to the global physics of
low n modes as opposed to high n micro-turbulence. However the basic mech-
anism is independent of scale, and the approach is in fact used for gyrokinetic
simulations of small scale instabilities such as the ion temperature gradient
driven (ITG) turbulence, or even those at electron gyro-radius scales, such as
the electron temperature gradient driven (ETG) turbulence through the use
of flux tube geometry[85].
The introduction of the concept of the flux tube, indeed allows the refor-
mulation of the problem of plasma turbulence using only a small portion of
the whole toroidal volume. First using magnetic flux coordinates, one switches
to a coordinate system in which the magnetic field is a straight line (at each
flux surface ψ), and then the dependence of the magnetic field pitch angle to
ψ through q (ψ), generates a “sheared slab” like coordinate system in these
variables. This means that following Ref. 85, we can define
x =
q
0
B
0
r
0
(ψ ψ
0
) , y =
r
0
q
0
(α α
0
) , z = θ
where q
0
= q (ψ
0
), B
0
is the field at the magnetic axis and r
0
is the distance
from the magnetic axis to the center of the box, in order to map the flux tube
coordinates x, y and z, to the magnetic geometry of the tokamak. In this system
the equations go back to being quasi-two dimensional, with the possibility of
a network formulation using k
x
, k
y
etc. as discussed in earlier sections.
3 Truncation
The wave-number space network formulation is particularly useful for com-
ing up with a reduction of the full system, when only a small number of
wave-number nodes and/or triads are involved in the interaction. While in the
general case, a proper reduction requires some kind of closure for the effects
of the modes that are dropped in the reduced system, be it in the form of
wave-number space nodes that do not contain much energy, higher moments
of a kinetic distribution function or damped eigenmodes of a fluid system, a
direct truncation of the system without any closure (or some kind of ad-hoc
closure) is sometimes the simplest solution. Truncation may focus on drop-
ping i) wave-number space nodes (i.e. Fourier space truncation), ii) triads (i.e.
considering only a subset of interactions) or iii) field variables (i.e. dropping
higher order moments, or damped eigenmodes), or a combination of those.
Springer Nature 2021 L
A
T
E
X template
22 Wave-number Space Networks in Plasma Turbulence
For example, quasi-linear theory is an example of dropping triadic interac-
tions while keeping only those interactions with the large scales (profiles or
zonal flows), without any attempt of closure for the effects of the rest of the
modes. In contrast, if one uses eddy damping in such a system, this choice may
represent an ad-hoc closure for the truncated modes. A straightforward trun-
cation of the Fourier space using a geometrically scaled subset of wave-vectors
is called the reduced wave vector approximation (sometimes abbreviated as
REWA) [86, 87] and is the archetypical example of the truncated models that
we discuss in this section. Such models tend to give very small intermittency
corrections, as they have built-in self-similarity, but they can be very powerful
for studying turbulence across a large range of scales.
3.1 Nested Polyhedra Models in 3D
Nested polyhedra models (NPMs) are self-similar truncations of Fourier space
based on nested polyhedra[88, 89], resulting naturally in a finite set of complete
triadic interactions at each scale. In these models, the wave-number space
is discretized using nested, alternating icosahedron-dodecahedron pairs that
are organized in such a way that the nodes of the resulting network form
complete triads with nodes of the polyhedra from neighboring scales (see figure
3) . Since the truncation is done at the level of the network, the underlying
system of equations remains unchanged, and as there is only a finite number of
nodes and links (i.e. three body interactions), the resulting truncated system
naturally respects the conservation properties of the original system. It is a nice
example of the use of a truncated network in order to describe certain aspects
(e.g. scale by scale energy transfer) of the complete system. It also illustrates
neatly the separation of the issues of network topology (i.e. the regular grid is
replaced by a Fourier space made up of vertices of the nested polyhedra) from
those of model reduction (i.e. here the equations are kept exactly the same, so
there is no attempt at introducing even an ad-hoc closure). Nested polyhedra
models can also be considered as an anisotropic generalization of shell models,
used in studies of turbulence[90]. Below, we discuss the NPM for both the
Navier-Stokes and the MHD cases. The latter being more relevant for plasmas
and providing a somewhat more general example than the Navier-Stokes case.
While in both cases, the model has the ability to represent anisotropy, since
there is no source of anisotropy, the resulting turbulence remains isotropic.
The Navier-Stokes equation in Fourier space:
t
u
i
k
+ ik
κ
δ
ij
k
i
k
j
k
2
X
p+q=k
u
κ
p
u
j
q
= 0. (22)
can be discretized using a logarithmic alternating icosahedral/dodecahedral
basis where k = k
n
ˆ
k
m
with k
n
= g
n
λk
0
is the logarithmically spaced wave-
number magnitude with g =
φ =
q
1 +
5
/2 and λ =
q
5/3 for an
icosahedron and λ = 1 for a dodecahedron. The unit vector can be written as
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 23
Figure 3 The numbering n
m
of the vertices of the a) icosahedron (for an even m) as in
table 1 on the left and b) dodecahedron (for an odd m) as in table 2 on the right.
ˆ
k
m
= e
j
m
= [sin θ
m
cos ϕ
m
, sin θ
m
sin ϕ
m
, cos θ
m
] where θ
m
and ϕ
m
are to be
picked from the angles corresponding to the icosahedral and the dodecahedral
vertices. This choice comes from imposing the condition of forming triads with
the vertices of three consecutive polyhedra. Defining
M
iκj
nm
= k
κ
nm
δ
ij
k
i
nm
k
j
nm
k
2
n
The nested polyhedra model can be written as
t
u
i
n,m
+ iM
κij
nm
X
{m
,m
′′
}
u
κ
n2,m
u
j
n1,m
′′
+ u
κ
n1,m
u
j
n+1,m
′′
+ u
κ
n+1,m
u
j
n+2,m
′′
(23)
where the nth polyhedron is either an icosahedron or a dodecahedron, with
M
iκj
nm
= M
iκj
nm
+ M
ijκ
nm
and the sum is computed over pairs of interacting nodes
m
, m
′′
of the consecutive shells as given in the interaction tables table 1 and 2.
When written in this form, the model has a clear resemblance to shell models
and can actually be transformed into one by choosing the phases in a particular
way. It can also be written using flattened indices and keeping only half of the
nodes in each scale due to Hermitian symmetry:
t
+ νk
2
u
α
=
X
,ℓ
′′
=i
j
M
αβλ
u
βc
u
λc
′′
′′
Springer Nature 2021 L
A
T
E
X template
24 Wave-number Space Networks in Plasma Turbulence
m
n
i
(n,m)
j
= {m
n2
, m
n1
} i
(n,m)
j
= {m
n1
, m
n+1
} i
(n,m)
j
= {m
n+1
, m
n+2
}
0

4, 0
,
5, 1
,
1, 2
,
2, 3
,
3, 4

5, 0
,
6, 1
,
7, 2
,
8, 3
,
9, 4

5, 4
,
6, 5
,
7, 1
,
8, 2
,
9, 3

1

3, 0
,
4, 4
,
5, 5
,
0, 7
,
2, 9

1, 0
,
3, 4
,
8, 5
,
2, 7
,
6, 9

1, 3
,
3, 4
,
8, 5
,
2, 0
,
6, 2

2

5, 0
,
4, 1
,
3, 5
,
1, 6
,
0, 8

4, 0
,
2, 1
,
7, 5
,
9, 6
,
3, 8

4, 5
,
2, 4
,
7, 3
,
9, 1
,
3, 0

3

1, 1
,
5, 2
,
4, 6
,
2, 7
,
0, 9

0, 1
,
3, 2
,
8, 6
,
5, 7
,
4, 9

0, 1
,
3, 5
,
8, 4
,
5, 2
,
4, 0

4

0, 5
,
1, 3
,
2, 2
,
3, 8
,
5, 7

0, 5
,
4, 3
,
1, 2
,
6, 8
,
9, 7

0, 0
,
4, 1
,
1, 2
,
6, 3
,
9, 5

5

0, 6
,
1, 8
,
2, 4
,
3, 3
,
4, 9

1, 6
,
5, 8
,
0, 4
,
2, 3
,
7, 9

1, 0
,
5, 1
,
0, 2
,
2, 3
,
7, 4

Table 1 = 8n + m
n
, where m
n
, which denotes the mth vertex of the nth polyhedron, is
shown on the leftmost column, interacts with
i
()
j
= {i
0
, i
1
} = {8n 16 + m
n2
, 8n 10 + m
n1
}, {8n 10 + m
n1
, 8n + 6 + m
n+1
}
and {8n + 6 + m
n+1
, 8n + 16 + m
n+2
} for an even n (i.e. an icosahedron node) where m
n
,
m
n±1
and m
n±2
are to be taken from their corresponding columns. Note that a bar over
the integer value m
n
indicates c
j
= 0 (i.e. not conjugated), whereas no bar means c
j
= 1
(i.e. conjugated).
=
N
X
j=0
M
αβλ
u
βc
j0
i
j0
u
λc
j1
i
j1
where the Greek indices denote vector components, i
j0
and i
j1
are the two
nodes
and
′′
that interact with the node in the interaction triad t
j
(i.e.
t
j
ℓ, i
j0
, i
j1
) that are to be taken from tables 1 and 2 that gives i
j
=
i
j0
, i
j1
. Here we also defined
u
αc
(
u
α
if c
= 1
u
α
if c
= 0
on the right hand side. This is needed because we one only keep half of the
nodes of each polyhedra as in Fig. 3, which consist of 6 nodes for the icosa-
hedron and 10 nodes for the dodecahedron, and the rest of the nodes can be
obtained by reflection with respect to the origin [since u (k) = u
(k)]. The
interaction tables are given in tables 1 and 2, which replaces the adjacency
matrix for this kind of network, and the flattened node index can be written as:
=
(
8n + m
n
n : even
= 8n + m
n
+ 2 n : odd
with m
n
being the node number within the polyhedron (basically the label of
the discretized angle denoted by
ˆ
k
m
). Here each node is connected to N
= 9
triads if it is an icosahedron node (i.e. n is even) or N
= 15 triads if it is
a dodecahedron node (i.e. n is odd). A python implementation of the nested
polyhedra model can be found at [91].
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 25
m
n
i
n,m
= {m
n2
, m
n1
} i
n,m
= {m
n1
, m
n+1
} i
n,m
= {m
n+1
, m
n+2
}
0

5, 0
,
1, 1
,
4, 2

4, 0
,
3, 1
,
5, 2

4, 5
,
3, 1
,
5, 4

1

6, 0
,
2, 2
,
0, 3

5, 0
,
4, 2
,
1, 3

5, 6
,
4, 2
,
1, 0

2

7, 0
,
3, 3
,
1, 4

1, 0
,
5, 3
,
2, 4

1, 7
,
5, 3
,
2, 1

3

8, 0
,
4, 4
,
2, 5

2, 0
,
1, 4
,
3, 5

2, 8
,
1, 4
,
3, 2

4

9, 0
,
3, 1
,
0, 5

3, 0
,
4, 1
,
2, 5

3, 9
,
4, 3
,
2, 0

5

8, 1
,
7, 2
,
0, 4

5, 1
,
3, 2
,
0, 4

5, 8
,
3, 7
,
0, 0

6

9, 2
,
8, 3
, (1, 5)

1, 2
,
4, 3
,
0, 5

1, 9
,
4, 8
,
0, 1

7

2, 1
,
5, 3
,
9, 4

0, 1
,
2, 3
,
5, 4

0, 2
,
2, 5
,
5, 9

8

3, 2
,
6, 4
,
5, 5

0, 2
,
3, 4
,
1, 5

0, 3
,
3, 6
,
1, 5

9

6, 1
,
4, 3
,
7, 5

2, 1
,
0, 3
,
4, 5

2, 6
,
0, 4
,
4, 7

Table 2 = 8n + m
n
+ 2 where m
n
is shown on the leftmost column, interacts with
i
= {i
0
, i
1
} = {8n 14 + m
n2
, 8n 4 + m
n1
}, {8n 4 + m
n1
, 8n + 12 + m
n+1
} and
{8n + 12 + m
n+1
, 8n + 18 + m
n+1
} for an odd n (i.e. a dodecahedron node) where m
n
,
m
n±1
and m
n±2
are to be taken from their corresponding columns. As in table 1, if the
integer value m
n
has a bar over it c
j
= 0 (i.e. not conjugated), whereas no bar means
c
j
= 1 (i.e. conjugated).
3.1.1 MHD
The three dimensional MHD equations can be written in the same wave-
number space network as
t
+ νk
2
u
α
=
N
X
j=0
M
αβλ
u
βc
j0
j0
u
λc
j1
j1
b
βc
j0
j0
b
λc
j1
j1
t
+ ηk
2
b
α
=
N
X
j=0
δM
αβλ
u
βc
j0
j0
b
λc
j1
j1
b
βc
j0
j0
u
λc
j1
j1
(24)
where M
αβλ
= M
αβλ
+ M
αλβ
, δM
αβλ
= M
αβλ
M
αλβ
and
M
αβλ
(k) = ik
β
δ
αλ
k
α
k
λ
k
2
.
Note that the ratio of the largest wave-number to the smallest in an interact-
ing triad in a nested polyhedra model is a constant around 62% (i.e. 1 where
φ =
1 +
5
/2 is the golden ratio). Applying random forcing only on the ve-
locity field at all the nodes of the shells n = 4 and n = 5 using a fixed time step
h
f
10
3
, larger than the maximum step size for the adaptive time stepping
normally used, gives a steady state wave-number spectrum that is consistent
with Kolmogorov’s k
5/3
spectrum as can be seen in figure 4. It is probably
worth noting that the model is self-similar by construction (i.e. mono-fractal),
shows no sign of intermittency in the sense that it follows the S
p
(k
n
) k
p/3
n
scaling in the inertial range, where S
p
(k
n
) =
D
1
N
P
P
i
|u
i
nℓ
|
2
p/2
E
with ⟨·⟩
denoting time average. The nested polyhedra approach for the MHD shows
that such a model can easily handle the spectral evolution over 6 decades
with a large possible variation of magnetic Prandtl number and describe one
Springer Nature 2021 L
A
T
E
X template
26 Wave-number Space Networks in Plasma Turbulence
Figure 4 The wave-number spectrum with P r
m
= 1, ν = 10
9
, N = 60 and h
f
= 10
3
where P r
m
ν is the magnetic Prandtl number, N is the number of polyhedra, and h
f
is
the random forcing time step. Kinetic (solid line) and magnetic (dotted line) energy spectra,
both follow the Kolmogorov’s k
5/3
spectrum. The result is averaged over the polyhedra
nodes and also from t = 460 to t = 500.
of the mechanisms by which the system can reach equipartition between ki-
netic and magnetic energies, when only the velocity field is stirred (i.e. dynamo
action)[89]. Interestingly the system reaches steady state only when the veloc-
ity field is stirred through random forcing, whereas if the forcing is coherent,
the large scale magnetic field keeps accumulating.
3.2 Spiral Chain Models for 2D
In two dimensions, the basic equation of turbulence is that of Navier-Stokes,
which can be written in wave-number space as a scalar equation for the Fourier
transform of the stream function:
t
Φ
k
=
X
ˆ
z × p ·q
q
2
p
2
k
2
Φ
p
Φ
q
D
k
Φ
k
where the sum over represents a sum over p and q such that k + p + q = 0
(with p < q, since the interaction coefficient is symmetrized).
If it happens that a particular triad with k < p < q satisfies the condition
that η
ln(q/k)
ln(p/k)
Q (i.e. is rational), we can write p = kg
and q = kg
m
(i.e.
η = m/ℓ and g > 1 with g R). Since not all triangles satisfy the condition
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 27
Figure 5 The spiral chain = 2, m = 3 with g =
ρ. The counter clockwise primary spiral
chain is shown in black dashed lines while the clockwise secondary spirals are shown in gray
dashed lines. Note that as the energy travels along the primary chain, it gets exchanged
between the 5 secondary chains. Finally an interacting triad with k = k
n
(black arrow,
pointing right), p = k
n2
(red arrow, pointing up) and q = k
n+1
(blue arrow, pointing
left) is shown (i.e. k + q p = 0).
η Q, especially if we consider low order rationals, only a select class of
triangles can be represented by p = kg
and q = kg
m
with and m integers.
If we consider a “triad triplet”, that is a set of three triads that are obtained
by scaling and rotating a given triad such that the wave-vector k appears first
as the smaller, then as the middle and finally as the larger leg of the resulting
triad and the contributions of each of those three triads to the equation for
the wave-vector k, we can write:
t
Φ
n
= k
2
n
sin α
qp
g
m+
g
2m
g
2
Φ
n+
Φ
n+m
+ g
m3
1 g
2m
Φ
n+m
Φ
n
+ g
3m
g
2
1
Φ
nm
Φ
nm+
. (25)
Springer Nature 2021 L
A
T
E
X template
28 Wave-number Space Networks in Plasma Turbulence
writing k = k
0
g
n
ˆ
k
n
so that Φ
k
Φ
n
. The equation (25) is what we call the
spiral chain model [92] of two dimensional turbulence and it appears as a direct
generalization of the two dimensional shell models, since m = 2, = 1 gives the
usual Gledzer-Okhitani-Yamada [93] model. Strictly speaking the spiral chain
models must have rotation by the same angle between its elements, which is
true only for certain very particular values of g, and m even though the
general form is also interesting but does not have a regular spiral structure in
the wave-number space.
Considering for example = 2, m = 3 in (25), with θ
n
= , so that
α
pk
= 2α, α
qp
= α and α
qk
= 3α. We can write the equations using the law of
cosines for each angle, which gives two polynomial relations for g, which can
be solved for example for the case k + p q = 0 with g =
ρ where ρ is the
plastic number defined as
ρ =
1
2
1/3
1
r
23
27
!
1/3
+
1 +
r
23
27
!
1/3
and an angle α = π arccos
g
3
/2
. This particular spiral chain is shown in
figure 5.
Generally speaking, spiral chain models are wave-number space networks
where the nodes are organized in the form of spirals, and interact through
exact triadic interactions. Since the structure is self-similar, if we take three
nodes that interact, when we scale and rotate the system, the scaled versions of
those three nodes also interact. If the spiral chain consists of a single set of , m
and g values, it means that the whole spiral is constructed from a single triad
that is rotated and scaled accordingly. However there are cases where multiple
values of m and can be possible for a given value of g. Note that spiral chains
can also be used to construct domain partitions instead of truncation.
Unfortunately as numerical models, the spiral chain models suffer the same
shortcomings as the shell models for describing 2D or quasi-2D systems. Phase
evolution tending towards random make these systems unable to reproduce
a proper equipartition, since they lack the dependence of the number of de-
grees of freedom to scale that the linear grid has, and therefore they fail, in
particular in reproducing the inverse cascade[94]. Note that this is a general
problem related to the use of logarithmic spacing without paying attention to
the statistical weight of each element, and not a problem related to the network
formulation. However this can be remedied by using a closure that somehow
gets rid of the phase dynamics, in which case the models become similar to
differential approximation models[95, 96], but local over a fractal structure in
wave-number space instead of the usual 1D k-space grid.
3.3 Self-consistent Quasi-Linear Models
The usual formulation of quasi-linear theory deals with the computation of
fluxes of transported quantities like particle density, temperature or angular
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 29
momentum using linear evolution equations for the fluctuations. When the
profiles of these transported quantities also evolve in time, and the linear rela-
tions are computed at each time step given the local gradients of these evolving
profiles (which may also include zonal flows and or other meso-scale flow
structures), the resulting system is what we call self-consistent quasi-linear the-
ory. One can also improve such a formulation by introducing renormalization
[29, 97] in order to allow for nonlinear saturation through mode coupling. Oth-
erwise, the saturation mechanism of the regular quasi-linear models are based
on profile flattening and/or corrugation coupled with turbulent generation and
dissipation in different regions. In any case, self-consistent quasi-linear models
without renormalization are also examples of truncation of the Fourier space
and one can also formulate them using a single poloidal mode, where only a
single (usually most unstable) mode is considered[98, 99] whose evolution is
then coupled to the k
y
= 0 modes that represent the profiles and meso-scales.
In order to study general features of such a system, consider the generic
system of equations
t
ξ
α
k
+ L
αβ
k
ξ
β
k
= NL
α
k
(26)
where ξ
α
k
are different fields which could be Φ
k
and n
k
etc. for example in the
Hasegawa-Wakatani model, or may represents various moments of a kinetic
system and L
αβ
k
is a linear matrix operator. In general the wave-number space
network formulation focuses on computing, approximating or modeling the
nonlinear term NL
α
k
, which usually comes from an advection term in real space.
A single mode self consistent quasi-linear model consists of a formulation
where the wave-number space grid consists of a single k
y
(e.g. the most unsta-
ble) mode in addition to the k
y
= 0 (e.g. zonal) mode. The equations remain
the same even though depending on the formulation one may have different
equations for zonal and non-zonal modes to begin with. One interesting aspect
of such a system is that one can take a 2D pseudo-spectral solver and reduce
the resolution N
x
× N
y
N
x
× 1. Note that in practice a padded resolution
of N
py
= 6 is needed in order to have a single non-zonal mode that is resolved
(i.e. excluding the Nyquist frequency and the modes that are zeroed out as
part of the 2/3 padding rule). This is a major truncation of the initial system,
and the equations are not usually renormalized to take this truncation into ac-
count. The resulting self-consistent quasi-linear system describes the evolution
of zonal modes, or “profiles” in the presence of a single most unstable mode
that represents the fluctuations. The most unstable mode can not couple to it-
self, because it is impossible to satisfy triadic interactions with three modes all
having the same k
y
but it can couple to the zonal modes, which in this context
are indistinguishable from profiles even though the radial spatial resolution
can be very high and therefore their spatiotemporal evolution very detailed.
Note that the matrix L
αβ
k
contains the linear physics, and computing its
eigenvalues and eigenvectors we can perform a simple local linear stability
analysis. Once its N
f
eigenvalues, (N
f
is the number of independent fields), are
computed, and sorted with respect to their growth rates at each wave-number,
we can pick the most unstable mode k
y0
. We would then setup the system so
Springer Nature 2021 L
A
T
E
X template
30 Wave-number Space Networks in Plasma Turbulence
that it would have a single k k
y0
(by choosing the box size L
y
= 2π/k
y0
)
and run it as a nonlinear system given by (26). Assuming the nonlinear term
comes from E ×B advection (or advection by a velocity field obtainable from
a stream-function Φ) we can write the equations by computing the inverse
Fourier transforms in x as
t
ξ
α
+ L
αβ
(
x
) ξ
β
=
x
D
e
ξ
α
k
y
e
Φ
k
E
(27)
t
e
ξ
α
k
+
e
L
αβ
k
(
x
, ik)
e
ξ
β
k
=
x
Φ ·
y
ξ
α
k
+
y
e
Φ
k
x
ξ
α
(28)
where k k
y0
is used for simplicity. We can usually obtain Φ from ξ
α
k
using
something like F (ξ
α
k
, Φ) = 0, maybe through the relation between vorticity and
potential if one of ξ
α
k
represent the plasma vorticity as in Hasegawa-Wakatani
system, or through the Poisson’s equation when ξ
α
k
represent parts of the
distribution function. Here the linear operators for the “profiles”
L
αβ
(
x
) are
usually simple diagonal operators (unless we consider off diagonal transport
terms due to geometry) representing diffusion or friction terms. In contrast,
the linear operators
e
L
αβ
k
(
x
, ik) are usually not diagonal and contain the full
linear physics for the mode most unstable mode k. Nonlinearly, the quasi-linear
system contains i) the local advection (hence including the effects of shear
suppression) of the fluctuations by a background zonal flow profile, which is
represented by the first term on the RHS of Eqn. (28), ii) drive due to gradients
ξ
α
as well as the corrections by the corrugations in ξ
α
represented by
the second term on the RHS of Eqn. (28) and iii) the radial self-consistent
quasi-linear fluxes of the ξ
α
’s represented by the RHS of Eqn. (27). The roles
of these different terms can be made clear by considering ξ
α
to be the ion
temperature for example, i) advection and shear suppression of fluctuations,
ii) corrections due to T to the linear dynamics normally imposed by the
background temperature gradient and iii) the self-consistent quasi-linear fluxes
Q =
D
e
ξ
α
k
y
e
Φ
k
E
that can be computed using these fluctuations.
Considering multiple modes, (already considering only two modes k
y0
and
k
y1
) the system starts to include mode coupling, even though it remains to
be severely truncated. As a wave-number space model (i.e. going back to k
x
space), the truncation with respect to a full 2D system is rather clear. The same
also applies to going from 3D to 2D by truncation, which works exactly the
same way, or keeping only a few modes in the k
z
direction. All these examples
can be seen as truncations of the wave-number space.
Another interesting example of the wave-number space truncation is the
multi-mode self-consistent quasi-linear model, which results when we keep the
full k-space resolution (i.e. N
x
× N
y
modes) but consider only interactions
between zonal flows and fluctuations dropping interactions among fluctuations
themselves. This can be done by re-writing Eqns. (27) and (28) by putting a
sum over k in front of the RHS term of Eqn. (27) and interpreting Eqn. 28 as
an equation valid for all k k
y
instead of a single k
y0
. The linear operator
becomes a function of k and the physical meanings of the RHS terms remain
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 31
the same. The only difference is that the self-consistent quasi-linear fluxes now
include contributions from all the different k’s. The resulting multi-mode self-
consistent quasi-linear model is equivalent to what is called the generalized
quasi-linear theory in geophysical fluid dynamics. These models are truncations
of the primitive wave-number space networks not the in the sense that we
drop nodes from the network as in all the previous examples of this section
including the single mode self-consistent quasi-linear models, but in the sense
that we drop links (i.e. triad interactions). In fact we drop the overwhelming
majority of the interactions, and keep only one kind of interaction, that is
the interaction with the zonal modes hoping that this is the dominant kind of
interaction in the full network.
4 Reduction
Since the turbulent cascade involves the nonlinear transfer of a conserved quan-
tity (Energy, Enstrophy etc.) from one scale to another, a reduced description
of it can be based on the evolution of the relevant conserved quantity. The
budget for the conserved quantity in a primitive network is given in Eqn. (6).
We will use energy in Navier Stokes turbulence as an example, but the reduc-
tion procedure, which applies to the nonlinear term can be invoked for any
conserved quantity.
4.1 Energy transfer
Recalling Eqn. (6) explicitly for energy E
of node :
t
E
=
X
,ℓ
′′
i
T
ℓℓ
′′
+ P
D
(29)
where P
is the production and D
is the dissipation at node and T
ℓℓ
′′
is the
energy transfer from the nodes
,
′′
to the node . As the energy is conserved
nonlinearly, we have T
ℓℓ
′′
+T
′′
+T
′′
ℓℓ
= 0. Since Eqn. (29) is written on a
primitive network, each node represents a distinct discretized wave-number
and we only have three body interactions. In order to reduce it to a manageable
size one can consider a collection of subsets denoted by L that forms, what is
called an exact cover of the set of nodes of the initial primitive network. This
collection of subsets provides a partition of the wave-number space and the
energy in each subset L can be computed simply as the sum of the energies of
its constituents:
E
L
X
L
E
Similarly the total production P
L
=
P
L
P
and dissipation D
L
=
P
L
D
in the partition can be defined as the sum of production and dissipation terms
from each element contained within the partition. Note that for some sys-
tems with internal free energy source, it may be difficult to disentangle the
two, and one may instead have to write the net production as P
L
D
L
=
P
L
(P
D
). In order to write a budget equation for E
L
similar to Eqn.
Springer Nature 2021 L
A
T
E
X template
32 Wave-number Space Networks in Plasma Turbulence
a)
b)
c)
Figure 6 Example of a network reduction. Here a) shows the complete primitive network
with only triadic interactions, b) shows the collection of subsets, or lumps that are used
to construct the reduced network, and c) shows the resulting reduced network with both
triadic and direct connections. Notice how the highlighted triad (blue if in color), gives rise
to a direct interaction between the lumps L and L
since both of its legs denoted by and
are in L, while only one of its legs denoted by
′′
is in L
′′
.
(29) we also need to consider all possible channels of energy transfer to and
from the partition L. Considering ,
and
′′
, and noting that by definition
we have L, we have four distinct possibilities, i) all the nodes are in differ-
ent partitions, i.e.
L
and
′′
L
′′
, ii) one of the interacting nodes is in L
[e.g.
L
while
′′
L] iii) both
and
′′
are in the same partition, which
is different from L [e.g.
L
and
′′
L
] iii) both
and
′′
is also in L. In
the first case if T
ℓℓ
′′
is positive, the energy is transferred from L
and L
′′
to
the partition L as a three body interaction. The net contributions of all such
nodes gives us the net energy transfer between these three partitions, which
we can write as T
LL
L
′′
. The second and third cases give the transfer between
two partitions L and L
, and a sum over the transfers due to all such nodes
gives us the net energy transfer from L
to L, which can be denoted by T
LL
.
Note that unlike T
LL
L
′′
, this is a two body interaction term, and it appears
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 33
only in reduced networks. In the last case where all the nodes are in the same
subset, since the energy is transferred from one internal node to another, there
is no transfer in or out of L (see Figure 6 as an example).
The resulting equation for reduced energy E
L
can be written as:
t
E
L
P
L
+ D
L
=
X
L,L
′′
T
LL
L
′′
+
X
L
T
LL
(30)
where
T
LL
L
′′
=
X
L,ℓ
L
,ℓ
′′
L
′′
T
ℓℓ
′′
is the three body energy transfer from L
and L
′′
to L, while
T
LL
=
X
L
X
L
X
′′
L
T
ℓℓ
′′
X
L
X
L
X
′′
L
T
ℓℓ
′′
is the “direct” energy transfer from L
to L, which we can also write as:
T
LL
= T
LL
L
T
L
LL
so that it is obvious that T
L
L
= T
LL
.
Note that as the network is reduced, two-body interaction terms become
more and more prevalent, and in the limit of only 2 partitions, we end up with
a single direct energy transfer term between those two regions, which may
represent a generalized predator-prey model as the ultimate reduced network
dynamics. One interesting topological metric, which reveals the degree of re-
duction of a given k-space partition network is thus the ratio of the number of
two to three body interactions, which would range from 0, no reduction to 1,
complete reduction.
4.2 EDQNM Closure on Reduced Networks
While reduction through lumping the nodes together, or partitioning allows us
to lower the number of nodes and links in our wave-number space network, in
general it is not possible to write the transfer terms T
LL
L
′′
using the reduced
variables E
L
. This means that Eqn. (30) as written is not closed, and one has
to make an assumption in order to close it. Actually, this is also true for (29),
since we can not write T
ℓℓ
′′
in terms of E
, however one can always write an
equation for the phase, ϕ
similar to Eqn. (11) and couple it to the Energy
equation in order completely describe the system. In contrast since the reduced
transfer terms T
LL
L
′′
involve many different nodes, each of which having a
different phase, and while it is possible to add energies, it is not meaningful to
add phases in order to write single net phase for the whole partition, the closure
in the reduced network becomes a necessity as opposed to a convenience that
it was in the primitive network.
Springer Nature 2021 L
A
T
E
X template
34 Wave-number Space Networks in Plasma Turbulence
One of the simplest closures that is commonly used in the study of turbu-
lence is the so-called eddy damped quasi-normal Markovian (EDQNM) closure.
In the context of the wave-number space network formulation, the EDQNM
allows us to write:
T
ℓℓ
′′
=
Θ
ℓℓ
′′
|M
ℓℓ
′′
|
2
k
2
k
2
k
2
′′
E
E
′′
+
ℓℓ
′′
M
ℓℓ
′′
M
′′
k
2
′′
E
E
′′
through a set of assumptions equivalent to random phases and Markovian
statistics. This form allows us to write the transfer terms for the reduced
network as:
T
LL
L
′′
= Θ
LL
L
′′
a
LL
L
′′
E
L
E
L
′′
+ b
LL
L
′′
E
L
E
L
′′
+ b
LL
′′
L
E
L
E
L
(31)
where
a
LL
L
′′
P
L,ℓ
L
,ℓ
′′
L
′′
Θ
ℓℓ
′′
|M
ℓℓ
′′
|
2
k
2
k
2
k
2
′′
E
E
′′
Θ
LL
L
′′
E
L
E
L
′′
b
LL
L
′′
P
L,ℓ
L
,ℓ
′′
L
′′
Θ
ℓℓ
′′
M
ℓℓ
′′
M
′′
k
2
′′
E
E
′′
Θ
LL
L
′′
E
L
E
L
′′
and
Θ
LL
L
′′
X
L,ℓ
L
,ℓ
′′
L
′′
Θ
ℓℓ
′′
,
and using T
LL
= T
LL
L
T
L
LL
we can write:
T
LL
=
Θ
LL
L
a
LL
L
E
2
L
+
LL
L
b
LL
L
E
L
E
L
L
LL
b
L
LL
E
L
E
L
Θ
L
LL
a
L
LL
E
2
L
. (32)
Here the network version of the triad interaction time can be written using
its evolution equation:
t
Θ
LL
L
′′
+ [η
L
+ η
L
+ η
L
′′
] Θ
LL
L
′′
= 1 (33)
with Θ
LL
L
′′
(0) = 0 and with
η
L
= ν
L
+
1
2
X
L
L
′′
Θ
LL
L
′′
(b
LL
L
′′
E
L
′′
+ b
LL
′′
L
E
L
)
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 35
+
X
L
LL
L
b
LL
L
Θ
L
LL
b
L
LL
) E
L
1
2
X
L
Θ
L
LL
a
L
LL
E
L
(34)
Note that on an arbitrary network partitioning, we can not use the heuristic
estimate [i.e. θ
kpq
=
1e
(
η
k
+η
p
+η
q
)
t
η
k
+η
p
+η
q
with η
k
νk
2
+ c
q
R
k
0
E (k
) k
2
dk
] for the
eddy damping commonly used in this kind of closure since the integral from
0 to k is meaningless on an arbitrary network. Instead, if we really must, we
can use the even simpler one based on dimensional analysis:
η
L
ν
L
+ λE
1/2
L
k
3/2
L
(35)
together with the “solution” of (33) written as:
Θ
LL
L
′′
=
1 e
(η
L
+η
L
+η
L
′′
)t
η
L
+ η
L
+ η
L
′′
. (36)
The practical usefulness of the closure depend in practice on our ability to write
its coefficients a
LL
L
′′
, b
LL
L
′′
and Θ
LL
L
′′
in terms of some average variables
that we can associate with each subset L, such as the average wave-number
k
L
for example.
5 Ad-hoc Models
Turbulent cascade can also be modeled through simpler models that do not
necessarily come from a systematic reduction or truncation but are proposed
based on a general idea of the cascade process and some aspects of conservation
properties of the considered system. Dimensional analysis may be the simplest
such example, but applying it directly to networks is not likely to provide
us with any new insight. Shell models are another example, and they can
actually be reformulated from a network perspective and allow us to talk about
network-topology vs. cascade in a simple framework[100]. We have also shown
some examples of models that were obtained through truncation, such as the
nested polyhedra models or the spiral chain models that basically had the
same structure as the shell models, so the shell models can also be seen as
generic versions of these models. It is also possible to use a formulation in
terms of shells in order to describe the cascade processes of the full system as
characterized by the data from direct numerical simulations. In this sense, the
generic dynamical complex network model that uses shells as its nodes can be
seen as a prototype of the cascade processes of the full system.
Springer Nature 2021 L
A
T
E
X template
36 Wave-number Space Networks in Plasma Turbulence
5.1 Shell Models as Networks
A straightforward generalization of the shell model concept to arbitrary range
(i.e. m) interactions can be achieved by using a set of wave-numbers k
n
= k
0
g
n
,
where g is a logarithmic scaling factor (e.g. g = 2 ) and writing:
t
u
n
=
m
a
m
n
u
n+m
u
n+m+1
+ b
m
n
u
n+1
u
nm
+ c
m
n
u
n1
u
n1m
+ f
n
ν
n
u
n
, (37)
with
a
m
n
M
n,n+m,n+m+1
b
m
n
M
n,nm,n+1
c
m
n
M
n,n1m,n1
so that we have:
σ
n
a
m
n
+ σ
n+m
b
m
n+m
+ σ
n+m+1
c
m
n+m+1
= 0
as a consequence of the conservation law:
σ
n
M
n,n+m,n+m+1
+ σ
n+m
M
n+m,n,n+m+1
+σ
n+m+1
M
n+m+1,n,n+m
= 0 . (38)
Since the interaction coefficients are symmetrized, so that M
ℓnm
= M
ℓmn
by
construction, it is obvious that Eqn. (38) is directly equivalent to Eqn. (5).
The choice of M
ℓnm
(and therefore the σ
n
) define the kind of turbulence that
can be described by Eqn. (37). For example the choice
M
n,ℓ,ℓ
=
(
k
+ k
if n <
(1)
n
k
+ k
if < n
corresponds to the usual three dimensional Navier-Stokes turbulence which
conserves energy
E =
X
n
u_{n}
2
and what can be called helicity in the context of a shell model:
H =
X
n
(1)
n
k
1
n
u_{n}
2
.
The overall factor α
m
in front of the nonlinearity in Eqn. (37) denotes the
average contribution from the geometric factor, and can be taken to have the
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 37
form α
m
= g
m
(e.g. [101]), f
n
is the forcing usually localized to a few shells
and ν
n
is the viscosity, which may have the form of hyper or hypo viscosity
depending on the details of the problem.
Note that the three terms in Eqn. (37) come from a single triad form t
n
=
(n, n + m, n + m + 1) similarity transformed to t
nm
= (n m, n, n + 1) and
t
nm1
= (n m 1, n 1, n). Each of these triads introduce corresponding
terms in the equations for each of its nodes. In other words, the triad t
n
=
(n, n + m, n + m + 1) contributes a term (i.e. a
m
n
u
n+m
u
n+m+1
) to the equation
for u
n
, a second term (i.e. b
m
n+m
u
n
u
n+m+1
) in the equation for u
n+m
and a
third term (i.e. c
m
n+m+1
u
n
u
n+m
) in the equation for u
n+m+1
. By keeping all
the legs of all the triads, we can guarantee that we respect the conservation
properties of the system with no additional effort.
Eqn. (37) is actually a network model with the simple interac-
tion table i
n
= {(n + m, n + m + 1) , (n m, n + 1) , (n m 1, n 1)},
that is to be obtained from the triad list t
m
n
=
{(n, n + m, n + m + 1) , (n m, n, n + 1) , (n m 1, n 1, n)} for each n.
The way it is written Eqn. (37) has a single, fixed value of m, which defines
the elongation of the triangles involved in the reduction. Note that as we
commonly take g = 2, Eqn. (37) can not be a truncation (since it would be
impossible to satisfy k
n
+ k
n+m
+ k
n+m+1
= 0 with k
n
= k
0
g
n
ˆ
k
n
). It can be
reduction however, since the wave-numbers are consistent with k + p + q = 0
where k
n
< k < k
n+1
, k
n+m
< p < k
n+m+1
and k
n+m+1
< q < k
n+m+2
.
5.2 Small World Network Shell Models
Picking m = 1 in Eqn. (37) gives us only local interactions, and the resulting
model is exactly equivalent to the usual GOY model. The resulting regular
lattice of the GOY model can be represented as a list of N 2 local triads
[t
012
, t
123
, t
234
, ··· , t
N2,N2,N
]. Going over this list and either replacing some
of the local interactions with nonlocal ones (Watts-Strogatz strategy [40] ), or
simply adding nonlocal interactions on top of existing local ones (Newman-
Watts strategy [39] ), we get a small world network from a regular lattice (see
figure 7). In practice this is done by going over each of the elements of the list
of triads and replacing or adding a nonlocal interaction with a probability p.
Since the links are three-body interactions (and not two body as in simpler
networks) we also need to choose in which sense the triad is elongated. A
triad that is elongated “forward”, has the form t
m
n
= (n, n + m, n + m + 1)
and describes interactions with small scales where the scale n plays the role
of large scale. On the other hand a triad that is elongated backward has the
form t
m
n
= (n, n m, n 1) where the node n plays the role of one of the small
scale legs in an interaction involving a large scale mode at n m. In practice
each time we are adding a non-local interaction or replacing a local interaction
with a non-local one, we pick its direction to be forward with a probability p
f
or backward with a probability 1 p
f
. In the end we can choose p
f
= 0, 0.5 or
1 and it does not seem to matter since n itself is just a random node among
all the possible nodes, so apart from the slight difference it makes for the edge
Springer Nature 2021 L
A
T
E
X template
38 Wave-number Space Networks in Plasma Turbulence
Figure 7 The topology of a small world network generated using the Newman-Watts strat-
egy. The local triads are shown with shaded lines whereas the non-local triads are shown in
solid black. While the network is shown in a circular form, familiar in small world network
studies, here the nodes represents wave-number space shells with k
n
= k
0
g
n
and as a result
n = 0 is not connected back to n = 23 by a local interaction.
nodes, we can choose all the non-local nodes to be forward without loss of
generality. Note that the range of interaction m is also chosen to be a random
number between 2 and N n 2 for the forward interaction (or between 3
and n 2 for backward).
The act of reorganizing the nodes either by supplementing or replacing the
local interactions with non-local ones is called rewiring. Note that in the pro-
cess of rewiring one goes back to the initial regular lattice of local interactions,
and then applies the small world construction algorithm (either Watts-Strogatz
or Newman-Watts as described above) so that interactions do not accumulate
or turn all the interactions into non-local ones. Rewiring acts on the list of
triads. Once it is finished, we can compute the list of interactions i
= {
,
′′
}
and the nonlinear interaction coefficients M
ℓℓ
′′
for each node using the tri-
ads that contain that node. This can be done in practice by going over the list
of triads and when considering the triad t
m
n
= (n, n + m, n + m + 1) adding
{n + m, n + m + 1} to i
n
, {n, n + m + 1} to i
n+m
and {n, n + m} to i
n+m+1
with the interaction coefficients (i.e. weights) M
n,n+m,n+m+1
, M
n+m,n+m+1,n
and M
n+m+1,n,n+m
. As a result, one would have a different number of inter-
actions for each node, and some of those interactions would be non-local, yet
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 39
Figure 8 The steady state spectra from the dynamical complex network models based on
WS and NW rewiring strategies, compared with the GOY model, showing that all three
models capture the k
5/3
spectrum, while NW is slightly lower in amplitude as opposed to
the other two, probably as a result of its extra connections, and therefore higher transfer
efficiency.
the energy would be conserved trivially since we always consider all the con-
tributions from each triad. For the Watts-Strogatz case the model can be said
to go from a regular shell model for p = 0 to a shell model with random scale
interactions for p = 1.
As the network is being manipulated by wiring and rewiring, the variables
u
n
evolve according to the usual form of the equations, that we can write for
instance as:
t
+ νk
2
n
u
n
=
X
,ℓ
′′
=i
n
M
n,ℓ
,ℓ
′′
u
u
′′
+ f
n
. (39)
It may be argued that while Eqn. (39) represents the cascade of energy (or
some other conserved quantity) in a given network topology, the random-
ized evolution of the network topology itself can be a proxy for further phase
evolution that would enhance or impede energy transfer depending on the
momentary alignment and synchronization of the phases of the (primitive) tri-
ads involved in the interaction. Notice that the variables u
n
already has some
complex phases and their alignment already has consequences for the cascade,
which already randomizes the evolution of shell models. However since each
shell model node does not represent a single wave-number node, those phases
are ad-hoc replacements for some kind of average phase for a whole region.
Springer Nature 2021 L
A
T
E
X template
40 Wave-number Space Networks in Plasma Turbulence
Energy transfer is actually enhanced or not, depending on the phase relations
among many primitive triads that are involved in the interactions between
three regions that are represented as a single three body interaction among
three shells. The turning on and off of nonlocal interactions may thus model
a momentary phase synchronization in a nonlocal triad resulting in a burst of
nonlocal transfer.
Further studies are needed in order to clarify both the meaning and the
details of the evolution of wave number space network topology in turbulent
systems. Even if it is true that network topology can be used to represent
phase synchronization, it is not clear how to describe the phase evolution of a
turbulent system using rewiring rules. The question is similar to percolation,
but is more complex as it involves three body interactions. Here we use ran-
dom rewiring using one of the strategies described above in order to generate
regularly rewired small world network topologies, and solve Eqn (39) in such
an evolving network, choosing the network rewiring time step to be larger than
the time step for the evolution of the field variables. The results can be seen
in figure 8.
Another interesting option would be to use quadratic conserved quantities,
E
n
instead of the shell variables u
n
. This requires the use of a closure, -for
example the EDQNM closure discussed in Section 4.2- which gets rid of the
complex phases completely. In such a formulation the argument that the net-
work evolution describes the phase dynamics becomes more transparent. For
example for a two body interaction, If the phases of dominant triads involv-
ing nodes from the two regions L and L
are aligned, the edge between L and
L
would be activated. An activate two body interaction would add T
LL
as
defined in Eqn. (32) and the corresponding term (e.g. T
L
L
) in the equation
for E
L
. It can be seen that in the limit that most active interactions are two
body interactions, the problem of the turbulent cascade, becomes that of per-
colation through a dynamical network, as the conserved quantity tries to go
from the scale of injection to that of dissipation through a system represented
by mostly neighboring interactions that can be turned on and off. The caveat
is that the real system has an increasingly large number of triads as we go to
smaller scales, allowing many possible paths for the cascade, and this is a key
aspect of the statistical nature of the unreduced system.
6 Wave-number space Networks for Analysis
We can also use wave-number space networks for analysis of the data from
numerical simulations (and possibly experiment as long as enough high qual-
ity data is available). This usually involves projecting the full DNS data onto
some reduced wave-number space consisting of a partition denoted by L, in
order to study the injection, transfer, and dissipation of some conserved quan-
tity within that partition. For example we can take Eqn. (30) and compute
each of its terms from their definitions using the data that we obtain from
direct numerical simulations. Depending on the partitions, this may gives us
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 41
concrete information about the energy transfer between different parts of the
k–space. For a given problem, say the dissipative drift waves described by the
Hasegawa-Wakatani equations, different transfer terms may indicate energy
transfer from energy containing scales, corresponding to most unstable modes
to zonal flows and small scale isotropic perturbations. In a wave-number space
network formulation with three variables representing injection scales, zonal
flows and isotropic small scales, we can answer questions such as how much
energy is transferred among different modes, or if the tertiary instability is
actually important in zonal flow saturation.
We can also do a primitive wave-number space network analysis, in which
case we would compute each component of T
ℓℓ
′′
in Eqn. (29) for a reasonably
low resolution numerical simulation. However this is hard to represent graphi-
cally, and therefore can only be used to extract other useful information later
on. For example we can compute t
ℓℓ
, the net energy transfer between two
nodes from all possible triadic interactions, as defined in Section 2.2, which can
actually be represented graphically, or used in further reductions. An example
of the network representation of the transfer tensor t
ℓℓ
, for a Hasegawa-
Wakatani simulation with the minimum resolution of 16 × 16, is shown in
Figure 9. We can see that this is a zonal flow dominated case, where the en-
ergy is either transferred to zonal flows or is refracted to higher k
x
suggesting
that a generalized quasi-linear description where only the interactions with
the zonal flows are considered can be sufficient for describing this system in
this limit. However note that the implication that k
x
scattering comes mainly
from zonal flows need to be shown by looking explicitly at the t
′′
ℓℓ
where
′′
(i.e. the mediator) is zonal vs. non-zonal for each pair. ℓ,
that is significant
in the transfer network.
Recall that the two dimensional Hasegawa-Wakatani system[11] consists of
an equation of plasma vorticity
t
2
Φ +
ˆ
z × Φ ·∇∇
2
Φ = C
e
Φ en
+ D
Φ
2
Φ
(40)
coupled to an equation of continuity
t
n +
ˆ
z × Φ ·n + κ∂
y
Φ = C
e
Φ en
+ D
n
(n) , (41)
where Φ is the electrostatic potential normalized to T/e, and n is the fluctuat-
ing part of the plasma density normalized to a background density n
0
, x and
y are the radial and the (locally) poloidal directions normalized to the sound
Larmor radius ρ
s
, κ is the background density gradient that is assumed to be
constant, normalized to the speed of sound and C is the so called adiabaticity
parameter, which is a measure of the electron mobility in the parallel direc-
tion. With these, the E ×B velocity is defined as v
E
=
ˆ
z×Φ and the plasma
vorticity becomes
2
Φ. Here Φ = Φ denotes averaging in y (i.e. poloidal)
direction, so that
e
Φ Φ Φ . Finally D
Φ
and D
n
are functions representing
Springer Nature 2021 L
A
T
E
X template
42 Wave-number Space Networks in Plasma Turbulence
Figure 9 The primitive wave-number space network of a 16 × 16 k- space grid, where
each node represents a separate k vector. The colored boxes in the background show the
linear growth rate. The node size indicate the magnitude of kinetic energy and the node
color indicate kinetic energy injection at that node. The arrows indicate the direction, with
their width and color indicating the strength of energy transfer between the nodes. Energy
transfer is shown as an arrow only if it is above a certain threshold, and zonal nodes with
k
x
< 0 are removed because of Hermitian symmetry. We can see that most of the energy is
localized at the zonal mode around k
x
1, even though this mode is removing energy from
the system. We also see that the linearly most unstable mode is not the node that has the
highest energy injection nonlinearly, and the injected energy either couples directly to the
zonal flows or is refracted towards higher k probably by the action of the sheared flow.
viscosity and particle diffusion respectively. For the standard form of kine-
matic viscosity, we can take D
Φ
2
e
Φ
= ν
4
e
Φ for fluctuations. In contrast,
it makes sense to argue that for zonal flows, the dominant mechanism is large
scale friction, so that we can write D
Φ
2
Φ
= ν
ZF
2
Φ . Normally D
n
should be zero, but it is usually taken to have the same form with diffusion
D
n
(en) = D
2
en for numerical stability.
The Hasegawa Wakani model can be considered as a minimum non-trivial
model of plasma turbulence, as it has all the key ingredients such as the linear
instability and the ability to generate waves and zonal flows, but is otherwise
one of the simplest possible models in plasma physics that describes the evo-
lution of dissipative drift instabilities in the tokamak edge [102, 12]. It has
the same nonlinear structure as that of the two dimensional Navier Stokes
equations coupled with a passive scalar, or that of thermal convection in two
dimensions. Therefore, the study of the Hasegawa-Wakatani system from a
wave-number space network perspective is useful as a characteristic example
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 43
for the kind of turbulence that one finds in fusion plasmas, as well as in other
fields where similar equations are commonly used, such as solar or geophys-
ical fluid dynamics. A simple pseudo-spectral implementation of this model
provides us with a primitive network in the wave-number space.
Increasing the wave-number space resolution without increasing the wave
number space domain corresponds to increasing the box size and the number of
grid elements by the same factor, resulting in a larger box with an unchanged
real space grid element size as shown in top plot of figure 10. This allows us
to increase the resolution in wave number space, which can then be reduced
as in the bottom plot of figure 10 in order to “go back” to the low resolution
case shown in figure 9. Of course the details of the partitioning matters, and
the choice shown in figure 10 implies a reduction where each reduced node
represents a grid of 4 × 4 primitive nodes apart from the k
x
= 0 or k
y
= 0
nodes, which represent 1×4 and 4×1 primitive nodes respectively. The results
from this reduced network are shown in figure 11 and is qualitatively consistent
with the primitive network, with maybe more statistics and larger total energy
because of the larger spatial extent.
In order to compute the transfer term from DNS, consider a simulation
algorithm that advances the fields ξ
α
k
according to Eqn. (26), with the nonlinear
term written explicitly as
t
ξ
α
k
+ L
αβ
k
ξ
β
k
= NL (ξ
α
, Φ)
k
where L is the linear part of the problem and NL (ξ
α
, Φ)
k
is the nonlinear part,
which may for example represent the Poisson bracket as:
NL (ξ
α
, Φ)
k
(
ˆ
z × ξ
α
· Φ)
k
We would also usually have a relation between Φ
k
and ξ
α
k
for example as in the
case where one of the fields, say ξ
1
k
represents vorticity so that ξ
1
k
= k
2
Φ
k
.
In this formulation we can write
[
t
ξ
α
k
]
lin
= L
αβ
k
ξ
β
k
and
[
t
ξ
α
k
]
nl
= NL (ξ
α
, Φ)
k
Net injection of E
σ
k
Re
h
ξ
α
k
σ
αα
k
ξ
α
k
i
can then be written as:
P
σ
k
D
σ
k
Re
ξ
α
k
σ
αα
k
h
t
ξ
α
k
i
lin
whereas the total transfer of E
σ
k
from all the other wave-numbers to the wave-
number k takes the form
t
k
=
X
k
t
k
k
Re
ξ
α
k
σ
αα
k
h
t
ξ
α
k
i
nl
Springer Nature 2021 L
A
T
E
X template
44 Wave-number Space Networks in Plasma Turbulence
Figure 10 The box size and the resulting wave-number space resolution. The top plot
shows the final snapshots, at t = 200, of the vorticity on the left and the density on the right
of the two runs with the 64 × 64 grid as the full box and the 16 × 16 grid as the smaller
box at the lower left corner. In order to construct the same network as in figure 9 through
reduction, we partition the wave-number space grid as shown in the bottom plot (note that
only a part of the k-space is shown for clarity). Both the k
y
= 0 and the k
x
= 0 modes have
1 × 4 wave-number elements, whereas the rest of the nodes have 4 × 4 elements each. The
resulting network is shown in figure 11.
consistent with Eqn. (8). Most of the numerical algorithms that deal with the
Poisson bracket nonlinearity, does so in real space because of the efficiency
of doing so, either using finite differences directly using real space fields, or
through the pseudo-spectral formulation based on back-and-forth fast Fourier
transforms. This means that, we have a quantity that is already summed over
p and q as a result of the computation of
h
t
ξ
α
k
i
nl
. In order to separate this, we
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 45
Figure 11 The reduced wave-number space network of a 64 ×64 wave-number space grid.
Here each node represents a partition containing 4 × 4 wave-numbers (with the k
x
= 0 and
k
y
= 0 nodes containing 1 × 4 and 4 × 1 elements as shown in the bottom plot of figure 10.
As in Figure 9, the background color shows the linear growth rate and the node size and
color indicate the magnitude and production rate of kinetic energy at that node. Again as
in Figure 9, the arrows indicate the direction and the width and color of the lines indicate
the strength of energy transfer between the nodes. Note also that the arrow lines are always
curved in the counter clock-wise direction, so that we can infer their direction even when the
arrow tip is not visible because of crowding. We used suitable thresholds and normalizations
for both the nodes and the edges. Again, most of the energy is localized at the zonal mode
around k
x
1. The relative differences between the zonal modes (an the transfer to them)
and the others between figure 9 and the current figure, suggests that when the resolution is
increased by increasing the system size, so that the wave-number space resolution increases
without changing the wave-number domain, the zonal flow generation increases.
can define a “band pass” filter that contains a single Fourier mode as φ
(ξ
α
) =
ξ
α
e
ik
·x
, which sets all the other Fourier coefficients to zero. We can then
compute:
t
ℓℓ
Re
h
ξ
α
σ
αα
NL
ξ
α
e
ik
·x
, Φ
i
(42)
switching to the node notation Φ
Φ
k
ˆ
Φ (k
xℓ
, k
y
), as the net transfer of
E
σ
from
to (summed over all possible
′′
). In order to do the same for a
reduced network, we can either sum over t
ℓℓ
by writing t
LL
=
P
L,ℓ
L
t
ℓℓ
,
or use the “band pass” filter for the region L, defined as:
φ
L
(ξ
α
) =
X
L
ξ
e
ik
·x
Springer Nature 2021 L
A
T
E
X template
46 Wave-number Space Networks in Plasma Turbulence
so that we can write using the larger filter:
t
LL
=
X
L
Re
h
ξ
α
σ
αα
NL
φ
L
ξ
α
, Φ
i
. (43)
In other words if we have a nonlinear function NL (ξ, Φ) that computes the
nonlinear terms of our system, we basically send it the band pass filtered
ξ
L
= φ
L
(ξ) as the input variable NL
ξ
L
, Φ
in order to compute the effective
nonlinear term involving the modes in that region L
. Then we multiply by
σ and ξ
α
and sum over L as in Eqn. (43) to obtain the net transfer of
the conserved quantity from the region L
to region L. Repeating this for all
possible L
and L we get the full network as shown in figure 11.
6.1 Model Extraction
In some cases where a reduced wave-number space network can actually be
described by a closed system of equations of its reduced variables, say E
L
,
the numerical analysis can be used to “extract” that model directly from the
data. Extraction of reduced models from data, especially “identification” when
the data actually comes from a nonlinear dynamical system with a reasonable
number of degrees of freedom has been a subject of great interest recently and
is one of the more interesting methods of the booming field of data science,
especially from the perspective of the study of complex dynamical systems. In
this sense, the idea of using the data to in order to extract a reduced model
is simply an application of one of these methods to the problem in question
somehow formulated in terms of networks.
However methods such as SINDy[103, 104], applies directly to the data,
and not exploit the advantage of having a model that we can query in dif-
ferent ways. The data that comes out of a numerical simulation as discussed
in the previous section would be a combination of many terms, linear and
nonlinear terms, interactions with different fields, interactions with different
wave-number space nodes etc. All these are combined to form a single output
data, or a list of outputs equal to the number of evolving field variables at
each grid elements. For example we may have a model for the linear physics,
so that we can subtract that and concentrate on the nonlinear part, and try
to decipher it using different nonlinear forms such as polynomials or a library
of other nonlinear functions. The goal would be to minimize the difference be-
tween the [
t
ξ]
data
and the [
t
ξ]
model
= F
model
(ξ
data
), where F is a functional,
while keeping the system as sparse as possible.
From a network perspective, where the primary objective may be to reduce
a full system (i.e. the DNS) that we master completely, and can query as we
want, by sending it filtered input in order to compute data that is separated
into parts, the constraints on what can be done is quite different. We can for
instance separate the linear and the nonlinear parts of [
t
ξ] = [
t
ξ]
lin
+ [
t
ξ]
nl
explicitly. We can also write [
t
ξ]
nl
=
P
L
t
LL
and compute each of t
LL
using
filtered inputs as discussed earlier in this section as a function of time. We
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 47
can then do a linear regression, or nonlinear optimization in order to minimize
the difference between the data [t
LL
]
data
and the model for [t
LL
]
model
=
F
model
(ξ
L
, ξ
L
). Following SINDy, we can pose the optimization problem in
such a way as to favor a sparse model. The advantage of being able to separate
the data into its constituents that describe distinct interaction terms among
its elements is weighed against the usual complexity of the underlying system
and the fact that the reduction is almost never exact. Obviously a complex
nonlinear system, displaying multi-scale chaotic behavior, is unlikely to be
fitted nicely with a close-enough reduced model. The slight error between the
reduced model and the full system would probably result in a large difference
in terms of the actual time traces. Note that the optimization can be done
either by computing the differences in [d
t
ξ]
model
[d
t
ξ]
data
, or the differences
between ξ
model
ξ
data
by integrating the model in time with the given initial
conditions from the data using ξ
model
= ξ
dat
(t = 0) +
R
t
0
[
t
ξ]
model
dt.
Of course in cases where the system reduces itself to a nonlinear dynamical
system of a very small degrees of freedom, the difference between the above
methods where each interaction pair (or triple) is computed explicitly and
compared to a model of that interaction explicitly vs. the sum of all the terms
that contribute to a field variable is compared to a model of all the terms
on the right hand side for that variable is academic and both methods work
equally well. The simplest such example is probably the case of the reduction
of the data form four dimensional gyrokinetic simulations near the instability
threshold to a Lotka-Volterra predator prey model[54]. In this case, because
the system excites only a finite number of modes, which then interact with
the zonal flows and saturate rapidly, the many degrees of freedom involved in
these high resolution simulations of the gyrokinetic equation reduces itself to
an effective system of only two degrees of freedom corresponding roughly to
the most unstable mode, and the zonal flow. The reduction to a predator-prey
like system works also in the case of interchange, and can be performed within
the SINDy framework[105].
A similar case can be made for the use of machine learning for the de-
velopment of reduced network models. Consider for example the case of a
pseudo-spectral formulation with only two non-zonal modes in k
y
space (say
k and k
= 2k) , whose evolution can be written in real space as:
t
ξ
0
(x) = N
0k
(x) + N
0k
(x)
t
ξ
k
(x) = N
k0
(x) + N
kk
(x)
t
ξ
k
(x) = N
k
0
(x) + N
k
k
(x)
where
N
0k
=
k
, ξ
k
] , N
k0
=
Φ, ξ
k
+
Φ
k
, ξ
N
kk
= [Φ
k
, ξ
k
] +
k
, ξ
k
] , N
k
k
= [Φ
k
, ξ
k
] +
k
, ξ
k
]
Springer Nature 2021 L
A
T
E
X template
48 Wave-number Space Networks in Plasma Turbulence
If we are to reduce this to a single mode system
t
ξ
0
(x) = N
0k
(x) + F
0
(x, k, ξ
0
, ξ
k
, ···)
t
ξ
k
(x) = N
k0
(x) + F
k
(x, k, ξ
0
, ξ
k
, ···)
where F
0
and F
k
represent the effect of the effaced mode k
on the two re-
maining modes; the zonal mode and the mode with the poloidal wave-number
k. The argument for machine learning would be that we can train two “mod-
els” F
0
and F
k
, which would allow for a closure of the system, using the data
for N
0k
(x) and N
kk
(x) respectively. Furthermore, since such a simulation is
reasonably cheap, the training data can be generated on demand in chunks.
7 Summary and Conclusion
The goal of this review was to provide a different perspective into a well known
problem in nonlinear dynamics of fluids and plasmas. In order to achieve that,
the basic formulation of the plasma turbulence in the context of fusion plasmas
is revisited using wave-number space networks without going into particulars
of the models or the physical systems in question. In their primitive form,
these networks consist of nodes that represent elements in a discretized wave-
number space grid and three body interactions connecting those nodes. The
fields, generally representing the Fourier coefficients of various moments of
the distribution function are evolved on this three body interaction network
where each node interacts with all the pairs with whom it satisfies the triadic
interaction condition of k + p + q = 0 imposed by the convolutions in Fourier
space. Each link in such a three body network is represented as a triangular
node that is linked to three nodes that it links together, and each of its legs
has an associated weight which is the nonlinear interaction coefficient. It is
argued that the network formulation where we consider a list of nodes, and
list of triads, provides a solid framework for model reduction. It is shown for
instance that using a truncated wave-number space network that is obtained by
dropping nodes, or triads from the original primitive network, one still respects
all the same conservation laws as the original system without any effort.
Even though this review focuses on simple nonlinear models, such as two
dimensional Navier-Stokes equations or the Hasegawa-Wakatani system for
plasma turbulence, some issues about the application of the network formalism
to fusion plasmas, such as the effect of magnetic shear on mode localization
around rational surfaces, or the link between the wave-number space and the
flux tube formulation are briefly mentioned. Studies about the details of par-
ticular systems in realistic tokamak geometry is probably essential for future
studies, if the wave-number space network formulation is to be useful for fusion
applications.
Considering truncated networks, various examples from rather exotic,
nested polyhedra models, to the well-known example of quasi-linear theory are
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 49
evoked. It is clear that such models can reach incredible resolutions in compari-
son to direct numerical simulations on a regular rectangular grid. However due
to the fact that the number of triads increase as a function of scale on a regular
grid whereas it remains constant on a logarithmically scaled truncated net-
work, they tend to underestimate intermittency and tend toward an incorrect
statistical equipartition solution, which can be remedied by invoking simple
closures. A detailed discussion of network reduction clarifies various mecha-
nisms for the transfer of the quadratic conserved quantities (energy, enstrophy
etc.) of the original system. It is shown in particular that while a primitive
network only has three body interactions, a reduced network which groups to-
gether primitive nodes, may also have two body interactions as a result of two
of the nodes of an interacting triad ending up in the same partition. Ad-hoc
models that separate the evolution of the network topology and the evolution
of the field variable on that network are discussed, in particular in the form
of small-world network shell models, where the small-world aspect is achieved
by randomly introducing some nonlocal interactions to the usual shell models
of turbulence. It was argued that the evolution of the network topology could
be used to model phase coherences that appear randomly in the evolution of a
turbulent system. The (rather trivial) example of the EDQNM closure on a re-
duced network is also discussed, and it was shown that a simple reduced model
based on random phase approximation that assumes Markovian statistics can
be derived on a network. Finally, a discussion of how the transfer terms either
in terms of transfer between three nodes ,
and
′′
in a primitive network
or L and L
and L
′′
in a reduced network can be computed from a direct nu-
merical simulation by properly filtering the input fields in the nonlinear term.
An example of this is given for the Hasegawa-Wakatani system. Possible ap-
plication of the network formulation to model extraction or model training
using machine learning algorithms is briefly discussed at the end, leaving the
perspective open for future studies.
Declarations
Conflict of interest The author states that there is no conflict of interest.
References
[1] Diamond, P.H., S.-I. Itoh, Itoh, K., Hahm, T.S.: Zonal flows in plasma
- a review. Plasma Physics and Controlled Fusion 47(5), 35–161 (2005).
https://doi.org/10.1088/0741-3335/47/5/R01
[2] Gürcan, Ö.D., Diamond, P.H.: Zonal flows and pattern formation. Jour-
nal of Physics A: Mathematical and Theoretical 48(29), 293001 (2015).
https://doi.org/10.1088/1751-8113/48/29/293001
[3] Winsor, N., Johnson, J.L., Dawson, J.M.: Geodesic acoustic waves in
hydromagnetic systems. Physics of Fluids 11(11), 2448–2450 (1968).
https://doi.org/10.1063/1.1691835
Springer Nature 2021 L
A
T
E
X template
50 Wave-number Space Networks in Plasma Turbulence
[4] Conway, G.D., Smolyakov, A.I., Ido, T.: Geodesic acoustic modes in
magnetic confinement devices. Nuclear Fusion 62(1), 013001 (2021).
https://doi.org/10.1088/1741-4326/ac0dd1
[5] Mett, R.R., Mahajan, S.M.: Kinetic theory of toroidicity-induced alfvén
eigenmodes. Physics of Fluids B: Plasma Physics 4(9), 2885–2893 (1992).
https://doi.org/10.1063/1.860459
[6] Chen, L., Zonca, F.: Physics of alfvén waves and energetic particles in
burning plasmas. Rev. Mod. Phys. 88, 015008 (2016). https://doi.org/
10.1103/RevModPhys.88.015008
[7] Coppi, B., Rosenbluth, M.N., Sagdeev, R.Z.: Instabilities due to temper-
ature gradients in complex magnetic field configurations. Phys. Fluids
10(3), 582–587 (1967). https://doi.org/10.1063/1.1762151
[8] Romanelli, F.: Ion temperature-gradient-driven modes and anomalous
ion transport in tokamaks. Physics of Fluids B: Plasma Physics 1(5),
1018–1025 (1989). https://doi.org/10.1063/1.859023
[9] Horton, W., Hong, B.G., Tang, W.M.: Toroidal electron temperature
gradient driven drift modes. Phys. Fluids. 31, 2971 (1988). https://doi.
org/10.1063/1.866954
[10] Jenko, F., Dorland, W., Kotschenreuther, M., Rogers, B.N.: Electron
temperature gradient driven turbulence. Phys. Plasmas 7(5), 1904–1910
(2000). https://doi.org/10.1063/1.874014
[11] Hasegawa, A., Wakatani, M.: Plasma edge turbulence. Pys. Rev. Lett
50(9), 682–686 (1983). https://doi.org/10.1103/PhysRevLett.50.682
[12] Koniges, A.E., Crotinger, J.A., Diamond, P.H.: Structure formation
and transport in dissipative drift-wave turbulence. Physics of Fluids
B: Plasma Physics 4(9), 2785–2793 (1992). https://doi.org/10.1063/1.
860151
[13] Sarazin, Y., Ghendrih, P.: Intermittent particle transport in two-
dimensional edge turbulence. Physics of Plasmas 5(12), 4214–4228
(1998). https://doi.org/10.1063/1.873157
[14] Scott, B.D.: Drift wave versus interchange turbulence in tokamak geom-
etry: Linear versus nonlinear mode structure. Physics of Plasmas 12(6),
062314 (2005). https://doi.org/10.1063/1.1917866
[15] Zeiler, A., Biskamp, D., Drake, J.F., Guzdar, P.N.: Three-dimensional
fluid simulations of tokamak edge turbulence. Physics of Plasmas 3(8),
2951–2960 (1996). https://doi.org/10.1063/1.871630
[16] Beyer, P., Sarazin, Y., Garbet, X., Ghendrih, P., Benkadda, S.: 2d and
3d boundary turbulence studies. Plasma Physics and Controlled Fusion
41(3A), 757 (1999). https://doi.org/10.1088/0741-3335/41/3A/068
[17] D’Angelo, N.: Kelvin-helmholtz instability in a fully ionized plasma in
a magnetic field. The Physics of Fluids 8(9), 1748–1750 (1965). https:
//doi.org/10.1063/1.1761496
[18] Garcia, L., Diamond, P.H., Carreras, B.A., Callen, J.D.: Theory of
resistivity-gradient-driven turbulence. The Physics of Fluids 28(7),
2147–2158 (1985). https://doi.org/10.1063/1.865396
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 51
[19] Kadomtsev, B.B.: Plasma Turbulence. Academic Press, London, UK
(1965)
[20] Horton, W.: Drift waves and transport. Rev. Mod. Phys. 71(3), 735–778
(1999). https://doi.org/10.1103/RevModPhys.71.735
[21] Vedenov, A.A.: Quasi-linear plasma theory (theory of a weakly turbulent
plasma). Journal of Nuclear Energy. Part C, Plasma Physics, Accelera-
tors, Thermonuclear Research 5(3), 169 (1963). https://doi.org/10.1088/
0368-3281/5/3/305
[22] Bourdelle, C., Garbet, X., Imbeaux, F., Casati, A., Dubuit, N., Guirlet,
R., Parisot, T.: A new gyrokinetic quasilinear transport model applied
to particle transport in tokamak plasmas. Phys. Plasmas 14(11), 112501
(2007). https://doi.org/10.1063/1.2800869
[23] Beer, M.A., Hammett, G.W.: Toroidal gyrofluid equations for simu-
lations of tokamak turbulence. Physics of Plasmas 3(11), 4046–4064
(1996). https://doi.org/10.1063/1.871538
[24] Lee, W.W.: Gyrokinetic approach in particle simulation. Physics of
Fluids 26(2), 556–562 (1983). https://doi.org/10.1063/1.864140
[25] Brizard, A.J., Hahm, T.S.: Foundations of nonlinear gyrokinetic the-
ory. Rev. Mod. Phys. 79, 421–468 (2007). https://doi.org/10.1103/
RevModPhys.79.421
[26] Yoshizawa, A., Itoh, S.-I., Itoh, K., Yokoi, N.: Turbulence theories and
modelling of fluids and plasmas. Plasma Physics and Controlled Fusion
43(3), 1 (2001). https://doi.org/10.1088/0741-3335/43/3/201
[27] Sagdeev, R.Z., Galeev, A.A.: Nonlinear Plasma Theory. ed: T. M. O’Neil,
D. L. Book, W. A. Benjamin Inc., New York (1969)
[28] Frisch, U.: Turbulence: The Legacy of A. N. Kolmogorov. Cambridge
University Press, Cambridge (1995)
[29] Dupree, T.H.: Nonlinear theory of drift-wave turbulence and enhanced
diffusion. Physics of Fluids 10(5), 1049–1055 (1967). https://doi.org/10.
1063/1.1762220
[30] Krommes, J.A.: Fundamental statistical descriptions of plasma turbu-
lence in magnetic fields. Physics Reports 360(1-4), 1–352 (2002). https:
//doi.org/10.1016/S0370-1573(01)00066-7
[31] Kadomtsev, B.B.: Landau damping and echo in a plasma. So-
viet Physics Uspekhi 11(3), 328 (1968). https://doi.org/10.1070/
PU1968v011n03ABEH003837
[32] On the vibration of the electronic plasma. In: TER HAAR, D. (ed.)
Collected Papers of L.D. Landau, pp. 445–460. Pergamon (1965). https:
//doi.org/10.1016/B978-0-08-010586-4.50066-3
[33] Terry, P.W., Baver, D.A., Gupta, S.: Role of stable eigenmodes in satu-
rated local plasma turbulence. Physics of Plasmas 13(2), 022307 (2006).
https://doi.org/10.1063/1.2168453
[34] Gürcan, Ö.D., Li, Y., Morel, P.: Turbulence as a network of
fourier modes. Mathematics 8(4), 530 (2020). https://doi.org/10.3390/
math8040530
Springer Nature 2021 L
A
T
E
X template
52 Wave-number Space Networks in Plasma Turbulence
[35] Newell, A.C., Rumpf, B.: Wave turbulence. Annual Review of
Fluid Mechanics 43(1), 59–78 (2011). https://doi.org/10.1146/
annurev-fluid-122109-160807
[36] Harper, K.L., Bustamante, M.D., Nazarenko, S.V.: Quadratic invariants
for discrete clusters of weakly interacting waves. Journal of Physics A:
Mathematical and Theoretical 46(24), 245501 (2013). https://doi.org/
10.1088/1751-8113/46/24/245501
[37] Barabási, A.-L., Pósfai, M.: Network Science. Cambridge University
Press, Cambridge (2016). http://barabasi.com/networksciencebook/
[38] LeCun, Y., Bengio, Y., Hinton, G.: Deep learning. Nature 521(7553),
436–444 (2015). https://doi.org/10.1038/nature14539
[39] Newman, M.E.J., Watts, D.J.: Scaling and percolation in the small-world
network model. Phys. Rev. E 60, 7332–7342 (1999). https://doi.org/10.
1103/PhysRevE.60.7332
[40] Watts, D.J., Strogatz, S.H.: Collective dynamics of small-world networks.
Nature 393, 440–442 (1998). https://doi.org/10.1038/30918
[41] Junker, B.H.: 1. Networks in Biology, pp. 1–14. John Wiley &
Sons, Ltd, Hoboken, New Jersey, USA (2008). https://doi.org/10.1002/
9780470253489.ch1
[42] Landi, P., Minoarivelo, H.O., Brännström, Å., Hui, C., Dieckmann, U.:
Complexity and stability of ecological networks: a review of the the-
ory. Population Ecology 60(4), 319–345 (2018). https://doi.org/10.1007/
s10144-018-0628-3
[43] Wasserman, S., Faust, K.: Social Network Analysis: Methods and Appli-
cations. Structural Analysis in the Social Sciences. Cambridge University
Press, New York (1994). https://doi.org/10.1017/CBO9780511815478
[44] Broder, A., Kumar, R., Maghoul, F., Raghavan, P., Rajagopalan,
S., Stata, R., Tomkins, A., Wiener, J.: Graph structure in the web.
Computer Networks 33(1), 309–320 (2000). https://doi.org/10.1016/
S1389-1286(00)00083-9
[45] Barabási, A.-L., Albert, R., Jeong, H.: Scale-free characteristics of ran-
dom networks: the topology of the world-wide web. Physica A: Statistical
Mechanics and its Applications 281(1), 69–77 (2000). https://doi.org/
10.1016/S0378-4371(00)00018-2
[46] Boccaletti, S., Latora, V., Moreno, Y., Chavez, M., Hwang, D.-U.: Com-
plex networks: Structure and dynamics. Physics Reports 424(4), 175–308
(2006). https://doi.org/10.1016/j.physrep.2005.10.009
[47] Barabasi, A.-L.: The network takeover. Nature Physics 8, 14–16 (2011).
https://doi.org/10.1038/nphys2188
[48] Taira, K., Nair, A.G., Brunton, S.L.: Network structure of two-
dimensional decaying isotropic turbulence. Journal of Fluid Mechanics
795, 2 (2016). https://doi.org/10.1017/jfm.2016.235
[49] Taira, K., Nair, A.G.: Network-based analysis of fluid flows: Progress
and outlook. Progress in Aerospace Sciences 131, 100823 (2022). https:
//doi.org/10.1016/j.paerosci.2022.100823
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 53
[50] Smolyakov, A.I., Diamond, P.H.: Generalized action invariants for drift
waves-zonal flow systems. Phys. Plasmas 6(12), 4410–4413 (1999). https:
//doi.org/10.1063/1.873725
[51] Gürcan, Ö.D., Garbet, X., Hennequin, P., Diamond, P.H., Casati,
A., Falchetto, G.L.: Wave-number spectrum of drift-wave turbulence.
Phys. Rev. Lett. 102(25), 255002 (2009). https://doi.org/10.1103/
PhysRevLett.102.255002
[52] Malkov, M.A., Diamond, P.H., Rosenbluth, M.N.: On the nature of
bursting in transport and turbulence in drift wave–zonal flow sys-
tems. Phys. Plasmas 8(12), 5073–5076 (2001). https://doi.org/10.1063/
1.1415424
[53] Goldenfeld, N., Shih, H.-Y.: Turbulence as a problem in non-equilibrium
statistical mechanics. Journal of Statistical Physics 167(3), 575–594
(2017). https://doi.org/10.1007/s10955-016-1682-x
[54] Kobayashi, S., Gürcan, Ö.D., Diamond, P.H.: Direct identification of
predator-prey dynamics in gyrokinetic simulations. Physics of Plasmas
22(9), 090702 (2015). https://doi.org/10.1063/1.4930127
[55] Kim, E.-J., Diamond, P.H.: Zonal flows and transient dynamics of the
L-H transition. Phys. Rev. Lett. 90, 185006 (2003). https://doi.org/10.
1103/PhysRevLett.90.185006
[56] Miki, K., Diamond, P.H., Fedorczak, N., Gürcan, Ö.D., Malkov, M.,
Lee, C., Kosuga, Y., Tynan, G., Xu, G.S., Estrada, T., McDonald, D.,
Schmitz, L., Zhao, K.J.: Spatio-temporal evolution of the L->H and H-
>L transitions. Nuclear Fusion 53(7), 073044 (2013). https://doi.org/
10.1088/0029-5515/53/7/073044
[57] Terry, P.W., Horton, W.: Drift wave turbulence in a low-order k space.
The Physics of Fluids 26(1), 106–112 (1983). https://doi.org/10.1063/
1.863997
[58] Chen, L., Lin, Z., White, R.: Excitation of zonal flow by drift waves
in toroidal plasmas. Physics of Plasmas 7(8), 3129–3132 (2000). https:
//doi.org/10.1063/1.874222
[59] Gürcan, Ö.D., Anderson, J., Moradi, S., Biancalani, A., Morel, P.: Phase
and amplitude evolution in the network of triadic interactions of the
hasegawa-wakatani system. Physics of Plasmas 29(5), 052306 (2022).
https://doi.org/10.1063/5.0089073
[60] Kartashova, E.: Discrete wave turbulence. Europhysics Letters 87(4),
44001 (2009). https://doi.org/10.1209/0295-5075/87/44001
[61] Kartashova, E.: Nonlinear resonance analysis. Nonlinear Resonance
Analysis, by Elena Kartashova, Cambridge, UK: Cambridge University
Press, 2010 1 (2010)
[62] Waleffe, F.: Inertial transfers in the helical decomposition. Physics of Flu-
ids A: Fluid Dynamics 5(3), 677–685 (1993). https://doi.org/10.1063/1.
858651
[63] Alexakis, A., Biferale, L.: Cascades and transitions in turbulent flows.
Physics Reports 767-769, 1–101 (2018). https://doi.org/10.1016/j.
Springer Nature 2021 L
A
T
E
X template
54 Wave-number Space Networks in Plasma Turbulence
physrep.2018.08.001. Cascades and transitions in turbulent flows
[64] Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions
with Formulas, Graphs, and Mathematical Tables, ninth dover printing,
tenth gpo printing edn. Dover, New York City (1964)
[65] Kraichnan, R.H.: The structure of isotropic turbulence at very high
reynolds numbers. Journal of Fluid Mechanics 5(4), 497–543 (1959).
https://doi.org/10.1017/S0022112059000362
[66] Lesieur, M.: Turbulence in Fluids, Fourth edition edn. Springer, Dor-
drecht (1997). https://doi.org/10.1007/978-1-4020-6435-7
[67] Kraichnan, R.H.: Eddy viscosity in two and three dimensions. Journal
of Atmospheric Sciences 33(8), 1521–1536 (1976). https://doi.org/10.
1175/1520-0469(1976)033<1521:EVITAT>2.0.CO;2
[68] Holland, C., Diamond, P.H., Champeaux, S., Kim, E., Gurcan, O.,
Rosenbluth, M.N., Tynan, G.R., Crocker, N., Nevins, W., Candy, J.: In-
vestigations of the role of nonlinear couplings in structure formation and
transport regulation: experiment, simulation, and theory. Nuclear Fusion
43(8), 761 (2003). https://doi.org/10.1088/0029-5515/43/8/319
[69] Nazarenko, S.: Wave Turbulence vol. 825. Springer, ??? (2011). https:
//doi.org/10.1007/978-3-642-15942-8
[70] Connaughton, C., Nazarenko, S., Quinn, B.: Rossby and drift wave tur-
bulence and zonal flows: The charney-hasegawa-mima model and its
extensions. Physics Reports 604, 1–71 (2015). https://doi.org/10.1016/
j.physrep.2015.10.009
[71] Balk, A.M., Nazarenko, S.V., Zakharov, V.E.: New invariant for drift
turbulence. Physics Letters A 152(5), 276–280 (1991). https://doi.org/
10.1016/0375-9601(91)90105-H
[72] Dupree, T.H.: A Perturbation Theory for Strong Plasma Turbulence.
The Physics of Fluids 9(9), 1773–1782 (1966). https://doi.org/10.1063/
1.1761932
[73] Bustamante, M.D., Hayat, U.: Complete classification of discrete reso-
nant rossby/drift wave triads on periodic domains. Communications in
Nonlinear Science and Numerical Simulation 18(9), 2402–2419 (2013).
https://doi.org/10.1016/j.cnsns.2012.12.024
[74] Dubrulle, B., Nazarenko, S.: Interaction of turbulence and large-scale
vortices in incompressible 2d fluids. Physica D 110(1-2), 123–138 (1997)
[75] Diamond, P.H., McDevitt, C.J., Gürcan, Ö.D., Hahm, T.S., Naulin, V.:
Transport of parallel momentum by collisionless drift wave turbulence.
Physics of Plasmas 15(1) (2008). https://doi.org/10.1063/1.2826436
[76] Gürcan, Ö.D., Diamond, P.H., Hahm, T.S., Lin, Z.: Dynamics of tur-
bulence spreading in magnetically confined plasmas. Physics of Plasmas
12(3) (2005). https://doi.org/10.1063/1.1853385
[77] Grad, H.: On the kinetic theory of rarefied gases. Communications on
Pure and Applied Mathematics 2(4), 331–407 (1949). https://doi.org/
10.1002/cpa.3160020403
[78] Holloway, J.P.: Spectral velocity discretizations for the vlasov-maxwell
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 55
equations. Transport Theory and Statistical Physics 25(1), 1–32 (1996).
https://doi.org/10.1080/00411459608204828
[79] Plunk, G.G., Cowley, S.C., Schekochihin, A.A., Tatsuno, T.: Two-
dimensional gyrokinetic turbulence. Journal of Fluid Mechanics 664,
407–435 (2010). https://doi.org/10.1017/S002211201000371X
[80] Parker, J.T.: Gyrokinetic simulations of fusion plasmas using a
spectral velocity space representation. PhD thesis, University of
Oxford (2016). https://doi.org/10.48550/ARXIV.1603.04727. https://
arxiv.org/abs/1603.04727
[81] Mandell, N.R., Dorland, W., Landreman, M.: Laguerre-hermite pseudo-
spectral velocity formulation of gyrokinetics. Journal of Plasma Physics
84(1), 905840108 (2018). https://doi.org/10.1017/S0022377818000041
[82] D’haeseleer, W.D.: Flux Coordinates and Magnetic Field Structure [E-
Book] : A Guide to a Fundamental Tool of Plasma Theory. Springer
Series in Computational Physics, p. 241. Springer, Berlin, Heidelberg
(1991). https://doi.org/10.1007/978-3-642-75595-8
[83] Connor, J.W., Taylor, J.B.: Ballooning modes or fourier modes in a
toroidal plasma? The Physics of Fluids 30(10), 3180–3185 (1987). https:
//doi.org/10.1063/1.866493
[84] Connor, J.W., Hastie, R.J., Taylor, J.B.: Stability of toroidal plas-
mas: the influence of magnetic shear, periodicity and rotation. Plasma
Physics and Controlled Fusion 46(12B), 1 (2004). https://doi.org/10.
1088/0741-3335/46/12B/001
[85] Beer, M.A., Cowley, S.C., Hammett, G.W.: Field-aligned coordinates for
nonlinear simulations of tokamak turbulence. Physics of Plasmas 2(7),
2687–2700 (1995). https://doi.org/10.1063/1.871232
[86] Eggers, J., Grossmann, S.: Does deterministic chaos imply intermittency
in fully developed turbulence? Physics of Fluids A: Fluid Dynamics 3(8),
1958–1968 (1991). https://doi.org/10.1063/1.857926
[87] Grossmann, S., Lohse, D., Reeh, A.: Developed turbulence: From full
simulations to full mode reductions. Phys. Rev. Lett. 77, 5369–5372
(1996). https://doi.org/10.1103/PhysRevLett.77.5369
[88] Gürcan, Ö.D.: Nested polyhedra model of turbulence. Phys. Rev. E 95,
063102 (2017). https://doi.org/10.1103/PhysRevE.95.063102
[89] Gürcan, Ö.D.: Nested polyhedra model of isotropic magnetohydrody-
namic turbulence. Phys. Rev. E 97(6), 063111 (2018). https://doi.org/
10.1103/PhysRevE.97.063111
[90] Biferale, L.: Shell models of energy cascade in turbulence. Ann. Rev.
Fluid Mech. 35, 441–468 (2003). https://doi.org/10.1146/annurev.fluid.
35.101101.161122
[91] Gürcan, Ö.D.: nestp3d. GitHub (2017)
[92] Gürcan, Ö.D., Xu, S., Morel, P.: Spiral chain models of two-dimensional
turbulence. Phys. Rev. E 100, 043113 (2019). https://doi.org/10.1103/
PhysRevE.100.043113
[93] Ohkitani, K., Yamada, M.: Temporal intermittency in the energy cascade
Springer Nature 2021 L
A
T
E
X template
56 Wave-number Space Networks in Plasma Turbulence
process and local lyapunov analysis in fully-developed model turbulence.
Progress of Theoretical Physics 81(2), 329–341 (1989). https://doi.org/
10.1143/PTP.81.329
[94] Aurell, E., Boffetta, G., Crisanti, A., Frick, P., Paladin, G., Vulpiani,
A.: Statistical mechanics of shell models for two-dimensional turbulence.
Phys. Rev. E 50, 4705–4715 (1994). https://doi.org/10.1103/PhysRevE.
50.4705
[95] Lilly, D.K.: Two-dimensional turbulence generated by energy sources at
two scales. J. Atmos. Sci. 46(13), 2026–2030 (1989). https://doi.org/10.
1175/1520-0469(1989)046<2026:TDTGBE>2.0.CO;2
[96] L’vov, V.S., Nazarenko, S.: Differential model for 2D turbu-
lence. JETP Letters 83(12), 541–545 (2006). https://doi.org/10.1134/
S0021364006120046
[97] Gültekin, Ö., Gürcan, Ö.D.: Generalized curvature modified plasma dis-
persion functions and dupree renormalization of toroidal ITG. Plasma
Physics and Controlled Fusion 62(2), 025018 (2019). https://doi.org/10.
1088/1361-6587/ab56a7
[98] Bian, N., Benkadda, S., Garcia, O.E., Paulsen, J.-V., Garbet, X.: The
quasilinear behavior of convective turbulence with sheared flows. Physics
of Plasmas 10(5), 1382–1388 (2003). https://doi.org/10.1063/1.1566442
[99] Sarazin, Y., Dif-Pradalier, G., Garbet, X., Ghendrih, P., Berger, A.,
Gillot, C., Grandgirard, V., Obrejan, K., Varennes, R., Vermare, L.,
Cartier-Michaud, T.: Key impact of phase dynamics and diamagnetic
drive on reynolds stress in magnetic fusion plasmas. Plasma Physics
and Controlled Fusion 63(6), 064007 (2021). https://doi.org/10.1088/
1361-6587/abf673
[100] Gürcan, Ö.D.: Dynamical network models of the turbulent cascade.
Physica D: Nonlinear Phenomena 426, 132983 (2021). https://doi.org/
10.1016/j.physd.2021.132983
[101] Plunian, F., Stepanov, R.: A non-local shell model of hydrodynamic and
magnetohydrodynamic turbulence. New Journal of Physics 9(8), 294
(2007). https://doi.org/10.1088/1367-2630/9/8/294
[102] Scott, B.D.: An implicit simulation of drift-wave turbulence in a sheared
magnetic field. Journal of Computational Physics 78(1), 114–137 (1988).
https://doi.org/10.1016/0021-9991(88)90040-X
[103] Brunton, S.L., Proctor, J.L., Kutz, J.N.: Discovering governing equations
from data by sparse identification of nonlinear dynamical systems. Pro-
ceedings of the National Academy of Sciences 113(15), 3932–3937 (2016).
https://doi.org/10.1073/pnas.1517384113
[104] Kaptanoglu, A.A., Morgan, K.D., Hansen, C.J., Brunton, S.L.: Physics-
constrained, low-dimensional models for magnetohydrodynamics: First-
principles and data-driven approaches. Phys. Rev. E 104, 015206 (2021).
https://doi.org/10.1103/PhysRevE.104.015206
[105] Dam, M., Brøns, M., Juul Rasmussen, J., Naulin, V., Hesthaven, J.S.:
Sparse identification of a predator-prey system from simulation data of
Springer Nature 2021 L
A
T
E
X template
Wave-number Space Networks in Plasma Turbulence 57
a convection model. Physics of Plasmas 24(2), 022310 (2017). https:
//doi.org/10.1063/1.4977057