Scaling of Turbulent Viscosity and Resistivity: Extracting a Scale-dependent Turbulent Magnetic Prandtl Number

, , , , and

Published 2021 August 9 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Xin Bian et al 2021 ApJL 917 L3 DOI 10.3847/2041-8213/ac0fe5

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

2041-8205/917/1/L3

Abstract

Turbulent viscosity νt and resistivity ηt are perhaps the simplest models for turbulent transport of angular momentum and magnetic fields, respectively. The associated turbulent magnetic Prandtl number Prtνt/ηt has been well recognized to determine the final magnetic configuration of accretion disks. Here, we present an approach to determining these "effective transport" coefficients acting at different length scales using coarse-graining and recent results on decoupled kinetic and magnetic energy cascades. By analyzing the kinetic and magnetic energy cascades from a suite of high-resolution simulations, we show that our definitions of νt, ηt, and Prt have power-law scalings in the "decoupled range." We observe that Prt ≈ 1–2 at the smallest inertial-inductive scales, increasing to ≈5 at the largest scales. However, based on physical considerations, our analysis suggests that Prt has to become scale independent and of order unity in the decoupled range at sufficiently high Reynolds numbers (or grid resolution) and that the power-law scaling exponents of velocity and magnetic spectra become equal. In addition to implications for astrophysical systems, the scale-dependent turbulent transport coefficients offer a guide for large-eddy simulation modeling.

Export citation and abstract BibTeX RIS

1. Introduction

Magnetohydrodynamic (MHD) turbulence is central to our understanding of many astrophysical systems, including the solar wind, interstellar medium (ISM), and accretion disks.

Most of these systems are characterized by very large Reynolds numbers (Re). For example, $\mathrm{Re}\sim {10}^{5}\mbox{--}{10}^{7}$ in the cool ISM (Elmegreen & Scalo 2004), $\mathrm{Re}\sim 4\times {10}^{6}$ in the solar wind (Verma 1996), and $\mathrm{Re}\sim {10}^{14}$ in Type Ia supernovae (Kuhlen et al. 2006). High-Re turbulent flows involve a wide range of dynamical scales, called the "inertial-inductive" range, over which the evolutions of the flow and magnetic field are immune from the direct effects of external forcing and microphysical dissipation. Similar to hydrodynamic turbulence, it is widely expected that MHD turbulence over the inertial-inductive range has universal statistics with power-law spectra, although details of such a scaling remain a subject of debate (Goldreich & Sridhar 1995; Biskamp 2003; Verma 2004, 2019; Zhou et al. 2004; Boldyrev 2005; Schekochihin 2020). While the large scales in a high-Re MHD flow are immune from the direct effects of microphysical transport (Aluie 2017; Zhao & Aluie 2018), they are indirectly influenced by the microphysics due to the "catalytic" role of turbulence via the cascade process, which acts as a bridge between the large and microphysical scales. For example, it is widely believed that turbulence plays an important role in the outward transport of angular momentum in accretion disks for inward mass accretion (Balbus & Hawley 1998).

The simplest conceptual framework to think of turbulence is as an effective (or turbulent) viscosity νt , which leads to the "turbulent diffusion" of angular momentum at scales far larger than viscous scales, and has long shaped our thinking of accretion disk dynamics (Shakura & Sunyaev 1973). Similarly, magnetic fields, which are essential for launching and collimating jets (Blandford & Znajek 1977; Blandford & Payne 1982; Jafari & Vishniac 2018), can be transported outward by an effective (or turbulent) resistivity ηt . In this way, the magnetic field configuration in accretion disks may be influenced by a balance between the inward advection by accretion and the outward diffusion by turbulent resistivity (Lubow et al. 1994; Fromang & Stone 2009; Guan & Gammie 2009; Lovelace et al. 2009; Cao 2011). This balance between the competing effects of νt and ηt is captured by the turbulent magnetic Prandtl number Prt νt /ηt . Whether global-scale structures or turbulent stress dominates the overall angular momentum transport is still an open question and important for determining the budget of thermal versus nonthermal emission (Blackman & Nauman 2015).

For turbulent astrophysical flows, current computing resources are unable to solve all relevant scales. Large-eddy simulations (LESs) rely on subgrid-scale modeling to represent the small-scale effects on resolved scales (Meneveau & Katz 2000; Miesch et al. 2015). Müller & Carati (2002), Chernyshov et al. (2007), and Grete et al. (2015) studied different subgrid-scale (SGS) models. Renormalization group (RG) analysis was used to develop scale-dependent turbulent coefficients (Zhou 2010). However, the studies on MHD scale-dependent turbulent transport coefficients are few compared to hydrodynamic turbulence.

We remind readers that the turbulent magnetic Prandtl number is different from the microscopic magnetic Prandtl number Prm ν/η, where ν is the microscopic viscosity, and η is the microscopic resistivity. Prm is large in the ISM while being small in stellar interiors and liquid metals (Davidson et al. 2012). Many studies have focused on the effect of Prm (e.g., Lesur & Longaretti 2007; Fromang & Stone 2009; Brandenburg 2014; Brandenburg & Rempel 2019). The extent to which existing simulations accurately capture the physics of realistic extreme regimes of low and high Prm is uncertain.

In this paper we focus on Prt , not Prm . Turbulent transport coefficients have been studied both analytically and numerically. Estimates using mixing length theory νt ηt U /3 (characteristic velocity U and characteristic scale ; Yousef et al. 2003; Käpylä et al. 2020) are consistent to order of magnitude with ηt calculated with the test-field method (Käpylä et al. 2009) and shearing box simulations (Snellman et al. 2009). The quasilinear approximation (Kitchatinov et al. 1994; Yousef et al. 2003) and RG analysis (Forster et al. 1977; Fournier et al. 1982; Verma 2001a, 2001b) suggested that 0.4 < Prt < 0.8. Zhou et al. (2002) developed eddy and backscatter viscosity and resistivity using eddy-damped quasinormal Markovian statistical closure model (EDQNM).

Numerical studies have traditionally identified "turbulence" as fluctuations from a (temporal or ensemble) mean flow and have typically yielded Prt ≈ 1. Yousef et al. (2003) measured Prt from the decaying large-scale fields in forced turbulence simulations. The results showed that Prt is near unity and insensitive to Prm . These simulations were conducted with a fixed small magnetic Reynolds number. Several groups studied the turbulent transport coefficients using shearing box simulations (Fromang & Stone 2009; Guan & Gammie 2009; Lesur & Longaretti 2009). Guan & Gammie (2009) inferred ηt from the evolution of an imposed magnetic field perturbation in an already turbulent flow. Lesur & Longaretti (2009) imposed an external magnetic field and defined ηt using the electromotive force induced by the field. Fromang & Stone (2009) calculated ηt from the spatially varying magnetic fields induced by an electromotive term added in the induction equation. νt was defined using Reynolds and Maxwell stress tensors in these studies. Despite different definitions, numerical schemes, and magnetic field configurations among these studies, they all find Prt ≈ 1. Käpylä et al. (2020) computed νt using both Reynolds stress and the decay rate of a large-scale field, and ηt using the test-field method, where a set of test fields are used to calculate the components of turbulent diffusivity tensors (Schrinner et al. 2005, 2007). The results suggested that Prt increases with increasing Reynolds number and saturates at large Reynolds number with 0.8 ≤ Prt ≤ 0.95.

Other than the RG and EDQNM analyses, the aforementioned studies did not analyze νt and ηt as a function of length scales, which is not possible from a Reynolds (mean versus fluctuation) decomposition (e.g., Fromang & Stone 2009; Guan & Gammie 2009; Lesur & Longaretti 2009; Käpylä et al. 2020). Determining the scale dependence of transport coefficients can improve the fidelity with which we characterize astrophysical turbulence in cohort with its practical application to subgrid-scale modeling. For example, if Prt α with α > 0, Prt grows at larger scales, indicating that the large-scale component of a flow, which is still part of the "fluctuations," feels a stronger νt relative to ηt .

Our study aims to define and measure νt , ηt , and Prt at different scales using the coarse-graining approach (Eyink 2005; Aluie 2017) and the eddy-viscosity hypothesis (Boussinesq 1877). Our analytical and numerical results show power-law scaling of the turbulent transport coefficients in the "decoupled range" over which the kinetic and magnetic cascades statistically decouple and become conservative (Bian & Aluie 2019).

2. Methodology

2.1. Coarse-grained Energy Equations

We analyze the incompressible MHD equations with a constant density ρ:

Equation (1)

Equation (2)

Equation (3)

where u is the velocity and B is the magnetic field normalized by $\sqrt{4\pi \rho }$ to have Alfvén (velocity) units. p is pressure, J = × B is (normalized) current density, f is external forcing, and ν and η are microscopic viscosity and resistivity, respectively.

We use the coarse-graining method to analyze the flow and define the turbulent magnetic Prandtl number. A coarse-grained field in n dimensions ${\overline{f}}_{{\ell }}({\boldsymbol{x}})=\int {d}^{n}{\boldsymbol{r}}{G}_{{\ell }}({\boldsymbol{x}}-{\boldsymbol{r}})f({\boldsymbol{r}})$ contains modes at length scales greater than , where G ( r ) ≡ n G( r /) is a normalized kernel with its main weight in a ball of diameter . The coarse-grained MHD equations for ${\bar{u}}_{{\boldsymbol{\ell }}}$, ${\bar{B}}_{{\boldsymbol{\ell }}}$, and the quadratic MHD invariants were shown by Aluie (2017). Hereafter, we drop subscript when possible.

The coarse-grained kinetic energy (KE) and magnetic energy (ME) density balance (at scales > ) are

Equation (4)

Equation (5)

where ·[ ⋯ ] denotes spatial transport terms, S = ( u + u T )/2 is the strain-rate tensor, and $\overline{{\boldsymbol{f}}}{\boldsymbol{\cdot }}\overline{{\boldsymbol{u}}}$ is the energy injection rate at forcing scale f = 2π/kf (kf are the modes of the forcing f ). Microscopic dissipation terms $\nu | \overline{{\boldsymbol{S}}}{| }^{2}$ and $\eta | \overline{{\boldsymbol{J}}}{| }^{2}$ are mathematically guaranteed (Aluie 2017; Eyink 2018) and numerically demonstrated (Zhao & Aluie 2018; Bian & Aluie 2019) to be negligible at scales ${\ell }\gg \left({{\ell }}_{\nu },{{\ell }}_{\eta }\right)$, where ν and η are the viscous and resistive length scales, respectively.

The KE cascade term ${\bar{{\rm{\Pi }}}}_{\ell }^{u}\equiv -{\bar{{\boldsymbol{S}}}}_{\ell }:{\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$ in Equation (4) quantifies the KE transfer across scale , where ${\overline{\tau }}_{{ij}}\,\equiv {\tau }_{{\ell }}({u}_{i},{u}_{j})-{\tau }_{{\ell }}({B}_{i},{B}_{j})$ is the sum of subscale Reynolds and Maxwell stresses generated by scales < acting on the large-scale strain ${\overline{S}}_{{ij}}$, resulting in "turbulent viscous dissipation" to scales < . Subscale stress is defined as ${\tau }_{{\ell }}(f,g)={\overline{\left({fg}\right)}}_{{\ell }}-{\overline{f}}_{{\ell }}{\overline{g}}_{{\ell }}$ for any two fields f and g. Similarly, the ME cascade term ${\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\equiv -{\overline{{\boldsymbol{J}}}}_{{\ell }}{\boldsymbol{\cdot }}{\overline{{\boldsymbol{\varepsilon }}}}_{{\ell }}$ in Equation (5)) quantifies the ME transfer across scale , where the subscale electromotive force (EMF) ${\bar{{\boldsymbol{\varepsilon }}}}_{\ell }\equiv \bar{{\boldsymbol{u}}{\rm{\times }}{\boldsymbol{B}}}-\bar{{\boldsymbol{u}}}{\rm{\times }}\bar{{\boldsymbol{B}}}$ is (minus) the electric field generated by scales < acting on the large-scale current $\bar{{\boldsymbol{J}}}={\rm{\nabla }}\times \bar{{\boldsymbol{B}}}$, resulting in "turbulent ohmic dissipation" to scales < . Both ${\overline{{\rm{\Pi }}}}_{{\ell }}^{u}$ and ${\overline{{\rm{\Pi }}}}_{{\ell }}^{b}$ appear as sinks in the energy budgets of large scales > and as sources in the energy budgets of small scales < (Aluie 2017).

The term ${\overline{S}}_{{ij}}{\overline{B}}_{i}{\overline{B}}_{j}$ quantifies KE-to-ME conversion at all scales > and appears as a sink in Equation (4) and a source in Equation (5). Bian & Aluie (2019) showed that $\langle {\overline{S}}_{{ij}}{\overline{B}}_{i}{\overline{B}}_{j}\rangle $ (〈 ⋯ 〉 denotes a spatial average) is a large-scale process, which only operates at the largest scales in the inertial-inductive range (which was called the "conversion range") and vanishes at intermediate and small scales in the inertial-inductive range (which was called the "decoupled range"). In the decoupled range, $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ become constant as a function of scale (i.e., scale independent). The observation of constant KE and ME fluxes $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ is important because it indicates separate conservative cascades of each of KE and ME, which arises asymptotically at high Reynolds number regardless of forcing, external magnetic field, and microscopic magnetic Prandtl number.

2.2. Scaling of Turbulent Transport Coefficients

Oftentimes, turbulence is modeled as a diffusive process via effective (or turbulent) transport coefficients. For example, mixing length or eddy-viscosity models represent the subscale stress ${\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$, due to scales < , as ${\overline{\tau }}_{{ij}}=-2{\nu }_{t}^{{\boldsymbol{x}}}{\overline{S}}_{{ij}}$, where ${\nu }_{t}^{{\boldsymbol{x}}}$ is a turbulent viscosity 4 (e.g., Pope 2001 and references therein). Similarly, the subscale EMF can be modeled as $\bar{{\boldsymbol{\varepsilon }}}=-{\eta }_{t}^{{\boldsymbol{x}}}\bar{{\boldsymbol{J}}}+\alpha \bar{{\bf{B}}}$, where the ${\eta }_{t}^{{\boldsymbol{x}}}\overline{{\boldsymbol{J}}}$ term models the subscales as turbulent resistive diffusion (Miesch et al. 2015) and the $\alpha \overline{B}$ term is the α-effect of dynamo theory (Moffatt 1978). The α-effect is expected to play a role in flows where the driving mechanism is helical. To simplify our analysis and the presentation of our approach, we shall ignore the $\alpha \bar{{\bf{B}}}$ term and assume that the subscale EMF can be modeled solely as ohmic diffusion, $\overline{{\boldsymbol{\varepsilon }}}=-{\eta }_{t}^{{\boldsymbol{x}}}\overline{{\boldsymbol{J}}}$. Note that ${\nu }_{t}^{{\boldsymbol{x}}}({\boldsymbol{x}},t,{\ell })$ and ${\eta }_{t}^{{\boldsymbol{x}}}({\boldsymbol{x}},t,{\ell })$ are generally functions of space x , length scale , and time.

The main goal of this paper is extracting the turbulence transport coefficients, ${\nu }_{t}^{{\boldsymbol{x}}}$ and ${\eta }_{t}^{{\boldsymbol{x}}}$, as a function of length scale. However, we do not pursue a phenomenological analysis similar to that of Smagorinsky (1963) or of a mixing length framework (Tennekes & Lumley 1972) in part because we lack a consensus MHD turbulence theory analogous to that of Kolmogorov (1941). To achieve our goal, we shall instead analyze the energy budgets resulting from the eddy-viscosity model. Within our coarse-graining framework, this is equivalent to having the rate of energy cascading to scales smaller than equal a turbulent dissipation acting on scales > :

Equation (6)

Equation (7)

These two relations are definitions for νt and ηt . Note that unlike in relation ${\overline{\tau }}_{{ij}}=-2{\nu }_{t}^{{\boldsymbol{x}}}{\overline{S}}_{{ij}}$, the turbulence transport coefficients in Equations (6)–(7) are defined using scalar quantities $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $, $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $, $\langle | {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $, and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle $. For homogeneous turbulence considered in this study, we rely on spatial averages, 〈...〉, rendering νt and ηt independent of location x but still a function of scale .

Consistent with the eddy-viscosity hypothesis, Equation (6) (Equation (7)) models the kinetic (magnetic) energy cascading from scales > to smaller scales as effectively being dissipated by a turbulent viscosity (resistivity). From these, we can also extract a scale-dependent turbulent magnetic Prandtl number,

Equation (8)

What power-law scaling can we expect these turbulent transport coefficients to have? It is possible to relate νt and ηt to energy spectra. Indeed, the space-averaged turbulent dissipation can be expressed in terms of energy spectra:

Equation (9)

Equation (10)

where Eu (k) (Eb (k)) is the kinetic (magnetic) energy spectrum with (dimensionless) wavenumber k = L/ for a periodic domain of size L.

The scaling of spectra in turn is related to the scaling of velocity and magnetic field increments (Aluie 2017)

Equation (11)

where the increment δ f(x; ) = f(x + ) − f(x) (see details in Eyink 2005; Aluie & Eyink 2010; Aluie 2017). From Equation (11), the kinetic and magnetic energy spectra scale as

Equation (12)

The relation between increments and spectra does not make any assumptions about the specific exponent values, only that they are σu,b < 1 (Sadek & Aluie 2018). Scaling exponents σu and σb are a measure of smoothness of the velocity and magnetic fields, respectively (see Figure 1 and related discussion in Aluie 2017). A value of σ = 1 indicates that the field is very smooth (e.g., of a laminar flow) with a spectrum decaying as k−3 or steeper. Canonical hydrodynamic turbulence has intermediate smoothness, with σu = 1/3 according to Kolmogorov (1941, hereafter K41). The larger the value of σ, the smoother is the field.

Figure 1.

Figure 1. Panels (a)–(b) show νt , ηt , and Prt calculated using their respective definitions in Equations (6)–(8), at different scales k = L/. Panels (c)–(d) show an alternate calculation of νt and ηt from Equations (17)–(18). We use the highest-resolution runs of Run III and IV (Taylor–Green forcing) in Table 1. Three reference lines with a slope of −1/3, −5/3 (black dashed–dotted), and −4/3 (black solid) are added. Note the reference line of −1/3 and Prt use the right y-axis, while others use the left y-axis. Scales < d are not shown. Simulations with helical forcing are shown in Figure 7 in Appendix.

Standard image High-resolution image

For sufficiently high-Reynolds-number flows, Bian & Aluie (2019) showed that each of $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ becomes constant, independent of scale in the decoupled range. From definitions given by Equations (6)–(8), and considering the scaling relations discussed above, we can infer that the turbulent transport coefficients vary with scale as follows:

Equation (13)

for scales k in the decoupled range. It is possible to obtain the scaling relations in Equation (13) from either the scaling of spectra in Equations (9)–(10), or the scaling of coarse-grained strain and current, $| {\overline{{\boldsymbol{S}}}}_{{\ell }}| \sim \delta u({\ell })/{\ell }$ and $| {\overline{{\boldsymbol{J}}}}_{{\ell }}| \sim \delta B({\ell })/{\ell }$ (Eyink et al. 2013; Aluie 2017). Equation (13) highlights that Prt is independent of scale only if σu = σb .

Regardless of the specific value, and consistent with existing MHD turbulence phenomenologies, we expect that σu,b < 1. Indeed, a σu,b ≥ 1 would correspond to a smooth flow that is inconsistent with the qualitative expectation of a "rough" or "fractal" turbulent flow. Therefore, relations (13) indicate that νt and ηt decay as → 0 (or k). This is consistent with physical expectations because the "eddies" affecting the turbulent transport become weaker at smaller scales, yielding smaller transport coefficients.

We highlight a technical, albeit important, aspect of scaling relations in Equation (13). Our coefficients seem to scale with the inverse of coarse-grained strain and current magnitudes, ${\nu }_{t}\sim | {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{-2}\sim {{\ell }}^{2-2{\sigma }_{u}}$ and ${\eta }_{t}\sim | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{-2}\sim {{\ell }}^{2-2{\sigma }_{b}}$, but do not appear to depend on the subscale stress and EMF, ${\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$ and ${\bar{{\boldsymbol{\varepsilon }}}}_{\ell }$, respectively. At face value, this result seems counterintuitive wherein σ → 1 associated with smoother fields and weaker "eddies" lead to an increase rather than a drop in the turbulent coefficient values in Equation (13). However, a key assumption to arrive at the relations in Equation (13) is that fluxes $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ are constant, independent of scale. For scale-independent fluxes to be established, consistent with a persistent cascade to arbitrarily small scales (in the $\mathrm{Re}\to \infty $ limit), σu and σb have to take on fixed values that are yet to be determined and agreed upon by the community. If σu,b were to be somehow increased above those values, the cascade would shut down (fluxes would decay with k) before carrying the energy all the way to dissipation scales (Aluie 2017). For scale-dependent fluxes, the relations in Equation (13) have to be modified to also include the scaling of $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ (see Aluie 2017 for details).

To infer the scaling of turbulence transport coefficients, the approach we adopt in this paper circumvents using values of σu and σb (in the asymptotic $\mathrm{Re}\to \infty $ limit) from a specific MHD phenomenology—whether it exists or not—by relying on results from Bian & Aluie (2019) of scale-independent fluxes $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $.

Under the K41 scaling σu = 1/3 (Kolmogorov 1941), our scaling of ${\nu }_{t}\propto {{\ell }}^{2-2{\sigma }_{u}}\propto {{\ell }}^{4/3}$ is equivalent to that from mixing length theory, ${\nu }_{t}={{\ell }}^{2}| \overline{{\boldsymbol{S}}}| \propto {{\ell }}^{{\sigma }_{u}+1}\propto {{\ell }}^{4/3}$ (Smagorinsky 1963). Our analysis is also compatible with different scaling theories and observations in MHD turbulence (Iroshnikov 1963; Kraichnan 1965; Goldreich & Sridhar 1995; Boldyrev 2005, 2006; Boldyrev & Perez 2009; Boldyrev et al. 2011). For example, solar wind observations (Podesta et al. 2007; Borovsky 2012) suggest that Eu (k) ∼ k−3/2 for the kinetic energy spectrum, corresponding to δ u() ∼ 1/4, and Eb (k) ∼ k−5/3 for the magnetic energy spectrum, corresponding to δ B() ∼ 1/3, yielding

Equation (14)

for k in the decoupled scale range.

2.3. Alternate Measure of the Coefficients

Instead of analyzing the energy budgets to determine νt , ηt , Prt , and their scaling, we can alternatively focus on the budgets for vorticity and current. Similar to Equations (4)–(5), we can derive the budgets

Equation (15)

Equation (16)

Here, ${\bar{Z}}_{\ell }=\bar{{\boldsymbol{\omega }}}{\boldsymbol{\cdot }}{\boldsymbol{\nabla }}{\rm{\times }}({\boldsymbol{\nabla }}{\boldsymbol{\cdot }}\bar{{\boldsymbol{\tau }}})$ and ${\bar{Y}}_{\ell }=-\bar{{\boldsymbol{J}}}{\boldsymbol{\cdot }}{\rm{\nabla }}{\rm{\times }}{\rm{\nabla }}{\rm{\times }}\bar{{\boldsymbol{\varepsilon }}}$ are the only "scale-transfer" terms in the coarse-grained Equations (15)–(16) involving the interaction of subscale terms ${\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$ and ${\bar{{\boldsymbol{\varepsilon }}}}_{{\boldsymbol{\ell }}}$ with large-scale quantities (here, $\overline{{\boldsymbol{\omega }}}$ and $\overline{{\boldsymbol{J}}}$) to cause transfer across scale . From the models ${\overline{{\boldsymbol{\tau }}}}_{{\ell }}=-2{\nu }_{t}{\overline{{\boldsymbol{S}}}}_{{\ell }}$ and ${\bar{{\boldsymbol{\varepsilon }}}}_{{\boldsymbol{\ell }}}=-{\eta }_{t}{\bar{{\boldsymbol{J}}}}_{\ell }$, we have alternate definitions for the turbulent transport coefficients:

Equation (17)

Equation (18)

Note that unlike energy, vorticity and current density are not ideal invariants and, therefore, do not undergo a cascade in the manner energy does. Yet, to the extent that νt and ηt are able to capture the subscale physics embedded in ${\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$ and ${\bar{{\boldsymbol{\varepsilon }}}}_{{\boldsymbol{\ell }}}$, it is reasonable to expect that the turbulent transport coefficients are consistent with the budget of any quantity derived from the underlying dynamics.

In Figure 1, we compare νt and ηt when calculated from Equations (17)–(18) to those obtained from the energy budgets in Equations (6)–(7). We find that the two definitions yield fairly similar results with slight quantitative differences. This consistency lends support to our approach of using the energy budgets to calculate νt and ηt (Equations (6)–(7)) and make inferences about the turbulent diffusion or dissipation of quantities other than energy.

2.4. Implications to Subgrid Modeling

It is almost always the case that astrophysical systems of interest are at sufficiently high Reynolds numbers (both magnetic and hydrodynamic) that it is impossible to simulate the entire dynamic range of scales that exist (Miesch et al. 2015). In practice, most simulations are either explicit or implicit LES, resolving only the large-scale dynamics (Meneveau & Katz 2000). The former includes explicit terms in the equations being solved that model the unresolved subgrid physics, whereas the latter relies on the numerical scheme to act as a de facto model for such missing physics. Our analysis here can offer guidance for tuning the turbulent coefficients when conducting explicit LESs using eddy diffusivity models. It can also offer us insight into whether relying on a similar scheme and grid for simulating both the momentum and magnetic fields is justified.

In the inertial-inductive range, using Equation (6), $| {\overline{{\boldsymbol{S}}}}_{{\ell }}| \sim \delta u({\ell })/{\ell }$, and the ansatz (Aluie 2017)

Equation (19)

where urms = 〈∣ u 21/2 and L is a characteristic large scale such as the integral scale or that of the domain, and we ignore intermittency corrections. We then have

Equation (20)

In an LES with grid spacing Δx, the turbulent viscosity accounting for subgrid scales should be evaluated at a coarse-graining scale c = L/kc proportional to Δx (Pope 2001), where kc = L/c is a dimensionless cutoff wavenumber:

Equation (21)

for $\left({{\ell }}_{\nu },{{\ell }}_{\eta }\right)\ll {{\ell }}_{c}\ll L$, where the dimensionless constant Cu is defined as the proportionality factor of the relation

Equation (22)

Figure 6 in Appendix B shows that Cu is indeed a proportionality constant that is scale independent within the decoupled range, taking on values from 2 to 5 in various simulated flows.

Similarly, the turbulent resistivity at the cutoff wavenumber is

Equation (23)

where Brms = 〈∣BB021/2 (B0 is the uniform external magnetic field), and the dimensionless constant Cb is defined as

Equation (24)

Figure 6 in Appendix B shows that Cb is indeed a proportionality constant that is scale independent within the decoupled range, taking on values from 10 to 15 in various simulated flows.

If the grid is sufficiently fine to resolve some of the scales in the decoupled range, then Equations (21) and (23) simplify to

Equation (25)

Equation (26)

with scale-independent fluxes $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle ={\varepsilon }_{u}$ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle ={\varepsilon }_{b}$. These are the KE and ME cascade rates, which were found by Bian & Aluie (2019) to reach equipartition in the decoupled range, εu = εb = ε/2, half the total energy cascade rate, ε.

Equations (21) and (23) (and Equations (25) and (26)) connect the scaling of turbulent transport coefficients with the scaling of velocity and magnetic spectra and are compatible with different MHD scaling theories. For example, if Eu (k) ∼ k−5/3 (corresponding to δ u() ∼ 1/3) as in the theory by Goldreich & Sridhar (1995), Equation (21) reduces to the turbulent viscosity model of Verma (2001a) derived from an RG analysis (see also Verma & Kumar 2004).

3. Numerical Results

We conduct pseudo-spectral direct numerical simulations (DNS) of MHD turbulence using hyperdiffusion with grid resolutions up to 20483. Simulation parameters are summarized in Table 1 (see details in Table 2 in Appendix B). To discern trends in the high-Re asymptotic limit, each set of simulations is run under the same parameters but at different grid resolutions (Reynolds numbers). Our flows are driven with either nonhelical forcing (Runs III and IV) or helical forcing (Runs I, II, and IV). Because we do not account for the α-effect when modeling the turbulent EMF, $\bar{{\boldsymbol{\varepsilon }}}$, which may be important in helically driven flows, we focus on results from Runs III and IV in the main text while those driven with helical forcing (Runs I, II, and IV) are shown in Appendix B for completeness. We note that all simulations yield remarkably similar results, regardless of the type of forcing.

Table 1. Simulation Parameters

RunForcing kf Prm $| {B}_{0}| /{B}_{k}^{\max }$ Nmax 3
IABC21010243
IIABC211010243
IIITG11010243
IVTG12010243
IV(Prm = 0.1)TG10.105123
IV(Prm = 5)TG1505123
IV(Prm = 10)TG11005123
VABC21220483

Note. ABC (helical) and Taylor–Green (TG; nonhelical) forcing are applied at wavenumber kf . ${B}_{k}^{\max }=\sqrt{{\max }_{k}[{E}^{b}(k)]}$ is at the magnetic spectrum's [Eb (k)] peak. Each simulation set includes runs with the same parameters except grid resolution (Reynolds numbers). N3 max denotes the highest resolution in each set. Subscripts a, b, c, and d denote resolutions of 2563, 5123, 10243, and 20483, respectively. More details are given in Table 2 in Appendix B.

Download table as:  ASCIITypeset image

In our simulations, we observe a scaling of Eu (k) ∼ k−4/3 in Runs Ic , IIIc , and IVc , and Eu (k) ∼ k−3/2 in Runs IIc and Vd (see Figures 10 and 11 in Appendix C), corresponding to scaling exponents of σu = 1/6 and σu = 1/4, respectively. Eb (k) ∼ k−5/3 in all runs (at the highest resolution), corresponding to scaling exponents of σb = 1/3. The Eb (k) scaling is consistent with that reported in solar wind studies (Podesta et al. 2007; Borovsky 2012). The Eu (k) ∼ k−4/3 scaling is consistent with that reported by Grete et al. (2020b) using the code K-Athena (Grete et al. 2020; Stone et al. 2020) and is slightly shallower than the k−3/2 reported in other studies (Haugen et al. 2004; Borovsky 2012), possibly due to the pronounced bottleneck effect from using hyperdiffusion (Frisch et al. 2008).

Without placing too much emphasis on the specific values of σu and σb for now, we wish to check if the scaling we derived in Equation (13) is consistent with the σu and σb we observe in our simulations. Figures 1(a) and (b) (also Figure 7 in Appendix B) show the effective transport coefficients νt , ηt , and Prt as a function of scale calculated using their respective definitions in Equations (6)–(8). We can see that νt (k) ∼ k−5/3 (or ∼ k−3/2) and ηt (k) ∼ k−4/3, consistent with Equation (13) when σu = 1/6 (or σu = 1/4) and σb = 1/3 as in our simulations. Moreover, we see in Figures 1(a) and (b) (and Figure 7 in Appendix B) that Prt (k) ∼ k−1/3 (or ∼ k−1/6), which is also consistent with the derived scaling in Equation (13) with σu = 1/6 (or σu = 1/4) and σb = 1/3 in our simulated flows. Panels (c)–(d) in Figure 1 also show νt , ηt , and Prt but calculated from Equations (17)–(18). Turbulent resistivity is very similar to that in Figures 1(a) and (b) with an ηt k−4/3 scaling, whereas νt has a scaling that is slightly shallower than that in Figures 1(a) and (b). Because Prt = νt /ηt , it is sensitive to slight changes in the scaling with Prt k−1/3 only over the decoupled range k ∈ [50, 200] but not for smaller k.

Qualitatively, the scalings of transport coefficients in Figures 1(c) and (d) are consistent with those in Figures 1(a) and (b), generally increasing at larger scales. We believe that this agreement between the different definitions of transport coefficients will be enhanced as the dynamic range increases and more definitive power-law scalings emerge. Indeed, we will present evidence below that the dynamic range in simulations that are possible today, including ours here, does not yet have a converged power-law scaling.

The scaling of ηt k−4/3 agrees with our Equation (14) applicable to the solar wind, as does νt k−3/2 from Runs IIc and Vd . The scaling of νt k−5/3 in Runs Ic , IIIc , and IVc decays faster than the k−3/2 in Equation (14) because σu < 1/4 in those simulations, associated with a shallower spectrum. This may be attributed to the bottleneck effect from hyperviscosity (Frisch et al. 2008), which produces a pileup at the small scales (see the spectra in Figure 10 of Appendix).

Figures 1(a) and (b) also show a Prt larger than unity in the inertial-inductive range, decreasing to Prt ≈ 1 to 2 at the smallest inertial-inductive scales d in all cases (see also Table 3 in Appendix B), where d is defined as the scale where $ \langle {\bar{{\rm{\Pi }}}}_{\ell }^{u}+{\bar{{\rm{\Pi }}}}_{\ell }^{b} \rangle =\nu \langle | {\rm{\nabla }}\bar{{\boldsymbol{u}}}{| }^{2} \rangle +\eta \langle | {\rm{\nabla }}\bar{{\bf{B}}}{| }^{2} \rangle $. For non-unity Prm , ${{\ell }}_{d}\equiv \max ({{\ell }}_{\nu },{{\ell }}_{\eta })$. ν and η are defined as scales where $ \langle {\bar{{\rm{\Pi }}}}_{\ell }^{u} \rangle =\nu \langle | {\rm{\nabla }}\bar{{\bf{u}}}{| }^{2} \rangle $ and $ \langle {\bar{{\rm{\Pi }}}}_{\ell }^{b} \rangle =\eta \langle | {\rm{\nabla }}\bar{{\boldsymbol{B}}}{| }^{2} \rangle $.

Table 3. Prt at k/kd = 1, where kd = L/d

 Run IRun IIRun IIIRun IVRun V
2563 1.231.311.401.441.30
5123 1.501.291.451.681.45
10243 1.411.251.511.671.10
20483     1.36

Download table as:  ASCIITypeset image

Figure 2 (and Figure 8 in Appendix C) shows ratios $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $, the product of which yields Prt in Equation (8). $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ becomes constant in the decoupled range due to the conservative (constant) KE and ME cascades in this range. $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ is equal to ${\int }_{0}^{k}k{{\prime} }^{2}{E}^{b}(k^{\prime} ){dk}^{\prime} /{\int }_{0}^{k}k{{\prime} }^{2}{E}^{u}(k^{\prime} ){dk}^{\prime} $. The ratio $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ increases because Eb (k) < Eu (k) at forcing scales (forcing in velocity field) but Eb (k) catches up and exceeds Eu (k) at larger k. The ratio $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ decays after reaching a peak because (1) each of ${\overline{{\boldsymbol{J}}}}_{{\ell }}$ and ${\overline{{\boldsymbol{S}}}}_{{\ell }}$ is dominated by the largest wavenumbers below the cutoff k < L/, and (2) Eu (k) is shallower than Eb (k) at high k in the inertial-inductive range, making $| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}$ grow faster than $| {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}$ as → 0.

Figure 2.

Figure 2. Plots showing Prt and its two components $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ at the highest resolutions of Runs III and IV (Taylor–Green forcing) in Table 1. A reference line with a slope of −1/3 (black dashed) is added. The plots show that $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ approaches a constant in the decoupled range. Simulations with helical forcing are shown in Figure 8 in Appendix C.

Standard image High-resolution image

4. Discussion

We now provide the physical explanation for why Prt seems to increase at larger scales and discuss whether or not this trend is expected to persist for an arbitrarily wide dynamical range (${Re}\to \infty $). As we have mentioned above, σu and σb are a measure of the velocity and magnetic fields' smoothness, respectively (Aluie 2017). If σu < σb (corresponding to a shallower scaling of Eu (k) relative to Eb (k)) as in our simulations and many other independent reports from solar wind observations and simulations (e.g., Podesta et al. 2007; Mininni & Pouquet 2009; Borovsky 2012; Grappin et al. 2016), then the velocity field is rougher than the magnetic field. This implies that small-scale velocity "eddies" have a higher proportion of the overall kinetic energy compared to the proportion small-scale magnetic "eddies" contribute to the overall magnetic energy (i.e., $\tfrac{\text{small-scale}\ \mathrm{KE}}{\mathrm{total}\ \mathrm{KE}}\,\gt \tfrac{\text{small-scale}\,\mathrm{ME}}{\mathrm{total}\ \mathrm{ME}}$). Note that the latter statement is not based on comparing Eu (k) to Eb (k) at high k in absolute terms, where we see Eb (k) ≳ Eu (k). Rather, it is based on the strength of "eddies" relative to the overall velocity or magnetic field, respectively.

The coarse-grained strain and current, ${\overline{{\boldsymbol{S}}}}_{{\ell }}$ and ${\overline{{\boldsymbol{J}}}}_{{\ell }}$, are cumulative quantities, i.e., they include the contribution from all scales larger than , for any . It follows from the above paragraph that as the coarse-graining is made smaller, the relative contribution from scales near to $| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}$ is more significant than that to $| {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}$. From the definition of Prt in Equation (8) and with $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ being scale independent in the decoupled range, we have ${\Pr }_{t}\propto \langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ in the decoupled range. Clear evidence of this is shown in Figure 2 (and Figures 3 and 8). As decreases (or k increases), both $| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}$ and $| {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}$ increase because contributions from < are included. However, due to the larger roughness of the velocity field, the increase in $| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}$ is more pronounced than that in $| {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}$, leading to a decrease in the ratio $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $. This explains why Prt seems to decrease with larger k over the decoupled range (range over which each of $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ is scale independent).

Figure 3.

Figure 3. Plots showing Prt and its two components $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ using unity (Prm = 1) and non-unity microscopic Prandtl numbers (Prm = 0.1, 5, 10) on a 5123 grid. See parameters in Table 1. A reference line with a slope of −1/3 (black dashed line) is added.

Standard image High-resolution image

In the conversion range over which $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ are still varying with , the scaling of Prt depends on both $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $. On the one hand, $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle \gt 1$ because energy is input into the velocity field at the largest scales and more kinetic energy is cascading compared to magnetic energy, such that $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle \to \infty $ as f approaches the forcing scale f . On the other hand, we have the ratio $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ decreasing in that limit of f because the strain becomes relatively stronger at the forced scales. From Figure 2, we find that in our simulated flows, the Prt scaling over the conversion range either decaying weakly or flat as k increases. Because the conversion range is limited in extent and does not increase with an increasing dynamic range (Bian & Aluie 2019), it is not very meaningful to discuss a scaling of Prt over this range.

Crude estimates of the competition between large-scale magnetic flux advection and large-scale magnetic flux diffusion in accretion disks require Prt (R/H) > 1 (where H is the disk scale height and R is the disk radius) for the former to be competitive with the latter in the disk interior (Lubow et al. 1994; Blackman & Nauman 2015). 5 That we find values of Prt > 1 means that large-scale MHD flow may be more efficient at advecting large-scale magnetic flux while shedding angular momentum outward (via νt ) than would be the case for Prt ≤ 1. That said, pinning down the exact implications are difficult given the additional dependence of disk physics on stratification with the possibility of flux advection in surface layers (e.g., Lovelace et al. 2009; Zhu & Stone 2018).

4.1. Prt Scaling under Different Flow Conditions

We have tested the scaling of Prt under different microscopic Prm flow conditions. We remind the reader that our results here pertain to the decoupled range, which is within the inertial-inductive range. These scales are immune from the direct influence of both resistivity and viscosity. We do not expect our results here (and those of Bian & Aluie 2019 upon which this analysis is based) to hold in the viscous-inductive (Batchelor) range at high Prm or in the inertial-resistive range at low Prm . From practical modeling considerations, such as when simulating a galactic accretion disk at global scales, grid-resolution constraints render it virtually impossible to have Δx within the viscous-inductive range. Therefore, the restriction of our analysis to inertial-inductive scales is still pertinent to modeling as well as being of theoretical import.

Figures 1(a) and (b) show the case (Run IV) of Prm = 2, where we find a scaling of Prt similar to the case of unity microscopic Prm . We also conduct simulations (Table 1) at Prm = 0.1, 5, and 10, albeit at a lower resolution of 5123 due to the computational overhead required by non-unity Prm . Figure 3 shows that the scale dependence of Prt is consistent with that of unity-Prm runs, although the scaling is not as clear. Due to the lower resolution, the decoupled range is barely established in the non-unity Prm cases. Non-unity Prm simulations require even larger Reynolds numbers to achieve a significant decoupled range and still make an allowance for a viscous-inductive or an inertial-resistive range of scales. This is beyond our computing capability for this work.

Our results also suggest that within the limited dynamic range of our simulations, increasing the external B-field strength from 0 (Run I) to 2 (Run V) to 10 (Run II) seems to change the Prt scaling slightly from k−1/3 to k−1/6 due to σu increasing from 1/6 to 1/4 (see Figure 7 in Appendix B). However, we do not believe this trend will persist at asymptotically high-Re as we discuss in the following subsection. We also note that our analysis here does not distinguish the anisotropy in turbulent transport. Our effective transport coefficients in this paper are isotropic even though the underlying turbulent flow may be anisotropic such as in Runs II and V (see Figure 5 in Appendix B). We hope this work is extended to anisotropic turbulent transport in future studies.

4.2. Prt Scaling at Asymptotically High Re

Can we expect the scaling of Prt in Figures 1(a) and (b), which is in support of our relations in Equation (13), to extrapolate to the wide dynamical ranges (high Re) that exist in many astrophysical systems of interest?

Figure 4 (and Figure 9 in Appendix C) examines the scaling of Prt (k) at different Reynolds numbers. Each panel shows results from a suite of simulations under the same parameters except for Re (or grid resolution). The plots show that Prt (k) takes on a value between 1 and 2 at the smallest scales within the inertial-inductive range, regardless of Re (also Figure 16 and Table 3 in Appendix C). These scales near d are bordering the dissipation range. The reason Prt (kL/d ) ≈ 1–2 can be understood from the definition in Equation (8) of ${\Pr }_{t}=\left(\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle \right)\left(\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle \right)$. Due to the equipartition of the cascades in the decoupled range, we have $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle \approx 1$,whereas $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle \approx 1\,\mathrm{to}\,2$, as is clear from Figure 2 (and Figures 3 and 8). The latter can also be inferred from comparing the spectra ${E}^{u}(k^{\prime} )$ and ${E}^{b}(k^{\prime} )$ in Figures 1011 via Equations (9)–(10).

Figure 4.

Figure 4. Plots showing Prt at different Reynolds numbers (grid resolution) of Runs III and IV (Taylor–Green forcing) in Table 1. A reference line with a slope of −1/3 (black dashed) is added. Simulations with helical forcing are shown in Figure 9 in Appendix C.

Standard image High-resolution image

Based on these observations, it is physically reasonable to assume that for a sufficiently wide dynamical range (or large Re), $\langle | {\overline{{\boldsymbol{J}}}}_{{{\ell }}_{d}}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{{\ell }}_{d}}{| }^{2}\rangle $ converges to a constant when d (or kkd ), independent of the dynamical range extent (i.e., independent of Re). That is, the ratio $\langle | {\overline{{\boldsymbol{J}}}}_{{{\ell }}_{d}}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{{\ell }}_{d}}{| }^{2}\rangle $ has the same value under successive refinements of the grid. This effectively provides us with a conceptual boundary condition on Prt (k) at those smallest scales kkd .

These logical considerations, when combined with the scaling ${\Pr }_{t}(k)\sim {k}^{-2({\sigma }_{b}-{\sigma }_{u})}$ in Equation (13) (with empirical support in Figures 1(a), 1(b), and 7), may lead us at face value to the uncomfortable conclusion that at any fixed k in the inertial-inductive range, Prt (k) will keep increasing with increasing Re (or higher resolution) as Figure 4 (and Figure 9 in Appendix C) seems to indicate. That is unless the MHD dynamics eventually yields σu = σb in this asymptotic limit, i.e., at successively higher-resolution simulations. Indeed, there are indications from Figures 4 and 9 that the Prt (k) sensitivity to Re decreases with increasing Re as we now discuss.

At first glance, Figure 4 seems to indicate that Prt (k) at any fixed wavenumber, e.g., k = 50 within the decoupled range, Prt (k = 50), increases with increasing resolution. Yet, as we shall now argue, Figure 4 highlights how certain metrics such as Prt in our simulations, which are very high resolution by today's standards, are still not fully converged to the high-Re limit. From the definition of Prt in Equation (8), this increase can only be due to an increase of the cascade ratios, $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $, or the current-to-strain ratio, $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $, or both. We find in Figure 15 that the latter accounts for most of this increase. Figure 15 suggests that the cascade ratio $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ is fairly converged with resolution in our largest simulations at = L/50. Physically, we expect $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ to also converge because the ratio depends on the strain and current (or equivalently, the spectra) at scales larger than L/50. These should not remain sensitive to the smallest scales in a simulation once a sufficiently high resolution has been achieved. Figure 15 in Appendix C indicates that the high resolution of our simulations is still not sufficient for the convergence of these quantities ($\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle $). Ignoring convergence trends under the guise of "having conducted the highest-resolution simulation to date" can be rife with pitfalls. In general, when analyzing simulations of turbulent flows, it is vitally important to study trends as a function of Reynolds number and check if the phenomenon under study persists and can be extrapolated to the large Reynolds numbers present in nature.

What conclusion on Prt (k) scaling does these convergence considerations lead us to? If we accept that with increasing resolution, Prt (k*) has to converge to a specific value for any fixed k* within the inertial-inductive range, and if we also accept that at the smallest scales within the inertial-inductive range ≈d, Prt (kd ) also converges to a constant value, then as the gap between k* and kd widens with a wider dynamical range (kd ), we must have that ${\Pr }_{t}(k)\sim {k}^{-2({\sigma }_{b}-{\sigma }_{u})}\sim {k}^{0}$ approach a k-independent scaling with σb = σu in the asymptotic limit $\mathrm{Re}\to \infty $. Our Figures 4 and 9 lend some support to our assertion as they show that the Prt (k) is converging (but not converged) at the largest scales with increasing resolution.

Such a conclusion would have wide-ranging implications, foremost of all regarding the power-law scaling of spectra in MHD turbulence. However, it is important that our results are further verified by the community under different parameter conditions, e.g., B0 strength and Prm , and perhaps also from higher-resolution simulations.

5. Conclusion

In this paper, we are proposing a somewhat new method to measure turbulent transport coefficients (turbulent viscosity νt , resistivity ηt , and magnetic Prandtl number Prt ) at different scales using the coarse-graining approach. To our knowledge, this is the first determination of Prt as a function of length scale. From analyzing the kinetic and magnetic energy cascade rates, we infer power-law scaling in Equation (13) for νt , ηt , and Prt given our definitions of those transport coefficients. This approach circumvents relying on particular values for the spectral scaling exponents (σu and σb ) from a specific MHD phenomenology—whether it exists or not—by relying on results from Bian & Aluie (2019) of conservative KE and ME cascades. Our analysis here relied on high-resolution DNS under different forcing, external B-field strength, and microphysical Prm .

Our DNS results indicate that Prt ≈ 1 to 2 at the smallest inertial-inductive scales, increasing to Prt ≈ 5 to 10 at the largest scales. For accretion disks, conservative minimalist estimates for the advection of large-scale vertical magnetic fields to win over turbulent diffusion require Prt (H/R) > 1, so that larger values of Prt improve the efficacy of flux advection over diffusion (e.g., Lubow et al. 1994). This condition and the direct applicability of our specific results are both textured by detailed disk physics (e.g., Zhu & Stone 2018), including stratification, not studied here.

Nevertheless, based on physical considerations, our analysis suggests that Prt has to become scale independent and of order unity in the decoupled range at sufficiently high Reynolds numbers (or grid resolution) and that the power-law scaling exponents of velocity and magnetic spectra become equal.

If indeed the power-law scaling exponents of velocity and magnetic spectra (σu and σb ) become equal in the $\mathrm{Re}\to \infty $ limit, it would have wide-ranging implications, foremost of all regarding the power-law scaling of spectra in MHD turbulence (Politano & Pouquet 1998a, 1998b; Aluie 2017). However, as discussed above, our Prt scaling is not quite converged, despite showing a converging trend. It is important for our results to be further checked by the community using simulations of higher resolution and for a wider range of parameters, e.g., B0 strengths and Prm values.

Our results also suggest that the presence of a mean B field does not affect Prt significantly. However, we only consider Prt as a scalar in this study. Lesur & Longaretti (2009) considered an anisotropic turbulent resistivity tensor with an external B field. Under non-unity microphysical Prm , our results are consistent with those of Prm = 1, although we could not establish a clear decoupled range due to insufficient simulation resolution.

In addition to potential implications for astrophysical systems, our analysis of how νt , ηt , and Prt vary with length scale provides a practical model for these quantities that does not rely on any particular MHD turbulence phenomenology.

The simulations we conducted here are fairly idealized (incompressible flows in a periodic domain with artificial forcing). We hope that this work offers a path to analyzing more complicated flows because our method can be applied to more realistic simulations such as of global accretion disk flows. For the pursuit of isotropic diffusion coefficients, measuring νt and ηt at any length scale from Equations (6) and (7) does not require the existence of an inertial range or even turbulence, even though in the present paper we applied the method to a case of fully developed turbulence. For some applications, we believe that our approach complements existing approaches such as test-field methods (Schrinner et al. 2007; Käpylä et al. 2020) of measuring turbulent transport. These methods involve taking the velocities computed from a numerical simulation and then separately solving for the transport coefficients using an imposed test magnetic field. Traditionally, these have been restricted to the kinematic regime of a weak magnetic field (although see Käpylä et al. 2021).

Finally, our work should not be construed as an endorsement of the "eddy-viscosity/resistivity" model wherein turbulent processes ${\bar{{\boldsymbol{\tau }}}}_{{\boldsymbol{\ell }}}$ and ${\bar{{\boldsymbol{\varepsilon }}}}_{{\boldsymbol{\ell }}}$ representing scales < are modeled as purely diffusive. Our approach can be extended to models in which transport is not entirely diffusive, such as those that include the helical α-effect.

This research was funded by DOE FES grants DE-SC0014318 and DE-SC0020229. Partial funding for this research was provided by the Center for Matter at Atomic Pressures (CMAP), a National Science Foundation (NSF) Physics Frontier Center, under award PHY-2020249. H.A. was also supported by NASA grant 80NSSC18K0772, DOE grant DE-SC0019329, and NNSA grants DE-NA0003914 and DE-NA0003856. J.S. was also supported by DOE grant DE-SC0019329 and NNSA grant DE-NA0003914. E.B. was also supported by DOE grants DE-SC0001063, DE-SC0020432, and DE-SC0020103, and NSF grant AST-1813298. Computing time was provided by the National Energy Research Scientific Computing Center (NERSC) under contract No. DE-AC02-05CH11231 and by an award from the INCITE program, using resources of the Argonne Leadership Computing Facility, which is a DOE Office of Science User Facility supported under contract DE-AC02-06CH11357.

Appendix A: Numerical Setup

Our numerical simulations of mechanically forced turbulence are conducted in a periodic box ${{\mathbb{T}}}^{3}=[0,L{)}^{3}$, with L = 2π. We use a pseudo-spectral code with phase-shift dealiasing. The time integration method is a second-order Adam–Bashforth scheme. We solve the incompressible MHD equations with hyperviscosity (Borue & Orszag 1995) and hyperresistivity with a Laplacian of exponent α = 5:

Equation (A1)

Equation (A2)

Equation (A3)

where νh is the hyperviscosity and ηh is the hyperresistivity coefficients. Hyperdiffusivity is commonly used in MHD turbulence studies (Cho & Vishniac 2000; Kawai 2013; Beresnyak 2015; Meyrand et al. 2016; Kawazura et al. 2019) to reduce the dissipation range extent, thereby allowing for a longer inertial-inductive range of scales. The velocity and magnetic field are initialized in k space with ${E}^{u,b}\sim | {\boldsymbol{k}}{| }^{2}{{\boldsymbol{e}}}^{-| {\boldsymbol{k}}{| }^{2}/11}$ spectra and random phases.

Runs I, II, and V (see Table 2 for simulation details) are driven by ABC forcing (named after Arnold, Beltrami, and Childress):

Equation (A4)

where A = B = C = 0.25, kf is the forcing wavenumber, and ex , ey , and ez are unit vectors in x, y, and z, respectively. ABC forcing is helical, which injects kinetic helicity into the flow. Kinetic helicity is an example of a pseudoscalar that facilitates large-scale dynamos (e.g., Parker 1955; Moffatt 1978; Mininni & Montgomery 2005; Blackman 2016).

Taylor–Green (TG) forcing, which is nonhelical, is used to drive the flow in Runs III and IV:

Equation (A5)

where the force amplitude f0 = 0.25. TG forcing injects no global integrated kinetic helicity into the flow.

The simulations are conducted at different Reynolds numbers with different grid resolutions. Detailed parameters are shown in Table 2, where the subscripts a, b, c, and d (e.g., Run Va versus Vb versus Vc versus Vd ) denote simulations using the same parameters but at different grid resolutions and Reynolds numbers. Runs I–IV are conducted with grid resolution of 2563, 5123, and 10243. Run V is also conducted at 20483 resolution. For Run III, Prm = 0.1, Prm = 5, and Prm = 10 at grid resolution of 2563 and 5123 are added to study the effects of a non-unity microscopic Prandtl number.

Figure 5 visualizes the magnitude of the velocity and magnetic fields (∣ u ∣ and ∣ B ∣) in two simulations. The anisotropic structures are significant in the presence of an external magnetic field (Figures 5(c) and (d)).

Figure 5.

Figure 5. Slices of the magnitude of velocity field ∣ u ∣ and magnetic field ∣ B ∣. Panels (a) and (b) show results from Run Ic without an external B field ∣B0∣ = 0. Panels (c) and (d) show results from Run IIc with ∣B0∣ = 10. The plots show significant anisotropic structures in Run IIc .

Standard image High-resolution image

Figure 6 shows that Cu and Cb used in Equations (21) and (23) are indeed proportionality constants that are scale independent within the decoupled range.

Appendix B: Helical Forcing Results

The section shows numerical results from simulations with helical forcing. The α-effect of dynamo theory is believed to be important in helical turbulence. We show here the helically forced results for completeness, although ignoring the α term in Equation (7) may not be justified. Nevertheless, our results are remarkably similar to those in the main text.

Figure 7 shows νt , ηt , and Prt scaling at the highest resolution in helical forcing simulations, as a supplement to Figures 1(a) and (b). The results are νt (k) ∼ k−5/3 (or ∼ k−3/2) and ηt (k) ∼ k−4/3, and Prt (k) ∼ k−1/3 (or ∼ k−1/6), similar to the nonhelical simulation results. As we mention in the main section, σu is ≈ 1/4 rather than 1/6 in the presence of a strong external B field (Run II), leading to the change in the scaling of νt and Prt .

Figure 8 shows $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ at the highest resolution in helical forcing simulations, as a supplement to Figure 2. The results suggest constant $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ in the decoupled range and the same scaling of $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ and Prt in the decoupled range, similar to the nonhelical simulation results. The scaling of $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ is explained in the main section.

Figure 6.

Figure 6. Plots showing Cu and Cb of Runs I–V at the highest resolution. Cb is calculated with σb = 1/3. Cu is calculated with σu = 1/4 in Run IIc and Run Vd and σu = 1/6 in other cases.

Standard image High-resolution image

Table 2. Simulations Parameters: Prm is the Magnetic Prandtl Number

RunGridForcing kf Prm $| {B}_{0}| /{B}_{k}^{\max }$ νh ηh
Ia 2563 ABC2105 × 10−16 5 × 10−16
Ib 5123 ABC2102 × 10−21 2 × 10−21
Ic 10243 ABC2104 × 10−25 4 × 10−25
IIa 2563 ABC21105 × 10−16 5 × 10−16
IIb 5123 ABC21102 × 10−21 2 × 10−21
IIc 10243 ABC21104 × 10−25 4 × 10−25
IIIa 2563 TG1105 × 10−16 5 × 10−16
IIIb 5123 TG1102 × 10−21 2 × 10−21
IIIc 10243 TG1104 × 10−25 4 × 10−25
IVa 2563 TG1202 × 10−16 1 × 10−16
IVb 5123 TG1204 × 10−21 2 × 10−21
IVc 10243 TG1204 × 10−25 2 × 10−25
IVa (Prm = 0.1)2563 TG10.102 × 10−17 2 × 10−16
IVb (Prm = 0.1)5123 TG10.102 × 10−21 2 × 10−20
IVa (Prm = 5)2563 TG1501 × 10−16 2 × 10−17
IVb (Prm = 5)5123 TG1501 × 10−20 2 × 10−21
IVa (Prm = 10)2563 TG11002 × 10−16 2 × 10−17
IVb (Prm = 10)5123 TG11002 × 10−20 2 × 10−21
Va 2563 ABC2125 × 10−16 5 × 10−16
Vb 5123 ABC2122 × 10−21 2 × 10−21
Vc 10243 ABC2124 × 10−25 4 × 10−25
Vd 20483 ABC2121 × 10−27 1 × 10−27

Note. ${B}_{k}^{\max }=\sqrt{{\max }_{k}[{E}^{b}(k)]}$ is at the magnetic spectrum's [Eb (k)] peak. ABC (helical) and TG (nonhelical) forcing are applied at wavenumber kf .

Download table as:  ASCIITypeset image

Figure 7.

Figure 7. Plots showing turbulent viscosity νt , turbulent resistivity ηt , and turbulent magnetic Prandtl number Prt calculated using their respective definitions in Equations (6)–(8), at different scales k = L/. We use the highest-resolution runs of Runs I, II, and V (ABC forcing) in Table 1. Three reference lines with a slope of −1/3, −5/3 (black dashed–dotted), and −4/3 (black solid) are added. Note the reference line of −1/3 and Prt uses the right y-axis, while others use the left y-axis. Scales < d are not shown.

Standard image High-resolution image

Figure 9 shows Prt at different Reynolds number in helical forcing simulations, as a supplement to Figure 4. The results are similar to nonhelical simulation results.

Appendix C: Results at Different Reynolds Numbers

Figure 10 shows the kinetic energy spectrum at different Reynolds numbers (grid resolution). The slope becomes steeper as the Reynolds number increases. The slope is near −3/2 at the highest resolution in Run II and Run V (with external B field), while shallower than −3/2 in other simulations. Grete et al. (2021) observed a kinetic energy spectrum of −4/3, which is also shallower than −3/2.

Figure 8.

Figure 8. Plots showing the turbulent magnetic Prandtl number Prt and its two components $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ at the highest resolutions of Run I, II, and V (ABC forcing) in Table 1. A reference line with a slope of −1/3 (black dashed) is added. The plots show that $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ approaches a constant in the decoupled range. Note that with a strong external B field (Run II), we expect $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ to plateau at sufficiently high Reynolds numbers (Bian & Aluie 2019).

Standard image High-resolution image
Figure 9.

Figure 9. Plots showing the turbulent magnetic Prandtl number Prt at different Reynolds numbers (grid resolution) of Runs I, II, and V (ABC forcing) in Table 1. A reference line with a slope of −1/3 (black dashed) is added.

Standard image High-resolution image
Figure 10.

Figure 10. Plots showing the kinetic energy spectrum Eu (k) at different Reynolds numbers (grid resolution) of Runs I–V. We show two reference lines with a slope of −3/2 (black solid) and −5/3 (black dashed–dotted). The slope becomes steeper as the Reynolds number increases. The slope is near −3/2 at the highest resolution in Run II and Run V (with external B field) while shallower than −3/2 in other simulations. Note Grete et al. (2021) observed a kinetic energy spectrum of −4/3, which is also shallower than −3/2.

Standard image High-resolution image

Figure 11 shows the magnetic energy spectrum at different Reynolds numbers. The slope agrees well with −5/3 for all Reynolds numbers. Figure 12 shows the kinetic and magnetic energy spectra of Run IV with Prm = 0.1, 5, and 10.

Figure 11.

Figure 11. Plots showing the magnetic energy spectrum Eb (k) at different Reynolds numbers (grid resolution) of Runs I–V. We show two reference lines with a slope of −3/2 (black solid) and −5/3 (black dashed–dotted). The slope of the magnetic spectrum agrees well with solar wind observations (−5/3).

Standard image High-resolution image
Figure 12.

Figure 12. Plots showing kinetic and magnetic energy spectra, Eu (k) and Eb (k), of Run IV with Prm = 0.1, 5, and 10. We show two reference lines with a slope of −3/2 (black solid) and −5/3 (black dashed–dotted).

Standard image High-resolution image

Figure 13 shows that the scaling exponent of νt is near −5/3 (−3/2 in Run IIc and Vd ) at the highest resolution. As the Reynolds number increases, it becomes steeper and approaches −3/2. Figure 14 shows the scaling exponent of ηt at all Reynolds numbers is near −4/3, consistent with Equation (14).

Figure 13.

Figure 13. Plots showing the turbulent viscosity νt at different Reynolds numbers (grid resolution) of Runs I–V. Reference lines with a slope of −5/3 (black dash-doted) and −3/2 (black solid) are added.

Standard image High-resolution image
Figure 14.

Figure 14. Plots showing the turbulent resistivity ηt at different Reynolds numbers (grid resolution) of Runs I–V. A reference line with a slope of −4/3 (black solid) is added. The scaling exponent agrees well with the expected value −4/3.

Standard image High-resolution image

Figure 15 shows $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ at different Reynolds numbers (grid resolution).

Figure 15.

Figure 15. Plots showing $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ (blue lines) and $\langle | {\overline{{\boldsymbol{J}}}}_{{\ell }}{| }^{2}\rangle /\langle 2| {\overline{{\boldsymbol{S}}}}_{{\ell }}{| }^{2}\rangle $ (red lines) at different Reynolds numbers (grid resolution) of Runs I, III, and V. The x-axis in the bottom panels is normalized by kd = L/d . A reference line (black dashed) of 1 is added.

Standard image High-resolution image

Figure 16 shows the Prt at different Reynolds numbers with the x-axis normalized by kd = L/d , where d is defined as the scale at which $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}+{\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle =\nu \langle | {\rm{\nabla }}\overline{u}{| }^{2}\rangle +\eta \langle | {\rm{\nabla }}\overline{B}{| }^{2}\rangle $. For non-unity Prm , ${{\ell }}_{d}\equiv \max ({{\ell }}_{\nu },{{\ell }}_{\eta })$. ν and η are defined as scales where $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle =\nu \langle | {\rm{\nabla }}\overline{u}{| }^{2}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle =\eta \langle | {\rm{\nabla }}\overline{B}{| }^{2}\rangle $. Prt at different Reynolds numbers collapse at k = kd , as expected (see also Table 3).

Figure 16.

Figure 16. Plots showing the Prt at different Reynolds numbers (grid resolution) of Runs I–V with x-axis normalized by kd = 2π/d . A reference line with a slope of −1/3 (black dashed) is added. Prt at different Reynolds numbers collapse at k = kd , as expected (see also Table 3).

Standard image High-resolution image

Figure 17 shows $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ at different microscopic Prandtl numbers (Prm = 0.1, 1, 5, 10). Because the decoupled range, over which each of $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ becomes scale independent, is barely resolved, these plots neither reinforce nor conflict with the expectation of asymptotic equipartition of the kinetic and magnetic cascades predicted in Bian & Aluie (2019), irrespective of microscopic Prm . It is worth emphasizing that the observation of Brandenburg (2014) of a positive correlation between Prm and the ratio of kinetic dissipation to magnetic dissipation does not have a direct bearing on the ratio of the cascades. This is because the cascades $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ in the decoupled range are not necessarily equal to the kinetic and magnetic energy dissipation, respectively. This is especially true at non-unity Prm at scales smaller than d beyond the decoupled range, where kinetic-magnetic conversion is expected to occur (e.g., in the viscous-inductive range at high Prm ) before all energy is dissipated microscopically.

Figure 17.

Figure 17. Plots showing $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle /\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ at different microscopic Prandtl numbers (Prm = 0.1, 1, 5, 10). The parameters are detailed in Table 2. These simulations are conducted on a 5123 grid. A reference line (black solid) of 1 is added in all panels. Note our usage of fifth-order hyperdiffusion in the simulations. Another reference line (black dashed) of ${\nu }_{h}^{1/5}/{\eta }_{h}^{1/5}$ is added as an estimate for the microscopic magnetic Prandtl number corresponding to normal (Laplacian) diffusion. The estimate is 0.63, 1, 1.38, and 1.58 for Prm = 0.1, Prm = 1, Prm = 5, and Prm = 10, respectively. Because the decoupled range, over which $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{u}\rangle $ and $\langle {\overline{{\rm{\Pi }}}}_{{\ell }}^{b}\rangle $ become scale independent, is barely resolved, these plots neither reinforce nor conflict with the expectation of asymptotic equipartition of the kinetic and magnetic cascades predicted in Bian & Aluie (2019), irrespective of microscopic Prm .

Standard image High-resolution image

Footnotes

  • 4  

    Strictly speaking, the eddy-viscosity definition is ${\mathring{\overline{\tau }}}_{{ij}}=-2{\nu }_{t}{\overline{S}}_{{ij}}$, where ${\mathring{\overline{\tau }}}_{{ij}}={\overline{\tau }}_{{ij}}-{\delta }_{{ij}}{\overline{\tau }}_{{kk}}/3$ is the traceless part of the stress. For our incompressible flow analysis here, which is based on the energy flux, this distinction does not matter.

  • 5  

    The turbulent Prandtl number used in Lubow et al. (1994) is the inverse of Prt

Please wait… references are loading.
10.3847/2041-8213/ac0fe5