Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

MNRAS 000, 000–000 (0000) Preprint 17 August 2023 Compiled using MNRAS LATEX style file v3.

Shedding light on the MRI driven dynamo in a stratified shearing box


Prasun Dhang1,2 , ⋆ Abhijit Bendre3 , Kandaswamy Subramanian2,4
1 JILA, University of Colorado and National Institute of Standards and Technology, 440 UCB, Boulder, CO 80309-0440, USA
2 IUCAA, Post Bag 4, Ganeshkhind, Pune 411007, India
3 Institute of Physics, Laboratory of Astrophysics, École Polytechnique Fédérale de Lausanne (EPFL), 1290 Sauverny, Switzerland
4 Department of Physics, Ashoka University, Rajiv Gandhi Education City, Rai, Sonipat 131029, Haryana, INDIA
arXiv:2308.07959v1 [astro-ph.HE] 15 Aug 2023

17 August 2023

ABSTRACT
We study the magneto-rotational instability (MRI) dynamo in a geometrically thin disc (H/R ≪ 1) using stratified zero net flux
(ZNF) shearing box simulations. We find that mean fields and EMFs oscillate with a primary frequency fdyn = 0.017 (≈ 9
orbital period), but also have higher harmonics at 3fdyn . Correspondingly, the current helicity, has two frequencies 2fdyn and
4fdyn respectively, which appear to be the beat frequencies of mean fields and EMFs as expected from the magnetic helicity
density evolution equation. Further, we adopt a novel inversion algorithm called the ‘Iterative Removal Of Sources’ (IROS), to
extract the turbulent dynamo coefficients in the mean-field closure using the mean magnetic fields and EMFs obtained from the
shearing box simulation. We show that an α−effect (αyy ) is predominantly responsible for the creation of the poloidal field from
the toroidal field, while shear generates back a toroidal field from the poloidal field; indicating that an α − Ω-type dynamo is
operative in MRI-driven accretion discs. We also find that both strong outflow (v̄z ) and turbulent pumping (γz ) transport mean
fields away from the mid-plane. Instead of turbulent diffusivity, they are the principal sink terms in the mean magnetic energy
evolution equation. We find encouraging evidence that a generative helicity flux is responsible for the effective α-effect. Finally,
we point out potential limitations of horizontal (x − y) averaging in defining the ‘mean’ on the extraction of dynamo coefficients
and their physical interpretations.
Key words: accretion,accretion discs - dynamo - instabilities - magnetic fields - MHD - turbulence - methods: numerical.

1 INTRODUCTION the density scale height and R is the local radius. Therefore, this
approach is valid only for geometrically thin discs with H/R ≪ 1.
The problem of angular momentum transport is a key concept in a ro-
Depending on whether the vertical component of gravity (gz =
tationally supported accretion disc (for a review, see Balbus & Hawley
−Ω2 z) (producing a vertically stratified gas density) is considered
1998). The current consensus is that a weak magnetic field instability,
in the momentum equation or not, shearing box simulations are of
namely magneto-rotational instability (MRI; Velikhov 1959; Chan-
two types: stratified (gz ̸= 0) and unstratified (gz = 0). Further,
drasekhar 1960; Balbus & Hawley 1991; Balbus & Hawley 1992)
depending on whether the computational domain can contain net
is responsible for outward angular momentum transport and drives
magnetic fields, shearing box models can be classified into zero
mass accretion in a sufficiently ionized accretion disc. Although lin-
net flux (ZNF) and net flux (NF) models. Therefore, four possible
ear MRI ensures outward angular momentum transport, it must be
combinations of the shearing-box model are: i) unstratified ZNF, ii)
studied in the non-linear phase to account for different observable
unstratified NF, iii) stratified ZNF and iv) stratified NF. This work
phenomena in accretion discs.
considers a stratified ZNF shearing box model to explore the MRI
MRI in an accretion disc is either studied in a local set-up (shearing
dynamo in saturation.
box; Balbus & Hawley 1992; Brandenburg et al. 1995; Hawley et al.
1995; Davis et al. 2010; Shi et al. 2010; Bodo et al. 2014; Bhat et al. Shearing box simulations provide a wide range of behaviour (e.g.,
2016) or in a global simulation (Stone et al. 1999; Hawley 2001; convergence, turbulence characteristics etc.) depending on the shear-
Hawley et al. 2013; Beckwith et al. 2011; Parkin & Bicknell 2013; ing box model used (for details, we refer to readers to see Table 1 in
Hogg & Reynolds 2016; Dhang & Sharma 2019; Dhang et al. 2023). Ryan et al. (2017)). However, it is to be noted that we will restrict
While a global approach is more desirable, it is computationally our discussion to the isothermal (i.e. sound speed is constant) models
expensive. On the other hand, the shearing box approach offers an where there is no explicit dissipation and the numerical algorithms
alternate path which is computationally less costly and can provide provide the dissipation through truncation error at the grid scale. In
deep insights into the local processes in MRI-driven turbulence. the presence of an NF, unstratified shearing box simulations show
In the shearing-box approach (Goldreich & Lynden-Bell 1965), a convergence (in terms of accretion stresses) and sustained turbu-
we expand fluid equations to the lowest order of H/R, where H is lence (Hawley et al. 1995; Guan et al. 2009; Simon et al. 2009). On
the other hand, stratified NF simulations present different accretion
stresses depending on the net flux strength and sustained turbulence
⋆ E-mail:prasundhang@gmail.com (Guan & Gammie 2011; Bai & Stone 2013). Unstratified ZNF mod-

© 0000 The Authors


2 Dhang et al.
els showed intriguing behaviour. Earlier isothermal Unstratified ZNF sion (Shi et al. 2016), singular value decomposition (SVD; Dhang
studies (Fromang & Papaloizou 2007; Pessah et al. 2007) found de- et al. 2020) to calculate dynamo coefficients in post-process or sta-
creased accretion stress and turbulence with increased resolution, tistical simulations to carry out combined study of the large-scale
implying non-convergence. However, later Shi et al. (2016) recov- dynamo and angular-momentum transport in accretion discs (Mon-
ered convergence using a box with a larger vertical extent than the dal & Bhat 2023). In this work, we use a direct method, a variant of
radial extent. On the contrary, earlier stratified ZNF models (Davis the cleaning algorithm ( Högbom CLEAN method; Högbom 1974),
et al. 2010) suggested that the models are converged till the resolution called ‘Iterative Removal Of Sources’ (IROS; Hammersley et al.
128/H; however, recent studies (Bodo et al. 2011; Ryan et al. 2017) 1992); mainly used in astronomical image construction to analyse
found the model loses convergent properties at higher resolution. MRI-dynamo in the mean-field dynamo paradigm. We modified the
The convergence problem is closely related to the magnetic energy IROS method according to our convenience (for details, see section
generation process in the MRI-driven flow. For the ZNF (absence of 4, also see Bendre et al. 2023) and used it to determine the dynamo
net flux) models, an MRI-driven dynamo must act to overcome the coefficients by post-processing the data obtained from the stratified
diffusion and sustain the zero net flux in the accretion flow. Earlier ZNF shearing box simulation.
ZNF simulations in unstratified (Shi et al. 2016) and stratified (Davis The paper is organised as follows. In section 2, we describe details
et al. 2010; Bodo et al. 2014; Ryan et al. 2017) shearing boxes of shearing box simulations, basics of mean field closure used and
found MRI turbulence can self-generate large-scale magnetic fields techniques of the IROS method. Section 3 describes the evolution
attaining quasi-stationarity and sustaining turbulence. Riols et al. of MRI to a non-linear saturated state, spatio-temporal variations
(2013) suggested that the non-linear MRI does not behave like a linear of mean magnetic fields, EMFs and periodicities present in differ-
instability; rather, it provides a pathway for saturation via a subcritical ent observables. The spatial profiles of calculated turbulent dynamo
dynamo process. This leads to the question of what kind of dynamo coefficients, the reliability of the calculation method (using both
can be sustained in the MRI-driven accretion flow, small-scale or EMF reconstruction and a 1D dynamo model) and contributions of
large-scale? The lack of convergence in ZNF models was attributed each coefficient to the mean magnetic energy equation are described
to the low numerical Prandtl number (Fromang & Papaloizou 2007; in section 4. In section 5, we discuss the plausible reasons behind
however, see Simon et al. 2009) and hence the inefficiency of small- different periodicities present (in mean magnetic fields, EMFs and
scale dynamo to operate at small Prandtl number (Schekochihin et al. helicties), comparison of our work with the previous works, the pos-
2005; Bodo et al. 2011). However, it is unclear what happens when sible importance of a generative helicity flux and limitations of the
convergence is recovered in unstratified ZNF simulations with tall averaging scheme and mean-field closure used in decoupling contri-
boxes (Shi et al. 2016). butions from different dynamo coefficients. Finally we summarized
Studying MRI dynamo is also important for understanding the our key results in section 6.
generation of coherent large-scale magnetic fields determining the
level of transport (Johansen et al. 2009; Bai & Stone 2013) and
outflows in the accretion disc (von Rekowski et al. 2003; Stepanovs 2 METHOD
et al. 2014; Mattia & Fendt 2022). MRI, in principle, can generate
magnetic fields coherent over several scale-heights (Dhang et al. 2.1 Shearing-box simulation
2023) and acts locally as a mean field in the absence of any external We perform stratified zero net flux (ZNF) shearing box simulations to
flux influencing convergence and the disc dynamics. study the MRI driven dynamo in a geometrically thin disc (H/R ≪
Generally, stratified models generate a more coherent large-scale 1). To do that, we solve ideal MHD equations in a Keplerian shearing
field over the unstratified models (for a comparison, see Shi et al. box given by
2016). Cyclic behaviour of azimuthally averaged magnetic fields
∂ρ
(mean fields), popularly known as the butterfly diagram, is a typical + ∇. (ρv) = 0, (1)
feature observed in the stratified shearing box simulations (Branden- ∂t
∂ρv
burg et al. 1995; Gressel 2010; Bodo et al. 2014; Ryan et al. 2017; + ∇. (ρvv − BB) + ∇P = ρgs − 2Ωẑ × ρv, (2)
Gressel & Pessah 2022). However, note that the presence of a strong ∂t
∂B
magnetic net flux (Bai & Stone 2013; Salvesen et al. 2016), con- = ∇ × (v × B) (3)
vection (Hirose et al. 2014; Coleman et al. 2017) etc. can alter the ∂t
periodicity in the butterfly diagram. Although the cyclic behaviour using the PLUTO code (Mignone et al. 2007) with x, y, z as the
of mean fields can be explained by invoking the interplay between radial, azimuthal and vertical directions respectively. Here, ρ, P, v
shear and helicity (Brandenburg & Donner 1997; Gressel & Pessah and B denote density, thermal pressure, velocity and magnetic fields,
2015), some features, such as upward migration of the mean fields, respectively. The terms gs = Ω2 (2qxx̂ − z ẑ) and 2Ωẑ × ρv repre-
still demand an explanation. sent the tidal expansion of the effective gravity and the Coriolis force
Several studies attempted to understand the underlying mechanism respectively with Ω denoting orbital frequency. We use an isothermal
of MRI dynamo using different approaches. While some of the studies equation of state
(Lesur & Ogilvie 2008; Bai & Stone 2013; Shi et al. 2016; Begelman P = ρc2s , (4)
& Armitage 2023) invoked toy models to complete the generation
cycles of radial and azimuthal fields, others (local: Brandenburg et al. which makes the energy equation redundant. Additionally, we use
2008; Gressel 2010; Shi et al. 2016; Gressel & Pessah 2022; Mondal constrained transport (Gardiner & Stone 2005) to maintain diver-
& Bhat 2023, global: Dhang et al. 2020) used mean-field theory to gence free condition
investigate the large-scale field generation in the MRI-driven turbu-
∇.B = 0 (5)
lent accretion flow. Most of the studies characterising the turbulent
dynamo coefficients in the regime of mean-field dynamo theory used for magnetic fields. We use the HLLD solver (Miyoshi & Kusano
state of the art ‘Test Field’ (TF) method (Gressel 2010; Gressel & 2005) with second-order slope-limited reconstruction. Second-order
Pessah 2015), while a few used direct methods such as linear regres- Runge–Kutta (RK2) is used for time integration with the CFL number

MNRAS 000, 000–000 (0000)


MRI driven dynamo 3
0.33. Also note that despite our shearing-box model lacking explicit we obtain the mean-field equation
dissipation, we refer to it as the direct numerical simulation (DNS).
We initialize an unmagnetized equilibrium solution with density ∂ B̄ 
and velocity given by = ∇ × v̄ × B̄ + ∇ × Ē; (13)
∂t
 
z2
ρ = ρ0 exp − , (6) where we assume that microscopic diffusivity is vanishingly small
2H 2
(ideal MHD limit). Here mean EMF
v = −q Ω x ŷ (7)
where q = 1.5 and ρ0 is the mid-plane (z = 0) density and Ē = v ′ × B ′ (14)
cs
H= (8)
Ω appears as a source term in equation 13. The crux of the mean-field
is the thermal scale height. We set ρ0 = cs = Ω = 1, so that H = 1. dynamo theory is how to express mean EMF in terms of the mean
Unless stated otherwise, all the length and time-scales are expressed magnetic fields. In general, the usual mean-field closure (Raedler
in units of H and Ω−1 respectively. We initialize a ZNF magnetic 1980; Brandenburg & Subramanian 2005; Shukurov & Subramanian
field given by 2021) is given by,
r
2 2πx
 
B= sin ẑ (9) Ēi (z) = αij (z) B̄j (z) − ηij (z) J¯j , (15)
β0 Lx
with β0 defining the strength of the field and Lx , Ly , Lz denoting where we neglect higher than the first-order spatial derivatives and
the size of the shearing-box. time derivatives of mean magnetic fields and αij , ηij are the tur-
Our computational domain extends from −Lx /2 < x < Lx /2, bulent dynamo coefficients which characterize the dynamo; and
−Ly /2 < y < Ly /2 and −Lz /2 < z < Lz /2. It has been found J¯j = ϵjzl ∂z B̄l (z) is the current. Further, while calculating turbulent
in earlier studies that shearing box results depend on the domain dynamo coefficients using direct methods (e.g. SVD (Bendre et al.
size; larger boxes tend to capture dynamo better than their smaller 2020; Dhang et al. 2020), linear regression (Squire & Bhattacharjee
counterparts as well as smaller boxes demonstrate a transition to 2016; Shi et al. 2016), it is also assumed that αij , ηij are constant in
anomalous behaviour (e.g. see Simon et al. 2012; Shi et al. 2016). time. However, we find that in our simulation of MRI-driven accretion
To avoid these discrepancies, we choose a shearing box of size Lx × flow, the current helicity, which is potentially a primary component
Ly ×Lz = 3H ×12H ×8H with a grid resolution Nx ×Ny ×Nz = determining the αij , shows a reasonably periodic change over time
96 × 192 × 256 giving rise to a resolution of 32/H in the vertical with a time period half the dynamo-period (for details, we refer the
direction. However, we must admit that there exists an issue with the reader to section 3.3). This time-dependent feature of current helicity
convergence in stratified ZNF models as discussed in the section 1. leads to considering a heuristic mean field closure defined as
We reserve the dependence of MRI dynamo on numerical resolution
as a topic of future research investigation. 0
Ēi (z) = αij 1
+ αij cos(2Ωdyn t + ϕ) B̄j (z) − ηij J¯j

(16)
We use periodic and shearing-periodic (Hawley et al. 1995) bound-
ary conditions in the y and x− boundaries, respectively. Outflow 0 1
boundary conditions are implemented in the vertical (z) boundaries. to capture the time dependence in αij . Here αij and αij are the
A gradient-free condition is maintained for scalars and tangential time-independent and time-dependent parts of αij respectively and
components of vector fields at the boundaries. At the same time, Ωdyn = 2πfdyn = 2π/Tdyn , with fdyn and Tdyn are the dynamo
vz ⩾ 0 for z > 0 and vz ⩽ 0 for z < 0 is set to restrict mass inflow frequency and period respectively. Further, one expects that ηij to
into the domain at vertical boundaries. be dominated by a DC component, because ηij -s are generally deter-
mined by the turbulent intensity of the flow, not by helicities. Thus
for simplicity, we adopt a time-independent ηij .
2.2 Mean field closure
Before describing the details of mean field dynamo theory and the
closure used, we define what is meant by ‘mean’ and ‘fluctuation’ in
our work. We define mean magnetic fields (B) as the x − y-averaged
values as follows 2.3 Dynamo coefficient extraction method - IROS
Z Lx /2 Z Ly /2
1
B̄(z, t) = B(x, y, z, t) dx dy. (10) To extract the dynamo coefficients from the shearing box simulations
Lx Ly −Lx /2 −Ly /2 described in the previous section, we solve equation 16, in a least
0 1
Fluctuating magnetic fields are defined as square sense, for αij , αij and ηij . However, since its an under-
determined system of equations, we take advantage of the fact that
B′ (x, y, z, t) = B(x, y, z, t) − B̄(z, t). (11) these dynamo coefficients stay statistically unchanged during the
quasi-stationary phase of evolution, and extract the time series of
Mean and fluctuations of the x− and z− components of the velocity
length N , Ēi (z, t1 . . . tN ), B̄i (z, t1 . . . tN ) and J¯i (z, t1 . . . tN ) from
are defined in the same way as those for magnetic fields, while the
the DNS (i ∈ {x, y}). With these time series, we rewrite equation 16
mean and fluctuation of y−component of velocity are defined as
at any particular z = z ′ as
v̄y (x, y, z, t) = −qΩx, vy′ = vy − v̄y . (12)
If we decompose the magnetic and velocity fields into mean and fluc-
tuation; and insert them into the magnetic field induction equation, y(z ′ , t) = A(z ′ , t) x(z ′ ), (17)

MNRAS 000, 000–000 (0000)


4 Dhang et al.
where the matrices y, A and x are defined as, 0.04
 ′ ′  αRey αMax
Ex (z , t1 ) Ey (z , t1 )
0.03

αRey , αMax
 E (z ′ , t ) Ey (z ′ , t1 )
 x 2
y(z ′ , t) = 

.. ..  0.02
 . . 
Ex (z ′ , tN ) ′
Ey (z , t1 ) 0.01

0.00
0 50 100 150 200 250 300
B̄x (z ′ , t1 ) B̄x (z ′ , t2 ) . . . B̄x (z ′ , tN )
 
 B̄y (z ′ , t1 ) 0.015 B 02 B̄ 2
 B̄y (z ′ , t2 ) . . . B̄y (z ′ , tN ) 
 ′ ′ ′ 

B 02 , B̄ 2
⊺ ′
 C x (z , t 1 ) C x (z , t 2 ) . . . C x (z , t N )  0.010
A (z , t) = 

′ ′ ′

C
 y (z , t 1 ) C y (z , t 2 ) . . . C y (z , t N ) 

 ¯ ′ 0.005
−Jx (z , t1 ) −J¯x (z , t2 ) . . . − J¯x (z , tN )
′ ′

−J¯y (z ′ , t1 ) −J¯y (z ′ , t2 ) . . . − J¯y (z ′ , tN ) 0.000


0 50 100 150 200 250 300
t (Ω−1 )
0
(z ′ ) 0
(z ′ )
 
αxx αyx
α0 (z ′ ) α0 (z ′ ) Figure 1. Top panel: Time history of Reynolds (αRey ) and Maxwell (αMax )
 xy yx  stresses. Bottom panel: time history of the volume-averaged mean (B̄ 2 ) and
 1 ′ 1 ′ 
α xx (z ) αyx (z ) fluctuating (B ′2 ) magnetic energies.
x(z ′ , t) = 
 
α1 (z ′ ) α1 (z ′ ) . (18)

 xy yx 
 ηxx (z ′ ) ηyx (z ′ ) 
 

ηxy (z ′ ) ηyy (z ′ ) the EMF associated with the best fitted coefficient, also multiplied by
the ϵ is subtracted from the EMF component. For instance using the
Here the terms Ci (z ′ , tl ) = B̄i (z ′ , tl ) cos (2Ωdyn tl + ϕ) (∀i ∈ same example a factor of ϵ αxy (z ′ ) B̄y (z ′ , t1 . . . tN ) is subtracted
{x, y}). For simplicity, we assume ϕ to be zero. To then determine the from Ex (z ′ , t1 . . . tN ). This residual EMF is then used as an actual
dynamo coefficients (x) we pseudo-invert equation 17. This task is EMF component, and this process is repeated a suitable number
complicated firstly by the fact that both components of mean-field and of times until either all the dynamo coefficients converge to their
current have additive correlated noise and secondly by the fact that the respective constant values or all four chi-squared errors get smaller
y component of the mean-field is typically much stronger compared than a certain predefined threshold. All the aforementioned steps are
to the x component, due to the rotational shear (and by consequence then repeated at every z = z ′ .
the x component of current is much stronger than its y component). We apply this method with ϵ = 0.1 for five hundred refinement
Typical schemes of the least square minimisation in such cases tends loops to the time series of EMFs, mean-fields and currents obtained
to underestimate the dynamo coefficients that are associated with from the DNS data. While constructing these time series (from t =
x and y components of mean-field and current respectively (i.e. 100Ω−1 to 300Ω−1 ) with data dumping interval ∆tdump = 0.2 Ω−1
0 1
αix , αix and ηiy ). To circumvent these issues, we rely upon the we make sure that they correspond to the quasi-stationary phase of
IROS method (Iterative removal of sources) (Hammersley et al. 1992) the magnetic field evolution.
that we have recently adapted for such inversions in the dynamo IROS method does not provide an estimate of errors on the calcu-
context (Bendre et al. 2023). This method is based on Högbom lated coefficients directly. We, therefore, calculate a statistical error of
clean algorithm used in Radio Astronomy to construct an image the dynamo coefficient by considering the five different realizations
by convolving multiple beams, iteratively locating and subtracting of time series. We construct five different time series of mean fields,
out the strongest source to model the rest of the dirty image. It is currents and EMFs by skipping four data points in the time series.
particularly useful when the relative contribution of some of the Specifically, the time series (t1 , t2 , . . . tN ) (of all components of
beams to the final image happens to be negligible. Such a situation mean-field, current and EMF) are split into (t1 , t6 . . .), (t2 , t7 . . .),
is analogous to have only a few of the columns of A (the beams) (t3 , t8 . . .), (t4 , t9 . . .) and (t5 , t10 . . .). We use these time series to
largely contribute to y (an image). A brief outline of the method is calculate five sets of dynamo coefficients and calculate their standard
as follows. deviations to represent the errors on the calculated coefficients.
Firstly, at any particular z = z ′ we set all the dynamo coefficients,
0
αij (z ′ ), αij
1
(z ′ ) and ηij (z ′ ) to zero, i.e we set x(z ′ ) = 0. Then
to derive their zeroth order estimates we fit every ith column of
y(z ′ , t) (say yi (z ′ , t)), against the individual columns of A(z ′ )
(denoted as Ak (z ′ )) separately as lines. The best amongst these four
3 RESULTS: SATURATION OF MRI, MEAN FIELDS AND
linear fits is decided basedPon the least of chi-square errors of the
EMF-S
individual fits (χ2ik (z ′ ) = i (yi − Ak xik )2 ). Then the best fitted
dynamo coefficient is updated by adding to it, its zeroth order estimate We now turn to the results of our shearing box simulation of MRI in
multiplied by a small factor (ϵ < 1), called the loop-gain, while other a geometrically thin disc, investigate its dynamo action in addition to
coefficients are kept constant. For example if the chi-squared error discussing several important properties which illuminates the nature
associated with the line fit Ex (z ′ , t1 . . . tN ) versus B̄y (z ′ , t1 . . . tN ) of the MRI dynamo. Most of our analysis of magnetic field generation
(i.e. χ212 (z ′ )) is the least then x2,1 (z ′ ) (i.e. αxy
0
) is updated by a will focus on the saturated state of MRI, when the disc is in the quasi-
factor of the slope multiplied by ϵ. Subsequently, the contribution to stationary phase.

MNRAS 000, 000–000 (0000)


MRI driven dynamo 5
3.1 Saturation of MRI studies of MRI dynamo (Brandenburg et al. 1995; Davis et al. 2010;
Gressel 2010; Gressel & Pessah 2015; Ryan et al. 2017). This peri-
First, consider the time evolution of accretion stresses and magnetic
odicity is semi-transparent in the butterfly diagram of B̄x , while this
energies. This will also allow us to determine the quasi-stationary
is hardly apparent for Ēy . However, we note that periodicities exist in
phase of the MRI-driven turbulence. The top panel of Fig. 1 shows
all components of mean fields and EMFs as will become clear below
the time history of accretions stresses (Reynolds and Maxwell). Nor-
(see Fig. 4).
malized Reynolds and Maxwell stresses are defined as
⟨ρvx′ vy′ ⟩
αRey = , (19) 3.3 Evolution of kinetic and current helicities
⟨pg ⟩
⟨B̄x B̄y ⟩ + ⟨Bx′ By′ ⟩ The generation of large-scale magnetic fields by a dynamo action is
αMax = , (20) often associated with helicity in the fluid velocity field. Assuming
⟨pg ⟩
isotropic homogeneous turbulence, Krause & Raedler (1980) sug-
where the averages are done over the whole volume. Reynolds stress gested a kinetic α-effect defined by
is due to the correlation of fluctuating velocity fields, while Maxwell
τc
stress is composed of a correlation between the fluctuating compo- αkin = − v ′ .∇ × v ′ (21)
nents as well as that between the mean components of the magnetic 3
fields. Both the stresses grow exponentially during the linear regime responsible for magnetic field generation; where τc is the correlation
of MRI, and eventually saturate around an average value when MRI time, and Khel = v ′ .∇ × v ′ is the kinetic helicity. It is suggested
enters into the non-linear regime. In our simulation, we find the that αkin accounting for the effects of the helical velocity field, takes
time-averaged (within the interval t = (100 − 300) Ω−1 ) values the role of driver, while αmag (Pouquet et al. 1976) defined by
of Reynolds and Maxwell stresses are to be ⟨αRey ⟩ = 0.0048 and dyn τc ′ ′
⟨αMax ⟩ = 0.0167 respectively. The ratio of Maxwell to Reynolds αmag (z, t) = v .∇ × vA , (22)
3 A
stress is αMax ⟩/⟨αRey ⟩ = 3.5, close to 4, as predicted by Pessah
is the non-linear response arising due to the Lorentz force feed-
et al. 2006 for q = 1.5 and similar to what is found in earlier nu-
back, gradually increasing and ultimately quenching the kinetic-
merical simulations (Nauman & Blackman 2015; Gressel & Pessah
α (Blackman & Brandenburg 2002; Subramanian 2002). Here,
2015). ′
p
′ ′
vA = B ′2 /ρ is the Alfv̀en velocity and Chel = vA .∇ × vA
The bottom panel of Fig. 1 shows how the volume-averaged mean
is the current helicity. Ideally, the effective α− effect, responsible to
(⟨B̄ 2 ⟩) and fluctuating (⟨B ′2 ⟩) magnetic energies evolve over time.
poloidal field generation, is supposed to be αdyn = αkin + αmag .
Like accretion stresses, magnetic energies oscillate about an average
Fig. 3 shows the spatio-temporal variation of αkin and αmag .
value in the quasi-stationary phase after the initial exponentially
We assume τc to be same for both α-s and τc = Ω−1 . The αmag
growing phase. It is also worth noting that the mean part of the
changes sign with a time-period ≈ 5 orbital period (2π/Ω), roughly
magnetic field shows a larger time variation than the fluctuating
half of the dynamo period, with which the mean fields and EMFs
part of the magnetic field. We point out an important point that the
change sign, while αkin does not show any explicit periodicity. We
fluctuating magnetic field is stronger than the mean magnetic field,
will postpone a detailed discussion on the periodicity of helicities
and the implication of this will be discussed in the latter part of the
to section 3.5 where we discuss periodicities associated with all
paper.
important variables.
We see in Fig. 1 that the accretion stresses and magnetic energies
start saturating around t = 40 Ω−1 . However, to remain safer, we
consider the simulation in the time range t = (100 − 300) Ω−1 for 3.4 Co-existence of small and large scale dynamos
dynamo coefficient calculation in the quasi-stationary state.
Both kinetic and magnetic-α-s are small close to the mid-plane while
this is not true of the random kinetic and magnetic energies (see Fig.
3.2 Evolution of mean fields and EMFs 6). At the same time, the amplitudes of the helicities increase away
The most preliminary diagnostic of the dynamo is to look at the from the mid-plane. These features suggest that both small-scale
spatio-temporal variation of the mean magnetic fields, popularly and large-scale dynamo co-exist in MRI-driven dynamo (Blackman
known as the butterfly diagram (e.g. see the review by Branden- & Tan 2004; Gressel 2010). The MRI-driven small-scale dynamo
burg & Subramanian 2005). Fig. 2 shows the butterfly diagrams for dominates magnetic field generation close to the disc mid-plane
mean magnetic fields B̄x and B̄y along with the mean EMFs Ēx (where stratification is unimportant). In contrast, at larger heights
and Ēy . Here we note that the mean EMF acts as a source term in where stratification becomes important, a helicity-driven large-scale
the mean magnetic field energy evolution equation. In particular, Ēy dynamo governs the magnetic field generation (Dhang & Sharma
is responsible for the generation of poloidal field (here B̄x ) from a 2019; Dhang et al. 2020). However, it is to be noted that αmag is
toroidal one due to an α-effect, which itself naturally emerges by larger than αkin by one order of magnitude, and hence it is very
a combined action of stratification and rotation (Krause & Raedler likely that the effective-α will be predominantly due to αmag .
1980) in our stratified shearing box simulation. At an early stage of
evolution (around t ≈ 2 orbital period), both mean fields and EMFs
3.5 Power spectra of mean fields, EMFs and helicities
show lateral stretches with changing the sign in the vertical direction,
which is clearly due to channel modes of MRI (Balbus & Hawley The butterfly diagrams shown in the previous sections depict the
1992; Balbus & Hawley 1998). During saturation, both mean fields apparent periodicities of mean fields, EMFs and helicities. We look
and EMFs show a coherent vertical structure which changes signs in at the power spectrum defined by
time with a definite period. We find magnetic field components By 2
z2
Z Z
and EMF Ēx show a very coherent spatio-temporal variation with 1
if t
Pq (f ) = dz q̄(z, t)e dt
(23)
a time period of ≈ 9 orbital period (2π/Ω), similar to the earlier z2 − z1 z1

MNRAS 000, 000–000 (0000)


6 Dhang et al.

4 4
B̄x Ēx
0.01 0.01
2 2

0.00 0.00
0 0
z

z
−2 -0.01 −2 -0.01

−4 −4
0 10 20 30 40 0 10 20 30 40
4 4
B̄y /10 10 × Ēy
0.01 0.01
2 2

0.00 0.00
0 0
z

z
−2 -0.01 −2 -0.01

−4 −4
0 10 20 30 40 0 10 20 30 40
t (2π/Ω) t (2π/Ω)
Figure 2. Spatio-temporal variation of mean magnetic fields, B̄x (top left panel), B̄y (bottom left panel) and mean EMFs Ēx (top right panel) and Ēy (bottom
right panel). Mean magnetic field component B̄y and y-component of EMF Ēx show a coherent change in space and time (with a time period ≈ 9 orbital period
(2π/Ω)), while the spatio-temporal patterns in B̄x and Ēy are less coherent.

4 mean EMFs Ēx , Ēy (middle panels) and helicities Khel , Chel (bottom
αkin panels). It is noticeable that power spectra for mean fields and spectra
0.02 peaks at the primary frequency fdyn = 0.017, which was also visible
2
in the butterfly diagrams. In addition to the primary frequency, the
power spectra also show the presence of higher harmonics (at 3fdyn ),
0.00
0 which got unnoticed in the earlier works of MRI dynamo. Similarly,
z

power spectra of current helicity Chel also show the presence of


−2 -0.02
higher harmonics (at 4 fdyn ) in addition to the primary frequency
at 2fdyn . However, kinetic helicity does not show any periodicity.
Presence of a strong time variation in αmag and its dominance over
−4
0 10 20 30 40 αkin necessarily leads to the expectation that turbulent dynamo coef-
4 ficients (α− coefficients) should harbour a time-dependent part (αij 1
)
αmag
0
0.10 along with the traditional time-independent part (αij ) as discussed
2 in section 4.

0.00
0
z

−2 -0.10

4 RESULTS: DYNAMO COEFFICIENTS FROM IROS


−4
0 10 20 30 40 We obtained mean fields (B̄x , B̄y ), EMFs (Ēx , Ēy ) from the shearing-
t (2π/Ω) box simulation and use a modified version of IROS method (see
section 4) to calculate time-independent and time-dependent turbu-
Figure 3. Spatio-temporal variation of αdyn dyn
kin (z, t) and αmag (z, t) assuming lent dynamo coefficients characterizing the MRI dynamo. However,
τc = Ω−1 . Both the helicities are small close to the mid-plane and become
larger at larger heights. The αmag flips sign with a time period ≈ 5 orbital
we find the x − y-averaging cannot remove all the signatures of the
period (2π/Ω), roughly half the dynamo period, while αkin does not show small-scale dynamo. The small-scale dynamo is expected to have a
any periodicity. shorter correlation time of order few Ω−1 and contribute noise at the
higher frequency end compared to the large-scale dynamo. There-
fore, we further smooth the mean fields and EMFs using a low-pass
where q̄(z, t) is any generic quantity to investigate the periodicities in Fourier filter and remove contributions from the frequencies f > fc .
greater detail. Here the spatial average is done over different heights, We consider three cases: (i) fc = 0.05 (≈ 3fdyn ), (ii) fc = 0.12
namely z = 0 − H, z = H − 2H and z = 2H − 3.5H. (≈ 6fdyn ) and (iii) fc → ∞ (unfiltered) to assess the effects of the
Fig. 4 shows the power spectra of mean fields B̄x , B̄y (top panels), small-scale dynamo on the dynamo coefficient extraction.

MNRAS 000, 000–000 (0000)


MRI driven dynamo 7

101
B̄x 103
0−H B̄y 0−H
100 H − 2H H − 2H
2H − 3.5H 102 2H − 3.5H
10−1 101
P

10−2 100

10−3 10−1

10−2
10−4 fdyn 3fdyn fdyn 3fdyn
10−3
10−2 10−1 100 10−2 10−1 100
−1
10
Ēx 0−H Ēy 0−H
100 H − 2H −2 H − 2H
2H − 3.5H
10 2H − 3.5H
10−1
10−3
10−2
P

10−4
10−3

10−4 10−5
fdyn 3fdyn fdyn 3fdyn
10−5 10−6
10−2 10−1 100 10−2 10−1 100
f (Ω) f (Ω)
4
10
0−H 0−H
104
103 H − 2H H − 2H
2H − 3.5H 103 2H − 3.5H
102
102
P

1
10
101
0
10 100
10−1 10−1
Khel 2fdyn 4fdyn Chel 2fdyn 4fdyn
10−2 10−2
10−2 10−1 100 10−2 10−1 100
f (Ω) f (Ω)
Figure 4. Power spectra of mean fields B̄x , B̄y (top panels), mean EMFs Ēx , Ēy (middle panels) and helicities Khel , Chel (bottom panels). Spatial averages
are done over different heights: z = 0 − H (black lines), z = H − 2H (green lines), and z = 2H − 3.5H (red lines). The zeroth frequency values are denoted
by ‘asterisks’. Vertical dashed lines denote the dynamo frequency fdyn = 0.017 and its multiples.

4.1 Time independent dynamo coefficients local (Brandenburg 2008; Gressel 2010; Gressel & Pessah 2015) and
global (Dhang et al. 2020) frameworks also found a similar trend in
Fig. 5 shows the vertical profiles of time-independent dynamo coef- 0
αyy . However, it is to be noted that our study suggests a stronger
0
ficients αij and ηij for different values of fc . Four panels at the top negative αyy0
in the upper-half plane compared to that in the earlier
0 0
illustrate the vertical profiles of coefficients (αxx , αxy , ηxx , ηxy ) studies. The negative sign in the upper half plane is attributed to
associated with the x-component of EMF Ēx , while four panels at the buoyant rise of magnetic flux tubes under the combined action
0 0
the bottom show profiles of those (αyx , αyy , ηyx , ηyy ) associated of magnetic buoyancy and shear (Brandenburg & Schmitt 1998;
with the y-component of EMF Ēy . Brandenburg & Subramanian 2005; see also Tharakkal et al. (2023)).
0
The ‘coefficient of most interest’ out of the calculated ones is αyy , Brandenburg & Schmitt (1998) also suggested that negative αyy is
which plays a vital role in producing the poloidal field (here B̄x ) responsible for the upward propagation direction of dynamo waves
out of the toroidal field (B̄y ) (also see section 4.4). The coefficient seen in the butterfly diagrams of MRI-driven dynamo simulations
0
αyy shows an anti-symmetric behaviour about the z = 0 plane, (e.g. see Fig. 2). Another different way of looking at the origin of the
with a negative (positive) sign in the upper (lower)-half plane (for effective α is to link it to the helicity flux as envisaged by Vishniac
0
|z| < 2). For |z| > 2, the sign of αyy tends to be positive (negative)
in the upper (lower)-half plane. Earlier studies of MRI dynamo in

MNRAS 000, 000–000 (0000)


8 Dhang et al.

fc = 0.05 fc = 0.12 Unfiltered


20 20
/10−2

/10−2
10 10
0 0
αxx

αxy
−10 −10

0
0

−20 −20
−3 −2 −1 0 1 2 3 2 −3 −2 −1 0 1 2 3
3
ηxx /10−2

ηxy /10−2
1
2 0
1 −1
−2
0
6 −3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
1.0
/10−2

/10−2
3 0.5
0 0.0
αyx

αyy
−0.5
0
0

−3
−1.0
−6
0.4 −3 −2 −1 0 1 2 3 1.0 −3 −2 −1 0 1 2 3
z z
ηyx /10−2

ηyy /10−2

0.3 0.5
0.2 0.0
0.1 −0.5
0.0 −1.0
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
z z
Figure 5. Vertical profiles of time-independent turbulent dynamo coefficients (α0ij , ηij
0 ) in MRI simulation calculated using IROS method. A low-pass Fourier

filter with a cut-off frequency fc removes the contribution from the small-scale dynamo. We used two values of fc : fc = 0.05 and fc = 0.12. The results are
compared to the case when IROS is applied to the unfiltered data obtained from DNS.

2015 and Gopalakrishnan & Subramanian 2023. We discuss this noted that MRI turbulence in a stratified medium is neither isotropic
possibility in section 5.3. nor homogenous. Minimal τ -approximation (MTA) suggests that in
The off-diagonal terms of the α-coefficients are related to turbu- a stratification and rotation-induced anisotropic turbulent medium,
lent pumping. This effect is responsible for transporting large-scale which includes the quasi-linear back reaction due to Lorentz forces,
magnetic fields from the turbulent region to the laminar region. We
0 0 0 0 1 1
found αxy and αyx to be antisymmetric and αxy > αyx unlike γzMTA = − τ ∇z (v ′2 − B ′2 ) − τ 2 Ωẑ × ∇z (v′2 + B′2 ), (24)
the previous studies (Brandenburg 2008; Gressel & Pessah 2015) 6 6
0 0
which found αyx ≈ αxy . This resulted in a strong turbulent pump- where τ is the correlation time and it is assumed that ρ = 1 (see
0 0
ing γz = (αyx − αxy )/2, transporting large-scale magnetic fields equation (10.59) in Brandenburg & Subramanian 2005). The last
from the disc to the corona as shown in the top panel of Fig. 6. term in equation 24 vanishes because all the variables are functions
We also compare the relative importance of turbulent pumping (γz ) of z alone. Therefore, equation 24 and the bottom panel of Fig. 6
′2
and wind (v̄z ) in advecting the magnetic field upward (in the upper illustrating the vertical profiles of v ′2 and vA imply that sign of
half-plane) at different heights. Vertical profiles of γz and v̄z in the turbulent pumping obtained from MTA supports that obtained from
top panel of Fig. 6 shows that at low heights (|z| < 2.5), turbulent extracted dynamo coefficients.
pumping is the dominant effect over the wind, while the effects of We found turbulent diffusion tensor ηij to be anisotropic with
wind become comparable or larger over the pumping term at large ηxx > ηyy and having a significant contribution from the off-
scale-heights (see also Fig. 11). diagonal components ηxy and ηyx . Different values of diagonal
The theory of isotropic kinematically forced turbulence predicts components of ηij imply that mean field components B̄x and B̄y
that γz is supposed to be in the direction of negative gradient of are affected differently by the vertical diffusion (also see section
turbulent intensity (v ′2 ) (Krause & Raedler 1980), that is, in the 4.4). It is worth mentioning that ηyy ≈ 0 for the fc = 0.05 case,
negative z-direction (in the upper-half plane) in our simulation. This while it is slightly negative for the other two cases. This is somewhat
is opposite to what has been found in Fig. 6. However, it is to be different from the earlier studies (Gressel 2010; Gressel & Pessah

MNRAS 000, 000–000 (0000)


MRI driven dynamo 9
4.3 Verification of method
v̄z γz , fc = 0.12
γz , fc = 0.05 γz , Unfiltered To verify the reliability of the determined dynamo coefficients we
0.1
reconstruct the EMFs using the calculated coefficients and run a 1D
dynamo model.
γz , v̄z

0.0

4.3.1 Reconstruction of EMFs


−0.1
Fig. 8 shows butterfly diagrams of the EMFs (Ēx,f , Ēy,f ) used
to determine the turbulent dynamo coefficients and the EMFs
−3 −2 −1 0 1 2 3 (Ēx,r , Ēy,r ) reconstructed using calculated coefficients and mean
z
fields for fc = 0.05. Here it is to be noted that Ēx,f , Ēy,f are the
02
smoothed EMFs obtained by filtering (by using a low-pass filter)
0.2 v 02 vA
EMFs Ēx , Ēy from DNS respectively. We can see a close match
between the broad features, such as the dynamo cycle period, in the
smoothed and reconstructed EMFs, implying the goodness of fit.
v 02, vA02

Further, we investigate the residual of the filtered and reconstructed


0.1 EMFs, defined by
δ Ēi = Ēi,f − Ēi,r , i ∈ x, y. (25)
Fig. 9 shows the histograms of the normalised residuals δ Ēx /|Ēx | and
0.0 δ Ēy /|Ēy | calculated within the region of different heights, namely
−3 −2 −1 0 1 2 3
z between 0−H, H −2H and 2H −3H, for the fc = 0.05 case. All the
histograms peak about the region close to zero. However, a Gaussian
Figure 6. Top panel: profiles of turbulent pumping (γz ) and mean vertical fit of the histograms shows that the mean of the distribution always
outflow (v̄z ). They act in the same direction, transporting mean fields verti- deviates from zero. Additionally, a careful comparison of histograms
cally outward. Bottom panel: vertical profiles of average fluctuating velocity
′2 = B ′2 /ρ. Minimal τ approximation
of δ Ēx /|Ēx | and δ Ēy /|Ēy | tells that fit is better for Ēx than that for
(v ′2 ) and fluctuating Alfven speed vA
Ēy , especially at larger scale-heights. Better quality fit for Ēx over Ēy
′2 suggest similar sign of γ as calculated using IROS.
and profiles of v ′2 , vA z is expected as Ēx obtained from DNS shows a more regular, coherent
space-time variation when compared to Ēy .

4.3.2 1D dynamo model


2015), which calculated dynamo coefficients using the TF method We additionally run a 1D dynamo model using the calculated dy-
and found ηxx ≈ ηyy > 0. Out of the two off-diagonal terms of the namo coefficients and mean velocity field v̄z . In particular we solve
diffusion tensor, ηyx is of particular interest. It is suggested that a equation 13, or in component form
negative value of ηyx can generate poloidal fields by the shear-current
effect (Squire & Bhattacharjee (2016)). However, we find ηyx to be ∂ B̄x ∂  0 0 ∂ B̄x ∂ B̄y 
= − (v̄z + αyx )B̄x − αyy B̄y + ηyy − ηyx
always positive, nullifying the presence of a shear-current effect in ∂t ∂z ∂z ∂z
our stratified MRI simulation.
∂ B̄y ∂  0 0 ∂ B̄y ∂ B̄x 
= − (v̄z − αxy )B̄y − αxx B̄x + ηxx − ηxy
∂t ∂z ∂z ∂z
+ qΩ B̄x . (26)
0
for B̄x and B̄y with αij and ηij obtained using the IROS method.
4.2 Time dependent dynamo coefficients
We note that B̄z = 0 as a consequence of the zero net flux (ZNF) as-
In the previous sections, We discussed the time-dependent nature sumption in our model. The initial profiles of B̄x and B̄y at are taken
of αmag . Effective α-effect is expected to be determined by αmag , directly from the DNS, at time t = 100Ω−1 roughly consistent with
especially at the larger scale heights where it is of larger amplitude. the beginning of the quasi-stationary phase in the DNS. v̄z profile
While α tensor is expected to have the time-dependent part, η-tensor is taken as a constant throughout the evolution and is also extracted
is supposed to have only the time-independent part, as it only depends from the direct simulations by averaging it over time throughout the
on the turbulent intensity (see section 4). Fig. 7 shows the vertical quasi-stationary phase, over which it roughly stays constant. Addi-
1 0
profiles of components of time-dependent α-tensor. We find that αij - tionally, for the profiles of dynamo coefficients αij (z) and ηij (z),
s in the fiducial case (fc = 0.05) are relatively smaller than the other we first smooth them with a box filter and also cut them off above and
two cases (fc = 0.12 and Unfiltered). It is also interesting to note below three scale heights, and use them in the 1D dynamo model.
that the coefficients αxx and αyx associated with B̄x in the mean We do this mainly to avoid the instability at boundaries noting that
field closure (equation 16) have stronger time-dependence compared these profiles are sharply flayed outside of that range. Note that only
to those coefficients (αxy and αyy ) associated with B̄y . Additionally, the time independent parts of the dynamo coefficients are used in the
1
it is to be noted that the amplitudes of the time-dependent part of α-s mean field equations, since the contributions of αij are negligible
are higher at larger scale heights. However, overall, the amplitudes of compared to the time-independent part.
1 0
αij are much smaller than the αij implying that the time-independent Furthermore, it must be noted that there is a contribution to the
α−s are predominantly governing the dynamo action. diffusion from the mesh grids. We do a rough estimation of numerical

MNRAS 000, 000–000 (0000)


10 Dhang et al.

fc = 0.05 fc = 0.12 Unfiltered

20 5
/10−2

/10−2
10
0 0
αxx

αxy
1
1

−10
−20 −5

6 −3 −2 −1 0 1 2 3 0.5 −3 −2 −1 0 1 2 3

3
/10−2

/10−2
0 0.0
αyx

αyy
1
1

−3

−6 −0.5
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
z z
Figure 7. Vertical profiles of time-independent turbulent dynamo coefficients (α1ij ) in MRI simulation calculated using IROS method. A low-pass Fourier filter
with a cut-off frequency fc is used to remove the contribution from the small-scale dynamo. We used two values of fc : fc = 0.05 and fc = 0.12. The results
are compared to the case when IROS is applied to the unfiltered data obtained from DNS.

Ēx 10 × Ēy
-0.01 0.00 0.01 -0.01 0.00 0.01

3 Ēx,f 3 Ēy,f

0 0
z

−3 −3

20 30 40 20 30 40
3 Ēx,r 3 Ēy,r

0 0
z

−3 −3

20 30 40 20 30 40
t (2π/Ω) t (2π/Ω)

Figure 8. Left panels: Comparison between x-component of EMF Ēx,f used to determine the turbulent dynamo coefficients,and EMF Ēx,r reconstructed using
the turbulent dynamo coefficients. Right panels: Same as figures in left panels, but for the y-component of EMF.

MNRAS 000, 000–000 (0000)


MRI driven dynamo 11
fc = 0.05 z =0−H z = H − 2H z = 2H − 3H z =0−H z = H − 2H z = 2H − 3H
µ = −0.01, σ = 0.42 µ = −0.02, σ = 0.37 µ = 0.01, σ = 0.37 µ = −0.03, σ = 0.39 µ = −0.08, σ = 0.37 µ = 0.08, σ = 0.51
16.0% 16.0% 16.0% 16.0% 16.0% 16.0%
Percentage

12.0% 12.0% 12.0% 12.0% 12.0% 12.0%

8.0% 8.0% 8.0% 8.0% 8.0% 8.0%

4.0% 4.0% 4.0% 4.0% 4.0% 4.0%

0.0% 0.0% 0.0% 0.0% 0.0% 0.0%


−1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0
δEx /|Ex | δEx /|Ex | δEx /|Ex | δEy /|Ey | δEy /|Ey | δEy /|Ey |

Figure 9. Histograms of the residual EMFs, δ Ēi = Ēi,f − Ēi,r calculated within region of different heights for fc = 0.05 case. We normalise δ Ēi with the
absolute values of the corresponding EMFs at the respective points. The Red dashed line shows a normal distribution fitting the histogram.

4.4 Mean magnetic energy equations


3
B̄x
2 0.01 It is challenging to calculate dynamo coefficients uniquely in the pres-
ence of both shear and rotation (Brandenburg et al. 2008) as there
1 are many unknowns (see also discussion in section 5.4). Therefore, it
0.00 is worth seeing how different terms involving turbulent dynamo co-
0
z

efficients contribute to the mean magnetic energy equation to make


−1
physical sense. The mean magnetic energy evolution equation is ob-
-0.01
−2 tained by taking the dot product of the mean-field equation (equation
13) with the mean magnetic field B and given by
−3
∂ 1 2
 
20 30 40 50 60
3 B̄x = TBx ,vz + Tαyx + Tαyy + Tηyx + Tηyy , (27)
B̄y /10 ∂t 2
0.01
∂ 1 2
 
2
B̄y = TBy ,vz + Tαxy + Tαxx + Tηxy + Tηxx + TS ,
∂t 2
1
(28)
0.00
0
z

where
−1 1 ∂ 
-0.01 TBx ,vz = − B̄x v̄z B̄x ,
−2 2 ∂z
1 ∂ 
−3 Tαyx = − B̄x αyx B̄x ,
2 ∂z
20 30 40 50 60
1 ∂ 
t (2π/Ω) Tαyy = − B̄x αyy B̄y ,
2 ∂z
1 ∂ ∂
 
Figure 10. Butterfly diagrams of the mean magnetic fields B̄x and B̄y Tηyx = − B̄x ηyx B̄y ,
obtained by running 1D dynamo model. Both B̄x and B̄y flip sign regularly 2 ∂z ∂z
with a cycle of ≈ 9 orbital period, similar to that found in shearing box 1 ∂ ∂
 
Tηyy = B̄x ηyy B̄x ,
simulations (see Fig. 2). 2 ∂z ∂z
1 ∂ 
TBy ,vz = − B̄y v̄z B̄y , (29)
2 ∂z
1 ∂ 
Tαxy = B̄y αxy B̄y ,
2 ∂z
1 ∂ 
Tαxx = B̄y αxy B̄y ,
2 ∂z
1 ∂ ∂

 
diffusion as follows η0 = vrms ∆x, where we consider the smallest Tηxy = − B̄y ηxy B̄x ,
one among the relevant velocities (vrms ′
, cs , vA ) in the problem. 2 ∂z ∂z
Therefore, we add a correction term η0 ≈ 10−3 (with ∆x = 1/32 1 ∂ ∂
 

Tηxx = B̄y ηxx B̄y ,
and vrms = 0.1) to the diagonal components of diffusivity tensor 2 ∂z ∂z
ηij to consider the contribution from the mesh to the magnetic field 1
TS = qΩ B̄x B̄y .
diffusion. This also helps us to stabilize the 1D dynamo solution. 2
With this setup, we solve the system of equations with a finite dif- Fig. 11 shows the space-time plots of different terms involving
ference method over a staggered grid of resolution ∆z = 1/32, same mean flow (v̄z ) and turbulent dynamo coefficients (αij , ηij ) in the
as the z resolution of DNS. Outcome of this analysis are presented in mean magnetic energy evolution equations. The top six panels in
Fig. 10, where top and bottom panels show the butterfly diagrams of Fig. 11 describe the terms in the x-component of the magnetic energy
B̄x and B̄y obtained using the 1D dynamo model respectively. We equation (equation 27), while the bottom seven panels illustrate terms
find both x and y-components of mean fields flip sign regularly with a in the y-component of the magnetic energy equation ( 28) at different
cycle of ≈ 9 orbital period, similar to what is found in DNS (see Fig. heights and times.
2). Thus, applying calculated coefficients to the 1D dynamo model Fig. 11 provides a fairly complicated picture to account for the
successfully reproduces broad features of spatiotemporal variations generation-diffusion scenario of the mean magnetic fields. Broadly
mean magnetic fields. speaking, the poloidal field (B̄x ) is predominantly generated by an

MNRAS 000, 000–000 (0000)


12 Dhang et al.

1 2
2 B̄x TBx ,vz Tαyx
2 2 2
50.0 2.0 2.0

0 0 0
z

0.0 0.0 0.0


−2 −2 −2
-50.0 -2.0 -2.0
20 30 40 20 30 40 20 30 40
Tαyy t (2π/Ω) Tηyx t (2π/Ω) Tηyy
2 2.0 2 2.0 2 2.0

0 0 0
z

0.0 0.0 0.0


−2 −2 −2
-2.0 -2.0 -2.0
20 30 40 20 30 40 20 30 40
t (2π/Ω) t (2π/Ω) t (2π/Ω)
1 2
2 B̄y TBy ,vz Tαxy
2 50.0 2 4.0 2 4.0

0 0.0 0 0.0 0 0.0


z

−2 −2
-50.0 -4.0 −2 -4.0

20 30 40 20 30 40 20 30 40
Tαxx t (2π/Ω) Tηxy t (2π/Ω) Tηxx
2 4.0 2 4.0 2 4.0

0 0.0 0 0.0 0 0.0


z

−2 -4.0−2 -4.0 −2 -4.0

20 30 40 20 30 40 20 30 40
TS
t (2π/Ω) t (2π/Ω) t (2π/Ω)
2 4.0

0 0.0
z

−2 -4.0

20 30 40
t (2π/Ω)
Figure 11. Contributions of different terms involving mean flow (v̄z ) and turbulent dynamo coefficients (αij , ηij ) to the x-(top six panels) and y-(bottom seven
panels) components of mean magnetic energy evolution equation (equations 27 and 28). Poloidal field (B̄x ) generation is primarily attributed to an α-effect (the
term Tαyy ), while shear (the term TS ) dominates the toroidal field generation; thus implying and α − Ω type of dynamo. Winds carry mean fields out of the
computational box and contribute largely as the sink term in the mean magnetic energy evolution equation.

α-effect (the term Tαyy in Fig. 11). However, there is a significant of the dynamo in an MRI-driven geometrically thick accretion disc
contribution from αyx (the term Tαyx in Fig. 11) in generating B̄x (Dhang et al. 2020), implying universal action of α − Ω-dynamo in
in larger scale-heights. Toroidal field generation is mainly due to the MRI-driven accretion flows.
presence of shear, here differential rotation, (TS in Fig. 11) which
converts poloidal fields to the toroidal fields. However, it is worth
noting that there is a minute contribution from the αxx , generating a
toroidal field out of the poloidal field by an α-effect, (as in an α2 -Ω
dynamo). The dominance of α-effect in generating a poloidal field Generally, it is expected that diagonal components of the diffusion
and that of Ω-effect (shear) in generating a toroidal field imply the tensor, ηyy and ηxx are primarily responsible for the diffusion of B̄x
presence of an α − Ω type dynamo in MRI-driven geometrically thin and B̄y respectively. However, our simulation finds that winds carry
accretion disc. This is similar to what has been found in the study mean fields out of the computational box and act as a sink in the
mean magnetic energy evolution equation, not the η-s.

MNRAS 000, 000–000 (0000)


MRI driven dynamo 13
5 DISCUSSION can be realised by comparing αyy profiles in our work (also in Shi
et al. (2016)) and in Gressel (2010). Additionally, we find stronger
5.1 Periodicities in the dynamo cycle
turbulent pumping (compared to that in the TF method), transporting
Investigations of spatio-temporal variation of different variables in large-scale magnetic fields from the disc to the corona, similar to that
our stratified shearing box simulations show a diverse range of peri- found in global MRI-dynamo studies (Dhang et al. 2020).
odicities. We observed that mean magnetic fields and EMFs oscillate Additionally, for the first time, we ventured to calculate the time-
with a primary frequency fdyn = 0.017 (equivalent to 9 orbital dependent part of αij inspired by the periodic behaviour of αmag .
periods), similar to what was found in earlier studies (Brandenburg However, we found that the amplitudes of the time-dependent part
1
et al. 1995; Gammie 1996; Davis et al. 2010; Gressel 2010; Ryan of α−s (αij ) are much smaller than that of the time-independent
0
et al. 2017). The primary frequency is determined by the effective α−s (αij ). Therefore, we suspect that the time-independent α−s are
dispersion relation of the α-Ω dynamo (Brandenburg & Subrama- predominantly governing the dynamo action.
nian 2005; Gressel & Pessah 2015) with the α dominated by the Diffusivity coefficients ηij in our work are found to be quite differ-
time-independent (DC) value of αyy . The plausible origin of this ent from that in the earlier local studies (Brandenburg 2008; Gressel
DC value of αyy is discussed below in section 5.3. 2010; Gressel & Pessah 2015; Shi et al. 2016), with ηxx ̸= ηyy and
Additionally, we observed the presence of higher harmonics at ηyy ≈ 0. Several earlier studies (Shi et al. 2016; Zier & Springel
3fdyn , which got unnoticed in earlier MRI simulations (see section 2022) found ηyx < 0) in their unstratified and stratified MRI sim-
3.5). Unlike the mean fields and EMFs, current helicity shows peri- ulations after imposing a few constraints on the coefficients (e.g.
odicities at different frequencies 2fdyn and 4fdyn , respectively. The ηyy = ηxx , ηxy = 0 etc.) and they proposed shear-current effect
presence of the frequencies in the mean EMFs, mean fields, and (Raedler 1980; Rogachevskii & Kleeorin 2004; Squire & Bhattachar-
current helicities can be understood better if we follow the mag- jee 2016) generating poloidal fields in addition to α-effect. Recently,
netic helicity density (hb ) evolution equation (e.g.see Blackman & Mondal & Bhat (2023) carried out statistical simulations of MRI
Field (2000); Subramanian & Brandenburg (2006); Kleeorin & Ro- in an unstratified ZNF shearing box and found ηyx < 0 propos-
gachevskii (2022); Gopalakrishnan & Subramanian (2023)), ing ‘rotation- shear-current effect’ and the ‘rotation-shear-vorticity
effect’ responsible for generating the radial and vertical magnetic
1 ∂hb 1 fields, respectively. However, like some other studies (TF: Branden-
= −Ē.B̄ − η0 Chel − ∇ · FH , (30)
2 ∂t 2 burg 2008; Gressel 2010; Gressel & Pessah 2015), SPH:Wissing
where FH is the helicity flux and the component of the EMF along the et al. 2022), we find ηyx ⩾ 0, unless we impose a constraint on ηyy
mean magnetic field generates mean magnetic and associated current being a positive fraction of ηxx . If we assume ηyy = fη ηxx while
helicities. Now consider the effect of the DC term in αyy , which leads calculating the coefficients, we find negativity of ηyx is an increasing
0
to a source αyy B̄y2 in equation 30. This leads to the generation of function of the factor fη (see Fig. A1 and Appendix). However, we
magnetic and current helicities at a primary frequency, twice that of find that the quality of fit is compromised slightly and histograms of
B̄y , i.e. 2 fdyn . This current helicity can now add to the α-effect, the residual of filtered (input) and reconstructed EMFs get broader
which combined with the mean field in the dynamo equation 13, can (with higher standard deviation) with the assumption ηyy = fη ηxx .
lead to secondary EMF and mean fields components oscillating at We refer the reader to see Appendix for details.
3fdyn which in turn sources helicity components at 4fdyn and so
on. These primary and secondary frequency components, limited by
noise, are indeed seen from the analysis of our simulations. 5.3 Helicity flux and the DC α−effect
In order to understand the DC value (time-independent) of the α-
effect, we take the time average of equation 30. The term ∂hb /∂t
5.2 Dynamo coefficients, comparison with earlier studies
averages to zero, and one gets the well-known constraint (Blackman
Earlier studies calculating turbulent dynamo coefficients using the 2016; Shukurov & Subramanian 2021)
simulation data and the mean field closure (equation 16) in the local 1
(Brandenburg et al. 1995; Brandenburg 2008; Gressel 2010; Gressel ⟨Ē · B̄⟩ = −η0 ⟨Chel ⟩ − ∇ · ⟨FH ⟩, (31)
2
& Pessah 2015; Shi et al. 2016) and global (Flock et al. 2012; Hogg where ⟨⟩ indicates a time average. This shows that in the absence
& Reynolds 2018; Dhang & Sharma 2019; Dhang et al. 2020) simu- of helicity fluxes, the average EMF parallel to the mean field, re-
lations of MRI-driven accretion discs used different methods. Earlier sponsible for the generation of poloidal from the toroidal mean field,
local (Brandenburg et al. 1995; Davis et al. 2010) and most of the is resistively (or catastrophically) quenched. Of the several helicity
global (Flock et al. 2012; Hogg & Reynolds 2018) studies calculated fluxes discussed in the literature, the generative helicity fluxes as
only the “coefficient of interest” αϕϕ (α− effect) by neglecting the envisaged in Vishniac (2015) and in Gopalakrishnan & Subrama-
contributions of other terms in the mean-field closure. Many of the nian (2023) can source the DC component of Ē · B̄ without the
local studies (Brandenburg 2008; Gressel 2010; Gressel & Pessah preexistence of any mean field or initial helicities. Using equation
2015) use the linear Test Field (TF) method during the run-time to (17) of Gopalakrishnan & Subramanian (2023), with mean vorticity
calculate all the coefficients. A few local (e.g. Shi et al. 2016; Zier Ω(2 − q)ẑ and noting that αyy B̄y2 dominates Ē · B̄, we estimate
& Springel 2022; Wissing et al. 2022, the current work) and global " 
(Dhang et al. 2020) studies used direct methods to quantify dynamo 0
 Ωτ 2 ′2 ′2 λ2 ∂b′2
coefficients. However, it is important to note that while several au- αyy ≈− C1 vA + C3 v + C4 2
hc 4⟨B̄y2 ⟩ τ ∂z
thors used a linear regression method assuming few constraints on # (32)
the diffusion coefficients (namely, ηxx = ηyy ), we use the IROS ′2 ∂v
′2

method without any constrained on the coefficients. +C2 b ,


∂z
Like most of the earlier local and global studies, we find a neg-
ative αyy close to the mid-plane in the upper half-plane. However, where (C1 , C2 , C3 , C4 ) = (7/45, −203/5400, 403/8100, −1/6)
direct methods seem to capture negative signs better than TF; which and we have taken q = 3/2. Adopting estimates for the correlation

MNRAS 000, 000–000 (0000)


14 Dhang et al.
0
10 × αyy 0
(αyy ) hc mean field closure reduces to equation 15. The symmetrised coeffi-
0.06 cients in equation 33 and non-symmetrised coefficients in equation
15 are related as
0.04
α̃xx = αxx ,
0.02 α̃yy = αyy ,
1
0.00 γ̃z = (αyx − αxy ) ,
α

2 (34)
−0.02 η̃xx + κ̃xyz = ηxx ,
η̃yy − κ̃yxz = ηyy ,
−0.04
1 1
δ̃z = (ηxy − ηyx ) + (κ̃xxz + κ̃yyz ) .
−0.06 2 2
−3 −2 −1 0 1 2 3
z Therefore it is evident from equation 34 that it is impossible to
decouple a few coefficients (coefficients in the last three identities)
Figure 12. Vertical profiles of α0yy (for fc = 0.05 case) obtained from IROS as there are more unknown coefficients than independent variables
inversion and (α0yy )hc , expected from helicity flux. (B̄, Ē) and the actual diffusion coefficients (η̃ij ) might be different
from the calculated ones (ηij ).

time τ ∼ Ω−1 , correlation length λ ∼ H/2 and using the vertical


profiles of various physical variables
 from the simulation, we calcu-
0
late the vertical profile of αyy hc
due to the generative helicity flux.
This is shown as a solid line in Fig. 12 and for comparison, we also
0 6 SUMMARY
show 10αyy (for fc = 0.05 case) from the IROS inversion. It is en-
0
couraging that the (αyy )hc predicted by the generative helicity flux We carried out stratified zero net flux (ZNF) simulations of MRI and
is negative in the north and has a qualitatively similar vertical profile characterised the MRI-driven dynamo using the language of mean
as that determined from IROS inversion. The amplitude, however, is field dynamo theory. The turbulent dynamo coefficients in the mean-
larger, which perhaps indicates the importance also of the neglected field closure are calculated using the mean magnetic fields and EMFs
diffusive and advective helicity fluxes which act as sink terms in obtained from the shearing box simulation. For this purpose, we used
equation 31. a cleaning (or inversion) algorithm, namely IROS, adapted to extract
the dynamo coefficients. We verified the reliability of extracted co-
efficients by reconstructing the EMFs and reproducing the cyclic
5.4 Vanishing ηyy , missing information? pattern in mean magnetic fields by running a 1D dynamo model.
In section 4.4 we pointed out that wind carries away the mean mag- Here we list the key findings of our work:
netic field and acts as the effective sink of its energy. However, the
• We find mean fields and EMFs oscillate with a primary fre-
poloidal field is also expected to be diffused by ηyy , and a positive
quency fdyn = 0.017 (≈ 9 orbital period). Additionally, they have
ηyy is required for diffusion. Instead, we find a vanishingly small
higher harmonics at 3fdyn . Current helicity αmag has two frequen-
(in some regions even negative) ηyy , which leads us to two possible
cies 2fdyn and 4fdyn respectively. These frequencies can be un-
thoughts: either it is impossible to recover ηyy in the direct meth-
derstood from mean-field dynamo effective dispersion relation and
ods, or there is incompleteness in the closure we used to retrieve the
helicity density evolution equation, respectively (for details, see sec-
coefficients. Here we discuss both possibilities.
tion 5.1).
It is clear from equation 16 that the turbulent diffusion coefficients
• Our unbiased inversion and subsequent analysis show that an
are associated with the currents, which are calculated by taking the z-
α−effect (αyy ) is predominantly responsible for the generation of
derivative of mean magnetic field components. Calculating derivative
poloidal field (here B̄x ) from the toroidal field (B̄y ). The differential
makes the currents noisy, especially Jy , as it involves a derivative of
rotation creates a toroidal field from the poloidal field completing
B̄x , which is fairly incoherent over space and time, as can be seen
the cycle; indicating that an α − Ω-type dynamo is operative in
from the butterfly diagram of B̄x (Fig. 2). Additionally, also note
MRI-driven accretion disc.
that the y-component of EMF is also noisy. Thus, the coefficients
• We find encouraging evidence that the effective DC α−effect
associated with Jy and Ēy turned out to be error-prone and difficult
can be due to a generative helicity flux (section 5.3).
to calculate. This pattern has been noticed by earlier works (Squire
• We find strong wind (v̄z ) and turbulent pumping (γz ) carry out
& Bhattacharjee 2016), which used direct methods other than IROS,
mean fields away from the mid-plane. Interestingly, they act as the
used in the current work.
principal sink terms in the mean magnetic energy evolution equation
In general, mean EMF can be expressed in terms of symmetric,
instead of the turbulent diffusivity terms.
anti-symmetric tensors and mean fields as follows,
• The unbiased inversion finds an almost vanishing ηyy , while ηxx
  sym and ηyx are positive. However ηyx and ηyy are strongly correlated;
Ē = α̃◦ B̄+ γ̃ × B̄− η̃ ◦ ∇ × B̄ − δ̃ × ∇ × B̄ − κ̃◦ ∇B̄ , if one imposes an arbitrary prior that ηyy = fη ηxx , then one finds
(33) an increasingly negative ηyx for increasing fη .
where we neglect the higher than first-order spatial derivatives and • We point out that defining mean fields by planar averaging can
time derivatives of mean fields (Raedler 1980; Brandenburg & Sub- necessarily introduce degeneracy in determining all the turbulent dy-
ramanian 2005; Schrinner et al. 2007; Simard et al. 2016). If we namo coefficients uniquely. This may have important consequences
define mean fields and EMFs as the x − y-averaged quantities, then for the physical interpretation of the dynamo coefficients.

MNRAS 000, 000–000 (0000)


MRI driven dynamo 15
ACKNOWLEDGEMENTS Bodo G., Cattaneo F., Mignone A., Rossi P., 2014, ApJ, 787, L13
Brandenburg A., 2008, Astronomische Nachrichten, 329, 725
We thank Prateek Sharma, Oliver Gressel and Dipankar Bhattacharya Brandenburg A., Donner K. J., 1997, MNRAS, 288, L29
for valuable discussions on numerical set-up, dynamo and IROS. All Brandenburg A., Schmitt D., 1998, A&A, 338, L55
the simulations are run using the Computing facility at IUCAA. Brandenburg A., Subramanian K., 2005, Phys. Rep., 417, 1
Brandenburg A., Nordlund A., Stein R. F., Torkelsson U., 1995, ApJ, 446,
741
Brandenburg A., Rädler K. H., Rheinhardt M., Käpylä P. J., 2008, ApJ, 676,
DATA AVAILABILITY
740
The data underlying this article will be shared on reasonable request Chandrasekhar S., 1960, Proceedings of the National Academy of Sciences,
to the corresponding author. 46, 253
Coleman M. S. B., Yerger E., Blaes O., Salvesen G., Hirose S., 2017, MNRAS,
467, 2625
Davis S. W., Stone J. M., Pessah M. E., 2010, ApJ, 713, 52
APPENDIX A: Dynamo coefficients with constraints on ηyy Dhang P., Sharma P., 2019, MNRAS, 482, 848
Dhang P., Bendre A., Sharma P., Subramanian K., 2020, MNRAS, 494, 4854
Diffusivities are challenging to calculate in any direct methods (SVD,
Dhang P., Bai X.-N., White C. J., 2023, ApJ, 944, 182
linear regression, IROS), as they involve the spatial derivatives of the Flock M., Dzyurkevich N., Klahr H., Turner N., Henning T., 2012, ApJ, 744,
mean fields. Primarily, we find that ηyy and ηxy are noisy as they 144
are related to spatial derivatives of B̄x , which is itself quite noisy Fromang S., Papaloizou J., 2007, A&A, 476, 1113
(e.g. see butterfly diagram in Fig. 2. Some earlier studies (Squire & Gammie C. F., 1996, ApJ, 457, 355
Bhattacharjee 2016; Shi et al. 2016) put constraints on calculating Gardiner T. A., Stone J. M., 2005, Journal of Computational Physics, 205,
η-s, try to alleviate this issue. E.g. Shi et al. (2016) imposed the 509
constraint that ηyy = ηxx in the shearing box simulation of MRI and Goldreich P., Lynden-Bell D., 1965, MNRAS, 130, 125
found a negative ηyx , implying the presence of a shear-current effect. Gopalakrishnan K., Subramanian K., 2023, ApJ, 943, 66
We have, on the other hand, done an unbiased inversion, as it Gressel O., 2010, MNRAS, 405, 41
Gressel O., Pessah M. E., 2015, ApJ, 810, 59
is not clear if such constraints are actually obeyed by MRI-driven
Gressel O., Pessah M. E., 2022, ApJ, 928, 118
turbulence. Nevertheless, for completeness, we explore here a more
Guan X., Gammie C. F., 2011, ApJ, 728, 130
generalized constraint on ηyy , given by ηyy = fη ηxx , and calculate Guan X., Gammie C. F., Simon J. B., Johnson B. M., 2009, ApJ, 694, 1010
only those coefficients that appear in the mean-field closure for Ēy , Hammersley A., Ponman T., Skinner G., 1992, Nuclear Instruments and
as those related to Ēx remain unaffected. Fig. A1 shows the vertical Methods in Physics Research A, 311, 585
0 0
profiles of αyx , αyy , ηyx and ηyy for different values of fη and for Hawley J. F., 2001, ApJ, 554, 534
fc = 0.05. The coefficients αij remain almost unaffected, while ηij Hawley J. F., Gammie C. F., Balbus S. A., 1995, ApJ, 440, 742
change significantly with change in fη . There is a clear trend that the Hawley J. F., Richers S. A., Guan X., Krolik J. H., 2013, ApJ, 772, 102
more positive the ηyy (or larger the imposed fη ), the more negative Hirose S., Blaes O., Krolik J. H., Coleman M. S. B., Sano T., 2014, ApJ, 787,
is ηyx . This implies a clear correlation between ηyx and ηyy . 1
Further, we investigate the histograms of the residual EMFs δ Ēi Högbom J. A., 1974, A&AS, 15, 417
Hogg J. D., Reynolds C. S., 2016, ApJ, 826, 40
to check the goodness of the fits. The x−components of the residual
Hogg J. D., Reynolds C. S., 2018, ApJ, 861, 24
EMFs remain unaffected as expected, while histograms for δ Ēy get
Johansen A., Youdin A., Klahr H., 2009, ApJ, 697, 1269
slightly broader with the increase in fη . This implies that the impo- Kleeorin N., Rogachevskii I., 2022, MNRAS, 515, 5437
sition of constraints on ηyy compromises the quality of fits, but not Krause F., Raedler K. H., 1980, Mean-field magnetohydrodynamics and dy-
greatly because αij are the significant contributors in the fitting of namo theory
EMFs, not the ηij . Lesur G., Ogilvie G. I., 2008, A&A, 488, 451
Mattia G., Fendt C., 2022, ApJ, 935, 22
Mignone A., Bodo G., Massaglia S., Matsakos T., Tesileanu O., Zanni C.,
Ferrari A., 2007, ApJS, 170, 228
REFERENCES
Miyoshi T., Kusano K., 2005, Journal of Computational Physics, 208, 315
Bai X.-N., Stone J. M., 2013, The Astrophysical Journal, 767, 30 Mondal T., Bhat P., 2023, arXiv e-prints, p. arXiv:2307.01281
Balbus S. A., Hawley J. F., 1991, ApJ, 376, 214 Nauman F., Blackman E. G., 2015, MNRAS, 446, 2102
Balbus S. A., Hawley J. F., 1992, The Astrophysical Journal, 392, 662 Parkin E. R., Bicknell G. V., 2013, MNRAS, 435, 2281
Balbus S. A., Hawley J. F., 1998, Reviews of Modern Physics, 70, 1 Pessah M. E., Chan C.-K., Psaltis D., 2006, Physical Review Letters, 97,
Beckwith K., Armitage P. J., Simon J. B., 2011, MNRAS, 416, 361 221103
Begelman M. C., Armitage P. J., 2023, MNRAS, 521, 5952 Pessah M. E., Chan C.-k., Psaltis D., 2007, ApJ, 668, L51
Bendre A. B., Subramanian K., Elstner D., Gressel O., 2020, MNRAS, 491, Pouquet A., Frisch U., Leorat J., 1976, Journal of Fluid Mechanics, 77, 321
3870 Raedler K. H., 1980, Astronomische Nachrichten, 301, 101
Bendre A. B., Schober J., Dhang P., Subramanian K., 2023, arXiv e-prints, Riols A., Rincon F., Cossu C., Lesur G., Longaretti P. Y., Ogilvie G. I., Herault
p. arXiv:2308.00059 J., 2013, Journal of Fluid Mechanics, 731, 1
Bhat P., Ebrahimi F., Blackman E. G., 2016, Monthly Notices of the Royal Rogachevskii I., Kleeorin N., 2004, Phys. Rev. E, 70, 046310
Astronomical Society, 462, 818 Ryan B. R., Gammie C. F., Fromang S., Kestener P., 2017, ApJ, 840, 6
Blackman E. G., 2016, in Balogh A., Bykov A., Eastwood J., Kaastra J., Salvesen G., Simon J. B., Armitage P. J., Begelman M. C., 2016, MNRAS,
eds, , Vol. 51, Multi-scale Structure Formation and Dynamics in Cosmic 457, 857
Plasmas. pp 59–91, doi:10.1007/978-1-4939-3547-5_3 Schekochihin A. A., Haugen N. E. L., Brandenburg A., Cowley S. C., Maron
Blackman E. G., Brandenburg A., 2002, ApJ, 579, 359 J. L., McWilliams J. C., 2005, ApJ, 625, L115
Blackman E. G., Field G. B., 2000, ApJ, 534, 984 Schrinner M., Rädler K.-H., Schmitt D., Rheinhardt M., Christensen U. R.,
Blackman E. G., Tan J. C., 2004, Ap&SS, 292, 395 2007, Geophysical and Astrophysical Fluid Dynamics, 101, 81
Bodo G., Cattaneo F., Ferrari A., Mignone A., Rossi P., 2011, ApJ, 739, 82 Shi J., Krolik J. H., Hirose S., 2010, ApJ, 708, 1716

MNRAS 000, 000–000 (0000)


16 Dhang et al.

fη = 1.0 fη = 0.5 fη = 0.2 fη = 0.0


6
1.0
3
/10−2

/10−2
0.5
0 0.0
αyx

αyy
0
0

−3 −0.5
−1.0
−6
−3 −2 −1 0 1 2 3 1.5 −3 −2 −1 0 1 2 3
0.2
ηyx /10−2

ηyy /10−2
0.1 1.0
0.0
−0.1 0.5

−0.2
0.0
−3 −2 −1 0 1 2 3 −3 −2 −1 0 1 2 3
z z
Figure A1. Vertical profiles of the time-independent αij and ηij related to the Ēy calculated imposing the constraint ηyy = ηxx for fc = 0.05 case. As
expected, αij are not affected with the change in fη , but ηij s are. A clear trend has been found; more positive the ηyy is more negative is ηyx .

ηyy = 0 z =0−H z = H − 2H z = 2H − 3H z =0−H z = H − 2H z = 2H − 3H


µ = −0.01, σ = 0.42 µ = −0.02, σ = 0.37 µ = 0.01, σ = 0.37 µ = −0.05, σ = 0.39 µ = −0.06, σ = 0.39 µ = 0.08, σ = 0.55
16.0% 16.0% 16.0% 16.0% 16.0% 16.0%
Percentage

12.0% 12.0% 12.0% 12.0% 12.0% 12.0%

8.0% 8.0% 8.0% 8.0% 8.0% 8.0%

4.0% 4.0% 4.0% 4.0% 4.0% 4.0%

0.0% 0.0% 0.0% 0.0% 0.0% 0.0%


−1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0
δEx /|Ex | δEx /|Ex | δEx /|Ex | δEy /|Ey | δEy /|Ey | δEy /|Ey |

ηxx = ηyy z =0−H z = H − 2H z = 2H − 3H z =0−H z = H − 2H z = 2H − 3H


µ = −0.01, σ = 0.42 µ = −0.02, σ = 0.37 µ = 0.01, σ = 0.37 µ = −0.01, σ = 0.44 µ = −0.02, σ = 0.45 µ = 0.04, σ = 0.50
16.0% 16.0% 16.0% 16.0% 16.0% 16.0%
Percentage

12.0% 12.0% 12.0% 12.0% 12.0% 12.0%

8.0% 8.0% 8.0% 8.0% 8.0% 8.0%

4.0% 4.0% 4.0% 4.0% 4.0% 4.0%

0.0% 0.0% 0.0% 0.0% 0.0% 0.0%


−1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0 −1.0 −0.5 0.0 0.5 1.0
δEx /|Ex | δEx /|Ex | δEx /|Ex | δEy /|Ey | δEy /|Ey | δEy /|Ey |

Figure A2. Histograms of the residual EMFs δ Ēx (top panels) and δ Ēy (bottom panels) for feta = 0 and fη = 1. We consider the fc = 0.05 case. Imposition
of constraints on ηyy compromises the quality of fits, but not significantly because αij are the main contributors in the fitting of EMFs, not the ηij .

Shi J.-M., Stone J. M., Huang C. X., 2016, MNRAS, 456, 2273 e-prints, p. arXiv:2305.03318
Shukurov A., Subramanian K., 2021, Astrophysical Magnetic Fields: From Velikhov E., 1959, Sov. Phys. JETP, 36, 995
Galaxies to the Early Universe. Cambridge Astrophysics, Cambridge Vishniac E. T., 2015, in American Astronomical Society Meeting Abstracts
University Press, doi:10.1017/9781139046657 #225. p. 229.08
Simard C., Charbonneau P., Dubé C., 2016, Advances in Space Research, 58, Wissing R., Shen S., Wadsley J., Quinn T., 2022, A&A, 659, A91
1522 Zier O., Springel V., 2022, MNRAS, 517, 2639
Simon J. B., Hawley J. F., Beckwith K., 2009, ApJ, 690, 974 von Rekowski B., Brandenburg A., Dobler W., Dobler W., Shukurov A., 2003,
Simon J. B., Beckwith K., Armitage P. J., 2012, MNRAS, 422, 2685 A&A, 398, 825
Squire J., Bhattacharjee A., 2016, Journal of Plasma Physics, 82, 535820201
Stepanovs D., Fendt C., Sheikhnezami S., 2014, ApJ, 796, 29
Stone J. M., Pringle J. E., Begelman M. C., 1999, MNRAS, 310, 1002
Subramanian K., 2002, Bulletin of the Astronomical Society of India, 30, 715
Subramanian K., Brandenburg A., 2006, ApJL, 648, L71
Tharakkal D., Shukurov A., Gent F. A., Sarson G. R., Snodin A., 2023, arXiv

MNRAS 000, 000–000 (0000)

You might also like