<<

Astronomy & Astrophysics manuscript no. main ©ESO 2021 September 8, 2021

Corona and XUV emission modelling of the and Sun-like Munehito Shoda1 and Shinsuke Takasao2

1 National Astronomical Observatory of Japan, National Institutes of Natural Sciences, 2-21-1 Osawa, Mitaka, Tokyo, 181-8588, Japan e-mail: [email protected] 2 Department of Earth and Space Science, Graduate School of Science, Osaka University, Toyonaka, Osaka 560-0043, Japan

Received month dd, yyyy; accepted month dd, yyyy

ABSTRACT

The X-ray and extreme- (EUV) emissions from the low-mass stars significantly affect the evolution of the planetary at- mosphere. However, it is, observationally difficult to constrain the stellar high-energy emission because of the strong interstellar extinction of EUV photons. In this study, we simulate the XUV (X-ray+EUV) emission from the Sun-like stars by extending the solar coronal heating model that self-consistently solves, with sufficiently high resolution, the surface-to-coronal energy transport, turbulent coronal heating, and coronal thermal response by conduction and radiation. The simulations are performed with a range of loop lengths and magnetic filling factors at the stellar surface. With the solar parameters, the model reproduces the observed so- lar XUV spectrum below the Lyman edge, thus validating its capability of predicting the XUV spectra of other Sun-like stars. The model also reproduces the observed nearly-linear relation between the unsigned magnetic flux and the X-ray luminosity. From the simulation runs with various loop lengths and filling factors, we also find a scaling relation, namely log LEUV = 9.93 + 0.67 log LX, where LEUV and LX are the luminosity in the EUV (100 Å < λ ≤ 912 Å) and X-ray (5 Å < λ ≤ 100 Å) range, respectively, in cgs. By assuming a power–law relation between the Rossby number and the magnetic filling factor, we reproduce the renowned relation be- tween the Rossby number and the X-ray luminosity. We also propose an analytical description of the energy injected into the corona, which, in combination with the conventional Rosner–Tucker–Vaiana scaling law, semi-analytically explains the simulation results. This study refines the concepts of solar and stellar coronal heating and derives a theoretical relation for estimating the hidden stellar EUV luminosity from X-ray observations. Key words. Sun: corona – Stars: coronae – Ultraviolet: stars – X-rays: stars

1. Introduction netic field of the (Pevtsov et al. 2003; Vidotto et al. 2014; Kochukhov et al. 2020), stellar EUV emissions are poorly char- The Sun has an aura of hot plasma called the corona, which has acterised because they are difficult to observe. EUV photons suf- a temperature of a few million Kelvin (Edlén 1943). The image fer from strong absorption by the interstellar medium (Rumph of the corona has been captured by the Atmospheric Imaging et al. 1994), for which stellar EUV spectra are observable only Assembly (Lemen et al. 2012) of the NASA’s Solar Dynamics for nearby stars in the limited range of wavelength (≤ 360 Å, Observatory (Pesnell et al. 2012). The corona is not unique to the Ribas et al. 2005; Johnstone et al. 2021). One thus needs to in- Sun and has been observed to surround low-mass main-sequence directly estimate or reconstruct the EUV spectrum of a star from stars in general (Güdel et al. 1997). Due to its high temperature, other observables. The proposed reconstruction methods include the corona is the principal source of stellar XUV (X-ray + EUV: the inversion of the differential emission measure from UV extreme ultraviolet) emissions. and/or X-ray observations (Sanz-Forcada et al. 2011; Duvvuri Stellar XUV emissions drive the expansion and thermal et al. 2021), the empirical correlation between other observable escape of planetary atmospheres (Vidal-Madjar et al. 2003; lines and XUV emission (Linsky et al. 2014; Youngblood et al. Lecavelier des Etangs et al. 2012; Owen & Wu 2013; Ehren- 2017; France et al. 2018; Sreejith et al. 2020), and/or a combi- reich et al. 2015; Airapetian et al. 2017). Therefore, describing nation of them (Diamond-Lowe et al. 2021). However, the em- the stellar XUV emission in terms of the stellar fundamental pa- pirical estimations of stellar EUV emission are based on obser- rameters (luminosity, mass, radius, etc.) is essential in under- vations with large uncertainty and a limited number of samples, arXiv:2106.08915v2 [astro-ph.SR] 7 Sep 2021 standing the evolution of a planet and its habitability. Since stel- and therefore require further validation from different perspec- lar XUV emissions decrease over time (Güdel et al. 1997; Ribas tives. et al. 2005; Telleschi et al. 2005; Claire et al. 2012; Guinan et al. To circumvent the intrinsic difficulty of stellar EUV obser- 2016) presumably in response to the stellar spin-down (Kraft vation, this study uses the solar-stellar connection to theoreti- 1967; Skumanich 1972; Barnes 2003; Irwin & Bouvier 2009; cally estimate the stellar XUV emission. The solar-stellar theo- Matt et al. 2015) caused by magnetised stellar wind (Weber & retical connection is a natural strategy to predict the stellar prop- Davis 1967; Kawaler 1988; Shoda et al. 2020), the long-term erties. For example, by extending the solar coronal theory, Shi- evolution of the XUV activity of the host star needs to be de- bata & Yokoyama(2002) modelled the X-ray characteristics of scribed as well (Tu et al. 2015; Johnstone et al. 2021). the stellar coronae, which was later extended by Takasao et al. While the observational characteristics of stellar X-ray emis- (2020) considering the size distribution of active regions. In this sions have been established, including their correlations with the study, by extending the solar atmospheric model, the structure rotation (Pallavicini et al. 1981; Wright et al. 2011) and mag- of the upper and the XUV emission are pre-

Article number, page 1 of 25 A&A proofs: manuscript no. main dicted. To this end, the coronal heating problem must be explic- notation meaning itly solved. To solve the coronal heating problem, the following three r radial distance from the stellar centre issues must be addressed (Klimchuk 2006). 1. Energy generation and transfer: the source of the coronal s coordinate along the flux-tube axis thermal energy probably comes from the magneto- G gravitational constant convection on the surface (Steiner et al. 1998). The energy generation and transfer to the corona, possibly in the form k Boltzmann constant of Alfvén waves (Alfvén 1947; Osterbrock 1961; Kudoh & B Shibata 1999; De Pontieu et al. 2007; McIntosh et al. 2011; mH hydrogen mass Srivastava et al. 2017), needs to be solved. 2. Energy dissipation in the corona: the magnetic (and kinetic) me electron mass energy in the corona needs to dissipate to sustain the high-temperature corona. The field-braiding process must be h Planck constant considered as a promising mechanism of magnetic-energy dissipation (Parker 1972, 1988). M solar mass 3. Thermal response to the heating: the density and temper- R solar radius ature of a are determined by the energy bal- ance among heating, conduction and radiation. The Rosner- T solar effective temperature Tucker-Vaiana (RTV) scaling law (Rosner et al. 1978) origi- nates from the thermal response to coronal heating. Thus, the Ω solar angular rotation rate RTV scaling law or its generalised form (Serio et al. 1981; Zhuleku et al. 2020) should be reproduced by the model (An- Table 1. Notations of the coordinates (r and s) and constant parameters. tolin & Shibata 2010). Note that subscript denotes the solar fundamental parameter. For these issues, a model of the coronal heating should 1. in- clude the (stellar surface) and , 2. appropriately consider the field-braiding process in the corona, XUV emission that includes a significant contribution from the and 3. implement the thermal conduction and radiative cooling. transition region. Classically, numerical models of a corona have often focused Considering the difficulty of the TR problem, we perform on the thermal responses to heating events by one-dimensional a series of 1D, high-resolution magnetohydrodynamic (MHD) (1D) (expanding) flux-tube models (Antiochos & Sturrock 1978; simulations for a wide range of parameters. This 1D model fa- Peres et al. 1982; Antiochos et al. 1999). A zero-dimensional cilitates a coronal loop simulation with sufficiently high numeri- model of the coronal thermal evolution is also proposed (Klim- cal resolution. The field-braiding process (or turbulence), which chuk et al. 2008). The advancement of numerical techniques and is essentially 3D, must be appropriately modelled when solving increase in computational power have made models highly so- the coronal heating problem by 1D simulation. In this study, an phisticated. Several models deal with the realistic energy gen- approximated formulation of turbulent dissipation, developed in eration by explicitly solving the magneto-convection (Hansteen previous studies (Dmitruk et al. 2002; Shoda et al. 2018), is em- et al. 2015; Rempel 2017). Other models have focused more ployed as it is likely to reproduce the average heating rate of the on energy dissipation with simpler numerical settings (Moriyasu coronal loop (van Ballegooijen et al. 2011). For simplicity, we et al. 2004; Rappazzo et al. 2008; van Ballegooijen et al. 2011; focus on the Sun-like stars that exhibit solar mass, luminosity, Dahlburg et al. 2016). These models have predicted that the radius, and metallicity. signature of coronal heating could be explained by convection- The rest of the manuscript is structured as follows. In Sec- driven energy injection. By extending these studies, we aimed tion2, the coronal model and numerical methods are detailed. to construct a stellar coronal model that would satisfy the three Section3 produces the numerical results, focusing on the depen- requirements. dencies of coronal properties on the loop length and magnetic In deriving the X-ray and EUV spectra from simulations, filling factor that are likely to vary with the star (Reale & Micela care needs to be taken in the spatial resolution at the transition 1998; Reale et al. 2004; Reiners et al. 2009; See et al. 2019). The region (the temperature-jump region between the chromosphere XUV spectra obtained from the simulations are also presented. and the corona). The numerical resolution around the transition In Section4, analytical arguments on the energy flux injected region is found to significantly affect the coronal density (Brad- into the corona are presented. In Section5, the interpretations shaw & Cargill 2013), and the coronal emission measure dis- and limitations of the proposed model are discussed. Section6 tribution (EMD). Because the coronal emission is proportional summarises the study. Several details are provided in the Ap- to the EMD, it means that the XUV emission predicted by sim- pendix, including the resolution dependence of the model (see ulation significantly depends on the numerical resolution. The AppendixB). required resolution is in the order of km or less, which is im- practical in realistic three-dimensional (3D) simulations. A pre- vious study attempted to solve this “transition-region problem” 2. Model by introducing the artificial broadening of the transition region 2.1. Model overview and notation by tuning the magnitude of radiative cooling and thermal con- duction (Johnston et al. 2017; Johnston & Bradshaw 2019; Iijima As mentioned earlier, the aim of this study is to model the XUV & Imada 2021; Johnston et al. 2021). However, this treatment (X-ray+EUV) emissions from the Sun and Sun-like stars (in- may yield an unrealistic EMD in the transition-region temper- cluding young Sun) with a range of magnetic activity level. To ature, and therefore is inappropriate for the calculation of the this end, the luminosity, mass, radius, and metallicity of the stars

Article number, page 2 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

2010), solar and stellar coronal heating (Moriyasu et al. 2004; Washinoue & Suzuki 2019), and acceleration (Suzuki & Inutsuka 2005; Shoda et al. 2018). Hereinafter, the two perpendicular directions shall be de- noted by x and y. Thus, the local xy plane is perpendicular to the axis of the loop. The flux-tube expansion is incorporated through the scale factors in the x and y directions: hx,y. For simplicity, the loop is assumed to expand isotropically in the perpendicular directions. In terms of scale factors, the isotropic expansion is represented by p hx = hy ∝ A(s), (2)

where A(s) is the cross section of the coronal loop. As hs = 1 by definition, Eq. (2) results in 1 ∂ ∇ · X = (X A(s)) , A(s) ∂s s 1 ∂  p  ∇ × X = √ Xx A(s) ey (3) A(s) ∂s 1 ∂  p  − √ Xy A(s) ex A(s) ∂s Fig. 1. A schematic picture of the system. One-dimensional dynamics along the axis of the closed flux tube is simulated. The axis is indicated for any vector field X, where ex,y represent the unit vectors in by the dotted line, while the flux tube surface is denoted by green lines. the x, y directions. The 1D spherical coordinate system is repro- The flux tube is intended to be nearly vertical and super-radially ex- duced as a special case of A(s) = s2. panding. The geometry of the loop is defined by the filling factor f and The flux tube expands in the chromosphere as a response to radial distance r as functions of field-aligned distance s. the exponential decrease in the ambient gas pressure (Cranmer & van Ballegooijen 2005; Ishikawa et al. 2021). As a result of are fixed to the solar value, i.e., this expansion, the filling factor of the magnetic field (flux tube) f should increase nearly exponentially with altitude. Under this L = L , M = M , R = R , Z = Z . (1) assumption, we model the filling factor f as " !# Dependence on metallicity is considered in the radiative loss r − R f = min 1, f∗ exp , Hmag = cmagH∗, (4) function (Section 2.5) and spectrum calculation (Section 3.1). Hmag We study the dependence of coronal properties on the coronal loop length and the filling factor of magnetic fields. Because the where f∗ is the magnetic filling factor on the photosphere and filling factor tends to increase with the stellar rotation rate (or 2 the Rossby number, Saar 2001; Reiners et al. 2009), the coronal kBT R H∗ = (5) dependence on the stellar rotation rate is implicitly investigated. GM mH We model a single coronal loop rooted in the stellar surface, is the pressure scale height at the photosphere. By this formula- and self-consistently solve the energetics and dynamics inside tion, we assume that the loop expands only in the chromosphere the loop. In other words, we solve the time-dependent, 1D MHD and exhibits a uniform cross section in the corona, which is sup- equations for an expanding coronal loop. The differential emis- ported by some solar observations (Klimchuk et al. 1992, but sion measure (DEM) of a single coronal loop is directly obtained see a recent discussion by Malanushenko et al.(2021)). Given from the simulation, which is then converted to the XUV spec- that the pressure scale height is uniform from the photosphere trum by prescribing the chemical composition and integrating up to the chromosphere, c = 2 yields a flux-tube expansion the continuum and line emissions as a function of wavelength mag with a constant plasma beta in altitude. In this work, we set using the CHIANTI atomic database version 10.0 (Dere et al. c = 2.5 that realizes a slightly low-beta chromosphere. We 1997; Del Zanna et al. 2021). mag have confirmed that the choice of cmag does not have a signifi- Notations of the coordinates and constant parameters used in cant influence over the simulation results. this study are listed in Table1. X denotes the solar value and X∗ the value of X measured at the photosphere. When the coronal loop extends to a region far above the sur- face, the flux tube also undergoes the radial expansion ∝ r2, where r is the radial distance from the stellar centre. Consid- 2.2. Model of the closed flux tube ering the chromospheric and radial expansions, the cross section A is expressed as A single coronal loop is modelled by a 1D expanding flux tube rooted in the photosphere. Figure1 illustrates a schematic of the A ∝ r2 f. (6) model. As shown in Table1, the coordinate along the axis of the flux tube is denoted by s. The spatial variation only along We define the inclination of the flux tube by prescribing r as the loop is assumed to be nonzero, that is, ∂/∂s , 0. Similar 1D a function of s. We consider nearly vertical flux tubes, so that flux tube models were used in the models of solar spicules (Holl- the vertical line of sight is nearly aligned with the axis of the weg et al. 1982; Kudoh & Shibata 1999; Matsumoto & Shibata flux tube (Figure1). It is easier to calculate the DEM along the

Article number, page 3 of 25 A&A proofs: manuscript no. main

1.0 eint denotes the internal energy per unit volume and is defined in Section 2.4. The source term S is given by  0   !  0.8  1 GM dr   p + ρv2 /L − ρ   ⊥ 2   2 r ds   1  B B    − s x v  0.6  ρvsvx + + ρDx   2L 4π ! 

loop    1 BsBy v   −ρvsvy + + ρD  S =  2L 4π y  , (12) z/l 0.4    1 p b   (vsBx − vxBs) + 4πρD   2L x   1     p b   vsBy − vyBs + 4πρDy  0.2  2L   GM∗ dr   −ρv + Q + Q  s r2 ds cnd rad

0.0 −1  2  0.0 0.2 0.4 0.6 0.8 1.0 where L = d/ds ln r f denotes the length scale of the flux- tube expansion. The conduction term Qcnd is defined in terms of x/lloop conductive flux qcnd as

1 ∂  2  Fig. 2. Shape of the flux-tube axis defined by Eq. (7). x and z denote Qcnd = − qcndr f . (13) the horizontal and vertical coordinates, repectively. r2 f ∂s Because the mean free path of an electron is generally smaller than the system size, the Spitzer–Härm flux (Spitzer & Härm 1953) is applied to qcnd: vertical line of sight for this structure. In particular, we set |Bs| 5/2 ∂T    qcnd = − κSHT , (14) 10 lloop − s |B| ∂s dr   = tanh   , r|s=0 = R (7) ds  lloop  −6 −1 −1 −7/2 where κSH = 10 erg cm s K . The radiative cooling Qrad v,b and turbulent dissipation Dx,y are described in Section 2.5 and where lloop is the half-loop length. The actual shape of the flux- tube axis is displayed in Figure2. Combining Eq.s (4)–(7), the 2.6, respectively. Because turbulence is considered not as an ex- cross section A(s) is well defined as a function of s. ternal force but as a dissipative process, the losses of the kinetic and magnetic energies by turbulence are locally balanced by the gain in the internal energy. Thus, the presence of the turbulence terms does not affect the conservation of the total energy. For the 2.3. Basic equations same reason, we do not explicitly consider the numerical dissi- pation of velocity and magnetic field in the energy equation. We The MHD equations with the equation of state of partially note that the numerical dissipation is unlikely to be the dominant ionised hydrogen, gravity, thermal conduction, radiative cooling, heating mechanism because the coronal Alfvén wave, which has and phenomenology of turbulent heating are selected as the ba- a typical wavelength of ∼ 100 Mm, is resolved by a sufficiently sic equations of the model, which are expressed in the form of fine grid in the corona (100 km). conservation law (for derivation, see AppendixA) ∂ 1 ∂   2.4. Equation of state U + Fr2 f = S. (8) ∂t r2 f ∂s We assume that the plasma consists of neutral hydrogen atoms, protons, and electrons. The internal energy per unit volume e The conserved variables U and the corresponding fluxes F are int is composed of the conventional thermal energy p/(γ − 1) and given by the latent heat of the ionised gas.      ρ   ρvs  p    2  eint = + nHχIH, nH = ρ/mH, (15)  ρvs   ρvs + pT  γ − 1    −   ρvx   ρvsvx BsBx/(4π)   ρv   −  U =  y  , F =  ρvsvy BsBy/(4π)  , (9) where χ is the ionisation degree and nH is the number density of  B   v B − v B  hydrogen atoms (proton + neutral hydrogen). I is the ionisation  x   s x x s  H  B   v B − v B  energy of the hydrogen atom (I = 13.6 eV). We assume that the  y   s y y s  H e (e + pT ) vs − Bs (v⊥ · B⊥) /(4π) ionisation degree could be determined from the approximated version of the Saha-Boltzmann equation, in which only the where ground state is considered as the bound state (low-temperature limit). v⊥ = vxex + vyey, B⊥ = Bxex + Byey, (10) 2 ! 2 2 χ 2 IH B⊥ 1 2 B⊥ − , pT = p + , e = eint + ρv + . (11) = 3 exp (16) 8π 2 8π 1 − χ nHλe kBT

Article number, page 4 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

] 21 1 10− mosphere exhibits isothermal behaviour in the absence of other −

s cooling/heating mechanisms.

3 The optically thin cooling is expressed in terms of the radia- 22 10− tive loss function Λ(T) by nenHΛ(T). In this work, we define the loss function over a wide range of temperature (103 K ≤ T ≤ 107 K) as follows. 23 10− 1. For simplicity, in the high-temperature range (T ≥ 1.5 × 104 K), the cooling rate is referred from the CHIANTI 24 10− atomic database with the photospheric abundance (no first ionisation potential (FIP) effect). CHINATI loss function 25 2. The cooling rate in the low-temperature range (T ≤ 1.0 × 10− 4 Λeff (T ) 10 K) is deduced by Goodman & Judge(2012), which par- Λ(T ) tially consider the non-LTE effect. 4 26 3. In the intermediate-temperature range (1.0 × 10 K < T <

radiative loss10 function [erg cm − 104 105 106 107 1.5 × 104 K), a bridging law between two loss functions asre T [K] employed following the method of Iijima(2016).

Fig. 3. Solid and dashed lines show the effective and original optically- The chromospheric heating effect by backward coronal radiation thin radiative loss functions Λeff (T) and Λ(T), respectively. Also shown is still missing in Λ(T) defined above. To introduce this effect, by crosses are the loss function from the CHIANTI atomic database we quench Λ(T) in the chromospheric temperature range, which thin with photospheric abundance. gives Qrad  T 2  thin  chr  Q = nenHΛeff(T), Λeff(T) = Λ(T) exp −  , (22) where λe is the thermal de Broglie wavelength of electron. rad  T 2  s 2 4 h where Tchr = 2.0 × 10 K. The effective radiative loss function λe = . (17) Λ (T) is displayed in Figure3, along with the original radiative 2πmekBT eff loss function Λ(T) and the CHIANTI loss function defined in When chromospheric hydrogen is no longer in thermal equi- T ≥ 104 K. librium, the ionisation degree will deviate from the Saha– Boltzmann value (Goodman & Judge 2012), which is beyond 2.6. Phenomenological model of coronal turbulence the scope of this study. Once the ionisation degree χ is obtained, the pressure and temperature are related by Although the mechanism of the coronal heating is debated, it is of no doubt that magnetic field feeds heat to the corona that p = (ne + nH) kBT = (1 + χ) nHkBT. (18) maintains the million-Kelvin temperature by dissipation. The formation of tangential discontinuities (electric current sheets) in 2.5. Radiation response to the continuous shuffling of the foot points of coro- nal magnetic fields is a plausible mechanism of magnetic-field The radiative cooling rate per unit volume Qrad is given by dissipation (Parker 1972; Sturrock & Uchida 1981; Parker 1983; van Ballegooijen 1986; Galsgaard & Nordlund 1996). The ubiq- thck thin Qrad = ξradQrad + (1 − ξrad) Qrad , (19) uitous current-sheet formation should lead to small-scale impul- ! p sive energy release, which is likely to feed a sufficient amount ξrad = 1 − exp − , prad/p∗ = 0.1, (20) of energy to the corona, possibly in the form of micro- and prad nano-flares (Parker 1988; Shimizu 1995; Aschwanden & Parnell 2002). where p is the pressure at the surface (photosphere) and Qthck ∗ rad The formation of current sheets can be interpreted as turbu- and Qthin approximate the optically thick and thin cooling rates, rad lent cascading (Rappazzo et al. 2007, 2008; Verdini et al. 2012). respectively. The optically thick and thin functions are seam- Because the coronal loop is threaded by a strong mean mag- lessly connected via ξrad. Instead of solving the radiative transfer, netic field, the MHD turbulence evolving in the coronal loop can we model Qthck and Qthin as follows. rad rad be accurately approximated by the reduced-MHD turbulence, The radiative heating and cooling are approximately bal- in which the energy cascades preferentially in the perpendicu- anced near the photosphere to maintain a nearly constant sur- lar direction (Shebalin et al. 1983; Cho & Vishniac 2000; Cho face temperature. The optically thick radiative loss is approxi- & Lazarian 2003). The energy-cascading (or heating) rate of mated by an exponential cooling function that forces the local the reduced-MHD turbulence, Qheat, is precisely approximated temperature to approach the reference value (Gudiksen & Nord- by the mean-field quantities as follows (Hossain et al. 1995; lund 2005): Matthaeus et al. 1999; Dmitruk et al. 2002; Verdini & Velli 2007) !−1/2 1   ρ + − 2 − + 2 thck ref − z⊥z⊥ + z⊥z⊥ Qrad = eint eint , τ = 0.1 s , (21) Qheat ≈ cdρ , (23) τ ρ∗ 4λ⊥ ref ± where ρ∗ is the photospheric mass density and eint is the refer- where z⊥ denotes the (root-mean-squared (RMS)) amplitude of ence internal energy density corresponding to the reference tem- the perpendicular Elsässer variables and λ⊥ is the correlation ref ref perature T . We simply assume T = T , i.e., the stellar at- length of the Elsässer variable (Alfvén wave) perpendicular to

Article number, page 5 of 25 A&A proofs: manuscript no. main the mean field. cd is a dimensionless parameter. By this approx- The photospheric magnetic field is known to form localised kilo- imation, we estimate the averaged heating rate using the coronal Gauss patches (Spruit & Zweibel 1979; Tsuneta et al. 2008). turbulence (field braiding). These patches are likely to be in thermal equipartition, which The approximated heating rate in Eq. (23) is implemented by equates the gas and magnetic pressures. For the non-magnetised v b photosphere, the thermal equipartition field is given by adding the source terms Dx,y, Dx,y given by Shoda et al.(2018). B = 1.34 × 103 G. (33) v cd  + − − +  eq Dx,y = − zx,y zx,y + zx,y zx,y , (24) 4λ⊥ In the magnetised photosphere, as the deeper region tends to b cd  + − − +  Dx,y = − zx,y zx,y − zx,y zx,y , (25) be observed (e.g. Keller et al. 2004), the ambient gas is likely 4λ⊥ to exhibit larger pressure than the non-magnetised photosphere. ± p Thus, the equipartition magnetic field should be larger than this where z = vx,y ∓ Bx,y/ 4πρ. The role of these terms is ex- x,y   equipartition value. Therefore, we set the photospheric axial plained below. Without the conservation part ∝ ∂ r2 f F /∂s, the magnetic field equal to perpendicular components of the equation of motion and induc- 3 tion equation are expressed as Bs,∗ = 1.5Beq = 2.01 × 10 G. (34)

∂   v ∂ p b ρvx,y = ρDx,y, Bx,y = 4πρDx,y, (26) ∂t ∂t We model the energy injection from the photosphere by im- In the limit of the reduced-MHD approximation (time- posing the velocity and magnetic-field fluctuations. Fluctuations independent density, ∂ρ/∂t = 0), Eqs. (24), (25), and (26) are are imposed at both ends of the simulation domain. The verti- reduced to cal and horizontal velocity fluctuations are modelled separately. The upward acoustic waves are excited on the photosphere by ∂ ± cd ∓ ± employing the time-dependent boundary conditions on the den- zx,y = − zx,y zx,y, (27) ∂t 2λ⊥ sity and axial velocity. ! which yields the energy conservation law of vs,∗ ρ∗ = ρ∗ 1 + , (35) 2 2 a∗ ∂ X z− z+ + z+ z− i i i i Z ωl e⊥ = −cdρ , (28) max h i ∂t 4λ l i=x,y ⊥ vs,∗ = dω v˜s (ω) sin ωt + ψ (ω) , (36) l ωmin where e⊥ is the sum of the kinetic and magnetic energies emerg- ing from the fluctuations of the perpendicular components: √where ρ∗ is the time-averaged photospheric density, a∗ = kBT∗/mH is the isothermal speed of sound on the photosphere, 1   1 B2 and ψl (ω) is a random phase function that ranges between 0 and e = ρ z+ 2 + z− 2 = ρv2 + ⊥ . (29) ⊥ 4 ⊥ ⊥ 2 ⊥ 8π 2π. In the numerical implementation of the integral in Eq. (36), the frequency range is evenly divided into 21 bins and the corre- Comparing Eq.s (23) and (28), one obtain sponding 21 components are summed. The time-averaged pho- tospheric density is given by equipartition on the photosphere: ∂ ≈ − e⊥ Qheat, (30) 2 ∂t Bs,∗ ρ∗kBT∗/mH = , (37) indicating that the energy dissipation by the reduced-MHD tur- 8π bulence is considered appropriately. which yields The perpendicular correlation length is assumed to increase −7 −3 with the flux-tube radius, i.e., ρ∗ = 4.22 × 10 g cm . (38) r s ρ is larger than the typical mass density on the solar surface be- A r f ∗ λ⊥ = λ⊥,∗ = λ⊥,∗ , (31) cause the magnetised photosphere should be deeper and denser. A∗ R f∗ The (time-averaged) Alfvén speed vA on the photosphere is then expressed as where the perpendicular correlation length at the photosphere is set equal to the typical width of the inter-granular lane: λ⊥,∗ = Bs,∗ −1 vA,∗ ≈ = 8.73 km s . (39) 150 km. However, the best possible free parameter cd is still de- p 4πρ∗ bated. In this study, we infer cd = 0.1 from the previous studies of the solar-wind turbulence (van Ballegooijen & Asgari-Targhi We arbitrarily setv ˜ (ω) ∝ ω−1/2 with 2017; Chandran & Perez 2019; Verdini et al. 2019). s l l 2π/ωmin = 300 s, 2π/ωmax = 100 s. (40) 2.7. Boundary condition and simulation setting The minimum frequency corresponds to the cut-off frequency of the acoustic wave at the photosphere (e.g. Felipe et al. 2018). Both boundaries of the simulation domain are located at the pho- The magnitude ofv ˜s (ω) is tuned such that the RMS amplitude tosphere, and the photospheric temperature is fixed to the effec- −1 of vs,∗ at the photosphere is 0.6 km s . tive temperature, i.e., q 3 2 −1 T∗ = T = 5.77 × 10 K. (32) vs,∗ = 0.6 km s , (41)

Article number, page 6 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars where the overline denotes the time average. Although the magnetic filling factor half-loop length longitudinal-wave excitation on the photosphere is explicitly (photosphere) [103 km] considered, the effect of the longitudinal-wave input is insignif- icant; the coronal temperature decreases by only 2% when the f∗ = 1 lloop = [20, 30, 40] longitudinal wave injection is terminated. f∗ = 0.5 l = [20, 30, 40] The horizontal velocity and magnetic field at the bottom loop boundary are expressed in terms of the Elsässer variables, which f∗ = 0.333 lloop = [20, 30, 40, 60] are defined as f∗ = 0.2 lloop = [20, 30, 40] ± Bx,y zx,y = vx,y ∓ p . (42) 4πρ f∗ = 0.1 lloop = [20, 30, 40, 60, 80]

The free boundary condition is imposed on the downward El- f∗ = 0.05 lloop = [20, 30, 40] sässer variables. lloop = [20, 30, 40, 60, f∗ = 0.0333 ∂ − 80, 120, 160, 240] zx,y = 0. (43) ∂s ∗ l = [20, 30, 40, 60, 80, 120, f = 0.01 = f loop The upward Elsässer variable is assumed to be non- ∗ 160, 240, 320, 480, 640] monochromatic with respect to the frequency. f∗ = 0.005 lloop = [20] t Z ωmax + ± h t i zx,y,∗ = dω z˜x,y (ω) sin ωt + ψ (ω) , (44) t ωmin Table 2. List of the simulation runs conducted in this study. The mag- netic filling factor on the solar photosphere is set to f = 0.01 (Cranmer where ψt(ω) is a random phase function that ranges between 0 2017). and 2π. As with Eq. (36), the frequency range is discretised into 21 bins when numerically implementing the integral in Eq. (44). ± −1/2 We setz ˜x,y (ω) ∝ ω with it reaches the maximum ∆smax. In particular, in 0 ≤ s ≤ lloop, the t t 2π/ωmin = 1000 s, 2π/ωmax = 100 s. (45) size of the i-th cell, ∆si, is iteratively defined as " " ## ± −1/2 2ε   z˜x,y (ω) ∝ ω corresponds to the 1/ω energy spectrum dis- ge ∆si = max ∆smin, min ∆smax, ∆smin + si−1 − sge , covered in previous solar simulations and observations (Van 2 + εge Kooten & Cranmer 2017). Given that the typical size of a gran- 1 ule is 1, 000 km and the typical speed of surface convection is si = si−1 + (∆si−1 + ∆si) , (48) 1 km s−1 (Chitta et al. 2012), the maximum wave period corre- 2 sponds to the turn-over time of granular motion. Similarly, given Letting N be the total number of cells, we express the cell size that the typical size of an inter-granular lane is 100 km, the min- in the latter half of the domain lloop ≤ s ≤ 2lloop by imum wave period corresponds to the turn-over time of inter- ± 1 granular motion. The magnitude ofz ˜x,y (ω) is tuned such that + ∆si = ∆sN+1−i, si = si−1 + (∆si−1 + ∆si) , (49) the root-mean-squared amplitude of zx,y,∗ at the photosphere is 2 1.2 km s−1: The maximum cell size is fixed to ∆smax = 100 km. The relation q q between the minimum cell size and f∗ is + 2 + 2 −1 zx,∗ = zy,∗ = 1.2 km s , (46) ∆smin = 5 km ( f∗ < 0.05), ∆smin = 2 km ( f∗ ≥ 0.05). (50) where the overline denotes the time average. Although some so- lar observations have observed the suppression of convective ve- This relation is used because a higher resolution is required locity in large-filling-factor regions (e. g. Katsukawa & Tsuneta at the transition region in the large- f∗ runs (for details, see 2005), we dismiss this effect for simplicity. AppendixB). The grid expansion rate also depends on f∗ as ≥ By this formulation, the energy flux of the upward Alfvén εge = 2.13 ( f∗ < 0.05) and εge = 1.89 ( f∗ 0.05). The grid wave at the footpoint of the flux tube is given by expansion height is fixed to sge = 10, 000 km, which is greater than the typical height of the transition region that needs to be 1   1   resolved with the minimum cell size. F = ρ z2 + z2 v ≈ ρ z2 + z2 v A,∗ 4 x,∗ y,∗ A,∗ 4 ∗ x,∗ y,∗ A,∗ In numerically solving Eq. (8), we rewrite the basic equa- = 2.65 × 109 erg cm−2 s−1, (47) tions in terms of the cross-section-weighted conserved variables U˜ and the corresponding fluxes F˜ defined by which is sufficiently larger than the energy flux required to sus-  2  tain the solar corona (Withbroe & Noyes 1977).  ρ˜   ρr f     2   ρv   ρvrr f   ˜˜r       ρv r2 f   ρ˜v˜x   x  2.8. Numerical method ˜    2  U =  ρ˜v˜y  =  ρvyr f  , (51)  ˜   p   Bx   Bxr f  A non-uniform grid system is used to resolve the computational  ˜   p   By   Byr f  domain 0 ≤ s ≤ 2lloop. A uniform cell size of ∆smin is used below     e˜  er2 f  the critical height s < sge, above which the cell size expands until

Article number, page 7 of 25 A&A proofs: manuscript no. main    ρ˜v˜r  thermal conduction, which reduce the numerical cost and time  2   ρ˜v˜r + p˜T  with minimum loss of accuracy.  ˜ ˜   ρ˜v˜rv˜x − Br Bx/(4π)   ˜ ˜  F˜ =  ρ˜v˜rv˜y − Br By/(4π)  , (52)  ˜ ˜  3. Simulation result  v˜r Bx − v˜xBr   ˜ ˜   v˜r By − v˜yBr  3.1. Fiducial (solar) case: atmosphere and spectrum     (e˜ + p˜T ) v˜r − B˜ r v˜⊥ · B˜ ⊥ /(4π) First, we discuss the simulation run with lloop = 20 Mm and −2 where f∗ = 1.0 × 10 as the fiducial case. In this case, the coronal field strength is ≈ 20 G, which is within the range of the solar ˜ 2 B⊥ coronal magnetic field strength measured by the coronal seis- p˜ = p˜ + = p r2 f. (53) T 8π T mology technique (Nakariakov & Ofman 2001; Verwichte et al. 2004; Jess et al. 2016), and thus the fiducial case is regarded as Using U˜ and F˜ , the basic equation is given by the solar case. ∂ ∂ Figure4 illustrates the time-averaged properties of a quasi- U˜ + F˜ = S˜, (54) steady coronal loop. Panels show the mass density (top) and tem- ∂t ∂s perature (middle) along the loop axis and the differential emis- where sion measure (DEM, bottom), defined by    0  2 dllos  !  DEM(T) = ne (T) , (56)  1 2 GM∗ dr  dT  p˜ + ρ˜v˜⊥ /L − ρ˜   2 r2 ds   !  where ne(T) is the electron density with temperature T and llos is  B˜ B˜   1 r x v  the length along the line of sight. Practically, dividing the tem-  −ρ˜v˜rv˜x + + ρ˜Dx   2L 4π  perature range into bins and considering the vertical line of sight  !  S˜ =  1 B˜ r B˜ y  . (55) (l = r), the DEM and associated emission measure distribution  −ρ˜v˜ v˜ + + ρ˜Dv  los  r y y  (EMD) are numerically obtained as follows  2L 4π   p b   4πρ˜Dx  2  p  DEM(Ti)∆Ti ≡ EMD(Ti) = n (Ti) ∆r(Ti), (57)  4πρ˜Db  e  y   GM∗ dr   −ρv Q r2 f Q r2 f  where ne(Ti) is the total number density of an electron that ex- ˜˜r 2 + cnd + rad r ds hibits a temperature in [Ti − ∆Ti/2, Ti + ∆Ti/2] and ∆r(Ti) is the total radial extension of where Ti − ∆Ti/2 ≤ T < Ti + ∆Ti/2. With this variable conversion, any MHD solver designed for the The DEM is calculated in the temperature range of 104 K ≤ T ≤ Cartesian coordinate system can be directly applied to Eq. (54). 107 K with an equal spacing in the logarithmic scale of T. The In this study, the Harten–Lax–van Leer discontinuities (HLLD) DEM in the range of T < 104 K is not calculated because, in approximated Riemann solver (Miyoshi & Kusano 2005) is used the low-temperature range, the atmosphere is not optically thin F˜ to calculate at the cell boundary. For spatial reconstruction, the and the DEM loses its meaning. Note that the unit of EMD is fifth-order accurate monotonicity-preserving method (Suresh & cm−5, while the “volume” EMD, which has often been used in Huynh 1997) is used to reconstruct the cross-section-weighted the literature (Güdel et al. 2003; Scelsi et al. 2005), represents U˜ s ≤ s l − s ≤ s conserved variables in ge and 2 loop ge, whereas the the distribution of an emission measure over the whole coronal monotonic upstream-centred scheme for the law of conservation volume and has a different unit (cm−3). (van Leer 1979) with a minmod flux limiter is used in s < s < ge The high-temperature (T > 106 K) corona is successfully re- 2l − s . loop ge produced in the fiducial case. Since we are imposing the phe- Thermal conduction and the other parts are solved indepen- nomenological, mean-field formulation of the coronal heating dently by the second-order operator-splitting procedure as fol- (field braiding/turbulence), the heating tends to be more con- lows. stant in time and more uniform in space than the actual three- dimensional case in which the heating is intermittent in time and 1. thermal conduction is solved for a half step ∆t/2: space. Nevertheless, the time-averaged heating rate should be n ∆t/2 ∗ U˜ −−−→ U˜ (thermal conduction only) similar between the 1D approximation and the 3D simulation, because the previous 3D simulation of solar wind yielded a simi- 2. the rest of the basic equations are solved for a full step ∆t: lar mean field to the 1D simulation with turbulence phenomenol- ∗ ∆t ∗∗ U˜ −→ U˜ (without thermal conduction) ogy (Shoda et al. 2019). i i Under the assumption that the upper atmosphere (T ≥ 104 K) is optically thin in the wavelength of interest, the XUV 3. thermal conduction is solved again for a half step ∆t/2: spectrum is obtained from the EMD using the open-source pack- ˜ ∗∗ ∆t/2 ˜ n+1 Ui −−−→ Ui (thermal conduction only) age ChiantiPy based on CHIANTI database ver 10.0 (Del Zanna et al. 2021). In particular, the specific intensity I was calcu- n λ where U˜ is the n-th step value of U˜ . With this procedure, we lated from the EMD using the ChiantiPy.core.Spectrum module avoid the severe constraints on the ∆t from thermal conduction with the coronal abundance given by Schmelz et al.(2012). Al- when updating the MHD equations. though the radiative loss function Λ(T) is constructed with pho- The third-order strong-stability-preserving (SSP) Runge– tospheric abundance, because the loss function is nearly inde- Kutta method is used in the time integration of the MHD equa- pendent of the FIP effect in the radiation-dominated temperature tions (Shu & Osher 1988; Gottlieb et al. 2001). The super-time- range T ≤ 3 × 105 K, the inconsistency in abundance do not vi- stepping method (Meyer et al. 2012, 2014) is used to solve the olate the simulation results. Assuming that the is

Article number, page 8 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

6 2 10− 10 ] time average 1 reconstructed from simulated DEM − 101 observed solar spectrum () 10 8 snapshot ˚ A − 1

− 100 s ] 2 3 10 −

− 1 10− 10−

2 10−

[g cm 12 10−

ρ 3 10− 14 10− 4 10−

5 16 10− 10− Lyman edge 0 10 20 30 40 flux at 1 au [erg cm 6 10− s [Mm] 0 200 400 600 800 1000 1200 ˚ 107 wavelength [A] TR TR Fig. 5. Observed (blue) and simulated (red) spectral flux density of the 106 Sun measured at 1 au. The observed spectrum is retrieved in the solar activity minimum. The two spectra are in a good agreement below the Lyman edge (≤ 912 Å).

[K] 105 T & Lightman 1979)

104 R 2 F = πI , (58) λ λ r

103 which yields the X-ray and EUV luminosities as 0 10 20 30 40 Z 100 Å Z 912 Å 2 2 s [Mm] LX = 4πr dλ Fλ, LEUV = 4πr dλ Fλ. (59) 5 Å 100 Å 1029 In terms of energy, X-ray photons are in the range of 0.12 − 2.48 keV. Caution must be exercised when calibrating the X- 27 ] 10 rays because a subtle difference in the bandpass of the instrument 1

− can result in large differences in the derived response function K 25 and X-ray luminosity (Zhuleku et al. 2020). For the procedure of 5 10

− translation between different instruments, see Judge et al.(2003). To test the capability of our model in the prediction of XUV 1023 spectrum, we compare in Figure5 the spectral flux density ob- tained from our simulation (red) and that from the observation in a solar activity minimum (blue). The observed spectrum is DEM [cm 21 10 obtained from the coordinated observation in the Whole Helio- sphere Interval (WHI, from March 20, 2008 to April 16, 2008 1019 Woods et al. 2009; Chamberlin et al. 2009). Figure5 shows that, 104 105 106 107 below the Lyman edge (≤ 912 Å), the simulated spectrum is in T [K] a good agreement with the observed spectrum. Above the Ly- man edge (≥ 912 Å), the continuum is underestimated, and the emission lines are overestimated in the simulated spectrum, pos- Fig. 4. Simulated loop properties of the fiducial case. Top: time- sibly because the optically thin approximation is inadequate in averaged (solid line) and snapshot (dashed line) profiles of mass den- sity along the loop axis. Middle: time-averaged (solid line) and snapshot this wavelength range. In this study, because the focus is on the (dashed line) profiles of temperature along the loop axis. Bottom: time- spectrum below the Lyman edge, the simulation is validated with averaged differential emission measure (solid line). Annotations “TR” respect to spectrum prediction. in the middle panel indicate the locations of the transition region in the snapshot profile. 3.2. Loop-length dependence: Density and temperature

The coronal loop length is a fundamental parameter that affects the coronal density and temperature. Here, we show the relation a uniformly bright sphere, the spectral flux density at the helio- between the time-averaged coronal properties and the coronal centric distance r is deduced by (see, e.g., Section 1.3 of Rybicki loop length. For simplicity, we fix the magnetic filling factor to

Article number, page 9 of 25 A&A proofs: manuscript no. main

101 Another factor that should be considered is the gravitational ]

2 stratification. The effective half-loop length often exceeds the − coronal pressure scale height. In such cases, the loop-top pres- sure in the original RTV scaling law is replaced by the coronal- base pressure (Serio et al. 1981). Given that the pressure is con- tinuous across the transition region, we define the coronal-base

[dyne cm 5 pressure pbase as the pressure measured at Tave = 10 K. For 0 the coronal electron density, both the loop-top value ne,top and base 10 p coronal-base value ne,base are measured. In contrast to pressure, the density is discontinuous across the transition region, and Ttop K] therefore the definition of the coronal-base density is not triv-

6 pbase ial. Here, we define ne,base as the value measured in the coronal eff 0.39 [10 l base: ∝ loop eff 0.25 Z scor,2 top lloop 1

T ∝ ne,base ≡ ne,ave ds, (61) 1 − 10− scor,2 scor,1 scor,1 101 102 103 leff [Mm] where we set scor,1 = 10 Mm and scor,2 = 15 Mm. The loop-top loop density and temperature are given by

10     10 ne,top = ne s = lloop , Ttop = T s = lloop . (62) ne,top ne,base Figure6 shows the leff -dependencies of the time-averaged 0.49 loop ] eff −

3 l ∝ loop coronal properties (density, temperature and pressure). The loop- − leff 0.10 top temperature and coronal-base pressure obey a power–law re- ∝ loop

[cm lation with respect to the effective half-loop length, which is for- mulated as 9 base 10 e, ∝ eff 0.39 n Ttop lloop , (63) 0.25 p ∝ leff . (64) top base loop e,

n The (generalised) RTV scaling law predicts that the loop-top temperature obeys the following relation

8  1/3 10 RTV ∝ eff 101 102 103 Ttop pbaselloop , (65) eff lloop [Mm] where we dismiss the exponential correction term as it is negligi- ble (Serio et al. 1981). A comparison of Eqs. (63), (64), and (65) reveals that the simulation results are consistent with the RTV Fig. 6. Top: relation between the effective half loop length leff (see loop scaling law. Eq. (60) for definition) and the time-averaged loop-top temperature (T , red circles) and coronal-base pressure (p , blue diamonds). Bot- An alternative form of the RTV scaling law predicts a rela- top base F leff tom: relation between the effective half loop length leff and the time- tion among the coronal energy flux cor, loop length loop, and loop RTV averaged loop-top electron density (ne,top, red circles) and coronal-base loop-top temperature Ttop , and is expressed as electron density (ne,base, blue diamonds). In both panels, lines represent the power-law fittings to the symbols. RTV  eff 2/7 Ttop ∝ Fcorlloop . (66) A comparison of Eqs. (63) and (66) indicates that the energy flux the fiducial (solar) value: f = f = 0.01. Hereinafter, the time- injected into the corona is larger for longer loops. In terms of the ∗ heating rate per unit volume Q, the RTV predictions, averaged value of X will be denoted by Xave. In the discussion on the behaviour of the coronal properties, RTV 2/7 eff 4/7 Ttop ∝ Q lloop , (67) the results must be compared with the analytical RTV scaling RTV 6/7 eff 5/7 law (Rosner et al. 1978). Note that the half-loop length in the pbase ∝ Q lloop , (68) RTV scaling law denotes the length from the transition region to the apex of the loop, whereas l stands for the length from and simulation results from Eqs. (63) and (64) indicate that loop eff eff the stellar surface to the apex of the loop. For better comparison, Q = Fcor/lloop is a decreasing function of lloop. These conclu- eff sions shall be directly validated in the following section. instead of lloop, we use the effective half-loop length lloop, which denotes the coronal length, and is defined as The bottom panel of Figure6 shows the variations in loop- top and coronal-base electron densities over leff . Given that the Z l loop loop RTV scaling law predicts a larger loop-top density at a higher leff = ds, (60) loop loop-top temperature for a uniform corona, the decrease in the sTR loop-top density is attributed to gravitational stratification. Note 5 where sTR (< lloop) is where Tave = 10 K. that the coronal-base density exhibits a weaker dependence on

Article number, page 10 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

107 ference in the position of the coronal base yields a significant er- 6 Fcnd measured at Tave = 10 K ror or uncertainty in Fcor. Therefore, instead of directly measur-

Fcnd averaged in [10Mm, 15Mm] ing Fcor, we measure the backward conductive flux Fcnd, which 0.51 should be balanced with F by energy conservation. ] cor lloop 1 ∝ − The top and bottom panels in Figure7 show the loop-length s eff 2 (lloop and lloop) dependence of the coronal conductive flux. The − factor 4 difference panels display the simulation runs with a fixed magnetic filling −2 106 factor of f∗ = 1.00×10 . The red circles and blue diamonds rep- 6 resent the conductive flux measured at Tave = 1.0×10 K and av-

[erg cm eraged over 10 Mm ≤ s ≤ 15 Mm, respectively. The black solid lines represent the power-law fittings to the blue diamonds. The cnd

F coronal conductive flux increases with the loop length. How- ever, the trend deviates from a simple power law. When the eff 1.5 loop length is sufficiently small (lloop . 10 Mm) or large . 5 eff 2 5 10 (lloop & 10 Mm), the conductive flux weakly depends on the 101.0 101.5 102.0 102.5 103.0 eff 2.0 loop length. The energy flux increases around lloop ∼ 10 Mm, lloop [Mm] with the minimum and maximum values differing by a factor of 4. 107 An approximate power-law fit to the blue diamonds yields 6 Fcnd measured at Tave = 10 K the following scaling law Fcnd averaged in [10Mm, 15Mm] 0.48 0.48 eff ] eff Fcnd ∝ l , (69)

1 l loop ∝ loop − s which, in combination with the RTV scaling law, predicts 2 −  2/7 factor 4 difference RTV ∝ eff ∝ eff 0.42 106 Ttop Fcorlloop lloop . (70)

eff [erg cm The simulated dependence of Ttop on lloop, Eq. (63), is repro- duced by the semi-analytical arguments. Thus, the results shall cnd

F be explained semi-analytically once the theoretical behaviour of Fcor (or equivalently Fcnd) has been derived. In Section4, we propose a simple model to produce Fcor. 105 The enhanced energy injection to the corona produces en- 101.0 101.5 102.0 102.5 103.0 hanced XUV emissions. Figure8 depicts the loop-length depen- leff [Mm] dence of the predicted XUV luminosity. For each coronal loop loop calculation, the XUV spectrum is calculated as in Figure5 and converted to XUV luminosity through Eq. (59). Both LX and Fig. 7. Top: half loop length versus coronal conductive flux measured LEUV exhibit increasing trends as those found in the coronal en- 6 at Tave = 10 K (red circles) and averaged in 10 Mm ≤ s ≤ 15 Mm ergy flux Fcnd. In particular, LX exhibits this trend significantly. (blue diamonds). Also shown by the solid line is the power-law fitting The inferred power–law relations are to the blue diamonds. Bottom: same as the top panel but with respect to eff 0.29 the effective half loop length lloop. eff LEUV ∝ lloop , (71) eff 0.65 LX ∝ lloop . (72) eff lloop than the coronal-base pressure, which is contradictory if eff eff the coronal-base temperature is constant in lloop. It may be in- The dependence of LX on lloop is further explained semi- eff analytically. The X-ray luminosity should be proportional to the terpreted that the coronal-base temperature increases with lloop eff emission measure of the corona, i.e., in response to the increasing Ttop with lloop. 2 eff LX ∝ ncorlloop, (73) 3.3. Loop-length dependence: Energy flux and XUV where ncor is the typical number density of the coronal loop that emission lies between nbase and ntop. Assuming that ncor follows the pre- For a better interpretation of the behaviour of coronal properties, diction of the RTV scaling law, the variation in the energy flux entering the corona Fcor with the RTV2 eff 4/7 eff −3/7 eff ncor ∝ Ttop /lloop ∝ Fcor lloop , (74) effective half-loop length lloop must be revealed. Directly measuring F from the simulation data is, how- cor the X-ray luminosity is connected to the coronal energy flux Fcor ever, not a trivial task. To obtain Fcor, one needs to measure the and loop length leff by energy flux at the base of the corona, which moves in time and loop is broadened after time averaging. Because a significant amount 1/7 0.69 L ∝ F 8/7leff ∝ leff , of energy is reflected back at the transition region, a slight dif- X cor loop loop (75)

Article number, page 11 of 25 A&A proofs: manuscript no. main

1029 107.0 ] 1 − 1028 106.5 [erg s [K] EUV top T L 27 10 LX 106.0 X LEUV L l 0.70 ∝ loop Ttop l 0.31 0.29 ∝ loop f 1026 ∝ ∗ 1.0 1.5 2.0 2.5 3.0 105.5 10 10 10 10 10 2 1 0 10− 10− 10 lloop [Mm] f ∗ 1029 1011 ] 1 − 1028

] 10

3 10 [erg s − [cm EUV

L L 1027 X top LEUV e, 9 X n 10 L leff 0.65 ∝ loop eff 0.29 ne,top lloop ∝ 0.49 1026 f 1.0 1.5 2.0 2.5 3.0 ∝ ∗ 10 10 10 10 10 108 eff 2 1 0 l [Mm] 10− 10− 10 loop f ∗

Fig. 8. (Effective) loop length versus X-ray luminosity LX (red circles) and EUV luminosity LEUV (blue diamonds). The solid and dashed lines Fig. 9. Magnetic filling factor on the photosphere f∗ versus time- show the power-law fittings to the simulation results. The top and bot- averaged loop-top temperature (Ttop, top panel) and loop-top electron tom panels show the relation with respect to the half loop length lloop density (ne,top, bottom panel). The half loop length is fixed to lloop = and the effective hal loop length leff (see Eq. (60)), respectively. 20 Mm. Red circles show the simulation results and the solid lines show loop the power-law fittings to them. where the simulated scaling law Eq. (69) is used in the second proportional relation. The actual scaling relation Eq. (72) is in pendence of the coronal properties on the filling factor is worth good agreement with the semi-analytical prediction Eq. (75), investigating as a proxy of the dependence of the stellar corona validating the RTV scaling law in describing the stellar coronal on the activity level. The simulation results with various mag- properties. netic filling factors shall be discussed below, with the half-loop l Note that EUV luminosity LEUV is weakly dependent on the length fixed to loop = 20 Mm. loop length compared with LX because a portion of the EUV Figure9 shows the filling-factor dependence of the time- photons originates from the upper chromosphere and transition averaged loop-top temperature Ttop and electron density ne,top. region, which are barely affected by the variation in the coronal The coronal density and temperature increase with an increase loop length. Within the investigated loop-length range, the ratio in the magnetic filling factor following a power–law relationship L /L decreases from ∼ 10 to ∼ 3 as the effective loop length EUV X 0.29 increases. Ttop ∝ f∗ , (76) 0.49 ne,top ∝ f∗ . (77) 3.4. Filling-factor dependence: density and temperature Although not explicitly demonstrated, the effective half-loop The magnetic filling factor of the photosphere appears to have length also weakly depends on f∗ as scaled with the Rossby number, or equivalently the magnetic ac- leff ∝ f 0.04. tivity level (Saar 1996, 2001; Reiners et al. 2009). Hence, the de- loop ∗ (78)

Article number, page 12 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

] The RTV scaling law predicts that the loop-top temperature 1 108 − Ttop depends on the coronal energy flux Fcor and effective half- s 6 Fcnd measured at Tave = 10 K eff 2 loop length lloop as Eq. (66). A comparison between Eqs. (66) − F averaged in [10 Mm, 15 Mm] cnd and (76) reveals the mostly proportional relationship between f 0.91 ∝ ∗ Fcor and f∗: 7 10 eff [erg cm Fcorlloop ∝ f∗, (82)

cnd which shall be further investigated in the following section. F

6 10 3.5. Filling-factor dependence: energy flux and XUV emission

The energetics of the corona with various magnetic filling factors shall be discussed here. The top panel of Figure 10 shows the 105 relation between the filling factor f and the coronal backward conductive flux: 2 1 0 ∗ 10− 10− 10 6 conductive flux measured at Tave = 10 K (red circles) and aver- f aged over 10 Mm ≤ s ≤ 15 Mm (blue diamonds). The two con- ∗ ductive fluxes exhibit different behaviours, with the red circles 18 appearing to saturate within the large f∗ limit. This difference 10 6 F measured at T = 106 K may arise from the different physical meanings of Tave = 10 K, cnd ave 6 where the values of the red circles are measured. Tave = 10 K

] Fcnd averaged in [10 Mm, 15 Mm] 1 0.95 is located near the loop top for the small f∗ cases and near the − 17 f s 10 ∝ ∗ transition region for the large ones. Therefore, the blue diamonds

1 may be interpreted as the coronal energy flux. − Given the dependence of the loop-top temperature and RTV 16 scaling law on the filling factor, the relation among the coronal 10 eff energy flux Fcor, effective loop length l , and filling factor f∗ is [erg cm loop predicted by Eq. (82). This prediction is confirmed in the bottom eff loop panel of Figure 10, which shows the following scaling relation: l 15 10 eff 0.95 cnd Fcorlloop ∝ f∗ . (83) F This consistency indicates that the RTV scaling law must be 1014 applicable to stellar coronae with varying activity levels. The 2 1 0 10− 10− 10 simulation results validate the previous studies on stellar coronal f energetics based on the RTV scaling law (Shibata & Yokoyama ∗ 2002; Takasao et al. 2020).

Fig. 10. Top: magnetic filling factor on the photosphere f∗ versus back- Figure 11 shows the spectra of XUV emission for differ- ward conductive flux measured in the corona. The half loop length is ent magnetic filling factors. For all wavelengths below the Ly- fixed to lloop = 20 Mm. The red circles and blue diamonds represent the man edge (912 Å), the XUV emissions increase as the magnetic 6 flux measured at Tave = 10 K and averaged in 10 Mm ≤ s ≤ 15 Mm. filling factor increases. This is a natural consequence of the in- The black solid line is the power-law fitting to the blue diamonds. Bot- creased coronal density shown in Figure9. In particular, as seen F leff tom: same as the top panel but the vertical axis denotes cnd loop. in the central panels of Figure 11, the enhancement of the en- ergy flux from f∗ = 0.01 to f∗ = 1 is centred on ∼ 10 with a variance of factor ∼ 5 in the EUV range (100 Å < λ ≤ 912 Å), whereas a significantly larger enhancement occurs in the X-ray We note that the RTV scaling law semi-analytically predicts a range (5 Å < λ ≤ 100 Å). This is because of the sensitivity of relation of the X-ray emissivity to the temperature. Also shown in the right q panels of Figure 11 are the spectra of the total photon number RTV ∝ eff ∝ 0.27 Ttop ne,toplloop f∗ , (79) emitted from the star per unit time (photon number luminosity) 2 given by 4πr λFλ/(hc). For all f∗, the spectrum of the photon which is close to the simulation result Eq. (76). Alternatively, in number luminosity is flatter than the energy spectrum, implying terms of the heating rate per unit volume Q, the RTV scaling law that the photon population with respect to wavelength is approx- coupled with Eq.s (76) and (78) yield imately uniform in the XUV range. 0.935 Q ∝ f , (80) The magnetic-flux dependencies of the X-ray luminosity LX and EUV luminosity LEUV are also investigated. Figure 12 re- while the RTV scaling law Eq.s (77) and (78) yield veals the variation in LX (blue diamonds) and LEUV (red circles) with the unsigned magnetic flux Φmag defined by Q ∝ f 0.868, (81) unsign Φmag = 4πR2 B f . (84) The two trends are mostly consistent with each other. unsign s,∗ ∗

Article number, page 13 of 25 A&A proofs: manuscript no. main

2 3 ] 40 10 10 1 10 − ] f = 0.01 f = 0.01 1 ˚ A ∗ 01 ∗ . −

f = 0.1 1 f = 0.1 ˚ A ∗ − ∗ 1 0 f = 1 = 0 38 f = 1 −

10 ∗ ∗ 10 ∗ s f

2 102 −

2 36 10− = 1 to 10 ∗ f 101 4 34 10− 10 flux at 1 au [erg cm flux ratio of 6 0 32

10− 10 photon number10 luminosity [s 0 200 400 600 800 1000 0 200 400 600 800 1000 0 200 400 600 800 1000 wavelength [A]˚ wavelength [A]˚ wavelength [A]˚

2 7 ] 38 10 10 1 10 − ] f = 0.01 f = 0.01 1 ˚ A ∗ 01 ∗ . −

f = 0.1 1 f = 0.1 ˚ A 0 ∗ − ∗ 1

10 = 0 f = 1 36 f = 1 −

∗ ∗ 10 ∗ s f

2 105 − 2 10− 34 = 1 to 10 ∗

4 f 10− 103 1032 6 10− flux at 1 au [erg cm flux ratio of 8 1 30

10− 10 photon number10 luminosity [s 100 101 102 100 101 102 100 101 102 wavelength [A]˚ wavelength [A]˚ wavelength [A]˚

Fig. 11. Left: spectral energy flux at 1 au for f∗ = 0.01 (orange, solid), f∗ = 0.1 (green, dashed), and f∗ = 1 (blue, dotted). Center: energy-flux ratio of f∗ = 1 to f∗ = 0.01. Right: photon-number luminosity spectrum for f∗ = 0.01 (orange, solid), f∗ = 0.1 (green, dashed), and f∗ = 1 (blue, dotted). Bottom panels are the same as top panels but with logarithmic x axis that emphasize the short-wavelength region. For better visualisation, spectra are averaged over 20 Å-bins in the top panels.

Because the photospheric magnetic field is fixed, the filling fac- tested by the Sun-as-a-star observations in the EUV range (e.g. mag tor f∗ is the only variable in the definition of Φunsign. We find that Toriumi et al. 2020). mag The filling-factor dependence of the EUV photon number lu- both LX and LEUV follow power laws with respect to Φ and unsign minosity ΦEUV is also investigated. ΦEUV is defined as the total are expressed as follows photon photon number of EUV photons emitted from the star, i.e., L ∝ mag 0.81, EUV Φunsign (85) Z 912 Å λFλ mag 1.19 EUV 2 L ∝ Φ , (86) Φphoton = 4πr dλ , (88) X unsign 100 Å hc which are represented by the solid and dashed lines in Fig- which acts as a proxy of the photo-ionised hydrogen number per ure 12, respectively. A scaling law is observationally discovered unit time. Figure 13 shows the EUV photon number luminosity EUV between the unsigned magnetic flux and the X-ray luminosity with varying magnetic filling factors. As with LEUV, Φphoton also (Pevtsov et al. 2003) follows a power law with respect to f∗, which is given by

mag 1.15 Lobs ∝ Φ , (87) EUV ∝ 0.78 ∝ mag 0.78 X unsign Φphoton f∗ Φunsign . (89) which is shown by the dotted line in Figure 12. Although the A comparison of Eq.s (85) and (89), reveals that LEUV and magnitudes of the X-ray luminosity from the simulation are EUV Φphoton follow marginally different power laws with respect to smaller than the observed values, possibly because the proposed f (or equivalently Φmag ). model reproduces the coronal properties in the activity mini- ∗ unsign mum (a quiet, quasi-steady corona without flaring activities), the power-law index from the simulation (1.19) is nearly identical EUV 3.6. LX–LEUV and LX–Φ relations to that from the observation (1.15). The reproduction of the ob- photon mag servational Φunsign–LX relation provides another support to our One of the objectives of this study is to propose a method to mag estimate the stellar EUV luminosity LEUV from other observ- model. The proposed Φunsign–LEUV relationship, Eq. (85), can be able quantities. Because the X-ray luminosity LX is measurable and should be correlated with LEUV (see e.g. Sanz-Forcada et al.

Article number, page 14 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

30 ]

10 1 − 39 [s 10 EUV 0.78

] Φphoton f 1 29 ∝ ∗ 10 EUV photon − [erg s 1028

EUV 38

L 10 LEUV LX X 27 10 mag 0.81 L Φ ∝ unsign Φmag 1.19 ∝ unsign relation by Pevtsov+ 1026 1023 1024 1025 1026 1027 mag 1037 Φunsign [Mx] 2 1 0 10− 10− 10 EUV photon number luminosity Φ f mag ∗ Fig. 12. Unsigned magnetic flux Φunsign versus X-ray luminosity LX (blue diamonds) and EUV luminosity LEUV (red circles). Results for the Fig. 13. Magnetic filling factor on the photosphere f∗ versus photon fixed half loop length lloop = 20 Mm are shown. Solid and dashed lines are the power-law fittings to the simulation results. Also shown by the number luminosity in the EUV range. Each simulation run corresponds dotted line is the observational relation by Pevtsov et al.(2003). to each red circle. Also shown by the black solid line shows the power- law fitting to the simulation results.

2011), the relationship between LX and LEUV over a wide range of filling factors and loop lengths warrants investigation. ported as transverse fluctuations of the velocity and magnetic The top panel of Figure 14 highlights the correlation between field, i.e., Alfvén waves, the efficiency of Alfvén-wave transmis- sion into the corona is key to the analysis. This section is devoted X-ray luminosity LX and EUV luminosity LEUV. The black cir- cles correspond to the simulation runs listed in Table2. Although to reviewing the linear theory of Alfvén-wave propagation in a the simulation runs are spread over wide ranges in the filling fac- stellar atmosphere and to provide analytical arguments for the tor and loop length, a single power-law fit adequately applies to origin of the scaling laws discovered through the simulation. the LX–LEUV relation, which is given by 4.1. Linear theory of Alfvén-wave propagation and reflection log LEUV = 9.93 + 0.67 log LX, (90) A significant amount of Alfvén waves excited at the photo- −1 where LEUV and LX are measured in the unit of erg s . The red sphere is reflected before reaching the solar corona (Cranmer points with the vertical error bars also indicate the observational & van Ballegooijen 2005; Verdini & Velli 2007). The relation- estimation by Sanz-Forcada et al.(2011). The pure theoretical ship between the Alfvén-wave transmission efficiency and the estimation of LEUV in this study conforms with the pure observa- loop properties in the linear regime remains unclear and shall be tional estimation of the same by Sanz-Forcada et al.(2011), Al- discussed in this section. though the power-law index of LX–LEUV relation is 0.86, accord- Figure 15 illustrates the schematic of the concerned system. ing to Sanz-Forcada et al.(2011), the coe fficient 0.67 in Eq. (90) The convection–magnetic field interaction excites the Alfvén is, in fact, close to the observational value (0.681) derived from waves propagating upward along the expanding flux tube. In our the Extreme Ultraviolet Explorer (EUVE) observations (John- simulation setting, the Alfvén speed is maintained at a nearly stone et al. 2021). The fact that the independent methods pro- constant value during the tube expansion (Eq.(4)). Because the duces a similar LX–LEUV relation validates the proposed model Alfvén-wave reflection is triggered by a gradient in the Alfvén of stellar coronae and their XUV emissions. speed (Stein 1971; An et al. 1989; Vasquez 1990), the Alfvén The correlation between the X-ray luminosity LX and the waves reflected in the flux-tube expansion region are feeble. EUV EUV photon-number luminosity Φphoton is also highlighted in Therefore, the bulk of the reflection occurs in a layer between the the bottom panel of Figure 14. They, too, follow a power–law flux-tube merging height (where the flux-tube expansion ends) relationship given by and the coronal base, which is in between the two vertical dashed EUV lines in Figure 15. log Φphoton = 20.40 + 0.66 log LX, (91) The Alfvén waves travelling through the reflection layer is

EUV −1 considered in the following analysis. For simplicity, the thick- where Φphoton is measured in unit of s . ness of the layer is disregarded. This approximation is validated by the fact that the reflection-layer thickness (∼ 100 Mm) is 4. Analytical arguments on coronal energy flux notably smaller than the typical wavelength of Alfvén waves therein (∼ 101−2 Mm). Then, the problem is simplified as a Several scaling relations with respect to the loop length and mag- linear-wave transmission through a discontinuity, as described netic filling factor are provided in Section3. In this section, we in the lower part of Figure 15. Hereinafter, we denote the vari- analytically describe the coronal energy injection process gov- ables in the region on the left/right of the discontinuity as region erning the scaling relations. Because energy is mainly trans- I/II, respectively.

Article number, page 15 of 25 A&A proofs: manuscript no. main

32

] 10 1 empirical estimation by Sanz-Forcada+ − simulation results 1031 log LEUV = 9.93 + 0.67 log LX [erg s 1030 EUV L 1029

1028

1027

EUV luminosity: 1026 1025 1026 1027 1028 1029 1030 1 X-ray luminosity: LX [erg s− ] ] 1 −

[s 1041 EUV log Φphoton = 20.40 + 0.66 log LX EUV photon

1040

1039

Fig. 15. A model of Alfvén-wave transmission across a free boundary 1038 1026 1027 1028 1029 1030 with density jump.

EUV photon number luminosity: Φ 1 X-ray luminosity: LX [erg s− ]

Fig. 14. Top: correlation between the X-ray luminosity LX and EUV luminosity LEUV. X-ray and EUV are defined in terms of wavelength The corresponding magnetic-field fluctuation is deduced from λ as 5 Å < λ ≤ 100 Å and 100 Å < λ ≤ 912 Å, respectively. Each the linearised induction equation, that is, black circle corresponds to the simulation run listed in Table2, and the ∂ ∂ black solid line shows the power-law fitting to them. Also shown by the δB = −B δv . (95) red points with vertical error bars are the empirical estimation by Sanz- ∂t I/II 0 ∂s I/II Forcada et al.(2011). Bottom: same as the top panel but for the X-ray EUV Because the fluctuations of the velocity and magnetic field are luminosity LX and photon number luminosity Φphoton. required to be continuous across the boundary, the amplitudes must satisfy the following relations. a + b = a + b , (96) The equation describing the Alfvén-wave propagation is I I II II − kIaI + kIbI = −kIIaII + kIIbII. (97) " # ∂2 ∂2 − v2 δv = 0, (92) Solving these two equations yield ∂t2 A,I/II ∂s2 I/II kI + kII kI − kII the solution of which is given by aI = aII + bII. (98) 2kI 2kI i(−kI/II s+ωt) i(kI/II s+ωt) δvI/II = aI/IIe + bI/IIe . (93)

The energy transmission rate TAW is defined as the ratio of Note that aI/II (bI/II) denotes the amplitudes of the upward (downward) propagating Alfvén wave in region I/II, respec- the forward-propagating Alfvén-wave energy in regions I and II. tively. Given the frequency ω, the dispersion relation yields the Namely, wavenumber k in each region. I/II ρ v |a |2 T II A,II II . AW = 2 (99) kI/II = ω/vA,I/II. (94) ρIvA,I |aI|

Article number, page 16 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

We obtain a general form of the energy transmission rate within 104 the limit of ρII/ρI  1 by combining Eqs. (98) and (99). Alfven-crossing time: τA

2 turbulent dissipation time: τdiss 4vA,I |aII| TAW = , (100)  2 ∗ ∗ 2 vA,II |aII| + aIIbII + aIIbII + |bII| 103 where X∗ denotes the complex conjugate of X. It is interesting to see that the energy transmission rate depends on the values of aII and bII, i.e., the wave population in region II (corona). The transmission rate in the two limiting cases are investi- 2 gated below. When bII = 0 (no Alfvén waves propagating to- timescale [s] 10 ward the chromosphere), the energy transmission rate is given by (Hollweg 1984) r vA,I ρII TAW = 4 = 4 . (101) 1 vA,II ρI 10 101 102 103 The other limiting case is bII = aII (forward and reflected Alfvén lloop [Mm] waves are equally abundant in the corona), where the transmis- sion rate is given by Fig. 16. Half loop length versus timescale of Alfvén-wave crossing r vA,I ρII (τA red circles) and Alfvén-wave turbulent dissipation (τdiss, blue dia- TAW = = . (102) monds). For the definitions of τA and τdiss, see Eq. (107). vA,II ρI

To summarise, the transmission rate of Alfvén wave is given The bottom panel of Figure6 shows that the mass density of by the coronal base remains mostly constant with respect to the loop r length. Then, according to Eq. (106), the change in the coronal ρII TAW = cref , 1 ≤ cref ≤ 4 (103) energy flux is attributed to the change in the value of cref. Be- ρI cause 1 ≤ cref ≤ 4, the coronal energy flux should saturate at approximately four times the basal value, as observed in Figure where the value of cref depends on the population ratio of the forward and backward Alfvén waves in region II. 7. The enhancement of the coronal energy flux at leff ≈ In the numerical model implemented in this study, the den- loop 2 sity in the region I should be the mass density at the flux-tube 10 Mm can be explained in terms of the timescales of Alfvén- merging height s = hmerge. In the flux-tube model employed in wave propagation and dissipation. The backward waves in the this model, hmerge is expressed as corona are originated from the other base of the corona. The backward waves are subject to the turbulent dissipation during ! hmerge their propagation, which results in a smaller amplitude at the re- f∗ exp = 1, Hmag = 2.5H∗, (104) flection layer than the amplitude of the forward waves. The pop- Hmag + 2 − 2 ulation ratio, z⊥ / z⊥ , should thus depends on the ratio of the where H∗ is the density scale height at the photosphere (see Eq. Alfvén-crossing time τA to the turbulent dissipation time τdiss, (5) for definition). Assuming that the density scale height is uni- which are respectively defined as form in 0 ≤ s ≤ hmerge, we can write the mass density at the Z s +2leff Z s +2leff merging height ρmerge as TR loop 1 TR loop λ τ = ds/v , τ = ⊥ ds. A A diss eff + ! sTR 2l sTR cdz⊥,rms hmerge 2/5 loop ρmerge ≈ ρ∗ exp − = ρ∗ f∗ . (105) (107) H∗ Then, using Eq. (99), we obtain the Alfvén-wave transmission Specifically, rate for our model as 1. when τA/τdiss < 1, the coronal Alfvén waves propagating r r ρcor ρcor −1/5 from one coronal base to the other hardly experience TAW ≈ cref = cref f∗ , (106) any turbulent dissipation. Therefore, the population ratio ρmerge ρ∗ + 2 − 2 z⊥ / z⊥ should tend to unity in this case. where ρ denotes the mass density at the coronal base. Note cor 2. When τA/τdiss > 1, the coronal Alfvén waves significantly that a large f∗ could lead to a small transmission rate, when ρcor dissipate during the propagation through the corona. Thus, 2/5 does not increase faster than f∗ . + 2 − 2 the population ratio z⊥ / z⊥ should be far from unity in this case. 4.2. Loop-length dependence of the coronal energy flux This argument is substantiated by Figure 16, in which τA (red In Section 3.3, it is shown that the coronal energy flux increases circles) and τdiss (blue diamonds) are plotted against the half- loop length l . The Alfvén crossing time is shorter than the with lloop but saturates at approximately four times the value of loop 2 the short-loop limit (see Figure7). This behaviour is naturally turbulent dissipation time in lloop < 10 Mm and vice versa in 2 2 explained by the theory of Alfvén-wave transmission. lloop > 10 Mm. Therefore, it is likely that, when lloop < 10 Mm,

Article number, page 17 of 25 A&A proofs: manuscript no. main

2 2 + − 3 the weak turbulent dissipation yields z⊥ / z⊥ ∼ 1 and cref ≈ 10− 2 1. In the opposite limiting case of lloop > 10 Mm, the backward Alfvén waves significantly decay in the coronal loop, resulting 2 2 in z+ / z−  1 and c ≈ 4. These arguments underscore the ⊥ ⊥ ref 10 4 importance of the timescale of turbulent dissipation in a coronal − loop.

5 4.3. Filling-factor dependence of the coronal energy flux /L 10− X

In Section 3.5, the coronal energy flux is found to follow a power L law as a function of f∗. 6 10− 0.91 simulation results Fcor ∝ f∗ , (108) relation by Pizzolato+ or alternatively relation by Pizzolato+ (scaled) 7 10− eff 0.95 2 1 0 Fcorlloop ∝ f∗ . (109) 10− 10− 10 Ro/Ro These power-law relations are explained by the combined ef- fect of the RTV scaling law and the energy transmission rate Eq.(106). In the absence of the Alfvén-wave dissipation in the Fig. 17. Rossby number versus normalized X-ray luminosity (blue dia- chromosphere, the energy flux injected into the corona is given monds). The Rossby number and the filling factor are connected by Eq. by (115). Results for the fixed half loop length lloop = 20 Mm are shown. Also shown by the solid and dashed lines are the observational relation 1/2 4/5 by Pizzolato et al.(2003) and its scaled-down version (1 /8). Fcor = FA,∗ f∗TAW ∝ crefρcor f∗ , (110) where FA,∗ is the upward Alfvén-wave energy flux at the foot- point of the flux tube, which is fixed for all the simulation runs. number (Saar 1996, 2001; Reiners et al. 2009). Given that the According to the RTV scaling law, the density and the coronal magnetic activity saturates at Ro/Ro ≈ 0.1 and f ≈ 0.01, the energy flux are related by power–law relationship is assumed as follows.  −2 −3/7 ! ρ ∝ F4/7leff .  Ro  cor cor loop (111) f∗ = min 1, f  . (115)  Ro  Eq.s (110) and (111) yield Figure 17 shows the relation between Ro/Ro given by Eq. 7/10 eff eff 28/25 1.15 (115) and L /L . The blue diamonds represent the simulation Fcorl ∝ crefl f∗ ∝ cref f∗ , (112) X loop loop results, and the black solid line represents the observational rela- where we use leff ∝ f 0.04. As addressed in the previous section, tion observed in Sun-like stars (Pizzolato et al. 2003). Although loop ∗ a gap exists between the simulation results and the observational the value of c depends on the balance between the Alfvén- ref relation, when the observational relation is scaled down by a fac- crossing timescale τ and the turbulent dissipation timescale τ A diss tor of 8 (represented by the dashed line), the simulation results in the coronal loop. Given that a larger f yields a smaller τ /τ ∗ A diff conform with the observations. Given that the observational data (as well as a smaller c ), the predicted dependence of c on f ref ref ∗ points have a large scatter (an order of magnitude in L /L ), is X and that the proposed model corresponds to the activity mini- −αref mum where the X-ray luminosity is 10–20 times smaller than cref ∝ f∗ , (113) the activity maximum (Johnstone & Güdel 2015), the simulation where αref > 0. Given the limited range of cref (1 ≤ cref ≤ 4), results adequately corroborate with the observations. the power index αref is small. Then, the semi-analytical scaling relation, 5.2. Missing physics and limitations eff ∝ 1.15−αref Fcorlloop f∗ , (114) Although our model self-consistently solves the physical pro- cesses of energy injection into the corona, turbulent coronal heat- conforms with the obtained scaling relation Eq. (109). ing, and thermal responses of the coronal loop, several other physical processes have not been considered. 5. Discussion 1. The impulsive nature of the coronal heating is disregarded. 5.1. Inferred Ro–L relation The phenomenological turbulent dissipation implemented X in this study yields the time-averaged heating rate, whereas The X-ray behaviour with respect to the stellar Rossby num- the actual coronal heating is highly intermittent in space and ber has been established in the literature (Pizzolato et al. 2003; time, as indicated by previous simulation studies (Einaudi Wright et al. 2011; Wright & Drake 2016; Wright et al. 2018; et al. 1996; Cassak et al. 2008; Dahlburg et al. 2016; Kanella Magaudda et al. 2020). By prescribing a relation between the fill- & Gudiksen 2018) and observations (Cirtain et al. 2013; ing factor f∗ and the Rossby number Ro, we compare the model Testa et al. 2014; Ishikawa et al. 2017; Antolin et al. 2021). prediction with the observational relation. The magnetic filling In the presence of localised heating, the DEM extends to factor follows a power-law relation with respect to the Rossby the high-temperature side, possibly forming the bi-modal

Article number, page 18 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

structure observed in active stellar coronae (Güdel et al. stellar population (Bochanski et al. 2010; Winters et al. 1997). The accuracy of the DEM obtained in this 1D 2019) and often host terrestrial planets within the habitable simulation must be tested in the future. zones (Bonfils et al. 2013; Dressing & Charbonneau 2015). 2. The coronal loop is assumed to be isolated from the ambient The applicability of the proposed theoretical predictions to plasma. However, previous studies have stated that the lower-mass stars (especially, M dwarfs) must be investigated interface between the coronal flux tube and its surrounding in the future. Given the lower metallicity of old M-type is a preferential region for heating by phase mixing and stars, the dependence of XUV emissions on metallicity must resonant absorption (Ionson 1978; Heyvaerts & Priest 1983; also be studied (Washinoue & Suzuki 2019). According to Hollweg & Yang 1988; Sakurai et al. 1991; Hood et al. Figure 14 in Johnstone et al.(2021), a single power law 2002; Pagano & De Moortel 2019; Van Damme et al. 2020). between LX and LEUV could be found in low-mass stars with In fact, resonant absorption is observed to occur in the solar various spectral types, and thus Eqs. (90) and (91) may be corona (Antolin et al. 2015; Okamoto et al. 2015). Although directly applicable to lower-mass stars. there is no scientific consensus on the ability of phase 2. The XUV emissions of pre-main-sequence stars significantly mixing and/or resonant absorption to feed sufficient thermal alter the evolution of proto-planetary discs and planet forma- energy to the corona, the relation between these processes tion (Gorti & Hollenbach 2009; Nakatani et al. 2018; Wang and XUV emissions must be investigated. et al. 2019). In the case of pre-main-sequence star accretion 3. Energy injection by the emergence of magnetic field from (classical T-Tauri stars), two effects of accretion on XUV the stellar interior (magnetic flux emergence, Takasao et al. emissions must be considered. First, the XUV emissions 2013) was dismissed. Recent observations of the solar from the accretion shocks could be comparable to or even corona indicate that the coronal loops are preferentially greater than those from the corona (Günther et al. 2007). heated near the bottom (Berghmans et al. 2021), possibly by Second, because the accretion to the stellar surface may ex- its interaction (magnetic reconnection) with adjacent small- cite velocity disturbances on the photosphere, the Alfvén- scale loops (Chen et al. 2021). Because the (small-scale) wave energy injected into the corona is likely to be enhanced coronal loops are formed by the magnetic flux emergence, (Cranmer 2008, 2009). A model that considers these accre- these observations imply that a fraction of the coronal tion effects is required to theoretically predict the XUV emis- energy must be supplied by the emergence. Indeed, a recent sions from pre-main-sequence stars. work on the solar wind indicates the considerable role of magnetic flux emergence in the injection of energy to the open-field regions (Wang 2020). 6. Conclusion 4. Throughout this study, the energy spectrum is reconstructed In this study, we theoretically investigate the coronal properties from EMD (DEM) under the optically thin approxima- of the Sun and Sun-like stars. The simulation domain extends tion. Although this approximation has been widely used from the solar surface to the corona, with energy transport from (Sanz-Forcada et al. 2011; Duvvuri et al. 2021), it must the surface and dissipation in the corona being solved in a self- be tested in future under a realistic treatment of radiative consistent manner. The thermal structuring of the stellar atmo- transfer. For example, the Lyman continuum appears to sphere is also reproduced by implementing thermal conduction be optically thick, with an extended source region in the and radiative cooling. The behaviours of the coronal properties upper chromosphere (Avrett & Loeser 2008). Other studies with respect to the loop length and magnetic filling factor are ex- have claimed that a simple optically thin treatment of the plained by a combination of the behaviour of the coronal energy coronal lines may produce incorrect results (Schrijver et al. flux and the RTV scaling law. With the parameter survey, we 1994). To solve this issue, the frequency-dependent radiative reproduce the nearly linear relation between the unsigned mag- transfer equation must be solved in the future. netic flux and the X-ray luminosity. Scaling relations between 5. In converting the single-loop properties retrieved from the the EUV properties (luminosity and photon number luminosity) simulation results to a global quantity such as LX, we assume and the X-ray luminosity are discovered, which will be useful that the stellar corona has a uniform brightness. The bright- in constraining the stellar EUV parameters from the observable ness of the actual stellar coronae, however, may be highly quantity LX. structured, as seen in the solar corona. Because emissions The authors thank Drs. Haruhisa Iijima, Kosuke Namekata, from a single active region depends on its size, the size distri- and Riouhei Nakatani for valuable comments. Numerical com- bution of active regions over the entire stellar surface affects putations were carried out on the Cray XC50 at the Center the measured distribution of the total coronal emissions from for Computational Astrophysics (CfCA), National Astronomi- the star (Takasao et al. 2020). The effect of coronal struc- cal Observatory of Japan. M.S. is supported by a Grant-in-Aid turing must be considered in the future by, for instance, the for Japan Society for the Promotion of Science (JSPS) Fellows global modelling of stellar coronae (van der Holst et al. 2014; and by the NINS program for cross-disciplinary study (grant Alvarado-Gómez et al. 2016; Oran et al. 2017; Airapetian Nos. 01321802 and 01311904) on Turbulence, Transport, and et al. 2021). Heating Dynamics in Laboratory and Solar/ Astrophysical Plas- mas: “SoLaBo-X.” S.T. is supported by JSPS KAKENHI grant Nos. JP18K13579 and JP21H04487. This work made use of mat- 5.3. Extension to other types of stars plotlib, a Python library for publication quality graphics (Hunter In this study, we focused on the main-sequence Sun-like stars 2007), and NumPy (van der Walt et al. 2011). that exhibit solar mass, solar radius, solar luminosity, and solar metallicity. Therefore, the application of Eq. (90) or Eq. (91) to other classes of stars would require additional considerations. References Airapetian, V. S., Glocer, A., Khazanov, G. V., et al. 2017, ApJ, 836, L3 1. M-type stars are of particular interest in the context of Airapetian, V. S., Jin, M., Lueftinger, T., et al. 2021, arXiv e-prints, exoplanet because they occupy a large fraction of the local arXiv:2106.01284

Article number, page 19 of 25 A&A proofs: manuscript no. main

Alfvén, H. 1947, MNRAS, 107, 211 Ishikawa, R., Trujillo Bueno, J., del Pino Aleman, T., et al. 2021, arXiv e-prints, Alvarado-Gómez, J. D., Hussain, G. A. J., Cohen, O., et al. 2016, A&A, 588, arXiv:2103.01583 A28 Ishikawa, S.-n., Glesener, L., Krucker, S., et al. 2017, Nature Astronomy, 1, 771 An, C. H., Musielak, Z. E., Moore, R. L., & Suess, S. T. 1989, ApJ, 345, 597 Jess, D. B., Reznikova, V. E., Ryans, R. S. I., et al. 2016, Nature Physics, 12, 179 Antiochos, S. K., MacNeice, P. J., Spicer, D. S., & Klimchuk, J. A. 1999, ApJ, Johnston, C. D. & Bradshaw, S. J. 2019, ApJ, 873, L22 512, 985 Johnston, C. D., Hood, A. W., Cargill, P. J., & De Moortel, I. 2017, A&A, 597, Antiochos, S. K. & Sturrock, P. A. 1978, ApJ, 220, 1137 A81 Antolin, P., Okamoto, T. J., De Pontieu, B., et al. 2015, ApJ, 809, 72 Johnston, C. D., Hood, A. W., De Moortel, I., Pagano, P., & Howson, T. A. 2021, Antolin, P., Pagano, P., Testa, P., Petralia, A., & Reale, F. 2021, Nature Astron- arXiv e-prints, arXiv:2106.03989 omy, 5, 54 Johnstone, C. P., Bartel, M., & Güdel, M. 2021, A&A, 649, A96 Antolin, P. & Shibata, K. 2010, ApJ, 712, 494 Johnstone, C. P. & Güdel, M. 2015, A&A, 578, A129 Aschwanden, M. J. & Parnell, C. E. 2002, ApJ, 572, 1048 Judge, P. G., Solomon, S. C., & Ayres, T. R. 2003, ApJ, 593, 534 Avrett, E. H. & Loeser, R. 2008, ApJS, 175, 229 Kanella, C. & Gudiksen, B. V. 2018, A&A, 617, A50 Barnes, S. A. 2003, ApJ, 586, 464 Katsukawa, Y. & Tsuneta, S. 2005, ApJ, 621, 498 Berghmans, D., Auchère, F., Long, D. M., et al. 2021, arXiv e-prints, Kawaler, S. D. 1988, ApJ, 333, 236 arXiv:2104.03382 Keller, C. U., Schüssler, M., Vögler, A., & Zakharov, V. 2004, ApJ, 607, L59 Bochanski, J. J., Hawley, S. L., Covey, K. R., et al. 2010, AJ, 139, 2679 Klimchuk, J. A. 2006, Sol. Phys., 234, 41 Boerner, P., Edwards, C., Lemen, J., et al. 2012, Sol. Phys., 275, 41 Klimchuk, J. A., Lemen, J. R., Feldman, U., Tsuneta, S., & Uchida, Y. 1992, Bonfils, X., Delfosse, X., Udry, S., et al. 2013, A&A, 549, A109 PASJ, 44, L181 Bradshaw, S. J. & Cargill, P. J. 2013, ApJ, 770, 12 Klimchuk, J. A., Patsourakos, S., & Cargill, P. J. 2008, ApJ, 682, 1351 Cassak, P. A., Mullan, D. J., & Shay, M. A. 2008, ApJ, 676, L69 Kochukhov, O., Hackman, T., Lehtinen, J. J., & Wehrhahn, A. 2020, A&A, 635, Chamberlin, P. C., Woods, T. N., Crotser, D. A., et al. 2009, Geophys. Res. Lett., A142 36, L05102 Kraft, R. P. 1967, ApJ, 150, 551 Chandran, B. D. G. & Perez, J. C. 2019, Journal of Plasma Physics, 85, Kudoh, T. & Shibata, K. 1999, ApJ, 514, 493 905850409 Lanzafame, A. C. 1995, A&A, 302, 839 Chen, Y., Przybylski, D., Peter, H., et al. 2021, arXiv e-prints, arXiv:2104.10940 Lecavelier des Etangs, A., Bourrier, V., Wheatley, P. J., et al. 2012, A&A, 543, Chitta, L. P., van Ballegooijen, A. A., Rouppe van der Voort, L., DeLuca, E. E., L4 & Kariyappa, R. 2012, ApJ, 752, 48 Lemen, J. R., Title, A. M., Akin, D. J., et al. 2012, Sol. Phys., 275, 17 Cho, J. & Lazarian, A. 2003, MNRAS, 345, 325 Linsky, J. L., Fontenla, J., & France, K. 2014, ApJ, 780, 61 Cho, J. & Vishniac, E. T. 2000, ApJ, 539, 273 Magaudda, E., Stelzer, B., Covey, K. R., et al. 2020, A&A, 638, A20 Cirtain, J. W., Golub, L., Winebarger, A. R., et al. 2013, Nature, 493, 501 Malanushenko, A., Cheung, M. C. M., DeForest, C. E., Klimchuk, J. A., & Rem- Claire, M. W., Sheets, J., Cohen, M., et al. 2012, ApJ, 757, 95 pel, M. 2021, arXiv e-prints, arXiv:2106.14877 Cranmer, S. R. 2008, ApJ, 689, 316 Matsumoto, T. & Shibata, K. 2010, ApJ, 710, 1857 Cranmer, S. R. 2009, ApJ, 706, 824 Matt, S. P., Brun, A. S., Baraffe, I., Bouvier, J., & Chabrier, G. 2015, ApJ, 799, Cranmer, S. R. 2017, ApJ, 840, 114 L23 Cranmer, S. R. & van Ballegooijen, A. A. 2005, ApJS, 156, 265 Matthaeus, W. H., Zank, G. P., Oughton, S., Mullan, D. J., & Dmitruk, P. 1999, Dahlburg, R. B., Einaudi, G., Taylor, B. D., et al. 2016, ApJ, 817, 47 ApJ, 523, L93 De Pontieu, B., McIntosh, S. W., Carlsson, M., et al. 2007, Science, 318, 1574 McIntosh, S. W., de Pontieu, B., Carlsson, M., et al. 2011, Nature, 475, 477 Del Zanna, G., Dere, K. P., Young, P. R., & Landi, E. 2021, ApJ, 909, 38 Meyer, C. D., Balsara, D. S., & Aslam, T. D. 2012, MNRAS, 422, 2102 Dere, K. P., Landi, E., Mason, H. E., Monsignori Fossi, B. C., & Young, P. R. Meyer, C. D., Balsara, D. S., & Aslam, T. D. 2014, Journal of Computational 1997, A&AS, 125, 149 Physics, 257, 594 Diamond-Lowe, H., Youngblood, A., Charbonneau, D., et al. 2021, arXiv e- Miyoshi, T. & Kusano, K. 2005, Journal of Computational Physics, 208, 315 prints, arXiv:2104.10522 Moriyasu, S., Kudoh, T., Yokoyama, T., & Shibata, K. 2004, ApJ, 601, L107 Dmitruk, P., Matthaeus, W. H., Milano, L. J., et al. 2002, ApJ, 575, 571 Nakariakov, V. M. & Ofman, L. 2001, A&A, 372, L53 Dressing, C. D. & Charbonneau, D. 2015, ApJ, 807, 45 Nakatani, R., Hosokawa, T., Yoshida, N., Nomura, H., & Kuiper, R. 2018, ApJ, Duvvuri, G. M., Pineda, J. S., Berta-Thompson, Z. K., et al. 2021, arXiv e-prints, 865, 75 arXiv:2102.08493 Okamoto, T. J., Antolin, P., De Pontieu, B., et al. 2015, ApJ, 809, 71 Edlén, B. 1943, ZAp, 22, 30 Oran, R., Landi, E., van der Holst, B., Sokolov, I. V., & Gombosi, T. I. 2017, Ehrenreich, D., Bourrier, V., Wheatley, P. J., et al. 2015, Nature, 522, 459 ApJ, 845, 98 Einaudi, G., Velli, M., Politano, H., & Pouquet, A. 1996, ApJ, 457, L113 Osterbrock, D. E. 1961, ApJ, 134, 347 Felipe, T., Kuckein, C., & Thaler, I. 2018, A&A, 617, A39 Owen, J. E. & Wu, Y. 2013, ApJ, 775, 105 Fontenla, J. M., Avrett, E. H., & Loeser, R. 1990, ApJ, 355, 700 Pagano, P. & De Moortel, I. 2019, A&A, 623, A37 France, K., Arulanantham, N., Fossati, L., et al. 2018, ApJS, 239, 16 Pallavicini, R., Golub, L., Rosner, R., et al. 1981, ApJ, 248, 279 Galsgaard, K. & Nordlund, Å. 1996, J. Geophys. Res., 101, 13445 Parker, E. N. 1972, ApJ, 174, 499 Goodman, M. L. & Judge, P. G. 2012, ApJ, 751, 75 Parker, E. N. 1983, ApJ, 264, 642 Gorti, U. & Hollenbach, D. 2009, ApJ, 690, 1539 Parker, E. N. 1988, ApJ, 330, 474 Gottlieb, S., Shu, C.-W., & Tadmor, E. 2001, SIAM Review, 43, 89 Peres, G., Serio, S., Vaiana, G. S., & Rosner, R. 1982, ApJ, 252, 791 Güdel, M., Audard, M., Kashyap, V. L., Drake, J. J., & Guinan, E. F. 2003, ApJ, Pesnell, W. D., Thompson, B. J., & Chamberlin, P. C. 2012, Sol. Phys., 275, 3 582, 423 Pevtsov, A. A., Fisher, G. H., Acton, L. W., et al. 2003, ApJ, 598, 1387 Güdel, M., Guinan, E. F., & Skinner, S. L. 1997, ApJ, 483, 947 Pizzolato, N., Maggio, A., Micela, G., Sciortino, S., & Ventura, P. 2003, A&A, Gudiksen, B. V. & Nordlund, Å. 2005, ApJ, 618, 1020 397, 147 Guinan, E. F., Engle, S. G., & Durbin, A. 2016, ApJ, 821, 81 Priest, E. 2014, Magnetohydrodynamics of the Sun Günther, H. M., Schmitt, J. H. M. M., Robrade, J., & Liefke, C. 2007, A&A, Rappazzo, A. F., Velli, M., Einaudi, G., & Dahlburg, R. B. 2007, ApJ, 657, L47 466, 1111 Rappazzo, A. F., Velli, M., Einaudi, G., & Dahlburg, R. B. 2008, ApJ, 677, 1348 Hansteen, V., Guerreiro, N., De Pontieu, B., & Carlsson, M. 2015, ApJ, 811, 106 Reale, F., Güdel, M., Peres, G., & Audard, M. 2004, A&A, 416, 733 Heyvaerts, J. & Priest, E. R. 1983, A&A, 117, 220 Reale, F. & Micela, G. 1998, A&A, 334, 1028 Hollweg, J. V. 1984, Sol. Phys., 91, 269 Reiners, A., Basri, G., & Browning, M. 2009, ApJ, 692, 538 Hollweg, J. V., Jackson, S., & Galloway, D. 1982, Sol. Phys., 75, 35 Rempel, M. 2017, ApJ, 834, 10 Hollweg, J. V. & Yang, G. 1988, J. Geophys. Res., 93, 5423 Ribas, I., Guinan, E. F., Güdel, M., & Audard, M. 2005, ApJ, 622, 680 Hood, A. W., Brooks, S. J., & Wright, A. N. 2002, Proceedings of the Royal Rosner, R., Tucker, W. H., & Vaiana, G. S. 1978, ApJ, 220, 643 Society of London Series A, 458, 2307 Rumph, T., Bowyer, S., & Vennes, S. 1994, AJ, 107, 2108 Hossain, M., Gray, P. C., Pontius, Duane H., J., Matthaeus, W. H., & Oughton, Rybicki, G. B. & Lightman, A. P. 1979, Radiative processes in astrophysics S. 1995, Physics of Fluids, 7, 2886 Saar, S. H. 1996, in Stellar Surface Structure, ed. K. G. Strassmeier & J. L. Hunter, J. D. 2007, Computing in Science and Engineering, 9, 90 Linsky, Vol. 176, 237 Iijima, H. 2016, PhD thesis, Department of Earth and Planetary Science, School Saar, S. H. 2001, in Astronomical Society of the Pacific Conference Series, Vol. of Science, The University of Tokyo, Japan 223, 11th Cambridge Workshop on Cool Stars, Stellar Systems and the Sun, Iijima, H. & Imada, S. 2021, arXiv e-prints, arXiv:2106.00864 ed. R. J. Garcia Lopez, R. Rebolo, & M. R. Zapaterio Osorio, 292 Ionson, J. A. 1978, ApJ, 226, 650 Sakurai, T., Goossens, M., & Hollweg, J. V. 1991, Sol. Phys., 133, 227 Irwin, J. & Bouvier, J. 2009, in The Ages of Stars, ed. E. E. Mamajek, D. R. Sanz-Forcada, J., Micela, G., Ribas, I., et al. 2011, A&A, 532, A6 Soderblom, & R. F. G. Wyse, Vol. 258, 363–374 Scelsi, L., Maggio, A., Peres, G., & Pallavicini, R. 2005, A&A, 432, 671

Article number, page 20 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

Schmelz, J. T., Reames, D. V., von Steiger, R., & Basu, S. 2012, ApJ, 755, 33 Schrijver, C. J., van den Oord, G. H. J., & Mewe, R. 1994, A&A, 289, L23 See, V., Matt, S. P., Folsom, C. P., et al. 2019, ApJ, 876, 118 Serio, S., Peres, G., Vaiana, G. S., Golub, L., & Rosner, R. 1981, ApJ, 243, 288 Shebalin, J. V., Matthaeus, W. H., & Montgomery, D. 1983, Journal of Plasma Physics, 29, 525 Shibata, K. & Yokoyama, T. 2002, ApJ, 577, 422 Shimizu, T. 1995, PASJ, 47, 251 Shoda, M., Suzuki, T. K., Asgari-Targhi, M., & Yokoyama, T. 2019, ApJ, 880, L2 Shoda, M., Suzuki, T. K., Matt, S. P., et al. 2020, ApJ, 896, 123 Shoda, M., Yokoyama, T., & Suzuki, T. K. 2018, ApJ, 853, 190 Shu, C.-W. & Osher, S. 1988, Journal of Computational Physics, 77, 439 Skumanich, A. 1972, ApJ, 171, 565 Spitzer, L. & Härm, R. 1953, Physical Review, 89, 977 Spruit, H. C. & Zweibel, E. G. 1979, Sol. Phys., 62, 15 Sreejith, A. G., Fossati, L., Youngblood, A., France, K., & Ambily, S. 2020, A&A, 644, A67 Srivastava, A. K., Shetye, J., Murawski, K., et al. 2017, Scientific Reports, 7, 43147 Stein, R. F. 1971, ApJS, 22, 419 Steiner, O., Grossmann-Doerth, U., Knölker, M., & Schüssler, M. 1998, ApJ, 495, 468 Sturrock, P. A. & Uchida, Y. 1981, ApJ, 246, 331 Suresh, A. & Huynh, H. T. 1997, Journal of Computational Physics, 136, 83 Suzuki, T. K. & Inutsuka, S.-i. 2005, ApJ, 632, L49 Takasao, S., Isobe, H., & Shibata, K. 2013, PASJ, 65, 62 Takasao, S., Mitsuishi, I., Shimura, T., et al. 2020, ApJ, 901, 70 Telleschi, A., Güdel, M., Briggs, K., et al. 2005, ApJ, 622, 653 Testa, P., De Pontieu, B., Allred, J., et al. 2014, Science, 346, 1255724 Toriumi, S., Airapetian, V. S., Hudson, H. S., et al. 2020, ApJ, 902, 36 Tsuneta, S., Ichimoto, K., Katsukawa, Y., et al. 2008, ApJ, 688, 1374 Tu, L., Johnstone, C. P., Güdel, M., & Lammer, H. 2015, A&A, 577, L3 van Ballegooijen, A. A. 1986, ApJ, 311, 1001 van Ballegooijen, A. A. & Asgari-Targhi, M. 2017, ApJ, 835, 10 van Ballegooijen, A. A., Asgari-Targhi, M., Cranmer, S. R., & DeLuca, E. E. 2011, ApJ, 736, 3 Van Damme, H. J., De Moortel, I., Pagano, P., & Johnston, C. D. 2020, A&A, 635, A174 van der Holst, B., Sokolov, I. V., Meng, X., et al. 2014, ApJ, 782, 81 van der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, Computing in Science and Engineering, 13, 22 Van Kooten, S. J. & Cranmer, S. R. 2017, ApJ, 850, 64 van Leer, B. 1979, Journal of Computational Physics, 32, 101 Vasquez, B. J. 1990, ApJ, 356, 693 Verdini, A., Grappin, R., & Montagud-Camps, V. 2019, Sol. Phys., 294, 65 Verdini, A., Grappin, R., & Velli, M. 2012, A&A, 538, A70 Verdini, A. & Velli, M. 2007, ApJ, 662, 669 Verwichte, E., Nakariakov, V. M., Ofman, L., & Deluca, E. E. 2004, Sol. Phys., 223, 77 Vidal-Madjar, A., Lecavelier des Etangs, A., Désert, J. M., et al. 2003, Nature, 422, 143 Vidotto, A. A., Gregory, S. G., Jardine, M., et al. 2014, MNRAS, 441, 2361 Wang, L., Bai, X.-N., & Goodman, J. 2019, ApJ, 874, 90 Wang, Y. M. 2020, ApJ, 904, 199 Washinoue, H. & Suzuki, T. K. 2019, ApJ, 885, 164 Weber, E. J. & Davis, Leverett, J. 1967, ApJ, 148, 217 Winters, J. G., Henry, T. J., Jao, W.-C., et al. 2019, AJ, 157, 216 Withbroe, G. L. & Noyes, R. W. 1977, ARA&A, 15, 363 Woods, T. N., Chamberlin, P. C., Harder, J. W., et al. 2009, Geophys. Res. Lett., 36, L01101 Wright, N. J. & Drake, J. J. 2016, Nature, 535, 526 Wright, N. J., Drake, J. J., Mamajek, E. E., & Henry, G. W. 2011, ApJ, 743, 48 Wright, N. J., Newton, E. R., Williams, P. K. G., Drake, J. J., & Yadav, R. K. 2018, MNRAS, 479, 2351 Youngblood, A., France, K., Loyd, R. O. P., et al. 2017, ApJ, 843, 31 Zhuleku, J., Warnecke, J., & Peter, H. 2020, A&A, 640, A119

Article number, page 21 of 25 A&A proofs: manuscript no. main     B   Appendix A: Derivation of basic equations  Bx ∂ p y ∂ p  (∇ × B) × B = − p r f Bx − p r f By  es  r f ∂s r f ∂s 

Bs ∂  p  Bs ∂  p  + p r f Bx ex + p r f By ey. r f ∂s r f ∂s In this Appendix, the basic equations are derived from the con- (A.12) ventional magnetohydrodynamic equations (e.g. Priest 2014) given by The s-component of Eq. (A.2) is given by

∂ρ ∂ ∂ 1  2 2 d  2  1 ∂ + ∇ · (ρv) = 0, (A.1) vs + vs vs − v + v ln r f + p ∂t ∂t ∂s 2 x y ds ρ ∂s   ∂v 1 1   B   + (v · ∇) v = − ∇p + (∇ × B) × B + g, (A.2) 1  Bx ∂ p y ∂ p  ∂t ρ 4πρ +  p r f Bx + p r f By  = gs, (A.13) 4πρ r f ∂s r f ∂s  ∂B + ∇ × (v × B) = 0, (A.3) ∂t which is, after some algebra, written in terms of conservation " # ∂e 1 law as + ∇ · (e + p) v − (v × B) × B = ρv · g + Q + Q . cnd rad  2 2   ∂t 4π ∂ 1 ∂  Bx + By   (ρv ) + ρv2 + p +  r2 f  (A.4) ∂t s r2 f ∂s  s 8π   " # From Eq. (3) and (6), for any vector field X, its divergence and 1   d   = p + ρ v2 + v2 ln r2 f + ρg , (A.14) rotation are given by 2 x y ds s 1 ∂   ∇ · X = X r2 f , (A.5) The transverse components (v⊥ = vxex + vyey) of Eq. (A.2) are r2 f ∂s s ∂ v ∂  p  B ∂  p  1 ∂  p  1 ∂  p  v + s r f v − s r f B = 0, (A.15) ∇ × X = − p Xyr f ex + p Xxr f ey. (A.6) ⊥ p ⊥ p ⊥ r f ∂s r f ∂s ∂t r f ∂s 4πρr f ∂s Under the assumption that gravity works only in the field-aligned which is written in terms of conservation law as " ! # direction, the gravitational force is simplified as ∂ 1 ∂ B B (ρv ) ρv v − s ⊥ r2 f ⊥ + 2 s ⊥ g = (g · es) es = gses, (A.7) ∂t r f ∂s 4π ! 1 Bs B⊥ d  2  where es is the unit vector in the s direction. Combining Eq.s = − ρvsv⊥ − ln r f . (A.16) (A.1), (A.4) and (A.5), the mass and energy conservation laws 2 4π ds are written as The actual basic equation is obtained by adding a phenomeno- v ∂ 1 ∂   logical turbulence term ρD⊥ on the right-hand side, i.e., ρ + ρv r2 f = 0, (A.8) ∂t r2 f ∂s s " ! # ∂ 1 ∂ Bs B⊥ 2  2 2    (ρv⊥) + ρvsv⊥ − r f ∂ 1 ∂  Bx + By  Bs    ∂t r2 f ∂s 4π e + e + p +  v − v B + v B  r2 f  ∂t r2 f ∂s  8π  s 4π x x y y   ! 1 Bs B⊥ d  2  v = − ρvsv⊥ − ln r f + ρD⊥. (A.17) = ρvsgs + Qcnd + Qrad, (A.9) 2 4π ds where we use The rotation of the electromotive force in Eq. (A.3) is given by " # 1 1 ∂ h p i (e + p) v − (v × B) × B · es ∇ × (v × B) = − p (vsBx − vxBs) r f ex 4π r f ∂s   B2 + B2   h  i  x y  Bs 1 ∂ p = e + p +  vs − vxBx + vyBy . (A.10) − p vsBy − vyBs r f ey. (A.18)  8π  4π r f ∂s

Then, the transverse components (B⊥ = Bxex + Byey) of the in- duction equation are ∂ 1 ∂ h p i The advection term and Lorentz force in Eq. (A.2) are written B⊥ + (vs B⊥ − v⊥Bs) r f = 0, (A.19) ∂t p ∂s as r f 1   which is rewritten in terms of conservation law as (v · ∇) v = (∇ × v) × v + ∇ v2 , 2 " # ∂ 1 ∂ h 2 i ∂ 1   d   B⊥ + (vs B⊥ − v⊥Bs) r f = v v − v2 + v2 ln r2 f e ∂t r2 f ∂s s ∂s s 2 x y ds s 1 d  2  = (v B⊥ − v⊥B ) ln r f . (A.20) vs ∂  p  vs ∂  p  2 s s ds + p r f vx ex + p r f vy ey, (A.11) r f ∂s r f ∂s

Article number, page 22 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

Riemann solver ∆smin filling factor f∗ 4.0

HLL 5 km [1, 0.33, 0.1, 0.05]

HLLD 5 km [1, 0.33, 0.1, 0.05] 3.0 ] 3

HLLD 2 km [1, 0.33, 0.1, 0.05] − cm

Table B.1. List of the simulation runs executed to study the transition- 9 2.0 region problem. Note that ∆smin corresponds to the spatial resolution of the transition region. [10 e n 1.0 HLL The basic equation is obtained after adding the phenomenologi- HLLD cal turbulence term: HLLD (high-reso) ∂ 1 ∂ h i 0.0 B + (v B − v B ) r2 f ∂t ⊥ r2 f ∂s s ⊥ ⊥ s 0 30 60 90 120 150 1 d   p t [min] = (v B − v B ) ln r2 f + 4πρDb . (A.21) 2 s ⊥ ⊥ s ds ⊥ Eq.s (A.8), (A.9), (A.14), (A.17) and (A.21) are the basic equa- 4.0 tions used in this study.

Appendix B: Transition-region problem and 3.0 dependence on numerical methods

A crucial difficulty in the simulation of coronal heating is that K] the transition region needs to be resolved with extremely fine 6 2.0 grids, otherwise the coronal density is significantly underesti- [10

mated (Bradshaw & Cargill 2013). This complex problem shall T be referred to as the “transition-region problem”. Here, we in- vestigate the relation between the coronal density and grid size 1.0 in the transition region and the numerical scheme. HLL The numerical settings of the test calculations are listed HLLD in Table B.1. Two types of the approximated Riemann solver HLLD (high-reso) are used in this test: Harten-van Leer-Lax (HLL) and HLL- 0.0 Discontinuity (HLLD) schemes. In exchange for its simplicity, 0 30 60 90 120 150 the HLL scheme has the disadvantage that the contact discon- t [min] tinuities (entropy modes) suffer from significant numerical dif- fusion. Meanwhile, by explicitly considering the substructures inside the Riemann fan, the contact (and rotational) discontinu- Fig. B.1. Time evolution of the loop-top electron density ne,top (top ities are better resolved in the HLLD scheme. Because the tran- panel) and temperature Ttop (bottom panel) for fixed filling factor f∗ = sition region is a type of contact discontinuity, its numerical dif- 0.05 and half loop length lloop = 20 Mm. Three lines show the results fusion in the HLL and HLLD schemes should behave differently with ∆smin = 5 km and HLL solver (red), ∆smin = 5 km and HLLD even with the same grid spacing. Accordingly, the transition- solver (green), and ∆smin = 2 km and HLLD solver (blue), respectively. region problem should also change. Therefore, dependence on the choice of the Riemann solver, as well as the spatial resolu- tion, is also investigated. the numerical convergence of coronal density, the test in this Figure B.1 shows the time evolution of the loop-top elec- study proves the numerical convergence for ∆smin = 5 km, when f = 0.05. This gap can be attributed to the different numeri- tron density (ne,top, top panel) and loop-top temperature (Ttop, ∗ bottom panel) for a fixed magnetic filling factor at the surface cal settings. Bradshaw & Cargill(2013) consider the thermal re- ( f∗ = 0.05). The orange-dashed, green-solid, and blue-dotted sponse to strong impulsive heating, whereas this study considers quasi-steady uniform coronal heating. lines correspond to the results of the HLL (∆smin = 5 km), HLLD (∆smin = 5 km), and HLLD (∆smin = 2 km) schemes, re- Figure B.2 is the same as Figure B.1 but for f∗ = 1. The dif- spectively. As indicated in the top panel of Figure B.1, the three ference in the loop-top density is more significant than f∗ = 0.05. lines exhibit subtle differences in behaviour. A low resolution The HLL and HLLD schemes produce largely different results with a large numerical diffusion at the transition region yields despite the same grid size at the transition region. The reason a small loop-top density, as reported in Bradshaw & Cargill is that, in the HLL method, the numerical diffusivity of the en- (2013). Meanwhile, the loop-top temperature is nearly indepen- tropy mode is roughly proportional to the speed of the fastest dent of the resolution, which is also consistent with Bradshaw & wave. At the transition region, the fastest wave speed is the fast Cargill(2013). Although Bradshaw & Cargill(2013) state that magneto-acoustic velocity that increases with the field strength. the transition region must be resolved within 1 km to ensure Thus, with HLL, f∗ = 1 yields a higher wave speed and larger

Article number, page 23 of 25 A&A proofs: manuscript no. main

2.0

K] 5

6 HLL

[10 HLLD

3 HLLD (high-reso) 1.5 / 1 ] 4 3  − eff loop l cm

1.0 top 10 p 3  [10 3 e 10 n

0.5 ×

HLL 4 2 . HLLD

HLLD (high-reso) = 1 0.0 0 30 60 90 120 150 RTV 1 T 1 2 3 4 5 t [min] 6 Ttop [10 K]

8.0 Fig. B.3. This figure compares the time-averaged loop-top temperature Ttop and the loop-top temperature predicted by the RTV scaling law TRTV. 6.0 numerical value monotonically approaches the RTV-predicted value. Note that the deviation of the results of the simulation K] from those of the RTV prediction is not necessarily evidence of 6 4.0 inaccuracy, although, as a general trend, a higher resolution of [10 the transition region should lead to a value closer to the RTV

T prediction. Deviations from the RTV value in a low-resolution run is prominent at higher coronal temperatures. However, it re- 2.0 mains to be seen if the large discrepancy from the RTV value HLL at high coronal temperatures is due to the large numerical diffu- HLLD sion at the transition region or the increase in the required res- HLLD (high-reso) olution. The coronal-temperature dependence of the transition- 0.0 region problem should be investigated in the future. 0 30 60 90 120 150 Finally, we note that there are physical mechanisms that can t [min] produce the broadening of the transition region, including am- bipolar diffusion (Fontenla et al. 1990; Lanzafame 1995). The transition-region problem should also be revisited considering Fig. B.2. Same as Figure B.1 but for f∗ = 1 and lloop = 20 Mm. the additional diffusive processes in future. numerical diffusion at the transition region. Such an enhanced Appendix C: XUV spectrum from different numerical diffusion of the entropy mode is suppressed in the atmospheric layers HLLD method because the contact discontinuity inside the Rie- The XUV photons of a specific wavelength (band) do not mann fan is resolved. Thus, the difference between the HLL and originate from the plasma with a specific temperature. Indeed, HLLD schemes is more remarkable for f = 1, as shown in Fig- ∗ narrow-band filters in the Atmospheric Imaging Assembly of the ure B.2. The weak dependence of the loop-top temperature on Solar Dynamics Observatory have broad response functions in the numerical scheme is also confirmed. terms of temperature (Boerner et al. 2012). Although there is To test the numerical convergence of the physical variables no direct correlation between temperature and wavelength, it is in the coronal loop on the actual values, the simulation result is still useful to understand the energy spectrum emanating from compared with the results of the RTV scaling law. Figure B.3 plasma with a specific temperature. displays the loop-top temperatures obtained in the simulation Figure C.1 shows the correlation between the emission mea- and those predicted by the RTV scaling law, i.e., sure distribution in a given temperature bin (top panel) and the corresponding spectral energy flux normalised at 1 au (bottom 3  eff 1/3 TRTV = 1.4 × 10 K ptoplloop , (B.1) panel) for the fiducial case ( f∗ = 0.01, lloop = 20 Mm). For bet- ter visualisation, the spectral energy fluxes averaged over 10 Å eff where ptop is the loop-top pressure and lloop is the effective half- are shown in the bottom panel. The black dotted line in the top loop length, both of which are measured in the cgs unit. In Fig- panel represents the total emission measure distribution, while ure B.3, as the resolution of the transition region increases, the the black dotted line represents the corresponding spectral en-

Article number, page 24 of 25 Munehito Shoda and Shinsuke Takasao : Corona and XUV emission modelling of the Sun and Sun-like stars

] sion is also non-negligible, especially in the shorter wave- 5 1032 − length. As a general trend, the low-energy photons tend to origi- nate from the low-temperature plasma and vice versa. Note that 1030 a non-negligible fraction of EUV photons originates from the chromosphere and the transition region, and therefore including them in the model is essential in predicting the XUV spectrum.

1028

1026

1024 emission measure distribution [cm 104 105 106 107 temperature [K] 102 spectrum from T 104 K plasma ≈ spectrum from T 105 K plasma ˚ A] ≈ 6

1 spectrum from T 10 K plasma 100 ≈ − integrated spectrum s 2 − 2 10−

4 10−

6 10− flux at 1 au [erg cm

8 10− 0 200 400 600 800 1000 wavelength [A]˚

Fig. C.1. This figure shows the emission measure in the limited range of temperature and corresponding spectral energy flux at 1 au. For better visualisation, the spectral energy fluxes in the bottom panel are aver- aged over 10 Å. Orange, green and blue lines correspond to the emis- sion measures and spectral energy fluxes of T ≈ 104 K (chromosphere), T ≈ 105 K (transition region) and T ≈ 106 K (corona), respectively. Black dotted lines show the full emission measure distribution and cor- responding full spectral energy flux. ergy flux in the bottom panel. We consider three temperature bins centred on T = 104 K, T = 105 K, and T = 106 K corresponding to the upper chromosphere, transition region, and corona, respec- tively. The corresponding spectral energy fluxes are highlighted with the same colours in the bottom panel, which showcase the following three properties. 1. Emissions from the upper chromosphere (T ≈ 104 K) dominate the Lyman continuum but negligible in ≤ 600 Å. 2. Emissions from the transition region (T ≈ 105 K) are nearly uniform in the middle of the EUV wavelength (200 Å ≤ λ ≤ 800 Å) but negligible in the X-ray range. 3. Emissions from the corona (T ≈ 106 K) are the dominant sources of stellar X-rays. The contribution to the EUV emis-

Article number, page 25 of 25