1. Introduction
The possibility to deal with a satisfactory confinement of the plasma in a Tokamak machine [
1] is strictly related to the existence of closed magnetic surfaces [
2,
3] (for a discussion on the influence of dissipation effects on this feature, see Ref. [
4]). However, the anomalous transport of particles and energy toward the machine walls is an intrinsic phenomenon in a Tokamak (or in medium or large size devices), and it deals with the subtle question of the power exhaust [
5]. Thus, the small region of plasma between the last closed magnetic surface and the walls or the divertor (commonly dubbed Scrape-off-Layer (SoL)) plays a very critical role in present and future Tokamak experiments. Such a plasma portion has very different properties with respect to the plasma in the Tokamak core and it possesses significant collisionality, up to admit (at list for low enough frequencies) a quasi-neutral two fluid representation, in which ions and electrons are properly described in interaction, mainly via the electromagnetic force [
6].
In the SoL, the magnetic field lines are always open and the presence of an X-point in the magnetic configuration, where the poloidal magnetic field identically vanishes, creates two additional “legs” in the magnetic configuration. The region just between the two legs, called “private zone”, is unavoidably particularly affected by a turbulent behavior of all the fundamental (electromagnetic and thermodynamical) quantities and, over the years, it increased attention to provide a satisfactory representation for the resulting turbulent transport [
7,
8,
9,
10,
11,
12,
13].
Here, we derive a local model for the turbulent dynamics of a plasma in the vicinity of an X-point, focusing attention to the basic ingredients of a predictive scenario. In addition to a constant magnetic field (taken along the toroidal z-direction of a Cartesian set of coordinates), we introduce the morphology of a small poloidal magnetic field, as it appears in the magnetic configuration of a region very close to the X-point and lying in the poloidal plane. The background configuration is also characterized by an equilibrium density and pressure, the former taken as an homogeneous contribution, while all the other dynamical variables live on the perturbation level only. Neglecting diamagnetic effects (ion and electron pressure gradient), the perpendicular current is then expressed by means of the ion polarization drift velocity only. We arrive to set up a coupled dynamical system, i.e., we have to deal with a partial differential non-linear set of equations in which the unknowns (density, electric and magnetic potentials and temperature) influence each others via coupling terms.
The theoretical and numerical analysis of the turbulence profile is then restricted to the two-dimensional (ortogonal) plane
, which dynamical features naturally emerge as soon as the drift coupling (induced by the parallel divergence of the parallel current) is neglected. Despite this reduction, the emerging electrostatic turbulence, in the presence of ion viscosity, is then still well-individualized via a mapping with the Euler equation for a viscous incompressible fluid (for a seminal paper, see Ref. [
14]). Such a correspondence concerns a direct isomorphism between the electric field potential and the so-called stream function for the fluid.
We analyze in some detail the morphology of the obtained dynamics, especially in its (truncated) Fourier representation and, for the inviscid regime, we derive an analytical solution for the (asymptotic) steady spectral morphology. This specific solution is recognized to mimic the Kolmogorov-Kraichnan enstrophy cascade spectrum proper the inertial range [
15], for which the energy per unit wave-number behaves as the inverse cubed wave-number (the enstrophy flux being constant). We then utilize the so-called Arnold criterion [
16] to demonstrate the stability of the analytical spectrum derived in the inviscid case, offering also more general hints on the stability in the viscous case.
The numerical analysis confirms that, if we fix the free parameters of the obtained Euler equation in the correspondence to the operation conditions of medium and large size Tokamak machines, the enstrophy cascade is dominating in the inertial range, with respect to the opposite phenomenon of an inverse cascade transferring energy from larger to smaller wave-number values. We stress that the behavior of the large scale modes resembles a phenomenon of condensation [
17,
18]. It is important to stress that the analytical solution is obtained by imposing an upper wave-number cut-off. In the model, we consider (as also outlined from the observed SoL turbulence phenomenology [
19]) the natural cut-off coinciding with the ion Larmor radius: the typical scale of the turbulence (from millimeters to few centimeters) is larger than this characteristic ion scale (0.1 mm).
The paper is structured as follows. In
Section 2, we outline the morphology properties of the magnetic configuration near the X point. In
Section 3, we derive the equations describing the electromagnetic turbulence using a quasi-neutral two-fluid approach. In
Section 4, we construct a local model for the turbulence dynamics near the X point in the form of a closed equation for the electric potential. The stability of the evolutionary dynamics is then discussed by means of the linear dispersion relation. In
Section 5, we provide the reduced two-dimensional dynamical equation for electrostatic turbulence, outlining the analogy with the theory of incompressible flow and we study the turbulence spectral properties in terms of the vorticity dynamics. The relevant scaling of the steady energy spectrum as a function of the wave number are also analyzed. We finally study the stability of the obtained analytical spectrum specified for the inviscid case. In
Section 6, the dynamics of the two dimensional electrostatic turbulence is investigated by means of a numerical code evolving the truncated Fourier expansion of the fluctuating potentials for three relevant cases. Concluding remarks follow.
2. Magnetic Configuration of the Equilibrium
In this Section, we fix the basic features of the local equilibrium on which we develop our turbulence analysis and the main assumptions regulating the addressed plasma scenario. We determine the morphology of the toroidal flux function, as considered sufficiently close to the X-point of a Tokamak device, while the toroidal component of the magnetic field is taken constant.
In what follows, restricting our attention to a spatial region of size much smaller than the major radius of the machine, we can neglect toroidal curvature effects and then we adopt the Cartesian coordinates
. The major magnetic field contribution is along the
z direction, mimicking the toroidal magnetic field of a Tokamak: we denote this component by
(here the suffix 0 denotes background quantities). Thus, the total magnetic configuration takes the following form:
where
is taken constant and
denotes the magnetic flux function.
It is easy to realize that, if the plasma is sufficiently cold and low dense near the X-point, then the function
must obey the (Ampère) equation [
20]
As solution of the equation above, we consider the expression
, where
denotes an assigned constant. This form of the magnetic flux function is depicted in
Figure 1, in arbitrary units, in order to represent the X-point morphology.
In what follows, we will thus deal with the background magnetic field:
having the directional versor
. Hence, the operator
takes the following expression:
where
.
We also deal with two other basic assumptions: (i) since the Debye length is much smaller than the turbulence scale, we assume the validity of quasi-neutrality, i.e., ( and denoting the ion and electron number density, respectively and we deal with a Hydrogen-like plasma); (ii) we implement the so-called “drift ordering”, i.e., the smallness of the fluctuations does not prevent that their gradients are comparable (or greater) to that of the background while their second gradients dominate the dynamics. Finally, we are assuming that no velocity fields are present in the local equilibrium, so that the velocity and electric fields are pure fluctuations. Only the number density and the pressure will contain a background contribution, here denoted by and , respectively, while their fluctuation components are denoted by barred quantities, namely and , respectively.
3. Dynamics of the Turbulence Fluctuations
We derive here the fundamental system of equations governing the turbulence dynamics and corresponding to a typical scheme for the non-linear low energy drift response. Our analysis is based on a two-fluid description of the plasma, in which the only relevant contribution to the ortogonal current density is provided by the ion polarization drift velocity. The starting point of the derivation is the balance law for both the parallel and the perpendicular momentum for ions and electrons, respectively. Here and in the following, we refer to parallel/perpendicular with respect to the background magnetic field versor . A basic assumption, present in the model since from the very beginning, is that the plasma is characterized by a negligible parallel ion velocity.
We start by stressing that the perpendicular electron momentum conservation provides the orthogonal electron velocity as coinciding with the
velocity (we neglect the perpendicular pressure gradient, as well as other smaller effects), namely
where
c is the speed of light (we adopt Gaussian units) and
denotes the electric field potential fluctuation. Since, the divergence of
contains only first order gradients of the electric potential, we will neglect it with respect to the divergence of the ion drift polarisation velocity. Thus, from now on, we consider
as a divergence-less vector. Moreover, we note that, for fully fluctuating quantities (which have no background counterparts), we omit the bar notation.
The dynamics of the ion perpendicular velocity
is governed by the following equation:
where
and we have neglected parallel components of the viscous stress (
e denotes the elementary charge,
the ion mass and
the kinematical (specific) ion viscosity). In the present analysis, the Lagrangian derivative reads
Since the turbulence is, in general, observed at frequencies much smaller than the ion gyro-frequency, we can set
(where
is a small correction to the fluctuating velocity). Hence, Equation (
6) gives at the leading order
where
denotes the ion gyro-frequency, calculated with the magnetic field intensity
.
If we neglect the diamagnetic effects, i.e., the pressure gradient contribution to the electron and ion velocities, the orthogonal current reads as
where we also implemented the quasi-neutrality condition for the plasma and we considered
as the only relevant difference between the two species velocity fields. In this respect, it is worth stressing how, for the case of a constant magnetic field (as considered below), the diamagnetic velocities would be divergenceless, so that their absence has no consequences on the charge conservation, which is a basic dynamical equation in the addressed scenario. Hence, the charge conservation equation
is rewritten via Equation (
8) as follows:
where
denotes the background Alfvén velocity, calculated with
. We now account for the parallel electron momentum in the presence of a constant parallel conductivity coefficient
and a parallel potential vector
only. Such a momentum balance explicitly reads
where
denotes the electron pressure. Now, by calculating from this equation the quantity
, and by implementing the Lorentz gauge condition
then, Equation (
10) rewrites as
Moreover, adopting the perfect gas law for the electron fluid and remembering that, according with the drift ordering the second gradients are dominant, we can write the expression
where
denotes the Boltzmann constant and
the electron temperature.
To solve the electric potential dynamics, we need to couple the density and electron temperature evolution to Equation (
13). The density dynamics is provided by the continuity equation (the same for ions and electrons due to the charge conservation equation, see
Appendix A), i.e.,
where we neglected the parallel ion velocity (
) and the diffusion coefficient
is a phenomenological tool to model different transport regimes [
12]. In this same approximation scheme, the (ideal) electron temperature evolution is governed by the following equation:
In the limit in which we can neglect the ion thermal conductivity, the thermal equilibrium holds and we can speak of an equal ion and electron temperature formally both governed by the equation above.
We conclude by observing that the assumptions underlying the system of equation above, describing the X-point turbulence, are well-justified from a phenomenological point of view. Clearly, we have to think of the X-point region as that one out of the plasma separatrix, which therefore has the same qualitative morphology of the SoL. Now, the spatial scale of the turbulence runs from few millimeters to ten centimeters, a scale surely much greater than the plasma Debye length (justifying the quasi-neutrality assumption) and significantly greater than the Larmor radius of the ions (allowing to speak of low frequency dynamics). Furthermore, the mean free path of the plasma constituents is about one meters, i.e., two orders of magnitude smaller than the parallel connection length of the background magnetic field [
21]. Thus, the plasma has a collisional nature and the two-fluid approximation is appropriate to describe the dynamics.
In view of these considerations, it is natural to apply, as done above, the drift ordering assumption [
12,
22,
23]. This paradigm can be summarized by saying that the amplitude of the fluctuations is proportional to their scale and it is, in absolute value, of the order of the ratio between the electrostatic energy and the thermal one, times background values. If we denote by
and
the fluctuation value of a quantity and its background one, respectively (the former having a spatial scale
, while the latter
), then the drift ordering approximation can be stated as
The relation above says that the fluctuation gradients ∼ are of the same order of magnitude of the background gradients ∼, while the second fluctuation derivative ∼ clearly provides the largest contribution.
Summarizing, the crucial point in order the approximations made in constructing the dynamical system derived in this Section are valid in the SoL is that its lower temperature and higher density make the plasma therein much more collisional than the one in the core of a Tokamak.
4. Relevant Reduced Model
In this Section, we construct, under suitable hypotheses, a closed equation in the electric potential field only, starting from the dynamical system fixed above. The first relevant assumption we implement on the dynamics is the possibility to neglect, in Equation (
15), the spatial gradients of the background density
. Then, comparing the resulting equation to Equation (
10) for the electric fluctuation
, we can easily get the following relation:
where we have assumed to model the phenomenological parameter as
[
12], which corresponds to the second hypothesis underling the present reduced model.
Now, focusing on a constant temperature model
, then Equation (
13) can be rewritten in the simplified closed form for the electric potential
as
If we now introduce the dimensionless quantities
,
,
and
(here
L denotes a given periodicity length of the system), the equation above reads in the following dimensionless form:
where
here
denotes the ion-electron collision frequency and
is the normalized orthogonal Laplace operator. Furthermore, retaining the dominant contribution in the second derivatives, we have
where
. Clearly, close enough to the X-point, i.e.,
, we have
and
.
Linear Dispersion Relation
In order to reduce the dynamics above to a coupled system of ordinary differential equation, let us now implement the Fourier representation of the field
. Using the vector notation
, with
, for the dimensionless wave-numbers, we can write
where the integral is extended to the whole space
and, since
is a real field, we have
. Substituting the expansion above into Equation (
20), we get, close enough to the X-point, the following equation for the Fourier component
:
where we have drop the
dependence.
Let us now investigate the behavior of the linearized system, i.e., we investigate the evolution of
neglecting the quadratic term regulated by the parameter
in Equation (
24). To this end, we set
, which, once substituted in the linearized equation, provides the following dispersion relation:
If we now separate the frequency into its real part and its imaginary one , respectively (i.e., ), it is easy to check that, for , we get , where . Conversely, for , we easily get and . We clearly see that there is always a damping of the modes since (no linear instability is present), but two different regimes can be distinguished: (i) when , we deal with pure damping; (ii) when , we have a damped oscillation of the mode. We observe that, since is a function of k and , the two regimes above can, in principle, co-exist but on different spatial scales.
In the case , the linear evolution is reduced to a pure damping behavior with . In this respect, we observe that, introducing the turbulent behavior of the poloidal magnetic field, via the dynamics of , we significantly affect the linear regime, simply because for small enough values of the parameter (in correspondence to a fixed value of ), we can have a very small negative value of , i.e., a much less damped mode appears. In the opposite case, i.e., when takes sufficiently large values (a less realistic situation than the previous one), we see the emergence of weakly damped (this is a k-dependent feature) oscillating modes.
6. Numerical Analysis of the Reduced Two-Dimensional Electrostatic Turbulence
The spectral properties of the two dimensional turbulence can be analyzed by means of the reduced model described in
Section 4 by numerically integrating Equation (
26) in the (truncated)
k-space. In the following, we adopt the re-definition
and we recall that
. We also expand the electric potential fluctuation in Fourier series as follows:
where
ℓ and
m are integer (positive and negative) numbers but not both zero and reversing the sign of
ℓ or
m (or both) corresponds to complex conjugation. In this scheme, Equation (
26) rewrites
The numerical scheme for integrating such equations is a Runge–Kutta algorithm (4th order) evolving in time a single component with or . For each cycle, the reality constraint is implemented for the corresponding counterpart. The summation is technically evaluated by cycling out the components (or ) outside a chosen domain.
The energy and enstrophy introduced above (of course conserved only for inviscid fluids if
) read now as
respectively, and we remark how the summations over
are taken on the whole domain (in the non viscous simulations presented in this paper, they are both conserved at the order
)). We also introduce the relation with the physical wave-number
K (in cm
) provided by
. In particular, for each time step,
and
can be evaluated by averaging over the 8 components (or 4 if
) having the same value of
K, thus providing the effective quantities
and
, respectively (see, for example, the pioneering work Ref. [
14]). It is important to remark the, using the analysis presented in
Section 5, the mode energy
introduced here is defined as
. By means of Equation (
36) and of the relevant scaling
, it is easy to recognize that
which, as already discussed, corresponds to an upward (from small to large mode-numbers) enstrophy cascade.
Since we are analyzing electrostatic turbulence in a region close to a X-point of a magnetic configuration, let us now implement realistic physical quantities specified for a typical Tokamak machine. We consider a hydrogen like plasma with
eV,
T and
m
[
30]. The viscous parameter
in Equation (
21) is defined by means of the specific ion viscosity coefficient
which reads [
1,
9]
where we set
. Using this setup, we get the following relevant quantities:
s
and
cm.
It is important to remark that, since we are addressing a model locally developed in a portion of the plasma near the poloidal null, we implement the physical condition of having the two dimensional periodicity box with length L of the order of the centimeter. At the same time, it is well know that the physical prediction of the spectral Fourier analysis are dependent on the truncation order of the k-series. In this sense, since we are treating electrostatic turbulence, as previously discussed we introduce a physical cut-off at small spatial scales provided by the Larmor radius, i.e., .
For all the simulations, we initialize the amplitude of the electric perturbation using
eV (which is then scaled by the factor
of Equation (
21)) only for the modes having 3 distinct
K values; all the other ones are set to zero amplitude. This corresponds to initialize the system with 3 “rings” of modes in the squared space
(we run a squared matrix of mode numbers
). We also underline that the energy spectrum plot presented in this Section are not instantaneous, but the relevant quantity
is time averaged over 250
(the dimensionless time) in order to avoid fast statistical fluctuations. Moreover, the considered final time corresponds to the thermal equilibrium in the sense that no systematic deviations of the time averaged spectrum occur for the inviscid case.
6.1. Small Box (Case A)
As first case, we consider a squared box of length cm which, using the plasma parameters defined above, yields to a viscosity coefficient (and to a scaling parameter ).
We first introduce (case A1) a small scale cut-off of about two Larmor radius by considering modes with
cm (together with the obvious constraint
). This yields to consider:
. The initial non zero modes are set as
(cm
) which corresponds to run the (
) mode numbers: (1, 1), (3, 2), (4, 1) (and, of course, all the other equivalent combinations). In total, we simulate 112 modes. We first analyze the inviscid regime by setting
in Equations (
49) and (
50). In
Figure 3, the contour plots of the fluctuations
are presented for different times as function of the mode numbers. The initialization is clearly evident, while the spectral evolution indicates a more intense excitation of the small mode numbers. In
Figure 4, we instead show the evolution of the modes
and
for a small temporal scale in order to show the saturation mechanism. The small mode number excitation is highlighted by the energy spectrum analysis presented in
Figure 5 (left-hand panel), where we plot the time averaged mode energy (in the sense described above) normalized to the initial one, i.e.,
. The chosen time corresponds to the thermal equilibrium when deviations of the spectrum no longer take place. It is evident that, using this setup, the relation (
53) is properly satisfied. In the right-hand panel of
Figure 5, we instead plot the spectral scaling for the viscous case by including in the simulations the parameter
. The reduction of the amount of energy in the system is evident and we observe a deviation from the scaling
for the first few modes.
In this first case, the inviscid simulation in
Figure 5 are thus found to properly reproduce the behavior predicted by the analytical solution discussed in the previous section. In this scenario, there is no trace of a process of constant rate of energy transport and the spectral features are essentially fixed by the vorticity dynamics. As soon as we introduce viscosity, a weak deviation from a constant vorticity spectrum is instead observed. The dissipation effect associated to a non-zero ion-ion friction in the plasma is responsible for a moderate energy trapping (a condensation process), i.e., small wave-number modes are excited to some extent, but clearly, as time goes by, the whole spectrum is progressively depressed by the viscous damping rate.
Using the same setup described above, let us now extend (case A2) the small spatial scale cut-off to the Larmor radius, i.e., considering
cm. This implies
, and, in total, we run 684 modes. The contour plots of the fluctuation evolution are depicted in
Figure 6, while the early mode saturation is shown in
Figure 7. The higher excitation (with respect to case A1 in which a more stringent cut-off was present) of the low mode numbers is now evident. The energy spectrum is finally plot in
Figure 8.
In particular, we can analyze the behaviour of
for the two (non viscous) cases A1 and A2 by comparing the left-hand panels of
Figure 5 and
Figure 8. A deviation from the
behavior emerges when simulating a large set of modes in the case A2. In fact, having increased the maximum available values of
K, but maintaining fixed the ratio between the enstrophy and the energy, some modes in the low region of the wave-numbers deviate from the constant vorticity spectrum, i.e., a marked condensation phenomenon is now present according to the increase of the available
K value. In fact, we observe that, even if not initialized, such modes are pumped and their spectral behavior is not accounted by their analytical solution Equation (
36). Moreover, this analysis confirms the relevant issue of this approach represented by the fact that the physical predictions are strongly dependent on the Fourier series truncation order. The introduction of the viscosity in the simulations (right-hand panel of
Figure 8) emphasizes the shape predicted in the previous section and its stability features appear significantly confirmed by the numerical investigation in its whole set-up.
We conclude the numerical analysis of the case A2 by showing the contour plots of constant
at fixed times, constructed by inverting the Fourier transform in Equation (
48). In
Figure 9, we report the results at 4 distinct times for the inviscid regime. Given the form of the energy spectrum described above, the evolutive configuration of the electric potential well represents a commonplace effect of the discrete vortex theory (see the seminal paper Ref. [
31]): the presence of a pair of large scale vortices mainly saturating the periodicity box.
When turning on the viscosity (see
Figure 10), this feature is still present. It worth noting how the dynamics has now a negative forcing term which results in a different temporal evolution of
when compared to
Figure 9. The relevant morphology related to large scale vortices is clearly outlined in the viscous case, but a time phase shift is present with respect to the inviscid regime. We also stress how viscosity yields to smooth hillsides of the
profile, providing a clean picture of the vortices. This is due to the fact that viscosity damps small-scale ripples.
6.2. Large Box (Case B)
For this case we now consider the squared box to have length
cm. Implementing the plasma parameters defined above, we get, for this case,
(and the scaling parameter for the fluctuation results
). For computational reasons, we set the cut-off
cm by considering
. Differently from the previous cases, non zero modes are chosen as
(cm
) corresponding to the (
) mode numbers: (4, 1), (4, 2), (5, 2) (plus the other combinations). In total, we thus simulate 1058 modes and we plot in
Figure 11 the energy spectrum for this case (viscous and inviscid).
Such a scenario has been obtained by pushing the model parameters up to a limit case of validity for the underlying assumptions of the investigated dynamics. The relevance of such a simulation is that, due to the presence of a large number of modes, it corresponds to the more realistic situation of a nearly continuum spectrum. Here we are considering increasingly large values of K, and the deviations from the constant vorticity spectrum for the low K portion (due to the condensation phenomenon) is again present in both inviscid and viscous cases.
Furthermore, in
Figure 12 we plot the spectrum compared to the theoretical
expectation (as in
Figure 8, right-hand panel) and to the non-local logarithmic correction of Equation (
37). We remark that, in the relation
, we have considered here
, which corresponds to the bottom of the
K-range. It is evident from the figure that the non-local correction ensured by the logarithmic term better reproduces the simulated profile, especially for small
K values.
Conversely from the previous analyses, where we have shown the early mode evolution to underline the saturation mechanism, in
Figure 13 we now plot the late temporal behavior of the most energetic mode
in the viscous regime. The evolution points out a sort of intermittency-like feature characterized by different time intervals. This property is however not dominant and also present in the inviscid case. In fact, there are several confirmations, both experimental and numerical, that two-dimensional turbulence is not intermittent [
26,
32,
33,
34] or, if present [
27,
35], intermittency is found to be weak. Actually, large-scale intermittency is due to the developments of enhanced tails in the expected Gaussian probability density function of the perturbed field. A detailed analysis should be thus performed regarding the statistical displacements
, but this lies outside the scope of the present paper.
Also for this case B, we show in
Figure 14 the contour plots of
at fixed times when viscosity is turned on. The presence of the pair of rotating large scale vortices clearly emerges form the graphs, outlining again the match with respect to the discrete vortex model.
6.3. General Remarks
Nonetheless small deviations, we can firmly conclude that the behavior of the turbulent plasma in the operation conditions typical of the SoL of a medium or large size machine, is well represented by the analytical solution (
36) (also according to previous analyses [
9]). Therefore, the transport of enstrophy from small to large wave-numbers is found to be a natural feature of the Tokamak edge turbulence. According to our description of the inviscid turbulence that accounts for a Fourier representation naturally truncated by the ion Larmor radius cut-off, we can argue that the available wave-numbers are almost above the critical value
, in the sense previously described. By other words, the SoL of a Tokamak lives in the range of parameters such that
, corresponding to say that the “temperature” associated to the energy is much greater, in absolute value (negative values of
A are also available for the system), than the value of the “temperature” associated to the enstrophy.
This situation would be clearly significantly altered in a three dimensional analysis, since the enstrophy constant of motion would be lost. However, the fact that a Tokamak has an axial symmetry structure, could imply that, at least under certain operation conditions, the three dimensional nature of the electrostatic turbulence is a weak modification of the here discussed two dimensional spectrum, so attributing a more general character to the numerical results discussed above. This claim requires a further discussion in view of the peculiarities outlined by three-dimensional turbulence with respect to the two-dimensional one [
36,
37]. As stated above, in the former case, the enstrophy is no longer a conserved quantity. The spectral feature is thus governed by the Kolmogorov law associated to a constant rate of energy transfer and to a scaling
. Nonetheless, we can better understand the morphology of the solution of Equation (
20), by the following simple considerations. According to the topology of a Tokamak, we can naturally assume that the field
is periodic in
(actually, in order to reproduce a Tokamak toroidal profile,
would have to be normalized by
), so that we can consider the Fourier expansion
Substituting the expansion above into Equation (
20), we get the following dynamics for the Fourier harmonics
:
To this equation we must add the gauge condition for each harmonic of the parallel magnetic potential vector.
Despite in constructing Equation (
20) we neglected the number density gradients, it is a well-known result that such a term is responsible for the linear drift wave instability. The idea proposed in Ref. [
9] is that, in the left-hand side of Equation (
56) the
mode dominates. At the same time, in the right-hand side, the value of
n maximizing the linear growth rate is associated to this mode. In the spirit of this postulation we can thus expect that the three-dimensional turbulence spectrum remains close the one discussed above. However, since it is well-known [
10] that the non-linear drift response occurs when the linear growth rate is small enough, the investigation of the fully developed three-dimensional turbulence, as well as its deviation from the
law, constitute a valuable field for future investigations.
We also note that, in a real Tokamak device setting, the background magnetic field lines have non-zero curvature and this feature has a relevant impact on the turbulence morphology. In this respect, see the analysis developed in Ref. [
11] where the interaction between ballooning mode and the non-linear drift response is discussed.
7. Conclusions
We analyzed the basic dynamical scheme regulating the turbulent behavior of the plasma in a Tokamak SoL. We focused on low frequency phenomenology (i.e., the typical frequency of the fluctuations is smaller than the ion gyro-frequency) and on a local model nearby the X-point, so that the magnetic configuration of the equilibrium has been modelled via a dominant contribution along the z-axis and a smaller poloidal contribution, exactly vanishing in the X-point taken as origin of the plane.
After the general set up of the problem, the theoretical and numerical analysis has been devote to the simplified two-dimensional electrostatic turbulence, also in the presence of ion viscosity. This formulation was investigated using the natural mapping existing with a two-dimensional turbulent incompressible fluid: the electric potential field and the fluid stream function are described by the same dynamics.
The inviscid turbulence was discussed on a theoretical framework by outlining the existence of a steady analytical solution for the (truncated) spectral representation, which corresponds to a ∼ behavior. Then, the stability of this solution has been properly investigated and, using the Arnold criterion, we arrive to conclude that such a spectral profile is a stable configuration for the system evolution.
The numerical analysis has confirmed that, in the range of parameters available in the SoL of a medium or large size Tokamak, the enstrophy cascade toward larger wave-numbers dominates the asymptotic spectrum both in the inviscid and viscous cases. However, a certain energy trapping into the larger spatial scales is observed, especially when the viscosity contribution is considered.
Concluding, we stress how having fixed all the basic features of the turbulence due to the advective transport of the “vorticity” offers a dynamical and physical scenario in which the role of the non-linear drift response can be properly evaluated [
11].