1 Introduction
Turbulence in electrically conducting fluids and plasmas is of relevance to a variety of processes in geophysical and astrophysical situations, as well as in industry (Weiss & Proctor Reference Weiss and Proctor2014; Davidson Reference Davidson2016) and for nuclear fusion under magnetic confinement. For example, the solar wind is turbulent (Bruno & Carbone Reference Bruno and Carbone2013), convection-driven turbulence occurs in planetary cores and in the outer layers of stars (Jones Reference Jones2011), turbulence on ion and electron scales affects plasma confinement in magnetic confinement fusion reactors (Freidberg Reference Freidberg2007), and the heat transfer in liquid metal cooling applications is dependent on the level of turbulence in the flow (Davidson Reference Davidson1999). Even though these systems are very different in terms of features like the presence of strong background magnetic fields, the level of magnetic field fluctuations, temperature gradients, density fluctuations, domain geometry or the level of collisionality, they nonetheless share fundamental nonlinear processes that define energy conversion and inter-scale energy transfers, at least on scales where the fluid approximation is applicable. Even in the simplest case of magnetohydrodynamic (MHD) turbulence, despite considerable theoretical progress (e.g. Goldreich & Sridhar Reference Goldreich and Sridhar1995; Biskamp Reference Biskamp2003; Zhou, Matthaeus & Dmitruk Reference Zhou, Matthaeus and Dmitruk2004; Petrosyan et al. Reference Petrosyan, Balogh, Goldstein, Léorat, Marsch, Petrovay, Roberts, von Steiger and Vial2010; Brandenburg, Sokoloff & Subramanian Reference Brandenburg, Sokoloff and Subramanian2012; Tobias & Cattaneo Reference Tobias and Cattaneo2013; Beresnyak Reference Beresnyak2019; Oughton & Matthaeus Reference Oughton and Matthaeus2020; Schekochihin Reference Schekochihin2022), the physical nature of these processes remain opaque.
Moreover, the typical parameter ranges in which MHD turbulence develops in Nature are far from those attainable with direct numerical simulation (DNS) (Plunian, Stepanov & Frick Reference Plunian, Stepanov and Frick2013; Miesch et al. Reference Miesch, Matthaeus, Brandenburg, Petrosyan, Pouquet, Cambon, Jenko, Uzdensky, Stone and Tobias2015; Schmidt Reference Schmidt2015). As a consequence, the demand for approximations and subgrid-scale (SGS) models for large-eddy simulation (LES) of MHD turbulence that are able to capture the effects of unresolved small-scale fluctuations – that govern important processes such as magnetic reconnection and plasma heating – is increasing (Miesch et al. Reference Miesch, Matthaeus, Brandenburg, Petrosyan, Pouquet, Cambon, Jenko, Uzdensky, Stone and Tobias2015). However, constructing such models is a challenge due to small-scale anisotropy (Shebalin, Matthaeus & Montgomery Reference Shebalin, Matthaeus and Montgomery1983; Oughton, Priest & Matthaeus Reference Oughton, Priest and Matthaeus1994; Goldreich & Sridhar Reference Goldreich and Sridhar1997; Tobias & Cattaneo Reference Tobias and Cattaneo2013), strong intermittency in the magnetic fluctuations as observed in numerical simulations (e.g. Mininni & Pouquet Reference Mininni and Pouquet2009; Sahoo, Perlekar & Pandit Reference Sahoo, Perlekar and Pandit2011; Yoshimatsu et al. Reference Yoshimatsu, Schneider, Okamoto, Kawahara and Farge2011; Rodriguez Imazio et al. Reference Rodriguez Imazio, Martin, Dmitruk and Mininni2013; Meyrand, Kiyani & Galtier Reference Meyrand, Kiyani and Galtier2015) and in the solar wind (e.g. Veltri Reference Veltri1999; Salem et al. Reference Salem, Mangeney, Bale and Veltri2009; Wan et al. Reference Wan, Osman, Matthaeus and Oughton2012; Matthaeus et al. Reference Matthaeus, Wan, Servidio, Greco, Osman, Oughton and Dmitruk2015), insufficient magnetic-field amplification and dynamo growth observed in LES of MHD turbulence (Haugen & Brandenburg Reference Haugen and Brandenburg2006), and whether magnetic-field fluctuations are maintained by the flow or by an external electromagnetic force (Alexakis & Chibbaro Reference Alexakis and Chibbaro2022). For a summary of the SGS modelling effort and its challenges, we refer to the review articles by Miesch et al. (Reference Miesch, Matthaeus, Brandenburg, Petrosyan, Pouquet, Cambon, Jenko, Uzdensky, Stone and Tobias2015) and Schmidt (Reference Schmidt2015).
Here, we focus on energy transfer across scales in statistically stationary homogeneous MHD turbulence in a saturated nonlinear dynamo regime without a mean magnetic field, and with negligible levels or cross- and magnetic helicity. The total energy cascade in this case is direct (Aluie & Eyink Reference Aluie and Eyink2010), transferring energy from the large to the small scales in a scale-local fashion. Herein, we use the term cascade to indicate a mean flux and refer to pointwise upscale and downscale energy transfers as inverse and direct transfer events, respectively, or in the former case also as backscatter. The aims of this paper are: (i) to understand the physical mechanisms that govern the MHD energy cascade and (ii) to quantify their importance and provide guidance for SGS modelling (Johnson Reference Johnson2022). In terms of turbulence theory, this corresponds to understanding physical properties of the SGS stresses as a function of scale. Eyink (Reference Eyink2006) introduced a viable approach for this that involves expanding the SGS tensors in terms of vector field gradients. Here, we follow a filtering approach, generalising an exact gradient-based decomposition of the hydrodynamic (HD) energy fluxes (Johnson Reference Johnson2020, Reference Johnson2021) to coupled advection–diffusion equations and hence to the MHD equations. This methodology distinguishes between terms that are local in scale, corresponding to the first term in the gradient expansion and those which are truly multi-scale, providing a closed expression for the remainder of the series expansion. Expressing SGS stresses through vector-field gradients results in a decomposition of the energy fluxes in terms of different tensorial contractions between strain-rate, vorticity, current and magnetic strain, and as such facilitates the physical interpretation of such sub-fluxes. The provision of closed expressions allows for a quantification of the relative contribution of all terms to the energy cascade using data obtained by direct numerical simulation (DNS). In homogeneous and isotropic HD turbulence, where only the inertial term is present, the decomposition identifies three processes that transfer kinetic energy across scales, vortex stretching, strain self-amplification and strain-vorticity alignment, and quantifies their relative contribution to the energy cascade (Johnson Reference Johnson2020, Reference Johnson2021). Similarly, the direct cascade of kinetic helicity is carried by three different processes, vortex flattening, vortex twisting and vortex entanglement (Capocci et al. Reference Capocci, Johnson, Oughton, Biferale and Linkmann2023).
In MHD, the total energy transfer can be split into four subfluxes, Inertial, Maxwell, Dynamo and Advection,Footnote 1 with the former two originating from the Reynolds and Maxwell stresses in the momentum equation, and the latter two from stresses in the induction equation that result in the advection and bending/stretching of magnetic field lines by the flow. As Dynamo and Advection terms have a common physical origin, the electric field, often only their sum is considered in a priori analyses of DNS data (Aluie Reference Aluie2017; Offermans et al. Reference Offermans, Biferale, Buzzicotti and Linkmann2018; Alexakis & Chibbaro Reference Alexakis and Chibbaro2022) and a posteriori in LES (Zhou & Vahala Reference Zhou and Vahala1991; Müller & Carati Reference Müller and Carati2002; Grete et al. Reference Grete, Vlaykov, Schmidt and Schleicher2016; Kessar, Balarac & Plunian Reference Kessar, Balarac and Plunian2016; Vlaykov et al. Reference Vlaykov, Grete, Schmidt and Schleicher2016). However, in the present work, it will prove instructive to consider them separately. Here, we generalise and apply the aforementioned decomposition to each of these four subfluxes.
In doing so, we identify a single process, current-sheet thinning, to be the main contributor to the forward cascade of magnetic energy. Contributions from strain-induced current-filament stretching, the formal analogue to vortex stretching, are subdominant. Furthermore, we find that vortex-stretching and strain self-amplification are strongly suppressed at all length scales. Instead the back-reaction of current-sheet thinning on the flow through the Lorentz force constitutes the main transfer of kinetic energy from large to small scales, with contributions related to current-filament stretching turning out to be negligible.
The structure of the paper is as follows: we begin in § 2 with an outline of how the generalised method can be applied to obtain MHD energy subfluxes. In § 3, we discuss the numerical details and the associated datasets on which we performed the filtering analysis. In § 4, we consider each subflux decomposition, showing results for both mean terms and fluctuations. Ramifications of those results for SGS modelling are considered in § 5. In § 6, we discuss our main results and indicate future work directions. Several appendices flesh out some aspects of the derivations and analysis.
2 Theory
In this section, we begin by sketching the derivation of the coarse-grained energy equations for MHD and giving the definitions of the scale-space energy fluxes that appear in them. Subsequently, we show how each flux can be decomposed in terms of physically distinct contributions, and discuss their physical interpretations.
2.1 Coarse-grained energy equations
Our starting point is incompressible three-dimensional (3-D) homogeneous MHD turbulence. The primary dynamical variables are then the fluctuation velocity $\boldsymbol {u} (\boldsymbol {x}, t)$ and the fluctuation magnetic field $\boldsymbol {b} (\boldsymbol {x},t)$, where we measure the latter in Alfvén speed units: $\boldsymbol {b} / \sqrt {4{\rm \pi} \rho } \to \boldsymbol {b}$, with $\rho$ the uniform mass density. We consider situations with no mean magnetic field since these are more likely to exhibit global isotropy. The governing equations, with allowance for hyper-dissipation, are
Here, $p$ is the pressure, $\boldsymbol {F}$ is a (large-scale) velocity forcing, $\nu _\alpha$ and $\mu _\alpha$ are the hyper-viscosity and hyper-resistivity, and $\alpha$ denotes the power of the Laplacian operator employed in the hyper-dissipation. Standard Laplacian dissipation corresponds to the case $\alpha =1$.
The MHD variables, and equations, may be spatially coarse-grained using a suitable filtering field, $G^\ell (\boldsymbol {r} )$ (Germano Reference Germano1992; Aluie Reference Aluie2017). The role of $G^\ell (\boldsymbol {r} )$ is to strongly suppresses structure at scales less than the filtering scale $\ell$. For example, the filtered velocity field is
This can be interpreted as a weighted average of $\boldsymbol {u}$ centred on the position $\boldsymbol {x}$. The weighting function $G^\ell$ decays very rapidly to zero at distances greater than a few $\ell$ from $\boldsymbol {x}$ and satisfies some other weak restrictions, such as smoothness and having a volume integral of unity. Filtering is a linear operation and commutes with differentiation, properties of which we will make considerable use below. From § 2.2 onwards, we will specialise to a Gaussian filter, but in this section, a specific choice of filter is not needed.
Coarse-graining of (2.1)–(2.2) introduces four SGS stress tensors, $\tau ^\ell (\cdot, \cdot )$, associated with the advective-type nonlinear terms (those containing a $\partial _j$). These each have the form
where $\boldsymbol {f}$ and $\boldsymbol {g}$ are the solenoidal vectors appearing in a $g_j \partial _j \, f_i$ advective-type term. We remark that with this notation, the advecting field is the second argument in a $\tau ^\ell (\cdot,\cdot )$.
To obtain the equations governing the (pointwise) evolution of the coarse-grained kinetic energy $E^\ell _u (\boldsymbol {x},t) = \tfrac {1}{2} \bar {\boldsymbol {u}}^\ell \boldsymbol {\cdot } \bar {\boldsymbol {u}}^\ell$ and magnetic energy $E^\ell _b (\boldsymbol {x},t) = \tfrac {1}{2} \bar {\boldsymbol {b}}^\ell \boldsymbol {\cdot } \bar {\boldsymbol {b}}^\ell$, one filters (2.1)–(2.2) and then multiplies by $\bar {\boldsymbol {u}}^\ell$ and $\bar {\boldsymbol {b}}^\ell$, respectively (e.g. Zhou & Vahala Reference Zhou and Vahala1991; Kessar et al. Reference Kessar, Balarac and Plunian2016; Aluie Reference Aluie2017; Offermans et al. Reference Offermans, Biferale, Buzzicotti and Linkmann2018; Alexakis & Chibbaro Reference Alexakis and Chibbaro2022). The result can be written as
where the $\boldsymbol {{\mathcal {J}}}$ terms account for the spatial transport of energy and the $\varPi^\ell$ terms embody energy fluxes (i.e. transfer across scale $\ell$). Our sign convention for the definitions of the $\varPi^\ell$ (see below) means that $\varPi^\ell > 0$ corresponds to forward transfer of energy, i.e. to scales smaller than $\ell$. The ${\mathcal {D}}$ terms represent (hyper-)dissipative effects. Also present is the resolved scale conversion (RSC) term, ${\mathcal {W}}^\ell = \bar {b}^\ell_i \bar {b}^\ell_j \partial _j \bar {u}^\ell_i$, here expressed as in Aluie (Reference Aluie2017). This appears with opposite sign in each equation, and represents an exchange between large-scale kinetic and magnetic energies. An important point is that it is not an energy flux term since it does not involve energy transfer across scale $\ell$. It supports several interpretations including as: (i) the rate of work done on the large-scale flow by the large-scale Lorentz force and (ii) the energy gained by $\bar {\boldsymbol {b}}^\ell$ as it is distorted by $\bar {\boldsymbol {u}}^\ell$ or vice versa. Detailed forms for the spatial transport currents, $\boldsymbol {{\mathcal {J}}}_u, \boldsymbol {{\mathcal {J}}}_b$, depend on the form employed for ${\mathcal {W}}^\ell$ and are available elsewhere (e.g. Kessar et al. Reference Kessar, Balarac and Plunian2016; Aluie Reference Aluie2017; Offermans et al. Reference Offermans, Biferale, Buzzicotti and Linkmann2018; Alexakis & Chibbaro Reference Alexakis and Chibbaro2022), while the problem of Galilean invariance has been addressed by Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018).
Our primary interest herein centres on the pointwise energy flux (at scale $\ell$) terms, denoted by $\varPi ^\ell (\boldsymbol {x}, t)$, together with their volume averages, $\left \langle \varPi ^\ell \right \rangle$. For any choice of filter, these can be expressed in terms of contractions of filtered gradient tensors and SGS stress tensors:
Like the $\tau ^\ell ( \cdot, \cdot )$, these arise in connection with the four advection type nonlinearities in (2.1)–(2.2) that we refer to as the Inertial, Maxwell (meaning from the Lorentz force), Advection and Dynamo terms. Note the capitalisation. Taken together with (2.6) and (2.7), these definitions of the $\varPi^\ell$ terms mean that the interpretation of the direction of an energy flux does not depend on which flux it is. This is why (2.9) and (2.11) lack a leading minus sign. Specifically, a positive value for any one of these fluxes corresponds to transfer of energy from scales greater than $\ell$ to scales smaller than $\ell$. Clearly, $\varPi ^{I,\ell } + \varPi ^{M,\ell }$ is the net flux of $E^\ell _u$, and $\varPi ^{A,\ell } + \varPi ^{D,\ell }$ that for $E^\ell _b$. As is well known, $\varPi ^{A,\ell }$ and $\varPi ^{D,\ell }$ have a common origin and may be readily combined to obtain the magnetic energy flux associated with the curl of the induced electric field.
We remark that energy transfer in MHD turbulence has traditionally been discussed in a Fourier-space approach (e.g. Dar, Verma & Eswaran Reference Dar, Verma and Eswaran2001; Verma Reference Verma2004; Alexakis, Mininni & Pouquet Reference Alexakis, Mininni and Pouquet2005; Mininni, Montgomery & Pouquet Reference Mininni, Montgomery and Pouquet2005; Teaca et al. Reference Teaca, Verma, Knaepen and Carati2009; Linkmann et al. Reference Linkmann, Sahoo, McKay, Berera and Biferale2017; Verma Reference Verma2019). The filtering approach used here results in equivalent expressions for the mean inertial flux if the filter is chosen to be a Galerkin projector. However, in MHD, the two approaches differ in important details. In the Fourier-space approach, the transfer terms originating from the Lorentz force (momentum equation) and the field-line stretching term (induction equation) retain contributions that only involve a coupling among resolved scales. That is, the RSC term ${\mathcal {W}}^\ell$ is not separated out from the fluxes. This is the reason these terms do not vanish in the limit of $k\to \infty$ in the Fourier-based approach, in contrast to the terms defined in (2.8)–(2.11) using the SGS stresses, which therefore are genuine flux terms. As a consequence, the Fourier-based definitions can result in misinterpretations of the multi-scale nature of kinetic-to-magnetic energy conversion and of the role of the Lorentz force in inducing kinetic energy transfer across scales. For further details, see the works of Aluie (Reference Aluie2017) and Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018). An in-depth discussion of the SGS-based definition of energy flux and a comparison with the Fourier-based approach is supplied by Aluie & Eyink (Reference Aluie and Eyink2009).
2.2 Gaussian filter
Rather remarkably, the choice of a Gaussian filter enables an analytic determination of the SGS stresses and fluxes in terms of field gradients with contributions from the resolved scales and the subfilter scales. In the remainder of the paper, we therefore employ an isotropic Gaussian filter
2.3 Exact expressions for the $\tau ^\ell$ and $\varPi ^\ell$
Employing the Fourier transform of the Gaussian filter (2.12), Johnson (Reference Johnson2020, Reference Johnson2021) showed that the filtered version of an arbitrary field $\boldsymbol {u}(\boldsymbol {x},t)$ satisfies a diffusion equation,
where $\ell ^2$ is the time-like variable. It was further shown that the associated SGS stress tensor, $\tau ^\ell (u_i, u_j )$, obeys a forced version of this diffusion equation. The forcing term is $\bar {A}^\ell _{ik} \bar {A}^\ell _{jk}$, where $\bar {A}^\ell _{ik} = \partial _k \bar {u}^\ell _i$ is the gradient tensor for the filtered field. An exact solution for $\tau ^\ell ( u_i, u_j)$ was obtained that depends on $\bar {A}^\phi _{ik}$ for all scales $\phi \le \ell$. Substituting this into the equation for the SGS Inertial flux, (2.8), produces an exact solution for this flux corresponding to the exact summation of the perturbation series proposed by Eyink (Reference Eyink2006).
Happily, this approach is readily extended to MHD and may be used to calculate the elements contained in (2.8)–(2.11). Below, we outline how to achieve this for the particular case of the magnetic energy subflux (2.10) that originates with the advection term in the induction equation (i.e. $\boldsymbol {u} \boldsymbol {\cdot } \boldsymbol {\nabla } \boldsymbol {b}$). More details, plus the general case of three distinct solenoidal fields, are available in Appendix A; see also Appendix C and Capocci et al. (Reference Capocci, Johnson, Oughton, Biferale and Linkmann2023).
We seek an exact solution for $\tau ^\ell ( b_i, u_j) = \overline {b_i u_j}^\ell - \overline {b_i}^\ell \overline {u_j}^\ell$. Clearly, $\overline { b_i u_j }^\ell$ will also satisfy (2.13) since that equation holds for any (Gaussian) filtered field. Together with the product rule expansion of $\partial ( \bar {b}^\ell _i \bar {u}^\ell _j ) / \partial {\ell ^2}$, this yields
where $\bar {B}^\ell _{ik} = \partial _k \bar {b}^\ell _i$ is the gradient tensor for $\bar {b}^\ell _i$. The solution can be written in the form
where $\phi = \sqrt {\ell ^2 - \theta }$. The first term on the right-hand side is a ‘single-scale’ piece as it contains only resolved scale terms (cf. Clark, Ferziger & Reynolds Reference Clark, Ferziger and Reynolds1979). In terms of other work, it corresponds to the nonlinear model employed by Leonard (Reference Leonard1975), Borue & Orszag (Reference Borue and Orszag1998) and Meneveau & Katz (Reference Meneveau and Katz2000), and is the first-order term in the expansion of Eyink (Reference Eyink2006). It is also the leading-order term of the power law expansion in the filter limit going to zero relative of any filter kernel with finite moments (cf. § 13.4.4 of Pope Reference Pope2000).
The second term, involving an integral over all scales smaller than $\ell$, is manifestly a multi-scale contribution that includes subfilter-scale field gradients. Note that the integrand in (2.15) can itself be written as an SGS stress, but one based on the field gradients rather than the fields themselves: $\tau ^\phi ( \bar {B}_{ik}^{\sqrt {\theta }}, \bar {A}_{jk}^{\sqrt {\theta }} )$.
Contracting with $\bar {B}_{ij}^\ell$ provides an exact expression for (2.10):
where the subscripts $s$ and $m$ denote the single- and multi-scale contributions, respectively. It is evident that all the SGS energy fluxes, (2.8)–(2.11), and also other SGS fluxes (e.g. for helicities), can be written strictly in terms of (multi-scale) gradients of the velocity and magnetic vector fields. Appendix A contains further details. When discussing the individual SGS energy fluxes in § 4, we will make regular reference to (A3), the generalised form of (2.17).
The tensor contractions present in (2.17) can be expressed as the trace of the matrix products involved, after appropriate use of the transpose operation (superscript $t$). For example, $\bar {B}^\ell _{ij} \bar {B}^\ell _{ik} \bar {A}^\ell _{jk} = \mathrm {Tr}\left \{ { (\bar {\boldsymbol {B}}^\ell )^t \,\bar {\boldsymbol {B}}^\ell (\bar {\boldsymbol {A}}^\ell )^t} \right \}$.
Further insight into the physics of the scale-space flux $\varPi ^{A,\ell }$ may be extracted by expressing each gradient tensor as the sum of its index-symmetric and index-antisymmetric components. Let us write $S_{ij} = ( A_{ij} + A_{ji} ) /2$ and ${\varOmega }_{ij} = ( A_{ij} - A_{ji} ) / 2$, respectively the (velocity) rate-of-strain tensor and the rotation rate tensor, with ${\varOmega }_{ij} = -\epsilon _{ijk} {\omega }_k /2$ in terms of the vorticity ${\omega }_k = \epsilon _{ijk} \partial _i {u}_j$. Similarly, $B_{ij} = \varSigma _{ij} + {J}_{ij}$, where the non-zero elements of $J_{ij} = ( B_{ij} - B_{ji} ) / 2 = -\epsilon _{ijk} j_k/2$ are essentially the components of the electric current density $\boldsymbol {j} = \boldsymbol {\nabla } \times \boldsymbol {b}$ and the magnetic strain-rate tensor $\varSigma _{ij} = (B_{ij} + B_{ji})/2$. We will of course require the filtered versions of all of these quantities.
Ostensibly, this decomposition gives eight single-scale and eight multi-scale sub-fluxes; see (A7). However, properties of the trace of particular products of symmetric or antisymmetric matrices mean some of these may vanish, cancel or be equivalent. In the present case, one obtains, in connection with the single-scale contributions,
so that there are only three distinct subfluxes; see Appendix A for details. We write the single-scale flux for the Advection term as
where $\varPi ^{A,\ell }_{s, PQR} = -\ell ^2 \mathrm {Tr}\left \{ {(\bar {\boldsymbol {\mathsf{P}}}^\ell )^t \bar { \boldsymbol {\mathsf{Q}}}^\ell (\bar { \boldsymbol {\mathsf{R}}}^\ell )^t} \right \}$ and each of $\boldsymbol {\mathsf{P}}, \boldsymbol {\mathsf{Q}},\boldsymbol {\mathsf{R}}$ are either symmetric or antisymmetric tensors.
The multi-scale flux contributions which explicitly contain subfilter-scale fluctuations can also be so decomposed, with the details given in Appendix A. The particular forms for $\varPi ^{A,\ell }_m$ are discussed in § 4.3.
3 Methods and data
To quantify the four energy fluxes $\varPi ^{Y,\ell }$ present in (2.6)–(2.7), and their decompositions, we employ outputs from numerical simulations of the MHD equations (2.1)–(2.3). To obtain an appreciable inertial range, we use hyperdiffusion ($\alpha = 5$), a comparison with standard diffusive MHD ($\alpha = 1$) is provided in Appendix E, always with $\nu _\alpha = \mu _\alpha$. The fluctuation fields, $\boldsymbol {u}$ and $\boldsymbol {b}$, have zero means and there is no background magnetic field (i.e. $B_0 = 0$). The equations are solved using fully dealiased Fourier pseudospectral codes in a triply periodic $(2{\rm \pi} )^3$ domain (Patterson & Orszag Reference Patterson and Orszag1971; Canuto et al. Reference Canuto, Hussaini, Quarteroni and Zang1988). The time advancement is via a second-order Runge–Kutta scheme with dealiasing implemented using the two-thirds rule.
As table 1 indicates, we use up to $2048^3$ grid points. The spatial resolution of the simulations is quantified by both the grid spacing $\Delta {x} = 2{\rm \pi} / N$ and the hyper-diffusive Kolmogorov scales $\eta ^u_{\alpha } = (\nu _\alpha ^3 / \varepsilon _u)^{1/(6 \alpha -2)}$ and $\eta ^b_{\alpha } = (\mu _\alpha ^3 / \varepsilon _b)^{1/(6 \alpha -2)}$, where $\varepsilon _u$ and $\varepsilon _b$ are the mean kinetic and magnetic energy dissipation rates (Borue & Orszag Reference Borue and Orszag1995). For adequate resolution, we require $\eta ^u_\alpha / \Delta {x} \gtrsim 1.3$ and $\eta ^b_\alpha / \Delta {x} \gtrsim 1.3$ (e.g. Donzis, Yeung & Sreenivasan Reference Donzis, Yeung and Sreenivasan2008; Wan et al. Reference Wan, Oughton, Servidio and Matthaeus2010).
The forcing $f$ applied to the system is a drag-free Ornstein–Uhlenbeck process, active in the wavenumber band $k \in [2.5,5.0]$ for the MHD simulations, while the hydrodynamics dataset H2 is forced in the band $k \in [0.5,1.5]$. The snapshots, consisting of instantaneous velocity and magnetic fields, have been sampled about once per large-scale turnover time after the simulations reach statistically stationary states.
Figure 1 shows the time-averaged omnidirectional kinetic and magnetic spectra. The peak in the kinetic spectrum is due to the activity of forcing for $k \in [2.5,5.0]$, while the magnetic spectrum is considerably lower over that interval since the induction equation is not forced. In the (approximate) inertial range, both spectra have approximately power law scaling, with $E_b(k)$ close to $k^{-5/3}$ and $E_u(k)$ significantly shallower. As is typically seen in MHD simulations with no mean field, the Alfvén ratio, $E_u(k)\,/\,E_b(k)$, is less than unity in the inertial range, i.e. magnetic energy predominates at these scales.
At high $k$, there is a steep and roughly coincident decrease of both spectra. This is a consequence of two factors. First, the employment of hyperviscosity makes the dissipation range more concentrated in the small scales, leading to the sharp fall-off. Second, unit Prandtl number ensures that the dissipation wavenumber band is the same for the kinetic and magnetic spectra. Note the clear superequipartition in the entire scaling range of the spectra, with $E_b(k) > E_u(k)$ for $10 \leq k \leq 400$. Equipartition at large $k$ in the inertial range as previously predicted for decaying turbulence (Müller & Grappin Reference Müller and Grappin2004; Haugen & Brandenburg Reference Haugen and Brandenburg2006) is not observed, however, as $E_b(k)$ is steeper than $E_u(k)$ it does approach $E_u(k)$ resulting in approximate equipartition in the dissipative range.
To calculate an effective Reynolds number for a hyperdissipative system, we follow the approach described by Buzzicotti et al. (Reference Buzzicotti, Aluie, Biferale and Linkmann2018). There, the standard (Laplacian dissipation) integral-scale Reynolds number $Re=U L_u \, / \,\nu \propto \left(L_u \, / \, \eta_2 \right)^{4/3}$ (e.g. Batchelor Reference Batchelor1970; Pope Reference Pope2000) is replaced with one based on the ratio between the integral scale $L_u$ and the effective dissipation range scale $I_d$. Specifically, we employ
where $I_d = \pi\,/\,\text{argmax}\left( k^2 E_u(k) \right)$ is the scale where the dissipation spectrum $k^2 E_u(k)$ has a maximum. Here, $C$ is a fit parameter that has to be estimated by comparing (3.1) with the common definition of the Reynolds number in a standard-viscosity run. Making use of this procedure, we obtain $C=40$ and $Re=9931$ for run A1 (table 1). Generalisation of the Reynolds number for systems with hyper-dissipation has been discussed by Spyksma, Magcalas & Campbell (Reference Spyksma, Magcalas and Campbell2012).
Figure 2 displays the four MHD energy subfluxes, introduced in (2.6)–(2.7), and their sum. Also shown is $\mathcal{W}^\ell$, the resolved scale conversion of kinetic energy to magnetic energy (recall this does not represent energy transfer across $\ell$). The two panels present fluxes obtained through different filters, results shown in figure 2(a) correspond to Galerkin truncation and those shown in figure 2(a) to the Gaussian filter of (2.12). The data shown in figure 2(a,b) are qualitatively similar but with quantitative differences. Focusing on the similarities, we see that the Inertial term $\langle \varPi^{I,\ell} \rangle$ is relatively weak and is the only one to exhibit inverse transfer regions, in the intervals $1.5 \times 10^2 \lesssim k \eta_\alpha \lesssim 2 \times 10^{-1}$ and $4\times 10^{-3} \lesssim k \eta_\alpha \lesssim 6 \times 10^{-3}$. All the other subfluxes are associated with a direct cascade from the large to the small scales. The energy transfer from the momentum equation of (2.6) is almost entirely dominated by the Maxwell subflux ($-\Pi^{M,\ell}$) whose peak occurs in proximity to the forcing region. In contrast, the advection term from the induction equation ($\Pi^{A,\ell}$) is peaked at the small scales, close to the dissipative range. The conversion term, $\mathcal{W}^\ell$, is positive and increases monotonically with $k \eta_\alpha$. Particularly for the Fourier filter, it is roughly constant (i.e. scale independent) in the region $0.2 \lesssim k \eta_\alpha \lesssim 0.9$ where kinetic and magnetic subfluxes are in equipartition (Bian & Aluie Reference Bian and Aluie2019); see figure 11 in Appendix D. Also evident is a sudden increase of $\mathcal{W}^\ell$ in the dissipative range ($k\eta_\alpha \gtrsim 1$), due to hyperdiffusion, where the conversion rate saturates to $\varepsilon_b$ as already pointed out by Bian & Aluie (Reference Bian and Aluie2019).
Turning to the differences between the two kinds of filtering, we observe that the bandwidth of the inertial range plateau is narrower for the Gaussian filter case, roughly $k \eta_\alpha \in [0.04,0.4]$ versus $k\eta_\alpha \in [0.012,0.7]$. In general, the Gaussian filtering peaks are of lower amplitude and a little less localised. Linked to this is a more gradual roll-off of the fluxes at high $k\eta_\alpha$ and a slower convergence of $\mathcal{W}^\ell$ to $\varepsilon_b$ with increasing $k$. These effects arise because Gaussian filtering at scale $\ell$ retains some effects from scales $\leq \ell$, unlike the situation for the sharp Galerkin truncation of the Fourier filter. For $\Pi^{D,\ell}$, there is also a qualitative difference, with the high-$k$ local minimum and maximum seen with the Fourier filter essentially absent when the Gaussian filter is used.
4 Analysis and discussion
4.1 Inertial flux and comparison with hydrodynamics
The exact decomposition of the Inertial term $\Pi^{I,\ell}$, (2.8), is
where $\Pi^{I,\ell}_{s,S \Omega S}$ is not included as it is identically zero. This is the special case of (A3) where all the fields are the velocity and naturally it coincides with the original decomposition provided by Johnson (Reference Johnson2020) for Navier–Stokes turbulence. It is thus of interest to investigate whether there are differences between the HD and MHD instances of (4.1), and how these might arise.
Figure 3 compares these two cases. For the hydrodynamic case, panel (a), there is a relatively clear and extended plateau for the total flux (and some of the subfluxes) corresponding to an inertial range. In contrast, the MHD case shown in panel (b) lacks such a plateau and the individual subfluxes are mostly much smaller than their hydrodynamic counterparts. Indeed, only $\langle \Pi^{I,\ell}_{m,SS\Omega} \rangle$ has values comparable to its hydrodynamic counterpart, albeit with a different functional form, being negative for almost all $k$. Intriguingly, this term is the only one that does not vanish point-wise in two-dimensional (2-D) turbulence, as discussed by Johnson (Reference Johnson2021).Footnote 2 A depletion of the inertial flux in MHD turbulence has been observed by Alexakis (Reference Alexakis2013) and Yang et al. (Reference Yang, Linkmann, Biferale and Wan2021) for configurations with large-scale electromagnetic forcing and by Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018) for a saturated dynamo at lower Reynolds number. We have verified that the relation of Betchov (Reference Betchov1956), $\langle \Pi^{I,\ell}_{s,SSS} \rangle = 3 \langle \Pi^{I,\ell}_{s,S\Omega \Omega} \rangle$, holds for both datasets, as it must.
Also of interest is that there appears to be an approximate multi-scale analogue of the Betchov relation, with $\langle \Pi^{I,\ell}_{m,SSS} \approx \langle \Pi^{I,\ell}_{m,S\Omega \Omega} \rangle$. For hydrodynamics, this was already noted by Johnson (Reference Johnson2021) and justified by Yang et al. (Reference Yang, Zhou, Xu and He2023). Evidently, this approximate degeneracy also holds in this MHD situation, although the smallness of the terms makes this difficult to appreciate from figure 3(b).
Having discussed mean fluxes, we now examine some statistical properties of their pointwise contributions. Figure 4 presents standardised probability density functions (p.d.f.s) of the MHD Inertial subfluxes at $k\eta_\alpha = 5.4 \times 10^{-2}$. It is evident that the distributions are strongly non-Gaussian and exhibit very wide tails, with fluctuations at tens of standard deviations. (As we shall see, this is a common characteristic for all the MHD energy fluxes and subfluxes.) Most importantly, we observe strong backscatter in all subfluxes. That is, the depletion of the mean Inertial flux results from cancellations between forward and inverse transfer events made possible by an increase in backscatter events. Similar observations for the total Inertial term have been reported by Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018), where a comparison between the kinematic, nonlinear and saturated stages of the dynamo had been carried out. The variance, skewness and kurtosis for each p.d.f. are reported in table 2. Although the p.d.f. of $\Pi^{I,\ell}_{s,S\Omega \Omega}$, which corresponds to the scale-local vortex stretching, has more asymmetrical tails than those of $\Pi^{I,\ell}_{s,SSS}$, the skewness for the latter is nonetheless larger. The term $\Pi^{I,\ell}_{s,S \Omega S}$ does not show up in figure 4 because it is identically zero due to the symmetries of the tensors involved in the corresponding trace of (A3). In contrast, panel (b) shows that . is the most negatively skewed p.d.f. among the Inertial ones.
In connection with the approximate degeneracy between $\Pi^{I,\ell}_{m,SSS}$ and $\Pi^{I,\ell}_{m,S\Omega \Omega}$ discussed above, figure 4(b) reveals that their p.d.f.s coincide up to events with a standardised probability density of $10^{-7}$, potentially indicating that the approximate identity is true not just on average but also for higher-order moments, although further analysis is required.
4.2 Maxwell flux
The decomposition of the energy flux associated with the Lorentz force, which we refer to as the Maxwell term, $\Pi^{M,\ell}$ of (2.9), contains an extra (single-scale) term with respect to the Inertial flux. Specifically, from (A3), we obtain
Terms of type $S \Sigma \Sigma$ can be associated with strain rate amplification by magnetic shear, while terms of type $SJJ$ correspond to current-filament stretching that is analogous to vortex stretching in HD. The last two terms are of type $SJ\Sigma$ and describe the back-reaction of the magnetic field on the flow or, more specifically, how the velocity strain rate is modified (typically amplified) in connection with a current-sheet thinning process. As we shall see, this is by far the dominant process. It proceeds as follows (figure 5). First, a current sheet is stretched by large-scale straining motions into a magnetic shear layer, in a process similar to vortex thinning in HD (Kraichnan Reference Kraichnan1976; Chen et al. Reference Chen, Ecke, Eyink, Rivera, Wan and Xiao2006; Johnson Reference Johnson2021). This results in a stretching of the magnetic flux tubes in the sheet. By conservation of magnetic flux, the magnetic field strength at the thereby generated smaller scales must increase. That is, magnetic energy is transferred from large to small scales. (We will revisit this process in § 4.3 in the context of the inter-scale transfer of magnetic energy.) The magnetic rate-of-strain field associated with the resulting magnetic shear layer now accelerates fluid along its extensional directions and slows it down in the compressional directions, thereby generating a stronger rate-of-strain field across smaller scales.
It is instructive to consider the process in two dimensions, in analogy to the vortex thinning of 2-D HD. In the reference frame of the rate-of-strain tensor at scale $\ell$, the associated terms are
where $\pm \lambda_S$ and $\pm \lambda_\Sigma$ are the eigenvalues of the velocity and magnetic rate-of-strain tensors, respectively, $\psi$ the angle between the respective eigenvectors, and $j$ the out-of-plane component of the current density. As can be seen from these formulae, a maximum energy transfer occurs when the principal axes of the magnetic rate-of strain tensor have a $\pm 45^{\circ}$ angle to those of the velocity rate-of-strain tensor. Depending on the sign of the out-of-plane current density and $\sin(2\psi)$, (4.3) and (4.4) result in a direct or an inverse energy transfer. Figure 5 presents a schematic depiction of the process. A direct energy transfer occurs if the angle between the principal axes of velocity and magnetic strain-rate tensors is in the same rotational direction as the out-of-plane current. A similar, albeit less straightforward, assessment is possible in three dimensions, where
and similarly for the multi-scale term. Here, $\mu_j^\ell$ is the $j-$th eigenvalue of the symmetric part of the product matrix $\overline{{J}}^\ell_{ik}\overline{\Sigma}^\ell_{kj}$ (a contribution to the Maxwell SGS stress tensor) and $\psi_{ij}$ the angle between the $i-$th eigenvector of $\bar{\boldsymbol{\mathsf{S}}}^\ell$ and the $j-$th eigenvector of the symmetric part of $\overline{J}^\ell_{ik}\overline{\Sigma}^\ell_{kj}$. For a forward cascade of kinetic energy, $\langle \Pi^{M,\ell} \rangle > 0$, which implies that $\overline{S}^\ell_{ij} \, \overline{J}^\ell_{ik} \overline{\Sigma}^\ell_{kj}$ should be preferentially positive. Due to the presence of the cosine squared factor, this implies that the principal axes of the rate-of-strain tensor and those of the subscale stress must preferentially align, resulting in a stretching of the magnetic flux tubes along the extensional directions of the strain-rate tensor, as discussed above.
The subfluxes on the right-hand side of (4.2) and the total Maxwell flux are shown in figure 6, as a function of (non-dimensionalised) reciprocal $\ell$. We see immediately that the net energy transfer proceeds from large scales to small scales with the total Maxwell flux $\langle \Pi^{M,\ell} \rangle$ being the dominant energy subflux for MHD, carrying approximately $80\%$ of the total energy dissipation rate at its peak. At large scales, the major contribution is from the multi-scale term $\langle \Pi^{M,\ell}_{m,SJ\Sigma}\rangle$, switching to its single-scale partner, $\langle \Pi^{M,\ell}_{s,SJ\Sigma}\rangle$, as the dissipation scale is approached. All remaining terms in (4.2) are negligible. Summarising the mean Maxwell flux behaviour, we may say that the net kinetic energy transfer in MHD proceeds by the back-reaction of the magnetic field on the flow during the aforementioned current-sheet thinning process, while the contribution from current filament stretching and strain-amplification by magnetic shear are negligible.
The p.d.f.s for the Maxwell energy fluxes are shown in figure 7. The predominance of $\langle \Pi^{M,\ell}_{m,SJ\Sigma} \rangle$ in the direct cascade can be also appreciated by examining its p.d.f., which is the most positively skewed among the multi-scale terms of figure 7; see also table 2 for p.d.f. moments. As can be seen from the data shown in figure 7, while all terms of type $S\Sigma \Sigma$ and $SJJ$ nearly vanish on average, fluctuations of approximately 60 standard deviations are not uncommon, and their multi-scale contributions show considerable backscatter. That is, the mean effects of current-filament stretching and strain-amplification by magnetic shear on the flow are negligible because of cancellations between pointwise forward and inverse transfers, and both these processes result in extreme backscatter events.
In Appendix B, we show that the averages of the subfluxes $\Pi_{m,S\Sigma \Sigma}^{M,\ell}$ and $\Pi_{m,S J J}^{M,\ell}$ can be connected via an exact Betchov-like relation that holds for all homogeneous flows,
Note that the final term, $\Pi_{m,\Omega \Sigma J}^{M,\ell}$, does not appear in the Maxwell flux, (4.2), and our numerical results indicate $\Pi_{m,S\Sigma \Sigma}^{M,\ell} \approx \Pi_{m,SJJ}^{M,\ell}$, see figure 6. Physically, we may interpret this approximate identity as indicating that the net ‘strain-production’ by magnetic shear is almost equal to the net strain production by current-filament stretching. However, we stress again that these contributions to the interscale kinetic energy transfer are negligible. Further discussion on terms associated with strain production and current-filament stretching is provided in Appendix B.
4.3 Advection and dynamo fluxes
In this section, we focus on the decomposition of both the Advection term, $\Pi^{A,\ell}$, and the Dynamo term, $\Pi^{D,\ell}$, as defined in (2.10)–(2.11). As is well known, these two SGS fluxes share the same physical origin, namely the induced (fluctuation) electric field, and this is associated with certain symmetries and equivalences between the Advection and Dynamo subfluxes.
From the application of (A3), we find
where we do not list terms that vanish identically, see Appendices A and F.
Figure 8 displays most of the Advection and Dynamo (sub)fluxes in separate panels. For clarity, only the subfluxes relevant to the discussion below and to the net energy flux are shown. The cyclic property of the trace can be used to show that some terms vanish identically and that each Advection subflux (both single-scale and multi-scale) is equal to (plus or minus) a partner Dynamo subflux; see the subfluxes expressions in Appendix F. For this reason, in figure 8, the following subflux pairs are not displayed since they are opposite in sign, $\Pi_{s,\Sigma \Sigma S}^{A,\ell}$ and $\Pi_{s,\Sigma S \Sigma}^{A,\ell}$ as well as $\Pi_{s,\Sigma J \Omega}^{A,\ell}$ and $\Pi_{s,\Sigma \Omega J}^{A,\ell}$ together with their multi-scale counterparts. Hence, these cancel pairwise and make no contribution to the net magnetic energy flux $\Pi^{A,\ell} + \Pi^{D,\ell}$; see Appendix F. In contrast, $\langle \Pi^{A,\ell}_{s,JJS} \rangle$ and $\Pi^{D,\ell}_{s,JSJ}$ and the related multi-scale terms, are equal and thus do contribute to the net flux.
Note, too, that the two Betchov relations (B6)–(B10) can provide another source of symmetry, or approximate symmetry.
The physical interpretation of the respective terms is very similar to what has been discussed for the Maxwell flux, except that now the effect of the flow on the magnetic field must be considered. Recall from § 4.2 that the Maxwell flux terms of type $SJ\Sigma$ correspond to the stretching and, by incompressibility, thinning of current sheets into magnetic shear layers, and that the back-reaction on the flow induced by this process is responsible for the bulk of the kinetic energy transfer to smaller scales. As we shall see, and as expected from the discussion of current-sheet thinning in § 4.2, this process also transfers most magnetic energy from large to small scales. However, in contrast to the Maxwell flux situation (dominated by a multi-scale term), here it is two of the Advection single-scale terms, $\Sigma J S$ and $J \Sigma S$, that carry most of the magnetic energy flux.
Focusing on the Advection term, from figure 8(a), we observe that the net flux is everywhere positive and peaked at smaller scales, roughly at the end of the inertial range. Recall that the Maxwell flux is peaked at larger scales (figure 6). Due to the cyclic property of the trace, the terms $\Pi^{A,\ell}_{s,\Sigma J S}$ and $\Pi^{A,\ell}_{s,J\Sigma S}$ are equal while, because of the symmetry of the tensors involved, the subfluxes $\Pi^{A,\ell}_{s,JJ\Omega}$ and $\Pi^{A,\ell}_{s,\Sigma \Sigma \Omega}$ together with $\Pi^{D,\ell}_{s,J\Omega J}$ and $\Pi^{D,\ell}_{s, \Sigma \Omega \Sigma}$ are identically zero. It is also apparent from figure 8(a) that the remaining Advection terms make negligible contributions to the net flux. We note that $\Sigma \Sigma S$ type terms correspond to magnetic shear amplification due to straining motions and those of type $JJS$ to current filament stretching – the analogue to vortex stretching in HD – as depicted in figure 9. Terms of type $\Sigma J \Omega$ encode bending of magnetic field lines into current filaments by rotational flow, i.e. a change in the magnetic-field geometry induced by vortical motion. Thus, the net Advection term is primarily due to single-scale contributions, being approximately equal to $2\, \Pi^{A,\ell}_{s,\Sigma J S}$, and it carries approximately $40\%$ of the total energy flux at its peak.
For the Dynamo term, $\Pi^{D,\ell}$, we observe that the net flux is almost flat in the inertial range, substantially positive definite for these scales (figure 8b) and responsible for approximately $15\%$ of the total energy flux. All subfluxes, except $\Pi^{D,\ell}_{s,\Sigma SJ}$ shown in yellow and $\Pi^{D,\ell}_{s,JS\Sigma}$ indicated by the filled purple symbols, are negligible. However, using the cyclic property once again, one can show that $\Pi^{D,\ell}_{s,\Sigma SJ} = -\Pi^{D,\ell}_{s,J S \Sigma}$; hence, these two terms cancel out and do not contribute to the net flux. In contrast to the Advection term, the major contributions to the net Dynamo term are from subleading single and multi-scale subfluxes of various types, with each contributing only 1–2 % to the total energy flux adding up to a total of approximately $15\%$ of the total energy flux. In summary, single-scale current-sheet thinning is the dominant process transferring magnetic energy across scales. It solely originates from the advective term in the induction equation.
Consider now the p.d.f.s. As a consequence of the symmetries between Dynamo and Advection subfluxes, only the Advection p.d.f.s are shown in figures 10(a) and 10(c), respectively for the single-scale terms and the multi-scale terms. It is striking that the p.d.f. of $\Pi^{A,\ell}_{s,\Sigma J \Omega}$, which describes the bending of magnetic field lines into current filaments by rotational flow, is by far the most strongly fluctuating with huge fluctuations of more than 100 standard deviations, also showing strong magnetic backscatter. In contrast, the aforementioned current-sheet thinning process, that is mostly responsible for the forward cascade of magnetic energy, shows weak magnetic backscatter. The other Advection subflux p.d.f.s, both single-scale and multi-scale, span a range comparable with those associated with the p.d.f.s for the Inertial and Maxwell fluxes. Moreover, all the multi-scale fluxes (Advection and Dynamo) have quite similar p.d.f.s, and thus so do the net multi-scale flux p.d.f.s (figure 10d). In the case of the net single-scale p.d.f.s, the Advection–Dynamo agreement is still good in the cores of the distributions, but there is a significant difference at larger negative fluctuations (figure 10b).
4.4 Total energy flux, .
In the preceding subsections, we analysed the four contributions to the incompressible MHD energy flux, finding that each of them may be reasonably well approximated using just some of the subfluxes. These approximations may now be assembled to give an approximate form for the mean total MHD energy flux. Specifically, we suggest that
is a suitable expression. Note that, after using cyclic properties of the trace, this is expressed only in terms of two of the Maxwell subfluxes, one single-scale and one multi-scale. These are discussed in § 4.2 and we remind the reader that all subflux definitions can be found in Appendix F. Interestingly, the terms on the right-hand side of (4.11) are part of a group of terms that remain non-zero in 2-D MHD; see Appendix C. We intend to explore this intriguing feature in future work.
Equation (4.11) demonstrates that the mean total energy flux in MHD turbulence is largely given by the stretching and thinning of current-sheets into magnetic shear layers by large-scale strain, resulting in a transfer of magnetic energy from large to small scales. In addition, there is a back-reaction of this process on the flow, whereby the ensuing magnetic strain-rate field accelerates fluid along its extensional directions and slows it down in the compressional directions, thereby generating a stronger strain-rate field across smaller scales, as shown schematically in figure 5.
4.5 Current-sheet thinning and magnetic reconnection
As a means of inter-scale energy transfer, the current-sheet thinning process can only proceed as described in an inertial range, i.e. at scales where Joule and viscous dissipation are negligible. As the dissipative scales are approached, Alfvén's theorem ceases to hold and magnetic reconnection can occur, with associated changes to the topology of the magnetic field. Interestingly, models of MHD reconnection are geometrically very similar to the inertial-range energy transfer process described here. For example in the Sweet–Parker model (Parker Reference Parker1957; Sweet Reference Sweet1958), a current sheet is thinned by a flow that pushes magnetic field lines closer and closer together. Eventually, when the distance between the field lines approaches scales where Joule dissipation becomes important, the topological conservation of the magnetic field is broken and magnetic field-lines reconnect. That is, the continuation of the current-sheet thinning process described herein, and shown conceptually in figure 5, to smaller and smaller scales can lead naturally to magnetic reconnection.
5 Consequences for MHD subgrid-scale modelling
MHD LES modelling usually proceeds through variations on the Clark and Smagorinsky models, here shown for simplicity for the HD case,
for the Clark model (Clark et al. Reference Clark, Ferziger and Reynolds1979; Pope Reference Pope2000),Footnote 3 where we note that this model corresponds to what we have herein called single-scale flux contributions (see the first term on the right-hand side of (2.15)), and the Smagorinsky model
where . is the turbulent or eddy viscosity given by
with . being a free parameter, the Smagorinsky constant (Smagorinsky Reference Smagorinsky1963). This is the case for incompressible MHD (Zhou & Vahala Reference Zhou and Vahala1991; Müller & Carati Reference Müller and Carati2002; Kessar et al. Reference Kessar, Balarac and Plunian2016). Moreover, adaptations have been made for the compressible case (Chernyshov, Karelsky & Petrosyan Reference Chernyshov, Karelsky and Petrosyan2010; Grete et al. Reference Grete, Vlaykov, Schmidt and Schleicher2016; Vlaykov et al. Reference Vlaykov, Grete, Schmidt and Schleicher2016), for channel flow at high magnetic Reynolds number (Hamba & Tsuchiya Reference Hamba and Tsuchiya2010; Jadhav & Chandy Reference Jadhav and Chandy2021, Reference Jadhav and Chandy2023) and for extensions of MHD taking various levels of microphysics terms into account, such as Hall MHD (Miura, Araki & Hamba Reference Miura, Araki and Hamba2016) or Braginskii-extended (two-fluid) MHD (Miura, Hamba & Ito Reference Miura, Hamba and Ito2017), with the latter specifically focussed on the ballooning instability in stellarator devices. Most such approaches use the same SGS model for the inertial and Maxwell stresses based on velocity-field gradients with eddy viscosities involving either only the strain rate tensor or a weighted sum of the squared strain rate and the squared current, and similarly structured SGS models for the induction equation, with either heuristics or trial-and-error approaches to find suitable values for model constants. In what follows, we briefly discuss how the present results can be used to construct suitably structured SGS models for each SGS-stress in the MHD equations.
In terms of SGS modelling, our simulation-based results suggest that the Inertial terms can be neglected and a dissipative model, such as the Smagorinsky model, for the Maxwell stresses should suffice to capture the (leading-order) mean effects. In terms of fluctuations, we find that the observed mean Inertial flux depletion is caused by considerable backscatter in all Inertial subfluxes (see figure 4). This could suggest that a more sophisticated model would be required for the Inertial term. However, in 3-D HD turbulence, backscatter-free SGS models such as the standard static Smagorinsky closure perform well in capturing high-order statistics, that is, anomalous exponents and multifractal predictions for the correlation between velocity-field increments and SGS stresses (Linkmann, Buzzicotti & Biferale Reference Linkmann, Buzzicotti and Biferale2018). Furthermore, as the SGS stresses enter the filtered Navier–Stokes equation only through their divergence, this results in a degree of (gauge) freedom to determine model stress tensors that produce much less backscatter than those constructed using the standard definition (Vela-Martín Reference Vela-Martín2022). In fact, backscatter can be traced back to spatial fluxes disconnected from the (scale-space) energy transfer and as such does not require modelling (Vela-Martín Reference Vela-Martín2022).
For the magnetic energy transfers, we observe the net transfers are from large to small scales, suggesting again that dissipative models should suffice. Indeed, as multi-scale terms in our flux decomposition are negligible, a Clark-type model involving only the coupling between current and strain-rate tensors may work well for a nonlinear saturated non-helical (small-scale) dynamo. However, additional stabilising terms may be required as the Clark (or gradient) model is know to result in numerically unstable LES of a mixing layer in HD (Vreman, Geurts & Kuerten Reference Vreman, Geurts and Kuerten1996, Reference Vreman, Geurts and Kuerten1997) and for MHD (Müller & Carati Reference Müller and Carati2002; Kessar et al. Reference Kessar, Balarac and Plunian2016), as discussed in further detail below. Similar to the Inertial term results, we observe considerable backscatter in the magnetic energy fluxes (see figure 10), and it remains to be seen if the aforementioned results by Linkmann et al. (Reference Linkmann, Buzzicotti and Biferale2018) on the effect of SGS closures on high-order statistics carry over from HD to MHD.
For a non-helical saturated dynamo, as is the case here, Müller & Carati (Reference Müller and Carati2002) and Kessar et al. (Reference Kessar, Balarac and Plunian2016) carried out MHD LES with Clark-type models constructed from the full velocity and magnetic-field gradients for the sum of the Reynolds and Maxwell SGS stress in the momentum equation and for the magnetic stresses, resulting in unstable simulations as in the HD case. Using a priori analyses of DNS data, Kessar et al. (Reference Kessar, Balarac and Plunian2016) trace the instabilities back to the Clark terms transferring an insufficient amount of kinetic and magnetic energy to small scales, and to a production of backscatter. According to our analysis, for the momentum equation, the former effect is due to the Maxwell stress having a significant multi-scale component which is not captured in the Clark model. Concerning the latter, we also observe strong backscatter from the kinetic and magnetic single-scale subfluxes except for terms connected with current-sheet thinning, see figures 4, 7 and 10. Furthermore, a Clark model based on the full gradients by generalisation of (5.1) to MHD will introduce effects that are not present in the full MHD dynamics especially concerning the Maxwell and magnetic stresses, as all combinations of vorticity, strain, current and magnetic strain are included in the model and equally weighted,
Our analysis, however, shows that only terms stemming from the coupling between current and strain-rate tensors are significant, with all remaining contributions to the net Maxwell and magnetic fluxes being negligible. Thus, retaining these contributions in an SGS model may substantially (and inappropriately) affect the small-scale structure of the flow and the magnetic field, and lead to an overestimation of backscatter. It remains to be seen if backscatter-related instabilities can be suppressed if only the current-sheet thinning (CST) terms,
are retained. That is, both magnetic strain and current density tensors must be included in SGS models as recently discussed by Alexakis & Chibbaro (Reference Alexakis and Chibbaro2022).
A further challenge for LES modelling of MHD turbulence is to accurately capture the transfer of magnetic to kinetic energy (Haugen & Brandenburg Reference Haugen and Brandenburg2006) and vice versa. Using standard Smagorinsky models for both the momentum and induction equations, Haugen & Brandenburg (Reference Haugen and Brandenburg2006) showed that the Smagorinsky constant can be fine-tuned to obtain a good agreement between filtered DNS and LES for kinetic energy spectra, but only at the expense of a strong suppression of the (nonlinear) dynamo. Moreover, as discussed by Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018), the resolved-scale conversion term must either be fully accounted for in LES, resulting in the need to resolve all scales where this term is active, or in the present case of a saturated dynamo, a model including an extra term accounting for the under-resolved dynamo effect must be provided.
In this paper, we have only considered the no mean magnetic field situation. The presence of a strong background magnetic field is likely to require an SGS modelling approach that differs from those just discussed, as the ensuing anisotropy and two-dimensionalisation of magnetic- and velocity-field fluctuations may result in partial inverse fluxes. We will report results on configurations with strong background magnetic fields in due course. Similarly, the large-scale (helical) dynamo requires an investigation in its own right, and the magnetic-field growth at large scales is likely to require a different type of SGS modelling approach, as Kessar et al. (Reference Kessar, Balarac and Plunian2016) report that the Clark and even a standard static Smagorinsky model result in unstable LES.
The method discussed here can be readily extended to Hall and two-fluid MHD and other fluid models for plasma turbulence. For instance, the applicability of the Smagorinsky closure to Hall MHD has been assessed by an a priori analysis using sharp filtering (Miura & Araki Reference Miura and Araki2012) prior to the deployment of said closure (Miura et al. Reference Miura, Araki and Hamba2016). An a priori analysis and decomposition of the Hall flux in analogy to the results presented here could lead to a better understanding of the physics of the interscale magnetic energy transfer induced by the Hall effect and thus to a refinement of Hall-MHD SGS models. Finally, we point out that a new LES method, so-called physics-inspired coarsening (PIC), has been devised recently (Johnson Reference Johnson2022). In the homogeneous case, this approach reduces to Gaussian filtering and the representation of SGS stresses in terms of field gradients as generalised herein. In PIC, the velocity field advanced in LES is formally obtained by artificial viscous smoothing, with the required pseudo-diffusion being introduced through an auxiliary Stokes equation. This approach may be generalisable to MHD and more complex fluid models applicable to plasma turbulence.
6 Conclusions
Generalising a method introduced by Johnson (Reference Johnson2020) for Navier–Stokes turbulence, we have presented a general analytical method for obtaining exact forms for inter-scale fluxes in advection–diffusion equations through products of vector-field gradients, and applied it to kinetic and magnetic energy fluxes in homogeneous MHD turbulence. The aim was to provide expressions for subfluxes that are physically interpretable in terms of the action of the magnetic field on the flow and vice versa. A quantification thereof is of interest for the fundamental understanding of cascade processes in MHD turbulence, and also provides guidance as to what physics needs to be captured in subgrid-scale models and how such models should be constructed so that they preserve, at least approximately, empirical features of the mean energy fluxes and their fluctuations.
In MHD, scale-space energy fluxes are defined as the contraction of velocity- or magnetic-field gradients with the appropriate subgrid-scale stresses. Rewriting these in terms of symmetric and antisymmetric components of field-gradients tensors yields terms with clear physical meanings. For example, strain and vorticity in the case of the velocity field, and current and magnetic strain/shear for the magnetic field. Expressing the MHD SGS stresses in terms of vorticity, rate-of-strain, current and magnetic rate-of-strain results in an exact decomposition of magnetic and kinetic energy fluxes in terms of interactions between the symmetric and antisymmetric components of velocity- and magnetic-field gradients.
The kinetic energy flux comprises two terms, the Inertial flux (as in hydrodynamics) and a flux term associated with the action of the Lorentz force on the flow. The former is decomposed into terms associated with vortex stretching, strain self-amplification and strain-vorticity alignment. A term-by-term comparison between the Inertial fluxes in HD and MHD turbulence shows that all Inertial subfluxes are depleted due to cancellations between forward scatter and backscatter events, and are indeed almost negligible in MHD turbulence. That is, the physics of the kinetic energy cascade is very different in statistically steady MHD turbulence as compared with HD turbulence, as vortex stretching and strain self-amplification have, on average, very little effect. In MHD turbulence, almost all kinetic energy is transferred downscale by a current-sheet thinning process: in regions of large strain, current sheets are stretched by large-scale straining motion into regions of magnetic shear. This magnetic shear in turn drives extensional flows at smaller scales. The magnetic energy is mainly transferred from large to small scales by the aforementioned current-sheet thinning in regions of high strain, while the mean contributions from current-filament stretching – the analogue to vortex stretching – and bending of magnetic field lines in high magnetic strain regions into current filament by vortical motion are almost negligible. The latter effect, which results in a change in the magnetic field geometry at small scales is associated with strong magnetic backscatter. A decomposition into single- and multi-scale components of the subfluxes shows that the mean kinetic energy flux induced by the back-reaction of current-sheet thinning on the flow has a strong multi-scale component especially at large scales where the magnetic field is weak, while the multi-scale component of the magnetic energy flux is almost negligible. We consistently observe that the multi-scale components of the respective flux terms fluctuate less than the single-scale components.
Finally, we note that the method can be further expanded in various directions. Within MHD turbulence, for instance, decompositions of the magnetic helicity and cross-helicity fluxes could identify the physics driving these respective cascades with potential implication for selective decay for decaying MHD turbulence. Kinetic energy transfer across scales and its conversion to magnetic energy, and vice versa, depend on the value of the magnetic Prandtl number (Brandenburg & Rempel Reference Brandenburg and Rempel2019). The methodology introduced here could be used to quantify the .-dependence of the different physical processes involved in the energy cascade. Another possibility would be to include temperature fluctuations. An extension or application to compressible flows would be of interest especially for astrophysical plasmas. As flux terms associated with any advective nonlinearity can be analysed by this method, a decomposition of Hall MHD and of fluid models of ion- or electron-temperature-gradient turbulence (Ivanov et al. Reference Ivanov, Schekochihin, Dorland, Field and Parra2020; Adkins et al. Reference Adkins, Schekochihin, Ivanov and Roach2022; Ivanov, Schekochihin & Dorland Reference Ivanov, Schekochihin and Dorland2022) may be of interest for the magnetic confinement fusion community.
Acknowledgements
Editor S. Tobias thanks the referees for their advice in evaluating this article.
Funding
This work used the ARCHER2 UK National Supercomputing Service (https://www.archer2.ac.uk) with resources provided by the UK Turbulence Consortium (EPSRC grants EP/R029326/1 and EP/X035484/1) and received funding from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (grant agreement No 882340, L. Biferale).
Declaration of interests
The authors report no conflict of interest.
Appendix A. General formulation of advective-type SGS flux terms
In § 2.3, we derived, following Johnson (Reference Johnson2020, Reference Johnson2021), the form for the scale-filtered magnetic energy flux associated with the $\mathbf{u}\cdot \nabla \mathbf{b}$ term of the induction equation. Here, we outline how this approach is generally applicable for flux terms involving three distinct fields, connected with an (unfiltered) term of the form $\mathbf{x} \cdot \left( \mathbf{z} \cdot \nabla \right) \mathbf{y}$, where $\mathbf{x}$, $\mathbf{y}$ and $\mathbf{z}$ are solenoidal, but otherwise arbitrary, 3-D vector fields (they are not coordinate vectors). For appropriate mappings of $\mathbf{x}$, $\mathbf{y}$, $\mathbf{z}$ to $\mathbf{u}$ and $\mathbf{b}$, this will yield any of the desired MHD SGS energy fluxes, (2.8)–(2.11). Moreover, the SGS fluxes associated with helically decomposed hydrodynamics and MHD (Waleffe Reference Waleffe1993; Lessinnes et al. Reference Lessinnes, Plunian, Stepanov and Carati2011; Linkmann et al. Reference Linkmann, Berera, McComb and McKay2015; Alexakis Reference Alexakis2017; Alexakis & Biferale Reference Alexakis and Biferale2018; Yang et al. Reference Yang, Linkmann, Biferale and Wan2021) and the kinetic, magnetic and cross-helicities may be obtained using similar special cases, see Capocci et al. (Reference Capocci, Johnson, Oughton, Biferale and Linkmann2023) for a decomposition of the kinetic helicity flux in Navier–Stokes turbulence.
The SGS stresses at scale $\ell$ associated with $\mathbf{z} \cdot \nabla \mathbf{y}$ are
where the choice of the filter kernel is for now arbitrary. Clearly, $\left( \tau^{\ell}(y_i,z_j) \right)^t = \tau^{\ell}\left(z_i,y_j \right)$, where $(\cdot)^t$ denotes matrix transpose. Note that the advecting field is $\mathbf{z}$. Contracting the SGS stress against the gradient tensor of a third arbitrary field, $\mathbf{x}$, yields the general SGS flux term
Here, we have included a leading minus sign in (A2). However, if one wishes to have $\Pi^{\ell}_{xyz} > 0$ always correspond to forward transfer – as we have elected to do herein – this may not be correct. It depends on the sign the $\mathbf{z} \cdot \nabla \mathbf{y}$ term has when it is written on the right-hand side of the underlying advection–diffusion equation. For the MHD momentum equation, for example, the Lorentz force term has $\mathbf{y} = \mathbf{z} = \mathbf{b}$ and the minus sign for the associated kinetic energy flux (with $\mathbf{x} = \mathbf{u}$) should be absent. See §§ 2.1 and 4.2. When it is appropriate to do so, the minus sign and its propagation into other equations in this appendix is easily removed.
In the special case that $\tau^\ell(y_i,z_j)$ is index-symmetric, only the index-symmetric part of $\partial_j \overline{x}^\ell_i$ contributes. This is the situation for the kinetic energy flux in HD (Germano Reference Germano1992) and by analogy in MHD, see e.g. Zhou & Vahala (Reference Zhou and Vahala1991), Kessar et al. (Reference Kessar, Balarac and Plunian2016), Aluie (Reference Aluie2017), Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018) and Alexakis & Chibbaro (Reference Alexakis and Chibbaro2022). In general, however, the index-antisymmetric part of $\partial_j \overline{x}^\ell_i$ is also needed.
As shown by Johnson (Reference Johnson2021), (A2) may also be expressed entirely in terms of (products of) the gradient tensors for $\mathbf{x}, \mathbf{y}, \mathbf{z}$ and integrals over them. Denoting the respective gradient tensors as $\partial_j x_i = X_{ij}\,$, $\partial_j y_i = Y_{ij}$ and $\partial_j z_i = Z_{ij}$, we have
where $\phi = \sqrt{\ell^2 - \theta}$ and $\sqrt{\theta}$ correspond to all filter scales smaller than $\ell$, and the subscripts $s$ and $m$ stand for single-scale and multi-scale. Equivalently, expressed in terms of the matrix trace operation, this is
Thus, both the single-scale and the multi-scale terms may be expressed as (integrals of) the trace of the appropriate filterings and transposes of the product of the three gradient tensors.
Splitting the gradient tensors into their index symmetric and antisymmetric parts, e.g. $\overline{\boldsymbol{\mathsf{X}}}^\ell = \overline{\boldsymbol{\mathsf{S}}}_X^\ell + \overline{\mathbf{\Omega}}_X^\ell$, produces a decomposition of (A5) that facilitates physical interpretation of the subterms. For the single-scale terms, one has, modulo the $-\ell^2$ factor,
which in general does not simplify further.
Simplifications do ensue, however, for special cases when one or more of ${\boldsymbol{\mathsf{X}}}$, ${\boldsymbol{\mathsf{Y}}}$, ${\boldsymbol{\mathsf{Z}}}$ are equal. One makes use of matrix properties like $\boldsymbol{\Omega}_Y \boldsymbol{\mathsf{S}}_Y + \boldsymbol{\mathsf{S}}_Y^t \boldsymbol{\Omega}_Y^t$ is a symmetric matrix and the square of any (square) matrix is a symmetric matrix. For example, when $\boldsymbol{\mathsf{Y}}=\boldsymbol{\mathsf{Z}}$, as is relevant for the Inertial ($\Pi^{I,\ell}_s$) and Maxwell ($\Pi^{M,\ell}_s$) fluxes, we obtain
Examples with $\boldsymbol{\mathsf{X}} = \boldsymbol{\mathsf{Y}}$ and $\boldsymbol{\mathsf{X}} = \boldsymbol{\mathsf{Z}}$ relate to the Advection and Dynamo magnetic energy SGS fluxes. See § 4.3.
Turning to the multi-scale contributions in (A3), these may of course be similarly decomposed. Since the filtering operation is linear, the integrand can be split into the sum of four terms that each have the same structure as the original integrand, e.g.
After integration and contraction with $\left(\overline{\boldsymbol{\mathsf{X}}}^\ell\right)^t = \overline{\boldsymbol{\mathsf{S}}}^\ell_X - \boldsymbol{\overline{\Omega}}^\ell_X$, this gives eight, in general distinct, multi-scale contributions. Once again, special cases such as $\boldsymbol{\mathsf{Y}}=\boldsymbol{\mathsf{Z}}$ may mean some of these eight are zero, or equivalent or cancel. The needed particular instances are discussed in the subsections of § 4.
A.1 Special cases
Here, we list specific examples of (A2) that are relevant to the HD and/or MHD equations.
(i) $\boldsymbol{x}, \boldsymbol{y}, \boldsymbol{z} \rightarrow \boldsymbol{u}$. This yields the usual Navier–Stokes energy flux, $\Pi^{\ell} = - \overline{S}^\ell_{ij} \,\tau^\ell(u_i,u_j)$. Due to the index symmetry of the SGS stress tensor, only the symmetric part of the gradient tensor of . plays a direct role, as noted previously.
(ii) $\boldsymbol{y}, \boldsymbol{z} \rightarrow \boldsymbol{u}$; $\boldsymbol{x} \rightarrow \boldsymbol{\omega} = \nabla \times \boldsymbol{u}$, the vorticity. This corresponds to the Navier–Stokes helicity flux, $\Pi^{H,\ell} = - 2\,\overline{S}^\ell_{\!\omega \, ij} \,\tau^\ell(u_i,u_j)$. As for the previous case, the symmetry of the SGS stress means that the flux can be written in terms of just the symmetric part of the gradient tensor of vorticity, namely $\overline{S}^\ell_{\! \omega \,ij} = (\partial_j \overline{\omega}^\ell_i + \partial_i \overline{\omega}^\ell_j)\,/\,2$.
(iii) $\boldsymbol{y} \rightarrow \boldsymbol{u},\, \boldsymbol{z} \rightarrow \boldsymbol{b}, \, \boldsymbol{x} \rightarrow \boldsymbol{a}$ together with $\boldsymbol{y} \rightarrow \boldsymbol{b}, \, \boldsymbol{z} \rightarrow \boldsymbol{u}= \nabla \times \boldsymbol{a}$; $\boldsymbol{x} \rightarrow \boldsymbol{a}$, where $\boldsymbol{a}$ such that $\boldsymbol{b} = \nabla \times \boldsymbol{a}$ is the magnetic vector potential. Here, the flux is that for the MHD magnetic helicity, $\Pi^{B,\ell} = -2\, \partial_j \overline{a}_i^\ell \left( \tau^\ell(u_i,b_j) - \tau^\ell(b_i,u_j) \right)$.
(iv) $\boldsymbol{x}, \boldsymbol{y}, \boldsymbol{z} \in \{\boldsymbol{u}, \boldsymbol{b} \}$: MHD kinetic energy, magnetic energy and cross-helicity fluxes. Regarding the energy fluxes, their exact decompositions and quantifications are discussed in detail in the main body of this work. Decomposition of the MHD helicity fluxes will be examined in a future paper.
(v) As an additional level of analysis, we may also consider various projections of the fields onto subspaces of particular interest. For instance, after decomposing the velocity field – using a basis constructed from eigenfunctions of the curl operator – into positively and negatively helical fields $\boldsymbol{u}^{\pm}$, such that $\boldsymbol{u} = \boldsymbol{u}^+ + \boldsymbol{u}^-$ (Waleffe Reference Waleffe1993; Lessinnes et al. Reference Lessinnes, Plunian, Stepanov and Carati2011; Linkmann et al. Reference Linkmann, Berera, McComb and McKay2015; Alexakis Reference Alexakis2017; Alexakis & Biferale Reference Alexakis and Biferale2018; Yang et al. Reference Yang, Linkmann, Biferale and Wan2021), the following SGS stresses occur in the evolution equations for .:
(A10)\begin{gather} \tau_{ij}^\ell(\boldsymbol{u}^\pm, \boldsymbol{u}^\pm) = \overline{u_i^\pm u_j^\pm}^\ell - \overline{u_i^\pm}^\ell \overline{u_j^\pm}^\ell , \end{gather}(A11)\begin{gather}\tau_{ij}^\ell(\boldsymbol{u}^\pm, \boldsymbol{u}^\mp) = \overline{u_i^\pm u_j^\mp}^\ell - \overline{u_i^\pm}^\ell \overline{u_j^\mp}^\ell , \end{gather}(A12)\begin{gather}\tau_{ij}^\ell(\boldsymbol{u}^\mp, \boldsymbol{u}^\pm) = \overline{u_i^\mp u_j^\pm}^\ell - \overline{u_i^\mp}^\ell \overline{u_{j_{\vphantom{\frac{A}{A}}}}^\pm}^\ell . \end{gather}
Appendix B. Two extended Betchov relations for MHD
Here, we derive two MHD analogues of the exact kinematic relation between components of the velocity-gradient tensor introduced by Betchov (Reference Betchov1956) and use them to obtain relations between several MHD energy subfluxes.
Recall that Betchov (Reference Betchov1956) showed that $\langle A_{ij} \, A_{jk} \, A_{ki} \rangle$, where $A_{ij} = \partial_j u_i$, etc. As a first step, we wish to prove a similar relation between gradients of filtered fields, where the filtering scale on each field need not be the same. Specifically, we demonstrate that
where $\ell, m$ are two generic filtering scales, and $B_{ij} = \partial_j b_i$ is an (unfiltered) gradient tensor related to a solenoidal magnetic field. The above gradient tensors can be, in principle, calculated in different positions.Footnote 4 Using incompressibility and periodic boundary conditions one obtains
This yields (B1) since the averages of the gradients vanish when the boundary conditions are periodic (or the system is homogeneous), and we are left with a quantity equal to its negative.
The next step is to decompose each gradient tensor of (B1) in terms of its symmetric and antisymmetric parts:
Exploiting the symmetries of the tensors involved yields the identity
This can be considered as a generalised Betchov identity for MHD. As a special case, we note that if the magnetic field becomes equal to the velocity field and we remove the filters, then (B4) collapses to the standard Betchov relation.
Equation (B4) is multi-scale but not in the form of energy fluxes. To obtain such a relation, we calculate its convolution with the Gaussian filter, with filter scale $\phi = \sqrt{\ell^2 - m^2}$, and integrate over the filter scale $m$, following what was done in the right-hand side of (2.15). The result is
which corresponds to (4.6) in the main body of the paper. The single-scale version of this relation ($m$ subscripts replaced with $s$) also holds, as can be seen by setting $\ell=m$ in (B4), i.e.
Note that this mixes terms from the momentum equation with one from the induction equation, with the last term being equivalent to $\langle \Pi^{M,\ell}_{s,\Omega \Sigma J} \rangle$ by the cyclic property of the trace.
In § 4.2, we infer from the simulation p.d.f.s that $\Pi^{M,\ell}_{m,S \Sigma \Sigma} \approx \Pi^{M,\ell}_{m,S JJ}$, where both these terms appear as averaged quantities in (B5). We can recover the pointwise identity relative to the subfluxes appearing in (B5) taking into account the (not averaged) gradients from the right-hand side of (B2). As a consequence, $\Pi^{M,\ell}_{m,\Omega \Sigma J}$ should be cancelled by the contribution from the gradients.
In addition to (B1), we can prove another exact identity that reads
This is obtained by employing incompressibility and periodic boundary conditions on
Clearly, the structure of this term may be of interest for Advection and Dynamo subfluxes. Decomposing each gradient tensor in terms of the symmetric and antisymmetric parts yields
and, following manipulations similar to those yielding (B5), this can be mapped into a relation between subfluxes
whose single-scale counterpart coincides with (B6). These two relations may be used to write the decomposition of the total MHD energy flux more compactly and to assist with physical interpretations.
B.1 Further observations
From figure 6, it can be observed that the multi-scale terms $\langle \Pi^{M,\ell}_{m,S \Sigma \Sigma} \rangle$ and $\langle \Pi^{M,\ell}_{m,S J J } \rangle$ are approximately equal, albeit being very small compared with terms of type $SJ\Sigma$. This is reminiscent of the similar relation for two multi-scale Inertial subfluxes discussed in § 4.1. However, in the present case, the structure of the fields is different since the subfluxes of $\Pi^{M,\ell}$ are formed from one velocity gradient tensor and two magnetic gradient tensors.
Since our numerical results indicate that $\langle \Pi^{M,\ell}_{m,S \Sigma \Sigma} \rangle \approx \langle \Pi^{M,\ell}_{m,S J J} \rangle$, (4.6) implies that $\langle \Pi^{M,\ell}_{m,\Omega \Sigma J} \rangle \approx 0$, as is also seen for the $\Pi^{I,\ell}_{m, \Omega S \Omega}$ Inertial term (in both the HD and MHD cases) that has the same symmetric/antisymmetric tensorial structure. Equation (4.6) reveals that the difference between $\langle \Pi^{M,\ell}_{m,S \Sigma \Sigma} \rangle$ and $\langle \Pi^{M,\ell}_{m,S J J } \rangle$ is governed by $\langle \Pi^{M,\ell}_{m, \Omega \Sigma J} \rangle$, a term that does not contribute to the energy balance, because in (2.9), only the symmetric part of the gradient tensor survives after the contraction with the symmetric SGS stress $\tau^\ell(b_i,b_j)$. Thus, we may posit a physical explanation for why $\langle \Pi^{M,\ell}_{m, \Omega \Sigma J} \rangle \approx 0$ by arguing that the values of $\langle \Pi^{M,\ell}_{m,S \Sigma \Sigma} \rangle$ and $\langle \Pi^{M,\ell}_{m,S J J } \rangle$ are essentially determined by the energy balance of the system, and hence cannot be altered by a quantity that does not contribute to this.
Furthermore, we observe in figure 7(b) that the p.d.f.s for $\Pi^{M,\ell}_{m,S \Sigma \Sigma}$ and $\Pi^{M,\ell}_{m,S J J }$ are roughly coincident, especially along the tails. Recall that a similar feature was seen with the analogous Inertial multi-scale subfluxes. Further quantitative confirmation of this approximate congruence is given by the similarity of the relevant moments listed in table 2. These types of approximate identity hold when there is an interplay of either three velocity gradient tensors (Inertial term – HD and MHD) or one velocity and two magnetic gradient tensor (Maxwell term). However, they do not occur when we study the same structure of subfluxes associated with one vorticity and two velocity gradient tensors in the context of helicity flux (see Capocci et al. (Reference Capocci, Johnson, Oughton, Biferale and Linkmann2023)). This suggests that the approximate identity is unlikely to be of kinematic origin, although the exact version, (4.6), is a kinematic result.
Appendix C. Two-dimensional MHD
C.1 Algebraic setting
In the 2-D case, we can express the strain-rate and rotation-rate tensors associated with the incompressible field $\boldsymbol{\mathsf{X}} = (X_1,X_2)$ in the following way:
where we have already enforced the incompressibility on the trace of (C1) and defined $\omega^X = (\partial_1 X_2 - \partial_2 X_1)$ to make the notation more compact. Given an additional incompressible field $\boldsymbol{\mathsf{Y}}$, it is straightforward to verify that $\boldsymbol{\mathsf{S}}_Y$ and $\boldsymbol{\Omega}_X$ satisfy the commutator algebra:
where, unlike the general 3-D scenario, the product $\boldsymbol{\mathsf{S}}_Y \boldsymbol{\cdot} \boldsymbol{\varOmega}_X$ is a symmetric and traceless tensor. It is also useful to note that the product of two strain-rate tensors related to different gradient tensors can be decomposed as the sum of a symmetric tensor and an antisymmetric one:
where $\mathbb{I}$ is the $2 \times 2$ identity matrix and $\hat{\Omega}$ is the antisymmetric and traceless matrix that defines the 2-D rotation-rate tensor of (C2). The (scalar) auxiliary functions $\sigma$ and $\alpha$ embody the functional part multiplying $\mathbb{I}$ and $\hat{\Omega}$, respectively; moreover, they are respectively symmetric and antisymmetric under argument exchange symmetry, i.e.
Thus, if we swap the fields $\boldsymbol{\mathsf{X}}$, $\boldsymbol{\mathsf{Y}}$ in (C5), we obtain
In terms of tensor traces, the decomposition of (C5) and (C3) lead to three relevant identities, viz.:
where (C10) vanishes because $\boldsymbol{\Omega}_Y \, \boldsymbol{\Omega}_Z \propto \mathbb{I}$ and $\boldsymbol{\mathsf{S}}_X$ is traceless.
With these properties in mind, we are ready to specialise (2.8)–(2.9) to the 2-D MHD situation. This will involve appropriate simplifications of the single/multi-scale decompositions like (2.17).
C.2 SGS energy subfluxes
As a consequence of (C8)–(C10), in both single-scale and multi-scale cases, the terms involving the contraction of either three strain-rate tensors, three rotation rate tensors, or one strain-rate and two rotation rate tensors vanish. Hence, the 2-D Inertial SGS flux is solely due to $\Pi^{I,\ell}_{m, S \Omega S}$ (Johnson Reference Johnson2021):
As mentioned in § 4.1, this is the only Inertial term that ‘survives’ in 3-D MHD (i.e. is not approximately zero; see figure 3), especially if we add a background magnetic field in the equations of motion (not shown).
The 2-D Maxwell flux contains one single-scale and one multi-scale term,
and the Dynamo and the Advection fluxes formally contain two further multi-scale terms:
After straightforward algebraic manipulations, we obtain the expression for $\Pi^\ell$ that corresponds to the total energy flux for 2-D MHD filtered at the scale $\ell$:
Clearly, this has one single-scale and three distinct multi-scale contributions.
Appendix D. Equipartition subrange p.d.f.s
Here, we show some p.d.f.s for the MHD energy subfluxes for a filter scale that lies in the region where there is approximate equipartition between the magnetic and kinetic energy fluxes. The p.d.f.s presented in the main body of the paper are calculated for a larger scale.
Figure 11 displays the net kinetic energy flux, $\Pi^{I,\ell} + \Pi^{M,\ell}$, and the net magnetic energy flux, $\Pi^{A,\ell} + \Pi^{D,\ell}$, obtained using the Fourier filter and dataset A1. In essence, it is a rearrangement of figure 2. An equipartition region is evident for $0.2 \lesssim k\eta_\alpha \lesssim 0.6$, where the magnetic and kinetic energy subfluxes reach approximately $50\%$ of the total energy flux, as indicated by the green dashed line. See Bian & Aluie (Reference Bian and Aluie2019) for a discussion of this feature. Moreover, this equipartition region is also the region where the conversion term . of (2.7) saturates and becomes scale-independent.
Figure 12 displays energy (sub)fluxes for MHD dataset A1 for the filter scale $k\eta_\alpha = 0.27$. Comparing these figures with those presented in § 4, it is apparent that the p.d.f.s in the equipartition region have fluctuations that are some three times larger. Recall that the p.d.f.s in figures 4, 7 and 10 were calculated for the larger scale $k\eta_\alpha = 5.4 \times 10^{-2}$. As we approach the dissipative range, the p.d.f.s develop even broader tails (not shown).
Appendix E. Comparison with standard diffusive MHD
The aim of this section is to repeat part of the analyses from §§ 3–4 on dataset A4 (see table 3) that employs the standard diffusive MHD equations. The results are discussed and compared with those from the hyper-dissipative run.
In figure 15, the kinetic and magnetic spectra of figure 1 and the standard diffusive MHD counterparts have been displayed together. Apart from the similarities concerning the peaks and the large-scale behaviour (connected with the identical forcing scheme), we observe that spectra from standard viscosity dataset A4 present a shorter power-law scaling and a consequent smoother fall in the dissipative range. Nonetheless, corresponding spectra are qualitatively similar.
Figure 16 describes the net kinetic energy fluxes, $\Pi^{I,\ell}+ \Pi^{M,\ell}$, and the net magnetic energy fluxes, $\Pi^{A,\ell} + \Pi^{D,\ell}$, obtained via employment of the Fourier filter on both the standard-diffusivity dataset A4 and the hyperviscous-diffusivity dataset A1. This plot can be considered as a rearrangement of figure 2, with the RSC term omitted. The most relevant difference between the two datasets is a striking reduction of the bandwidth of the inertial range related to dataset A4. From a subfluxes perspective, this feature is accompanied by the absence of an equipartition range: the magnetic and kinetic energy subfluxes do not each reach $\approx 50\%$ of the total energy flux, the level indicated by the green line.
With regards to the exact decomposition of the SGS stresses, the panels of figure 17 illustrate the standard-dissipation counterparts of figures 3(b), 6 and 8, which are being used as a comparison for the following analysis. Starting from panel (a), we notice a weaker depletion in the averaged Inertial transfer whose peak, at $k\eta \approx 0.01$, accounts for 20$\%$ of the total energy flux. We consider this feature to be a low-$\text{Re}$ effect. To reinforce this view, we note that the maximum of $\langle \Pi^{I,\ell}\rangle$ increases as $Re$ decreases (not shown) and that a similar profile in the mean Inertial transfer was already observed by Offermans et al. (Reference Offermans, Biferale, Buzzicotti and Linkmann2018). However, an increase in the Inertial transfer is compensated by an overall decrease of the Maxwell flux in panel (b), where the value of the latter is slightly reduced (relative to A1), diminishing even more as $Re$ decreases (not shown). Moreover, $\Pi^{M,\ell}$ decays faster as $k\eta$ increases as a consequence of a shift towards the small-scales of the maximum of $\Pi^{M,\ell}_{s,SJ\Sigma}$ and to a less pronounced skewness of $\Pi^{M,\ell}_{m,SJ\Sigma}$. Thus, we conclude that these two effects are responsible for the lack of an equipartition range where the total kinetic subflux $\Pi^{I,\ell} + \Pi^{M,\ell}$ flattens reaching equipartition with the total magnetic flux $\Pi^{D,\ell} + \Pi^{A,\ell}$. Finally, both the total advection and dynamo subfluxes, respectively panels (c) and (d), present relatively decreased values at their peaks compared with the corresponding curves from hyperdissipative dataset A1; as expected, those maxima also keep decreasing with $Re$ (not shown). In addition, $\langle \Pi^{A,\ell}\rangle$ shows, on logarithmic scale, a remarkable even-symmetry with respect to $k\eta\approx 0.12$ that can be extended to the coincident averages of $\Pi^{A,\ell}_{s,\Sigma J S}$ and $\Pi^{A,\ell}_{s, J \Sigma S}$. In conclusion, it is worth emphasising that the absence of a flattening region of the total magnetic subflux $\Pi^{A,\ell} + \Pi^{D,\ell}$ is ostensibly caused by a shift towards the large scales of the mean advection flux combined with a substantial reduction of the averaged total dynamoFootnote 5 subflux.
Appendix F. Subfluxes definitions
In this section, we provide the definitions of all the subfluxes appearing in the decomposition of the MHD energy fluxes. As highlighted in § 3, for the subfluxes $\Pi^{I,\ell}_{s,S \Omega S}$, $\Pi^{I,\ell}_{m,S \Omega S}$ and $\Pi^{M,\ell}_{s,S J \Sigma}$, $\Pi^{M,\ell}_{m,S J \Sigma}$, there is an extra factor of two that arises from the symmetry of the corresponding SGS stress tensors. Because in (2.6) and (2.7) the fluxes appear with the same leading signs, both the Maxwell and the Dynamo subfluxes in the definition below acquires an additional minus sign.
F.1 Inertial
The following subfluxes are identical to the hydrodynamic counterpart from Johnson (Reference Johnson2020, Reference Johnson2021):
F.2 Maxwell
F.3 Advection
F.4 Dynamo