Preprint
Article

Generalized Chi and Eta Cross Helicities in Non-Ideal Magnetohydrodynamics

Altmetrics

Downloads

81

Views

22

Comments

0

A peer-reviewed article of this preprint also exists.

This version is not peer-reviewed

Submitted:

26 September 2023

Posted:

27 September 2023

You are already at the latest version

Alerts
Abstract
We study the generalized χ and η cross helicities for non-ideal non-barotropic magnetohydrodynamics (MHD). χ and η, the additional label translation symmetry group, have been used to generalize the cross helicity in ideal flows. Both new helicities are additional topological invariants of ideal MHD. To study there behaviour in non-ideal MHD, we derive the time derivative of both helicities using the non-ideal MHD equations taking viscosity, finite resistivity and heat conduction into account. Physical variables are divided into ideal and non-ideal quantities separately during the mathematical analysis for simplification. The analytical results indicate that χ and η cross helicities are not strict constant of motion in the non-ideal MHD and show the rate of dissipation which is compared to the dissipation of other topological constants of motion.
Keywords: 
Subject: Physical Sciences  -   Fluids and Plasmas Physics

1. Introduction

Topological invariants have always been useful for several decades in various physical systems, and there are such invariants in MHD flows. For example, the importance of two helicities i.e., magnetic helicity and cross-helicity have long been discussed in relation to the controlled nuclear fusion problem and in numerous astrophysical scenarios. In the absence of dissipation, cross helicity is conserved. Earlier work [1,2,3] have studied the relations between the helicities and symmetries of ideal MHD. MHD connects Maxwell’s equations with hydrodynamics of highly conductive flows to explain the macroscopic behaviour of conducting fluid such as plasma. Considering the fact that ideal MHD does not describe precisely the behaviour of real plasmas was the paramount motivation for the present work. Some important realistic processes are missing in the ideal description such as resistive heating, heat conduction and viscous effects. Viscosity plays an important role on dissipation scale while investigating the plasma turbulence in solar wind and elsewhere. Similarly magnetic diffusivity is one of the reasons for the happening of magnetic reconnection phenomena. Thermal conductivity is also a substantial process needed to understand real plasmas. It causes the perturbations of physical variables to spread out through the plasmas. These essential properties of all three dissipative processes are the stimulus for the authors to study this current analysis.
The mathematical expression for cross helicity is defined as [4,5]:
H C = B · v d 3 x
in which the integral is taken over the entire flow domain. Here H C is conserved for barotropic or incompressible MHD (but not for non-barotropic MHD) and is given a topological interpretation in terms of the knottiness of magnetic and flow field lines. This correlation is having great importance in case of Alfven waves. However, it shows relation with magnetosonic waves in the compressible case [6]. A generalization for non-barotropic MHD of this quantity was given by [7,8]. This resembles the generalization of barotropic fluid dynamics conserved quantities including helicity to non barotropic flows including topological constants of motion derived by [9]. Conservation law of cross helicity for non barotropic MHD has been discussed by [10] in a multi-symplectic formulation of MHD. A potential vorticity conservation equation for non-barotropic MHD was derived by [11] by using Noether’s second theorem.
Recently the non-barotropic cross-helicity was generalized using additional label translation symmetry groups ( χ and η translations) [12], this led to additional topological conservation laws, the χ and η cross-helicities. The functions χ and η are sometimes denoted ‘Euler potentials’, ‘Clebsch variables’ and also ‘flux representation functions’ [13].
The concept of metage as a label for fluid elements along a vortex line in ideal fluids was first introduced by Lynden-Bell & Katz [14]. A translation group of this label was found to be connected to the conservation of Moffat’s [15] helicity by Yahalom [3] using a Lagrangian variational principle. The concept of metage was later generalized by Yahalom & Lynden-Bell [1] for barotropic MHD, but now as a label for fluid elements along magnetic field lines which are comoving with the flow in the case of ideal MHD. Yahalom & Lynden-Bell [1] has also shown that the translation group of the magnetic metage is connected to Woltjer [4,5] conservation of cross helicity for barotropic MHD. Recently the concept of metage was generalized also for non barotropic MHD in which magnetic field lines lie on entropy surfaces [16]. This was later generalized by dropping the entropy condition on magnetic field lines [17]. In those papers the metage translation symmetry group was used to generate a non-barotropic cross helicity generalization using a Lagrangian variational principle.
Cross helicity is expected to play an important role in several MHD plasma phenomena such as global magnetic-field generation, turbulence suppression, etc. It provides a measure of the degree of linkage of the vortex tubes of the velocity field with the flux tubes of the magnetic field. Cross helicity plays an important role in turbulent dynamo [18]. The cross helicity density conservation law for barotropic flows is important in MHD turbulence theory [19,20,21]. Plasma velocity and magnetic field measurements from the Voyager 2 mission are used to study solar wind turbulence in the slow solar wind [22] and characterize its cross helicity. [23] has discussed MHD turbulence in his review paper in detail. He has examined the Alfvénic MHD turbulence with zero and non-zero helicities. The energy fluxes of MHD turbulence provide a measure for transfers of energy among velocity and magnetic fields [24,25].
Magnetic helicity is also one of the important topological invariants in fluid dynamics. [26] has already been showed the well conservation of magnetic helicity in turbulent flows. However when flux tubes diffuse through one another on resistive time scales, magnetic helicity dissipates[27]. [28] has studied the role of magnetic helicity in the plasmas stabilization in details by doing series of experiments and numerical simulations. Further, its important role in determining the structures, dynamics and heating of the solar corona has been well explained by [29].
The importance of the χ and η cross helicities is expected to be not less important than their more older predecessors the magnetic and generalized cross helicities, for stability and dynamics of MHD [12]. Thus it is of paramount importance to study the effects of non-ideal processes on their development.
The plan of this paper is as follows: There are three sections in the current paper. Section 2 describes the basic quantities and equations in non-ideal MHD. Calculations for time derivative of χ helicity is performed in section 3 and section 4 deals with the mathematics of η helicity.

2. Standard Formulation of Non-ideal Non-barotropic MHD

The standard set of equations solved for non-ideal non-barotropic MHD is given below (Here we use the EMU system of units):
v n t = ( v n · ) v n p n ρ n + J n × B n ρ n ϕ + 1 ρ n σ i k x k
ρ n t + · ( ρ n v n ) = 0
· B n = 0
B n t = × ( v n × B n ) + η v 4 π 2 B n
ρ n T n d s n d t = σ i k v n i x k + η v J n 2 + · ( k T n )
The following notations are utilized: t is the partial temporal derivative. has its standard meaning in vector calculus. We use the subscript n to describe non-ideal processes. Thus v is the ideal velocity and v n is the velocity for non-ideal fluid etc. . ρ n is the density, p n is the pressure respectively which depends through the equation of state on the density and entropy s n (the non-barotropic case), T n is the temperature and ϕ is a potential.
The stress tensor is defined as:
σ i k = μ v ( v n i x k + v n k x i 2 3 δ i k v n l x l )
in which μ v is a coefficient of kinematic viscosity (not to be confused with the metage μ defined in [12]). Notice that we take coefficient of second viscosity (or volume viscosity) is zero for the sake of simplicity. The entropy equation (6) depends on the heat conduction coefficient k. According to classical kinetic theory, viscosity arises from collisions between particles. The justification for those equations and the conditions under which they apply can be found in standard books on MHD (see for example [30,31,32,33]).
The current density J n and the magnetic field b n of MHD are related by Ampere’s law:
× B n = 4 π J n
in which the displacement current is neglected. Equation (5) depends on the non-ideal magnetic diffusivity η v (not to be confused with the label η introduced in [12]). In the limit of ideal flows all the non-ideal coefficients μ v , η v , k tend to zero and the ideal MHD equations are recovered we prime the difference between ideal and non-ideal quantities, for example: v = v n v . It is clear that in the ideal limit all primed quantities tend to zero, for example: lim ( μ v , η v , k ) 0 v = 0 .

3. Direct derivation of the constancy of non-barotropic χ cross-helicity

We introduce the abstract ‘magnetic fields’ as follows [34]:
B χ = μ × η
Non-barotropic χ cross-helicity is given by:
H C N B χ = v n t · B χ d 3 x
Here the topological non-ideal velocity field is defined as v n t = v n σ n s n [34], σ n is auxiliary variable, which depends on the Lagrangian time integral of the temperature i.e.:
d σ n d t = T n .
Please refer [16] for the detailed justification for the relation of non-barotropic cross helicity.
Taking the temporal derivative of the non-barotropic χ cross-helicity:
d H C N B χ d t = d 3 x ( v n t · B χ t + B χ · v n t t )
Now, find out the value of first term of RHS, we take the time derivative of equation (9):
B χ t = ( μ t ) × η + μ × ( η t )
Notice that both the labels are comoving and conserved under ideal material derivative [12], thus:
μ t + ( v · ) μ = 0 .
Similarly:
η t + ( v · ) η = 0 .
So:
B χ t = [ ( v · μ ) ] × η + μ × [ ( v · η ) ]
Using vector identity:
× ( ψ a ) = ψ × a + ψ × a
equation (16) takes the form:
B χ t = [ × { μ ( v · η ) η ( v · μ ) } ]
Now, with the help of identity:
A × ( B × C ) = B ( A · C ) C ( A · B ) ,
we obtain:
B χ t = [ × { v × ( μ × η ) } ] .
Substituting B χ defined in equation (9), we obtain:
B χ t = × ( v × B χ ) .
Next, we dot both sides with v n t :
v n t · B χ t = v n t · × ( v × B χ ) ,
and using a well known vector identity we obtain:
v n t · B χ t = · { ( v × B χ ) × v n t } + ( v × B χ ) · ω n t ,
where we define the topological vorticity of the non-ideal flow field as:
ω n t × v n t .
Next we calculate second term on the RHS of equation (12):
t v n t · B χ = B χ · t ( v n σ n s n ) = B χ · ( t v n t σ n s n σ n t s n ) .
Now we simplify right hand side of equation (25) in three steps. The first term will be calculated with the help of equation (2):
v n t = ( v n × ω n ) + J n × B n ρ n ( v n 2 2 ) w n + T n s n ϕ + 1 ρ n σ i k x k ,
in which the non-ideal vorticity is:
ω n × v n
and we have used the thermodynamical identity:
d w n = d ε n + d ( p n ρ n ) = T n d s n + 1 ρ n d p n w n = T n s n + 1 ρ n p n .
Thus:
B χ · v n t = B χ · ( v n × ω n ) ( v n 2 2 + w n ) + T n s n ϕ + B χ i ρ n σ i k x k + B χ · J n × B n ρ n .
In the second term we use equation (11) to obtain:
σ n t s n = ( v n · σ n T n ) s n .
In the third term we use equation for the rate of change of entropy in non-ideal MHD also known as the heat equation (6):
σ n s n t = σ n [ v n · s n 1 ρ n T n σ i k v n i x k η v ρ n T n J n 2 1 ρ n T n · ( k T n ) ] .
Combining equation (29), equation (30) and equation (31) we obtain:
B χ · v n t t = B χ · [ ( v n × ω n t ) + { σ n ( v n · s n ) v n 2 2 w n ϕ } σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] + B χ i ρ n σ i k x k + B χ · J n × B n ρ n
in which the current density is given by
J n = × B n 4 π · J n = 0 .
Now,
B χ · J n × B n ρ n = 1 ρ n J n · B n × B χ
B n × B χ = B n × ( μ × η ) = μ ( B n · η ) η ( B n · μ )
B n × B χ = μ ( B · η ) + μ ( B · η ) η ( B · μ ) η ( B · μ ) ,
however, using equations ( 15 ) and ( 17 ) of [12]: B · η = 0 and B · μ = ρ . And thus:
B n × B χ = μ ( B · η ) η ( B · μ ) ρ η
So:
B n × B χ = B × ( μ × η ) ρ η = B × B χ ρ η
It thus follows that:
B χ · J n × B n ρ n = ρ ρ n ( J n · η ) + J n · ( B × B χ ) ρ n = ρ ρ n · ( J n η ) + J n · ( B × B χ ) ρ n .
Now equation (32) can be written as:
B χ · v n t t = B χ · [ ( v n × ω n t ) + { σ n ( v n · s n ) v n 2 2 w n ϕ } σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] + B χ i ρ n σ i k x k ρ ρ n · ( J n η ) + J n · ( B × B χ ) ρ n .
Combining equation (23) and equation (40) and taking into account that:
B χ · ( v n × ω t ) = ( v n × B χ ) · ω t
We obtain:
v n t · B χ t + B χ · v n t t = · { ( v × B χ ) × v n t } + B χ · { σ n ( v n · s n ) v n 2 2 w n ϕ } + B χ i ρ n σ i k x k B χ · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ρ ρ n · ( J n η ) + J n · ( B × B χ ) ρ n + ω n t · ( B χ × v ) .
Now substituting equation (42) into equation (12) we obtain:
d H C N B χ d t = · { ( v × B χ ) × v n t } + B χ · { σ n ( v n · s n ) v n 2 2 w n ϕ } d 3 x + [ B χ i ρ n σ i k x k B χ · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x [ ρ ρ n · ( J n η ) J n · ( B × B χ ) ρ n ω n t · ( B χ × v ) ] d 3 x .
Using Gauss divergence theorem, we can write part of this integral as a surface integral:
d H C N B χ d t = [ ( v × B χ ) × v n t + B χ { σ n ( v n · s n ) v n 2 2 w n ϕ } J n η ] · d S + [ B χ i ρ n σ i k x k B χ · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x + [ ρ ρ n · ( J n η ) + J n · ( B × B χ ) ρ n + ω n t · ( B χ × v ) ] d 3 x .
Here, the surface integral encapsulates the volume for which the χ non barotropic cross helicity is calculated as well as a cut in case that η is multiple valued. If the surface integral vanishes:
d H C N B χ d t = [ B χ i ρ n σ i k x k B χ · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x + [ ρ ρ n · ( J n η ) + J n · ( B × B χ ) ρ n + ω n t · ( B χ × v ) ] d 3 x .
To conclude we notice that is not possible to partition the time derivative of the non barotropic χ cross helicity in accordance with the different non ideal processes: viscosity, finite conductivity and heat conductivity. Each of this processes may contribute to the primed quantities directly or indirectly. In the limit that all non-ideal coefficients tend to zero it is easy to see that the non barotropic χ cross helicity is indeed conserved:
d H C N B χ d t = 0 .
As is expected for an ideal flow.

4. Direct derivation of the constancy of non-barotropic η cross-helicity

We introduce the abstract ‘magnetic field’ as follows [34]:
B η = χ × μ
Non-barotropic η cross-helicity is given by:
H C N B η = v n t · B η d 3 x .
Taking the temporal derivative of the non-barotropic η cross-helicity
d H C N B η d t = d 3 x ( v n t · B η t + B η · v n t t )
B η t = ( χ t ) × μ + χ × ( μ t )
Again, both the labels are comoving and conserved under an ideal material derivative [12]:
μ t + ( v · ) μ = 0 , χ t + ( v · ) χ = 0 .
So:
B η t = [ ( v · χ ) ] × μ + χ × [ ( v · μ ) ]
Using vector identity:
× ( ψ μ ) = ψ × μ ,
equation (52) takes the form.
B η t = × { χ ( v · μ ) μ ( v · χ ) } .
Now, with the help of identity:
A × ( B × C ) = B ( A · C ) C ( A · B ) ,
we obtain:
B η t = × { v × ( χ × μ ) } = × ( v × B η ) .
Thus:
v n t · B η t = v n t · [ × ( v × B η ) ] = · { ( v × B η ) × v n t } + ( v × B η ) · ω n t
Next we calculate second term:
t v n t · B η = B η · t ( v n σ n s n ) = B η · ( t v n t σ n s n σ n t s n )
Now we simplify right hand side of equation (58) in three steps. First term will be calculated with the help of momentum equation (26) for the non ideal case:
B η · v n t = B η · ( v n × ω n ) ( v n 2 2 + w n ) + T n s n ϕ + B η ρ n σ i k x k + B η · J n × B n ρ n
For the second and third terms we use equation (30) and equation (31) to obtain:
B η · v n t t = B η · [ ( v n × ω n t ) + { σ n ( v n · s n ) v n 2 2 w n ϕ } σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] + B η i ρ n σ i k x k + B η · J n × B n ρ n
Now:
B η · J n × B n ρ n = 1 ρ n J n · B n × B η
And we have:
B n × B η = B n × ( χ × μ ) = χ ( B n · μ ) μ ( B n · χ ) = χ ( B · μ ) + χ ( B · μ ) μ ( B · χ ) μ ( B · χ ) ,
however, using equations ( 15 ) and ( 17 ) of [12] which dictate:
B · χ = 0 , B · μ = ρ ,
it follows that:
B n × B η = χ ( B · μ ) μ ( B · χ ) + ρ χ = B × ( χ × μ ) + ρ χ = B × B η + ρ χ
Inserting equation (64) into equation (61) leads to the following identity:
B η · J n × B n ρ n = ρ ρ n ( J n · χ ) + J n · ( B × B η ) ρ n = ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n
So equation (60) becomes:
B η · v n t t = B η · [ ( v n × ω n t ) + { σ n ( v n · s n ) v n 2 2 w n ϕ } σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] + B η i ρ n σ i k x k + ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n .
Combining equation (57) and equation (66) we obtain:
v n t · B η t + B η · v n t t = · { ( v × B η ) × v n t } + B η · { σ n ( v n · s n ) v n 2 2 w n ϕ } + B η i ρ n σ i k x k B η · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] + ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n + ω n t · ( B η × v ) .
Now substituting equation (67) into equation (49), we obtain:
d H C N B η d t = · { ( v × B η ) × v n t } + B η · { σ n ( v n · s n ) v n 2 2 w n ϕ } d 3 x + [ B η i ρ n σ i k x k B η · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x + [ ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n + ω n t · ( B η × v ) ] d 3 x .
Using Gauss divergence theorem, we obtain:
d H C N B η d t = [ ( v × B η ) × v n t + B η { σ n ( v n · s n ) v n 2 2 w n ϕ } + J n χ ] · d S + [ B η i ρ n σ i k x k B η · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x + [ ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n + ω n t · ( B η × v ) ] d 3 x .
Here, the surface integral encapsulates the volume for which the χ non barotropic cross helicity is calculated ( χ is usually single valued so no cut is required). If the surface integral vanishes:
d H C N B η d t = [ B η i ρ n σ i k x k B η · [ σ n { 1 ρ n T n σ i k v n i x k + η v ρ n T n J n 2 + 1 ρ n T n · ( k T n ) } ] ] d 3 x + [ ρ ρ n · ( J n χ ) + J n · ( B × B η ) ρ n + ω n t · ( B η × v ) ] d 3 x .
To conclude we notice that is not possible to partition the time derivative of the non barotropic η cross helicity in accordance with the different non ideal processes: viscosity, finite conductivity and heat conductivity. Each of this processes may contribute to the primed quantities directly or indirectly. In the limit that all non-ideal coefficients tend to zero it is easy to see that the non barotropic η cross helicity is indeed conserved:
d H C N B η d t = 0 .
As is expected for an ideal flow.

5. Conclusion

The current paper highlights on the constancy of two new topological invariants generalized χ and η cross helicities when fluid flow is not ideal, but the non ideal processes are less significant. We have showed that the helicities are not conserved and their time derivatives depends on dissipative processes present in the system. Non-zero cross helicity has an important consequences related to transport and [35] has been showed the close relation between momentum transfer and cross helicity. [36] has discussed its role in MHD turbulence by high resolution direct numerical simulations. Cross helicity is proportional to the correlation between velocity and magnetic field fluctuation and measures the relative importance of Alfvén waves in global fluctuation and [37] have been predicted the generation of interstellar turbulence caused by nonlinear interactions among Shear Alfvén waves. Many studies in the literature have shown that the cross helicity is correlated to the self production of turbulence.
The cascade process in MHD turbulence has been studied by [38] in details and concluded "cross helicity blocks the spectral energy transfer in MHD turbulence and results in energy accumulation in the system. This accumulation proceeds until the vortex intensification compensates the decreasing efficiency of nonlinear interactions". [39] examined the effect of cross helicity on the decay of isotropic MHD turbulence and concluded that an initial non zero cross helicity makes imbalanced MHD turbulence. The subtle anisotropic effect of cross helicity can be caused for the same. The technological (fusion) and astrophysical importance of cross helicity suggest that the newly discovered χ and η cross helicities may also play a pivotal rule. The reason for this is two fold. First, as we have shown above all processes that changes the quantities are slow dissipative processes thus any fast process conserve the χ and η cross helicities. Second, given that those quantities are approximately constant they serve as lower bounds to other quantities such as "energy":
H C N B χ = B χ · v t d 3 x 1 2 B χ 2 + v t 2 d 3 x ,
H C N B χ = B χ · v t d 3 x v t 2 d 3 x B χ 2 d 3 x ,
H C N B η = B η · v t d 3 x 1 2 B η 2 + v t 2 d 3 x ,
H C N B η = B η · v t d 3 x v t 2 d 3 x B η 2 d 3 x ,
And thus may prevent at least part of the most fast and dangerous instabilities which may provide some hope that controlled fusion is indeed feasible.

References

  1. Yahalom, A.; Lynden-Bell, D. Simplified variational principles for barotropic magnetohydrodynamics. Journal of Fluid Mechanics 2008, 607, 235. [Google Scholar] [CrossRef]
  2. Yahalom, A. Aharonov–Bohm effects in magnetohydrodynamics. Physics Letters A 2013, 377, 1898–1904. [Google Scholar] [CrossRef]
  3. Yahalom, A. Helicity conservation via the Noether theorem. Journal of Mathematical Physics 1995, 36, 1324–1327. [Google Scholar] [CrossRef]
  4. Woltjer, L. A theorem on force-free magnetic fields. Proceedings of the National Academy of Sciences of the United States of America 1958, 44, 489. [Google Scholar] [CrossRef]
  5. Woltjer, L. On hydromagnetic equilibrium. Proceedings of the National Academy of Sciences of the United States of America 1958, 44, 833. [Google Scholar] [CrossRef] [PubMed]
  6. Yokoi, N. Mass and internal-energy transports in strongly compressible magnetohydrodynamic turbulence. Journal of Plasma Physics 2018, 84. [Google Scholar]
  7. Webb, G.; Dasgupta, B.; McKenzie, J.; Hu, Q.; Zank, G. Local and nonlocal advected invariants and helicities in magnetohydrodynamics and gas dynamics I: Lie dragging approach. Journal of Physics A: Mathematical and Theoretical 2014, 47, 095501. [Google Scholar] [CrossRef]
  8. Webb, G.; Dasgupta, B.; McKenzie, J.; Hu, Q.; Zank, G. Local and nonlocal advected invariants and helicities in magnetohydrodynamics and gas dynamics: II. Noether’s theorems and Casimirs. Journal of Physics A: Mathematical and Theoretical 2014, 47, 095502. [Google Scholar] [CrossRef]
  9. Mobbs, S. Some vorticity theorems and conservation laws for non-barotropic fluids. Journal of Fluid Mechanics 1981, 108, 475–483. [Google Scholar] [CrossRef]
  10. Webb, G.; McKenzie, J.; Zank, G. Multi-symplectic magnetohydrodynamics: II, addendum and erratum. Journal of Plasma Physics 2015, 81. [Google Scholar] [CrossRef]
  11. Webb, G.; Mace, R. Potential vorticity in magnetohydrodynamics. Journal of Plasma Physics 2015, 81. [Google Scholar] [CrossRef]
  12. Yahalom, A. A new diffeomorphism symmetry group of non-barotropic magnetohydrodynamics. Journal of Physics: Conference Series. IOP Publishing 2019, 1194, 012113. [Google Scholar]
  13. Hazeltine, R.D.; Meiss, J.D. Plasma confinement; Courier Corporation, 2003. [Google Scholar]
  14. Lynden-Bell, D.; Katz, J. Isocirculational Flows and their Lagrangian and Energy principles. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 1981, 378, 179–205. [Google Scholar]
  15. Vladimirov, V.A.; Moffatt, H. On general transformations and variational principles for the magnetohydrodynamics of ideal fluids. Part 1. Fundamental principles. Journal of fluid mechanics 1995, 283, 125–139. [Google Scholar] [CrossRef]
  16. Yahalom, A. A conserved local cross helicity for non-barotropic MHD. Geophysical & Astrophysical Fluid Dynamics 2017, 111, 131–137. [Google Scholar]
  17. Yahalom, A. Metage symmetry group of non-barotropic magnetohydrodynamics and the conservation of cross helicity. Quantum Theory And Symmetries; Springer, 2017; pp. 387–402. [Google Scholar]
  18. Yokoi, N. Cross helicity and related dynamo. Geophysical & Astrophysical Fluid Dynamics 2013, 107, 114–184. [Google Scholar]
  19. Zhou, Y.; Matthaeus, W.H. Transport and turbulence modeling of solar wind fluctuations. Journal of Geophysical Research: Space Physics 1990, 95, 10291–10311. [Google Scholar] [CrossRef]
  20. Zhou, Y.; Matthaeus, W.H. Models of inertial range spectra of interplanetary magnetohydrodynamic turbulence. Journal of Geophysical Research: Space Physics 1990, 95, 14881–14892. [Google Scholar] [CrossRef]
  21. Zank, G.; Dosch, A.; Hunana, P.; Florinski, V.; Matthaeus, W.; Webb, G. The transport of low-frequency turbulence in astrophysical flows. I. Governing equations. The Astrophysical Journal 2011, 745, 35. [Google Scholar] [CrossRef]
  22. Iovieno, M.; Gallana, L.; Fraternale, F.; Richardson, J.; Opher, M.; Tordella, D. Cross and magnetic helicity in the outer heliosphere from Voyager 2 observations. European Journal of Mechanics-B/Fluids 2016, 55, 394–401. [Google Scholar] [CrossRef]
  23. Verma, M.K. Statistical theory of magnetohydrodynamic turbulence: recent results. Physics Reports 2004, 401, 229–380. [Google Scholar] [CrossRef]
  24. Verma, M.; Sharma, M.; Chatterjee, S.; Alam, S. Variable energy fluxes and exact relations in Magnetohydrodynamics turbulence. Fluids 2021, 6, 225. [Google Scholar] [CrossRef]
  25. Verma, M.K. Energy transfers in fluid flows: multiscale and spectral perspectives; Cambridge University Press, 2019. [Google Scholar]
  26. Faraco, D.; Lindberg, S. Proof of Taylor’s conjecture on magnetic helicity conservation. Communications in Mathematical Physics 2020, 373, 707–738. [Google Scholar] [CrossRef]
  27. Barnes, C.W.; Fernandez, J.; Henins, I.; Hoida, H.; Jarboe, T.; Knox, S.; Marklin, G.; McKenna, K. Experimental determination of the conservation of magnetic helicity from the balance between source and spheromak. The Physics of fluids 1986, 29, 3415–3432. [Google Scholar] [CrossRef]
  28. Candelaresi, S.; Del Sordo, F. Stability of plasmas through magnetic helicity. arXiv preprint arXiv:2112.01193 2021.
  29. Knizhnik, K.J.; Antiochos, S.K.; Klimchuk, J.A.; DeVore, C.R. The role of magnetic helicity in coronal heating. The Astrophysical Journal 2019, 883, 26. [Google Scholar] [CrossRef]
  30. Batchelor, G.K. An Introduction to Fluid Dynamics,(1967). Cambridge,: UP xviii 1967, 615. [Google Scholar]
  31. Sturrock, P.A. Plasma physics: an introduction to the theory of astrophysical, geophysical and laboratory plasmas; Cambridge University Press, 1994. [Google Scholar]
  32. Landau, L.; Lifshitz, E. CHAPTER V - THERMAL CONDUCTION IN FLUIDS. In Fluid Mechanics (Second Edition), Second Edition ed.; Landau, L., Lifshitz, E., Eds.; Pergamon, 1987; pp. 192–226. [Google Scholar] [CrossRef]
  33. Ogilvie, G.I. Astrophysical fluid dynamics. Journal of Plasma Physics 2016, 82. [Google Scholar] [CrossRef]
  34. Yahalom, A.; Qin, H. Noether currents for Eulerian variational principles in non-barotropic magnetohydrodynamics and topological conservations laws. Journal of Fluid Mechanics 2021, 908. [Google Scholar] [CrossRef]
  35. Heinonen, R.; Diamond, P.; Katz, M.; Ronimo, G. On the role of cross-helicity in β-plane magnetohydrodynamic turbulence. arXiv preprint arXiv:2103.08091 2021.
  36. Perez, J.C.; Boldyrev, S. Role of cross-helicity in magnetohydrodynamic turbulence. Physical review letters 2009, 102, 025003. [Google Scholar] [CrossRef] [PubMed]
  37. Goldreich, P.; Sridhar, S. Toward a theory of interstellar turbulence. 2: Strong alfvenic turbulence. The Astrophysical Journal 1995, 438, 763–775. [Google Scholar] [CrossRef]
  38. Mizeva, I.; Stepanov, R.; Frik, P. The cross-helicity effect on cascade processes in MHD turbulence. Doklady Physics. Citeseer, 2009, 54, 93–97. [Google Scholar] [CrossRef]
  39. Briard, A.; Gomez, T. The decay of isotropic magnetohydrodynamics turbulence and the effects of cross-helicity. Journal of Plasma Physics 2018, 84. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated