Unraveling the Trigger Mechanism of Explosive Reconnection in Partially Ionized Solar Plasma

Abdullah Zafar Yunnan Observatories, Chinese Academy of Science, Kunming, Yunnan 6502016 PR china Lei Ni Yunnan Observatories, Chinese Academy of Science, Kunming, Yunnan 6502016 PR china Center for Astronomical Mega-Science, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100012, PR China University of Chinese Academy of Sciences, Beijing 100049, PR China Lei Ni leini@ynao.ac.cn Jun Lin Yunnan Observatories, Chinese Academy of Science, Kunming, Yunnan 6502016 PR china Center for Astronomical Mega-Science, Chinese Academy of Sciences, 20A Datun Road, Chaoyang District, Beijing 100012, PR China University of Chinese Academy of Sciences, Beijing 100049, PR China Ahmad Ali Pakistan Tokamak Plasma Research Institute, Islamabad 3329, Pakistan
Abstract

Plasmoid instability is usually accounted for the onset of fast reconnection events observed in astrophysical plasmas. However, the measured reconnection rate from observations can be one order of magnitude higher than that derived from MHD simulations. In this study, we present the results of magnetic reconnection in the partially ionized low solar atmosphere based on 2.5D magnetohydrodynamics (MHD) simulations. The whole reconnection process covers two different fast reconnection phases. In the first phase, the slow Sweet-Parker reconnection transits to the plasmoid-mediated reconnection, and the reconnection rate reaches about 0.02. In the second phase, a faster explosive reconnection appears, with the reconnection rate reaching above 0.06. At the same time, a sharp decrease in plasma temperature and density at the principle X-point is observed which is associated with the strong radiative cooling, the ejection of hot plasma from the local reconnection region or the motion of principle X-point from hot and denser region to cool and less dense one along the narrow current sheet. This causes gas pressure depletion and the increasing of magnetic diffusion at the main X-point, resulting in the local Petschek-like reconnection and a violent and rapid increase in the reconnection rate. This study for the first time reveals a common phenomenon that the plasmoid dominated reconnection transits to an explosive faster reconnection with the rate approaching the order of 0.1 in partially ionized plasma in the MHD scale.

1 Introduction

Magnetic reconnection plays a key role in variety of process in the Universe. For example, in active galactic nuclei, it triggers highly energetic bursts in the accretion disk around the black hole (Liu et al., 2002). It has a significant impact on space weather conditions (Paschmann et al., 1979), and enables phenomena like sawtooth crashes and tearing instabilities in tokamaks (Furth et al., 1973). On the Sun, magnetic reconnection process sparks many remarkable events such as Ellerman bombs (EBs) (Ellerman, 1917; Georgoulis et al., 2002), Ultraviolet (UV) bursts (Peter et al., 2014; Tian et al., 2016), flares (Yokoyama & Shibata, 2001; Chen et al., 2023), surges (Yokoyama & Shibata, 1995), jets (Shibata et al., 2007) and so on.

A long-standing problem in the field of magnetic reconnection is to predict the rate at which the magnetic reconnection proceeds. Sweet (Sweet, 1958) and Parker (Parker, 1957) separately made the first attempt to describe magnetic reconnection. The Sweet-Parker (SP) model is characterized by a long thin current sheet between oppositely directed magnetic fields having length L, and thickness δL/Ssimilar-to𝛿𝐿𝑆\delta\sim L/\sqrt{S}italic_δ ∼ italic_L / square-root start_ARG italic_S end_ARG. Where S𝑆Sitalic_S is the Lundquist number and is related to the Length (L𝐿Litalic_L), the Alfven speed (VAsubscript𝑉𝐴V_{A}italic_V start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT) and the plasma resistivity (η𝜂\etaitalic_η) by an expression S=LVA/η𝑆𝐿subscript𝑉𝐴𝜂S=LV_{A}/\etaitalic_S = italic_L italic_V start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT / italic_η. The reconnection rate scales with the Lundquist number as   S0.5superscript𝑆0.5S^{-0.5}italic_S start_POSTSUPERSCRIPT - 0.5 end_POSTSUPERSCRIPT. The value of the Lundquist number is extremely large in most plasma environments in the universe (e.g., S is about 106101410{{}^{6}}-10^{14}10 start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT in the solar atmosphere). The reconnection rate predicted by the SP model is normally several orders of magnitude slower for explaining the fast energy release process in most astrophysical environments (Kulsrud, 1998). For example, the measured reconnection rate of a solar flare current sheet in the corona is about 0.01-0.1 (Lin et al., 2005; Takasao et al., 2011), but the predicated reconnection rate by the SP model is only about 10-5 for a typical Lundquist number of 1010 in the solar corona.

The discrepancy between observed and SP reconnection rates has shifted attention to other models in which the reconnection rate is not strongly dependent on the Lundquist number. Petschek (Petschek, 1964) proposed one such MHD model that seems capable of explaining observed reconnection events. In contrast to the SP model, the current sheet region in the Petschek model is much shorter, and most of the incoming fluid does not pass through it but is redirected by two pairs of slow-mode shocks on each side of the current sheet. The reconnection rate γ𝛾\gammaitalic_γ has a weaker dependence on the Lundquist number as γsimilar-to𝛾absent\gamma\simitalic_γ ∼ 1/logS. Therefore, fast magnetic reconnection is expected in a large Lundquist number system. However, numerical simulations indicate that the classical Petschek configuration cannot be sustained unless the local resistivity or viscosity is enhanced (Biskamp, 1986; Uzdensky & Kulsrud, 2000).

On the other hand, the classical steady-state SP and Petschek models cannot match the unstable and explosive reconnection process usually occur in the solar and astrophysical environments. Many observations show that plasmoid structures are formed in the reconnection current sheet in the solar atmosphere (Lin et al., 2005; Takasao et al., 2011; Li et al., 2016). The plasmoid is believed to correspond to magnetic island, which refers to a region of closed magnetic fields with an O-type magnetic null point located in the center in a two-dimensional plane. Recent studies have shown that reconnection usually takes place in a fragmented current sheet due to the presence of plasmoid instabilities and can be found in a variety of plasma models, including resistive magnetohydrodynamics (MHD) (Shibata & Tanuma, 2001; Huang & Bhattacharjee, 2012; Ali & Zhu, 2019), Hall MHD (Shepherd & Cassak, 2010; Huang et al., 2011) and particle-in-cell (PIC) simulations (Daughton et al., 2009; Fermo et al., 2012; Zhu et al., 2020a, b). The MHD simulations demonstrated that the reconnection rate depends weakly on the Lundquist number and is always increased to a value of order 0.01 (Bhattacharjee et al., 2009; Ni et al., 2015). However, this value is still one order of magnitude smaller than the upper limit of measured reconnection rates from the observed explosive events in the solar atmosphere (De Pontieu et al., 2007; Takasao et al., 2011; Moore et al., 2011). Though the observed width of the current sheet is in the MHD scale in the solar atmosphere (Lin et al., 2005), the reconnection rate of order 0.1 is rarely achieved in MHD simulations with an initial equilibrium state, unless the initial system is in an unstable stage (Mei et al., 2012).

Plasmas in many of astrophysical environments are usually partially ionized, such as the low solar atmosphere, protoplanetary nebulae, discs around young stellar objects and the interstellar medium. The neutral plasma species can strongly affect the magnetic reconnection process in different ways (Ni et al., 2020). Previous numerical simulations prove that collisions between ions and neutrals can directly increase magnetic diffusion and the reconnection rate in the photosphere (Zafar et al., 2023; Liu et al., 2023). The decoupling of ions and neutrals, called ambipolar diffusion, causes faster thinning of the current sheet in the case with zero guide field (Brandenburg & Zweibel, 1994, 1995), but it does not accelerate the reconnection process after plasmoid instability appears (Ni et al., 2015). The ionization-recombination effect decreases the plasma pressure inside the current sheet and accelerates magnetic reconnection (Leake et al., 2012; Murtas et al., 2021). But this effect is not obvious when the ionization is stronger than the recombination in the current sheet with a significant temperature increase (Ni et al., 2018). Moreover, recent two-fluid MHD simulations show that the ionization-recombination process stabilizes the current sheet and dampens its dynamics (Murtas et al., 2022).

In this work, we numerically investigate the non-stationary magnetic reconnection process at different altitudes in a highly stratified lower solar atmosphere. All numerical results indicate that the local unstable Petschek-like reconnection usually occurs after the plasmoid instability appears. More importantly, the reconnection process is sharply accelerated when the fragment current sheet with the main X-point shows a Petschek-like structure, the maximum reconnection rate always reaches above 0.06. These results suggest that an efficient explosive energy release process with a reconnection rate of order 0.1 can happen in the partially ionized low solar atmosphere without kinetic reconnection mechanisms.

2 Numerical Model

2.1 Single fluid MHD equations

In this study, we performed high-resolution 2.5D MHD simulations using NIRVANA (Ziegler, 2011) code. All species of the hydrogen-helium plasma, such as H𝐻Hitalic_H, H+superscript𝐻H^{+}italic_H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, He𝐻𝑒Heitalic_H italic_e, He+𝐻superscript𝑒He^{+}italic_H italic_e start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, and electrons, are strongly coupled and treated as a single fluid. We used the following single-fluid MHD equations in our numerical experiments:

ρt=(ρ𝐯),𝜌𝑡𝜌𝐯\displaystyle\frac{\partial\rho}{\partial t}=-\nabla\cdot(\rho\mathbf{v}),divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG = - ∇ ⋅ ( italic_ρ bold_v ) , (1)
(ρ𝐯)t=[ρ𝐯𝐯+(p+12μ0|𝐁|2)I1μ0𝐁𝐁]+τS,𝜌𝐯𝑡absentdelimited-[]𝜌𝐯𝐯𝑝12subscript𝜇0superscript𝐁2𝐼1subscript𝜇0𝐁𝐁missing-subexpressionsubscript𝜏𝑆\displaystyle\begin{aligned} \frac{\partial(\rho\mathbf{v})}{\partial t}&=-% \nabla\cdot\left[\rho\mathbf{vv}+\left(p+\frac{1}{2\mu_{0}}|\mathbf{B}|^{2}% \right)I-\frac{1}{\mu_{0}}\mathbf{BB}\right]\\ &+\nabla\cdot\tau_{S},\end{aligned}start_ROW start_CELL divide start_ARG ∂ ( italic_ρ bold_v ) end_ARG start_ARG ∂ italic_t end_ARG end_CELL start_CELL = - ∇ ⋅ [ italic_ρ bold_vv + ( italic_p + divide start_ARG 1 end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | bold_B | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_I - divide start_ARG 1 end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG bold_BB ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + ∇ ⋅ italic_τ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , end_CELL end_ROW (2)
𝐁t=×(𝐯×𝐁η×𝐁+𝐄AD),𝐁𝑡𝐯𝐁𝜂𝐁subscript𝐄𝐴𝐷\displaystyle\frac{\partial\mathbf{B}}{\partial t}=\nabla\times(\mathbf{v}% \times\mathbf{B}-\eta\nabla\times\mathbf{B}+\mathbf{E}_{AD}),divide start_ARG ∂ bold_B end_ARG start_ARG ∂ italic_t end_ARG = ∇ × ( bold_v × bold_B - italic_η ∇ × bold_B + bold_E start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT ) , (3)
et=[(e+p+12μ0|𝐁|2)𝐯]+[1μ0(𝐯𝐁)𝐁]+[𝐯τs+ημ0𝐁×(×𝐁)][1μ0𝐁×𝐄AD]+Qrad+,𝑒𝑡absentdelimited-[]𝑒𝑝12subscript𝜇0superscript𝐁2𝐯missing-subexpressiondelimited-[]1subscript𝜇0𝐯𝐁𝐁missing-subexpressiondelimited-[]𝐯subscript𝜏𝑠𝜂subscript𝜇0𝐁𝐁missing-subexpressiondelimited-[]1subscript𝜇0𝐁subscript𝐄𝐴𝐷missing-subexpressionsubscript𝑄𝑟𝑎𝑑\displaystyle\begin{aligned} \frac{\partial e}{\partial t}&=-\nabla\cdot\left[% \left(e+p+\frac{1}{2\mu_{0}}|\mathbf{B}|^{2}\right)\mathbf{v}\right]\\ &+\nabla\cdot\left[\frac{1}{\mu_{0}}(\mathbf{v}\cdot\mathbf{B})\mathbf{B}% \right]\\ &+\nabla\cdot\left[\mathbf{v}\cdot\tau_{s}+\frac{\eta}{\mu_{0}}\mathbf{B}% \times(\nabla\times\mathbf{B})\right]\\ &-\nabla\cdot\left[\frac{1}{\mu_{0}}\mathbf{B}\times\mathbf{E}_{AD}\right]\\ &+Q_{rad}+\mathcal{H},\end{aligned}start_ROW start_CELL divide start_ARG ∂ italic_e end_ARG start_ARG ∂ italic_t end_ARG end_CELL start_CELL = - ∇ ⋅ [ ( italic_e + italic_p + divide start_ARG 1 end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | bold_B | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) bold_v ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + ∇ ⋅ [ divide start_ARG 1 end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( bold_v ⋅ bold_B ) bold_B ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + ∇ ⋅ [ bold_v ⋅ italic_τ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + divide start_ARG italic_η end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG bold_B × ( ∇ × bold_B ) ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL - ∇ ⋅ [ divide start_ARG 1 end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG bold_B × bold_E start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT ] end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_Q start_POSTSUBSCRIPT italic_r italic_a italic_d end_POSTSUBSCRIPT + caligraphic_H , end_CELL end_ROW (4)

with

e=pγ1+12ρ|𝐯|2+12μ0|𝐁|2𝑒𝑝𝛾112𝜌superscript𝐯212subscript𝜇0superscript𝐁2\displaystyle e=\frac{p}{\gamma-1}+\frac{1}{2}\rho|\mathbf{v}|^{2}+\frac{1}{2% \mu_{0}}|\mathbf{B}|^{2}italic_e = divide start_ARG italic_p end_ARG start_ARG italic_γ - 1 end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ρ | bold_v | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | bold_B | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (5)

and

p=(1.1+XiH+0.1XiHe)ρ1.4mikBT,𝑝1.1subscript𝑋𝑖𝐻0.1subscript𝑋𝑖𝐻𝑒𝜌1.4subscript𝑚𝑖subscript𝑘𝐵𝑇\displaystyle p=\frac{(1.1+X_{iH}+0.1X_{iHe})\rho}{1.4m_{i}}k_{B}T,italic_p = divide start_ARG ( 1.1 + italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT + 0.1 italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT ) italic_ρ end_ARG start_ARG 1.4 italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T , (6)

where ρ𝜌\rhoitalic_ρ, 𝐯𝐯\mathbf{v}bold_v, p𝑝pitalic_p, 𝐁𝐁\mathbf{B}bold_B, e𝑒eitalic_e, T𝑇Titalic_T, mi, kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT are the plasma mass density, fluid velocity, plasma thermal pressure, magnetic field, total energy density, plasma temperature, proton mass, and Boltzmann constant, respectively. Whereas, XiHsubscript𝑋𝑖𝐻X_{iH}italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT represents the ionization fraction of hydrogen, while XiHesubscript𝑋𝑖𝐻𝑒X_{iHe}italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT denotes the helium ionization fraction. The total helium number density is set to 10% of that of hydrogen, and only primary helium ionization is considered. The adiabatic constant γ𝛾\gammaitalic_γ is set to 5/3. The stress tensor is τS=ξ[𝐯+(𝐯)T23(𝐯)I]subscript𝜏𝑆𝜉delimited-[]𝐯superscript𝐯𝑇23𝐯𝐼\tau_{S}=\xi[\nabla\mathbf{v}+(\nabla\mathbf{v})^{T}-\frac{2}{3}(\nabla\cdot% \mathbf{v})I]italic_τ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = italic_ξ [ ∇ bold_v + ( ∇ bold_v ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - divide start_ARG 2 end_ARG start_ARG 3 end_ARG ( ∇ ⋅ bold_v ) italic_I ] , where the parameter ξ𝜉\xiitalic_ξ is the coefficient of dynamic viscosity, expressed in the units of kg m-1 s-1. The current sheet considered in this study is parallel to the surface of the Sun, and the current sheet width thins down below 10 km during the main reconnection process, hence the gravity effect is ignored, and the initial plasma density and temperature are assumed to be uniform in the whole simulation domain.

2.2 Diffusions and viscosity coefficients

In the partially ionized plasma of the lower solar atmosphere, interactions between various plasma species are of key interest (Wargnier et al., 2023). Magnetic diffusion (η𝜂\etaitalic_η) is defined as the sum of diffusion caused by electron-ion collision (ηeisubscript𝜂𝑒𝑖\eta_{ei}italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT) and electron-neutral collision (ηensubscript𝜂𝑒𝑛\eta_{en}italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT), and is given as (Khomenko & Collados, 2012; Ni et al., 2022)

η=ηei+ηen=meνeiec2neμ0+meνenec2neμ0𝜂subscript𝜂𝑒𝑖subscript𝜂𝑒𝑛subscript𝑚𝑒subscript𝜈𝑒𝑖subscriptsuperscript𝑒2𝑐subscript𝑛𝑒subscript𝜇0subscript𝑚𝑒subscript𝜈𝑒𝑛subscriptsuperscript𝑒2𝑐subscript𝑛𝑒subscript𝜇0\displaystyle\eta=\eta_{ei}+\eta_{en}=\frac{m_{e}\nu_{ei}}{e^{2}_{c}n_{e}\mu_{% 0}}+\frac{m_{e}\nu_{en}}{e^{2}_{c}n_{e}\mu_{0}}italic_η = italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT + italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT = divide start_ARG italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG (7)

where mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the mass of electron, ecsubscript𝑒𝑐e_{c}italic_e start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the charge on electron, μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the permeability of free space, nesubscript𝑛𝑒n_{e}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the electron number density, and νensubscript𝜈𝑒𝑛\nu_{en}italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT, νeisubscript𝜈𝑒𝑖\nu_{ei}italic_ν start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT describe the collision frequencies of electron-neutral and electron-ion, respectively. The number density of electron and collision frequencies are (Ni et al., 2022)

ne=ρ(XiH+0.1XiHe)1.4misubscript𝑛𝑒𝜌subscript𝑋𝑖𝐻0.1subscript𝑋𝑖𝐻𝑒1.4subscript𝑚𝑖\displaystyle n_{e}=\frac{\rho(X_{iH}+0.1X_{iHe})}{1.4m_{i}}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = divide start_ARG italic_ρ ( italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT + 0.1 italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT ) end_ARG start_ARG 1.4 italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG (8)
νei=neec4Λ3me2ϵ02(me2πkBT)3/2,subscript𝜈𝑒𝑖subscript𝑛𝑒subscriptsuperscript𝑒4𝑐Λ3subscriptsuperscript𝑚2𝑒subscriptsuperscriptitalic-ϵ20superscriptsubscript𝑚𝑒2𝜋subscript𝑘𝐵𝑇32\displaystyle\nu_{ei}=\frac{n_{e}e^{4}_{c}\Lambda}{3m^{2}_{e}\epsilon^{2}_{0}}% \left(\frac{m_{e}}{2\pi k_{B}T}\right)^{3/2},italic_ν start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT = divide start_ARG italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_Λ end_ARG start_ARG 3 italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( divide start_ARG italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT , (9)

and

νen=nn8kBTπmenσen.subscript𝜈𝑒𝑛subscript𝑛𝑛8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑒𝑛subscript𝜎𝑒𝑛\displaystyle\nu_{en}=n_{n}\sqrt{\frac{8k_{B}T}{\pi m_{en}}}\sigma_{en}.italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT . (10)

where ΛΛ\Lambdaroman_Λ, ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, nnsubscript𝑛𝑛n_{n}italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, σensubscript𝜎𝑒𝑛\sigma_{en}italic_σ start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT denote the Coulomb logarithm, vacuum permittivity, neutrals number density, and collisional cross-section, respectively. The Coulomb logarithm is expressed as (Ni et al., 2022)

Λ=23.41.15log10ne+3.45log10T.Λ23.41.15subscript10subscript𝑛𝑒3.45subscript10𝑇\displaystyle\Lambda=23.4-1.15\log_{10}n_{e}+3.45\log_{10}T.roman_Λ = 23.4 - 1.15 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT + 3.45 roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_T . (11)

In the hydrogen-helium mixture, the collision frequency between electrons and neutrals (νensubscript𝜈𝑒𝑛\nu_{en}italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT) is contributed by both the collisions of electrons with neutral helium and neutral hydrogen. The νensubscript𝜈𝑒𝑛\nu_{en}italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT is

νen=nnHe8kBTπmeσenHe+nnH8kBTπmeσenH,subscript𝜈𝑒𝑛subscript𝑛𝑛subscript𝐻𝑒8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑒subscript𝜎𝑒𝑛subscript𝐻𝑒subscript𝑛𝑛𝐻8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑒subscript𝜎𝑒𝑛𝐻\displaystyle\nu_{en}=n_{nH_{e}}\sqrt{\frac{8k_{B}T}{\pi m_{e}}}\sigma_{e-nH_{% e}}+n_{nH}\sqrt{\frac{8k_{B}T}{\pi m_{e}}}\sigma_{e-nH},italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_n italic_H start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_n italic_H end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H end_POSTSUBSCRIPT , (12)

where nnHe=0.1ρ(1YiHe)/(1.4mi)subscript𝑛𝑛𝐻𝑒0.1𝜌1subscript𝑌𝑖𝐻𝑒1.4𝑚𝑖n_{nHe}=0.1\rho(1-Y_{iHe})/(1.4mi)italic_n start_POSTSUBSCRIPT italic_n italic_H italic_e end_POSTSUBSCRIPT = 0.1 italic_ρ ( 1 - italic_Y start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT ) / ( 1.4 italic_m italic_i ), nnH=ρ(1YiH)/(1.4mi)subscript𝑛𝑛𝐻𝜌1subscript𝑌𝑖𝐻1.4𝑚𝑖n_{nH}=\rho(1-Y_{iH})/(1.4mi)italic_n start_POSTSUBSCRIPT italic_n italic_H end_POSTSUBSCRIPT = italic_ρ ( 1 - italic_Y start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT ) / ( 1.4 italic_m italic_i ) are the number densities of neutral helium and the neutral hydrogen, respectively. In our simulations, the electron-neutral hydrogen collision cross-section is 2×10192superscript10192\times 10^{-19}2 × 10 start_POSTSUPERSCRIPT - 19 end_POSTSUPERSCRIPT m2, whereas electron-neutral helium collision cross-sections is σenH/3subscript𝜎𝑒𝑛𝐻3\sigma_{e-nH}/3italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H end_POSTSUBSCRIPT / 3 (Vranjes & Krstic, 2013). After simplification, the magnetic diffusivities becomes

ηei1.0246×108ΛT1.5similar-to-or-equalssubscript𝜂𝑒𝑖1.0246superscript108Λsuperscript𝑇1.5\displaystyle\eta_{ei}\simeq 1.0246\times 10^{8}\Lambda T^{-1.5}italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT ≃ 1.0246 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_Λ italic_T start_POSTSUPERSCRIPT - 1.5 end_POSTSUPERSCRIPT (13)

and

ηen0.0351T[0.13(1XiHe)+(1XiH)]XiH+0.1XiHesimilar-to-or-equalssubscript𝜂𝑒𝑛0.0351𝑇delimited-[]0.131subscript𝑋𝑖subscript𝐻𝑒1subscript𝑋𝑖𝐻subscript𝑋𝑖𝐻0.1subscript𝑋𝑖𝐻𝑒\displaystyle\eta_{en}\simeq 0.0351\sqrt{T}\frac{\left[\frac{0.1}{3}(1-X_{iH_{% e}})+(1-X_{iH})\right]}{X_{iH}+0.1X_{iHe}}italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT ≃ 0.0351 square-root start_ARG italic_T end_ARG divide start_ARG [ divide start_ARG 0.1 end_ARG start_ARG 3 end_ARG ( 1 - italic_X start_POSTSUBSCRIPT italic_i italic_H start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) + ( 1 - italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT ) ] end_ARG start_ARG italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT + 0.1 italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT end_ARG (14)

which are in units of m2 s-1.

The ambipolar electric field (EAD) in the induction equation (Eq. (3)) and the energy equation (Eq. (4)) is (Ni et al., 2021, 2022)

𝐄AD=1μ0ηAD[(×B)×B]×B,subscript𝐄𝐴𝐷1subscript𝜇0subscript𝜂𝐴𝐷delimited-[]𝐵𝐵𝐵\displaystyle\mathbf{E}_{AD}=\frac{1}{\mu_{0}}\eta_{AD}[(\nabla\times B)\times B% ]\times B,bold_E start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_η start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT [ ( ∇ × italic_B ) × italic_B ] × italic_B , (15)

where ηADsubscript𝜂𝐴𝐷\eta_{AD}italic_η start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT represents the coefficient of the ambipolar diffusion and is expressed as (Khomenko & Collados, 2012; Ni et al., 2020, 2022)

ηAD=(ρn/ρ)2ρiνin+ρeνensubscript𝜂𝐴𝐷superscriptsubscript𝜌𝑛𝜌2subscript𝜌𝑖subscript𝜈𝑖𝑛subscript𝜌𝑒subscript𝜈𝑒𝑛\displaystyle\eta_{AD}=\frac{(\rho_{n}/\rho)^{2}}{\rho_{i}\nu_{in}+\rho_{e}\nu% _{en}}italic_η start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT = divide start_ARG ( italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / italic_ρ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT end_ARG (16)

the ambipolar diffusion coefficient is measured in m3 s kg-1 units. For hydrogen-helium plasma, the ratio of the mass density of neutral species to total mass density is (Ni et al., 2022)

ρn/ρ=0.4(1XiHe)+(1XiH)1.4.subscript𝜌𝑛𝜌0.41subscript𝑋𝑖𝐻𝑒1subscript𝑋𝑖𝐻1.4\displaystyle\rho_{n}/\rho=\frac{0.4(1-X_{iHe})+(1-X_{iH})}{1.4}.italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / italic_ρ = divide start_ARG 0.4 ( 1 - italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT ) + ( 1 - italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT ) end_ARG start_ARG 1.4 end_ARG . (17)

The ion collision part is written as (Ni et al., 2022)

ρiνin=ρiHnnH8kBTπmi/2σiHnH+ρiHnnHe8kBT4πmi/5σiHnHe+ρiHennH8kBT4πmi/5σiHenH+ρiHennHe8kBT2πmiσiHenHe,subscript𝜌𝑖subscript𝜈𝑖𝑛absentsubscript𝜌𝑖𝐻subscript𝑛𝑛𝐻8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑖2subscript𝜎𝑖𝐻𝑛𝐻missing-subexpressionsubscript𝜌𝑖𝐻subscript𝑛𝑛𝐻𝑒8subscript𝑘𝐵𝑇4𝜋subscript𝑚𝑖5subscript𝜎𝑖𝐻𝑛𝐻𝑒missing-subexpressionsubscript𝜌𝑖𝐻𝑒subscript𝑛𝑛𝐻8subscript𝑘𝐵𝑇4𝜋subscript𝑚𝑖5subscript𝜎𝑖𝐻𝑒𝑛𝐻missing-subexpressionsubscript𝜌𝑖𝐻𝑒subscript𝑛𝑛𝐻𝑒8subscript𝑘𝐵𝑇2𝜋subscript𝑚𝑖subscript𝜎𝑖𝐻𝑒𝑛𝐻𝑒\displaystyle\begin{aligned} \rho_{i}\nu_{in}&=\rho_{iH}n_{nH}\sqrt{\frac{8k_{% B}T}{\pi m_{i}/2}}\sigma_{iH-nH}\\ &+\rho_{iH}n_{nHe}\sqrt{\frac{8k_{B}T}{4\pi m_{i}/5}}\sigma_{iH-nHe}\\ &+\rho_{iHe}n_{nH}\sqrt{\frac{8k_{B}T}{4\pi m_{i}/5}}\sigma_{iHe-nH}\\ &+\rho_{iHe}n_{nHe}\sqrt{\frac{8k_{B}T}{2\pi m_{i}}}\sigma_{iHe-nHe},\end{aligned}start_ROW start_CELL italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT end_CELL start_CELL = italic_ρ start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / 2 end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_ρ start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H italic_e end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG 4 italic_π italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / 5 end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H italic_e end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_ρ start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG 4 italic_π italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / 5 end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_ρ start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H italic_e end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG 2 italic_π italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H italic_e end_POSTSUBSCRIPT , end_CELL end_ROW (18)

where ρiH=ρXiH/1.4subscript𝜌𝑖𝐻𝜌subscript𝑋𝑖𝐻1.4\rho_{iH}=\rho X_{iH}/1.4italic_ρ start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT = italic_ρ italic_X start_POSTSUBSCRIPT italic_i italic_H end_POSTSUBSCRIPT / 1.4, ρiHe=0.4ρXiHe/1.4subscript𝜌𝑖𝐻𝑒0.4𝜌subscript𝑋𝑖𝐻𝑒1.4\rho_{iHe}=0.4\rho X_{iHe}/1.4italic_ρ start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT = 0.4 italic_ρ italic_X start_POSTSUBSCRIPT italic_i italic_H italic_e end_POSTSUBSCRIPT / 1.4 are the mass densities of ionized hydrogen and helium, respectively; σiHnHsubscript𝜎𝑖𝐻𝑛𝐻\sigma_{iH-nH}italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H end_POSTSUBSCRIPT represents the collision cross-section of ionized hydrogen and neutral hydrogen, σiHnHesubscript𝜎𝑖𝐻𝑛𝐻𝑒\sigma_{iH-nHe}italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H italic_e end_POSTSUBSCRIPT is the collisional cross-section between ionized hydrogen and neutral helium, σiHenHsubscript𝜎𝑖𝐻𝑒𝑛𝐻\sigma_{iHe-nH}italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H end_POSTSUBSCRIPT is the cross-section for the ionized helium and neutral hydrogen collision, and σiHenHesubscript𝜎𝑖𝐻𝑒𝑛𝐻𝑒\sigma_{iHe-nHe}italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H italic_e end_POSTSUBSCRIPT corresponds to that of the ionized helium-neutral helium collision cross-section. We consider σiHnH=1.5×1018subscript𝜎𝑖𝐻𝑛𝐻1.5superscript1018\sigma_{iH-nH}=1.5\times 10^{-18}italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H end_POSTSUBSCRIPT = 1.5 × 10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT m2, σiHnHesubscript𝜎𝑖𝐻𝑛𝐻𝑒\sigma_{iH-nHe}italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H italic_e end_POSTSUBSCRIPT = σiHenHsubscript𝜎𝑖𝐻𝑒𝑛𝐻\sigma_{iHe-nH}italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H end_POSTSUBSCRIPT = σiHenHesubscript𝜎𝑖𝐻𝑒𝑛𝐻𝑒\sigma_{iHe-nHe}italic_σ start_POSTSUBSCRIPT italic_i italic_H italic_e - italic_n italic_H italic_e end_POSTSUBSCRIPT = σiHnH/3subscript𝜎𝑖𝐻𝑛𝐻3\sigma_{iH-nH}/\sqrt{3}italic_σ start_POSTSUBSCRIPT italic_i italic_H - italic_n italic_H end_POSTSUBSCRIPT / square-root start_ARG 3 end_ARG (Vranjes & Krstic, 2013; Barata & Conde, 2010). The electron collision part reads (Ni et al., 2022)

ρeνen=ρennH8kBTπmeσenH+ρennHe8kBTπmeσenHe,subscript𝜌𝑒subscript𝜈𝑒𝑛absentsubscript𝜌𝑒subscript𝑛𝑛𝐻8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑒subscript𝜎𝑒𝑛𝐻missing-subexpressionsubscript𝜌𝑒subscript𝑛𝑛𝐻𝑒8subscript𝑘𝐵𝑇𝜋subscript𝑚𝑒subscript𝜎𝑒𝑛𝐻𝑒\displaystyle\begin{aligned} \rho_{e}\nu_{en}&=\rho_{e}n_{nH}\sqrt{\frac{8k_{B% }T}{\pi m_{e}}}\sigma_{e-nH}\\ &+\rho_{e}n_{nHe}\sqrt{\frac{8k_{B}T}{\pi m_{e}}}\sigma_{e-nHe},\\ \end{aligned}start_ROW start_CELL italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT end_CELL start_CELL = italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_ρ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_n italic_H italic_e end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 8 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG end_ARG italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H italic_e end_POSTSUBSCRIPT , end_CELL end_ROW (19)

where σenHsubscript𝜎𝑒𝑛𝐻\sigma_{e-nH}italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H end_POSTSUBSCRIPT, σenHesubscript𝜎𝑒𝑛𝐻𝑒\sigma_{e-nHe}italic_σ start_POSTSUBSCRIPT italic_e - italic_n italic_H italic_e end_POSTSUBSCRIPT are the collisional cross-sections of electrons with neutral hydrogen and neutral helium species, respectively. The collisional cross-sections contributed by electrons and neutrals are smaller than the collisional cross-section between ions and neutrals and are ignored. The coefficient of ambipolar diffusion (Eq. (16)) is simplified into ηAD=(ρn/ρ)2ρiνinsubscript𝜂𝐴𝐷superscriptsubscript𝜌𝑛𝜌2subscript𝜌𝑖subscript𝜈𝑖𝑛\eta_{AD}=\frac{(\rho_{n}/\rho)^{2}}{\rho_{i}\nu_{in}}italic_η start_POSTSUBSCRIPT italic_A italic_D end_POSTSUBSCRIPT = divide start_ARG ( italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT / italic_ρ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT end_ARG.

In partially ionized plasma, the dynamic viscosity is contributed by both ions and neutrals, and is given by (Ni et al., 2022)

ξ=ξi+ξn=nikBTνii+nnkBTνnn,𝜉subscript𝜉𝑖subscript𝜉𝑛subscript𝑛𝑖subscript𝑘𝐵𝑇subscript𝜈𝑖𝑖subscript𝑛𝑛subscript𝑘𝐵𝑇subscript𝜈𝑛𝑛\displaystyle\xi=\xi_{i}+\xi_{n}=\frac{n_{i}k_{B}T}{\nu_{ii}}+\frac{n_{n}k_{B}% T}{\nu_{nn}},italic_ξ = italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ξ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_ν start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT end_ARG , (20)

where ξisubscript𝜉𝑖\xi_{i}italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, ξnsubscript𝜉𝑛\xi_{n}italic_ξ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT are the viscosity coefficients contributed by ion and neutrals, respectively, while νnn,iisubscript𝜈𝑛𝑛𝑖𝑖\nu_{nn,ii}italic_ν start_POSTSUBSCRIPT italic_n italic_n , italic_i italic_i end_POSTSUBSCRIPT corresponds to the collision frequencies between similar species, such as collisions between neutral-neutral and ion-ion. The collision frequencies between the similar species are as below (Leake et al., 2013):

νnn=nnσnn16kBTπmnsubscript𝜈𝑛𝑛subscript𝑛𝑛subscript𝜎𝑛𝑛16subscript𝑘𝐵𝑇𝜋subscript𝑚𝑛\displaystyle\nu_{nn}=n_{n}\sigma_{nn}\sqrt{\frac{16k_{B}T}{\pi m_{n}}}italic_ν start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT square-root start_ARG divide start_ARG 16 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG start_ARG italic_π italic_m start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG end_ARG (21)

and

νii=niec4Λ3mi2ϵ02(mi2πkBT)3/2.subscript𝜈𝑖𝑖subscript𝑛𝑖subscriptsuperscript𝑒4𝑐Λ3subscriptsuperscript𝑚2𝑖subscriptsuperscriptitalic-ϵ20superscriptsubscript𝑚𝑖2𝜋subscript𝑘𝐵𝑇32\displaystyle\nu_{ii}=\frac{n_{i}e^{4}_{c}\Lambda}{3m^{2}_{i}\epsilon^{2}_{0}}% \left(\frac{m_{i}}{2\pi k_{B}T}\right)^{3/2}.italic_ν start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT = divide start_ARG italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_Λ end_ARG start_ARG 3 italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( divide start_ARG italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT . (22)

The contributions of hydrogen and helium are both considered when calculating the dynamic viscosity, which is assumed to be more realistic than the one used in previous study (Ni et al., 2022). The final form of dynamic viscosity used here is

ξ=ξi+ξn4.8692×1016ΛT2T+2.0127×107T.𝜉subscript𝜉𝑖subscript𝜉𝑛similar-to-or-equals4.8692superscript1016Λsuperscript𝑇2𝑇2.0127superscript107𝑇\displaystyle\xi=\xi_{i}+\xi_{n}\simeq\frac{4.8692\times 10^{-16}}{\Lambda}T^{% 2}\sqrt{T}+2.0127\times 10^{-7}\sqrt{T}.italic_ξ = italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_ξ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≃ divide start_ARG 4.8692 × 10 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Λ end_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT square-root start_ARG italic_T end_ARG + 2.0127 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT square-root start_ARG italic_T end_ARG . (23)

In our simulations, we use time-dependent ionization degrees for hydrogen and helium. The ionization degree of helium varies exponentially with temperature, X=iHe1100.3255710.0000596T{}_{iHe}=1-10^{0.325571-0.0000596T}start_FLOATSUBSCRIPT italic_i italic_H italic_e end_FLOATSUBSCRIPT = 1 - 10 start_POSTSUPERSCRIPT 0.325571 - 0.0000596 italic_T end_POSTSUPERSCRIPT. This expression is valid for both the photosphere and chromosphere layers, but it gives a negative value when the temperature is below 5413K. Therefore, we consider X=iHe0.00010084814{}_{iHe}=0.00010084814start_FLOATSUBSCRIPT italic_i italic_H italic_e end_FLOATSUBSCRIPT = 0.00010084814 when the temperature is less than that critical value. On the other hand, different approaches are employed to calculate the ionization degree of hydrogen (XiH) in the photosphere and chromosphere regions, respectively. The modified Saha and Boltzamann equation (Gan & Fang, 1990; Fang et al., 2002) is employed to calculate hydrogen ionization degree in the photosphere, which depends on plasma density and temperature. The temperature dependent hydrogen ionization degree for the chromospheric environment is adopted from Carlsson & Leenaarts work (Carlsson & Leenaarts, 2012). We refer interested readers to see Figure 1 of Ni et al. 2022 (Ni et al., 2022) and Liu et al. 2023 (Liu et al., 2023), which show temperature-dependent faction of hydrogen and helium. At the beginning, the uniform initial plasma density and temperature profiles produce a constant value of ionization degrees across the simulation domain; however, later on, these ionization degrees are updated based on the local plasma parameters.

2.3 Radiative Cooling models

Radiative transfer processes, or the energy-exchange processes between the solar atmosphere and the radiation field, have a significant impact on the properties and behavior of the plasma especially in the low solar atmosphere. In order to make our numerical experiments more realistic and better understand magnetic reconnection in the lower solar atmosphere, we included different radiative cooling models in the photosphere and chromosphere layers.

For effective radiative cooling process in the photosphere layer, the Gan & Fang (Gan & Fang, 1990) is applied in the photospheric magnetic reconnection process, read as

Qrad1=1.547×1042nenHαT1.5,subscript𝑄𝑟𝑎𝑑11.547superscript1042subscript𝑛𝑒subscript𝑛𝐻𝛼superscript𝑇1.5\displaystyle Q_{rad1}=-1.547\times 10^{-42}n_{e}n_{H}\alpha T^{1.5},italic_Q start_POSTSUBSCRIPT italic_r italic_a italic_d 1 end_POSTSUBSCRIPT = - 1.547 × 10 start_POSTSUPERSCRIPT - 42 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT italic_α italic_T start_POSTSUPERSCRIPT 1.5 end_POSTSUPERSCRIPT , (24)

where ne, nH are the electron and hydrogen number densities, respectively. The modified Saha and Boltzamann formula  (Gan & Fang, 1990; Fang et al., 2002) is used to calculate the number density of electron. The height dependent function α𝛼\alphaitalic_α is given by  (Gan & Fang, 1990)

α=10a1+2.3738×104ea2,a1=2.75×103Z5.445,a2=Z163𝛼absentsuperscript10𝑎12.3738superscript104superscript𝑒𝑎2𝑎1absent2.75superscript103𝑍5.445𝑎2absent𝑍163\displaystyle\begin{aligned} \alpha&=10^{a1}+2.3738\times 10^{-4}e^{a2},\\ a1&=2.75\times 10^{-3}Z-5.445,\\ a2&=\frac{-Z}{163}\end{aligned}start_ROW start_CELL italic_α end_CELL start_CELL = 10 start_POSTSUPERSCRIPT italic_a 1 end_POSTSUPERSCRIPT + 2.3738 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_a 2 end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_a 1 end_CELL start_CELL = 2.75 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_Z - 5.445 , end_CELL end_ROW start_ROW start_CELL italic_a 2 end_CELL start_CELL = divide start_ARG - italic_Z end_ARG start_ARG 163 end_ARG end_CELL end_ROW (25)

where Z𝑍Zitalic_Z represents height in km.

The second radiative cooling model used in the chromospheric simulations of this study is the model proposed by Carlsson & Leenaarts (Carlsson & Leenaarts, 2012), which is regarded as the widely acceptable model for the chromosphere. This model focuses on the effects of radiative cooling in the three chromospheric spectral lines, such as neutral hydrogen, singly ionized calcium, and singly ionized magnesium. The model is expressed as below

Qrad2=X=H,Mg,CaLXm(T)EXm(τ)NXmNX(T)AXNHρneρ,subscript𝑄𝑟𝑎𝑑2subscript𝑋𝐻𝑀𝑔𝐶𝑎subscript𝐿𝑋𝑚𝑇subscript𝐸𝑋𝑚𝜏subscript𝑁𝑋𝑚subscript𝑁𝑋𝑇subscript𝐴𝑋subscript𝑁𝐻𝜌subscript𝑛𝑒𝜌\displaystyle Q_{rad2}=-\sum_{X=H,Mg,Ca}L_{Xm}(T)E_{Xm}(\tau)\frac{N_{Xm}}{N_{% X}}(T)A_{X}\frac{N_{H}}{\rho}n_{e}\rho,italic_Q start_POSTSUBSCRIPT italic_r italic_a italic_d 2 end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_X = italic_H , italic_M italic_g , italic_C italic_a end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT ( italic_T ) italic_E start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT ( italic_τ ) divide start_ARG italic_N start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_ARG ( italic_T ) italic_A start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT divide start_ARG italic_N start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ end_ARG italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_ρ , (26)

where LXm(T)subscript𝐿𝑋𝑚𝑇L_{Xm}(T)italic_L start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT ( italic_T ) is the optically thin radiative loss function that varies with temperature, per electron, and per particle of element X𝑋Xitalic_X in the ionization stage m𝑚mitalic_m, EXm(τ)subscript𝐸𝑋𝑚𝜏E_{Xm}(\tau)italic_E start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT ( italic_τ ) denotes the escape probability that depends on optical depth τ𝜏\tauitalic_τ, NXmNX(T)subscript𝑁𝑋𝑚subscript𝑁𝑋𝑇\frac{N_{Xm}}{N_{X}}(T)divide start_ARG italic_N start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_ARG ( italic_T ) represents the fraction of element X𝑋Xitalic_X in the ionization stage m𝑚mitalic_m, AXsubscript𝐴𝑋A_{X}italic_A start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT denotes the element abundance X𝑋Xitalic_X, and NHρ=4.407×1023g1subscript𝑁𝐻𝜌4.407superscript1023superscript𝑔1\frac{N_{H}}{\rho}=4.407\times 10^{23}g^{-1}divide start_ARG italic_N start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ end_ARG = 4.407 × 10 start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT italic_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT shows the number of hydrogen particles per unit mass of the chromospheric material. The τ𝜏\tauitalic_τ in the above expression (eq. 26) is calculated by multiplying the total column density of neutral hydrogen with a constant number of 4.0×10144.0superscript10144.0\times 10^{-14}4.0 × 10 start_POSTSUPERSCRIPT - 14 end_POSTSUPERSCRIPT cm2. The RADYN (Carlsson & Stein, 1997, 2002) code computed LXmsubscript𝐿𝑋𝑚L_{Xm}italic_L start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT, EXmsubscript𝐸𝑋𝑚E_{Xm}italic_E start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT, and NXm/NXsubscript𝑁𝑋𝑚subscript𝑁𝑋N_{Xm}/N_{X}italic_N start_POSTSUBSCRIPT italic_X italic_m end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT for hydrogen in a 1D radiation hydrodynamic simulation with non-equilibrium ionization. However, the components contributed by Mg and Ca were obtained from a 2D radiation-MHD simulation with BIFROST, that gave the atmospheric structure and radiative transfer calculations employing MULTI3D (Leenaarts & Carlsson, 2009).

2.4 Simulation setup

In this study, we carried out high-resolution 2.5D MHD simulations to investigate the magnetic reconnection process at different altitude in the low solar atmosphere with an anti-parallel magnetic field configuration. The initial magnetic fields of the Harris current sheet are given by Bx0=b0tanh[y/(0.05L0)]subscript𝐵𝑥0subscript𝑏0𝑡𝑎𝑛delimited-[]𝑦0.05subscript𝐿0B_{x0}=-b_{0}tanh[y/(0.05L_{0})]italic_B start_POSTSUBSCRIPT italic_x 0 end_POSTSUBSCRIPT = - italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_t italic_a italic_n italic_h [ italic_y / ( 0.05 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ], By0=0subscript𝐵𝑦00B_{y0}=0italic_B start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT = 0 and Bz0=b0/cosh[y/(0.05L0)]subscript𝐵𝑧0subscript𝑏0𝑐𝑜𝑠delimited-[]𝑦0.05subscript𝐿0B_{z0}=b_{0}/cosh[y/(0.05L_{0})]italic_B start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_c italic_o italic_s italic_h [ italic_y / ( 0.05 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ] (b0subscript𝑏0b_{0}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, L0=2×105msubscript𝐿02superscript105𝑚L_{0}=2\times 10^{5}mitalic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_m are the characteristics of magnetic field strength and system scale length, respectively). The size of the simulation box is from 00 to L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT along the x-direction and 0.5L00.5subscript𝐿0-0.5L_{0}- 0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to 0.5L00.5subscript𝐿00.5L_{0}0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT along the y-direction which is large enough to allow nonlinear evolution and motion of plasmoids. The small initial magnetic perturbations that trigger the magnetic reconnection are bx1=bpertsin[2π(y+0.5L0)/L0]cos[2π(x+0.5L0)/L0]subscript𝑏𝑥1subscript𝑏𝑝𝑒𝑟𝑡𝑠𝑖𝑛delimited-[]2𝜋𝑦0.5subscript𝐿0subscript𝐿0𝑐𝑜𝑠delimited-[]2𝜋𝑥0.5subscript𝐿0subscript𝐿0b_{x1}=-b_{pert}sin[2\pi(y+0.5L_{0})/L_{0}]cos[2\pi(x+0.5L_{0})/L_{0}]italic_b start_POSTSUBSCRIPT italic_x 1 end_POSTSUBSCRIPT = - italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT italic_s italic_i italic_n [ 2 italic_π ( italic_y + 0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] italic_c italic_o italic_s [ 2 italic_π ( italic_x + 0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] and by1=bpertcos[2π(y+0.5L0)/L0]sin[2π(x+0.5L0)/L0]subscript𝑏𝑦1subscript𝑏𝑝𝑒𝑟𝑡𝑐𝑜𝑠delimited-[]2𝜋𝑦0.5subscript𝐿0subscript𝐿0𝑠𝑖𝑛delimited-[]2𝜋𝑥0.5subscript𝐿0subscript𝐿0b_{y1}=b_{pert}cos[2\pi(y+0.5L_{0})/L_{0}]sin[2\pi(x+0.5L_{0})/L_{0}]italic_b start_POSTSUBSCRIPT italic_y 1 end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT italic_c italic_o italic_s [ 2 italic_π ( italic_y + 0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] italic_s italic_i italic_n [ 2 italic_π ( italic_x + 0.5 italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]. Open boundary conditions are used at both directions in all simulation cases. The adaptive mesh refinement (AMR) technique with the highest level of 9 and a base-level grid of 192 ×\times× 192 are applied in this study. The initial plasma density (ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) and temperature (T0) given in Table 1 are taken from C7 model (Avrett & Loeser, 2008). The parameter Z in Table 1 represent the height above the solar surface in kilometers. The simulation at 400 km corresponds to the photosphere layer of the Sun. The cases at Z = 600 and 800 km relate to the lower chromosphere. The heights from 1000 km to 1400 km correspond to the middle chromosphere, while the 1700 km to 2000 km range represents the upper chromosphere region. The simulations in the middle chromosphere (Z = 1250 and 1400 km) are performed with different initial magnetic perturbations (bpert) and plasma β𝛽\betaitalic_β (β0)\beta_{0})italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). In our simulations, we changed β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT by varying the strength of the initial magnetic field.

Table 1: Input parameters used in the simulations.
{ruledtabular}
Z (km) ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT(kgm-3) T0 (K)
400 1.56×105absentsuperscript105\times 10^{-5}× 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 4590
600 2.40×106absentsuperscript106\times 10^{-6}× 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 4421
800 3.44×107absentsuperscript107\times 10^{-7}× 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT 5100
1000 6.32×108absentsuperscript108\times 10^{-8}× 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT 6223
1250 1.16×108absentsuperscript108\times 10^{-8}× 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT 6588
1400 4.51×1009absentsuperscript1009\times 10^{-09}× 10 start_POSTSUPERSCRIPT - 09 end_POSTSUPERSCRIPT 6610
1700 7.44×1010absentsuperscript1010\times 10^{-10}× 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 6641
2000 1.68×1010absentsuperscript1010\times 10^{-10}× 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT 6678
Refer to caption

(a)

Refer to caption

(b)

Refer to caption

(c)

Figure 1: Time evolution of the reconnection rate γ𝛾\gammaitalic_γ for different initial perturbations bpertsubscript𝑏𝑝𝑒𝑟𝑡b_{pert}italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT at Z = 1250 km (solid lines) and Z = 1400 km (dash-dotted lines) with β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.05 (a), for different vales of initial plasma β𝛽\betaitalic_β at Z = 1250 km (solid lines) and Z = 1400 km (dash-dotted lines) with bpert = 0.009 (b) and at different altitudes in the low solar atmosphere using the same initial plasma β𝛽\betaitalic_β (β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5) and initial perturbation (bpertsubscript𝑏𝑝𝑒𝑟𝑡b_{pert}italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT = 0.009). The dash lines correspond to photospheric reconnection cases (c).
Refer to caption

(a)

Refer to caption

(b)

Refer to caption

(c)

Figure 2: Temporal evolutions of the temperature, pressure, density at the main X-point and the reconnection rate during the explosive fast reconnection phase (a), diffusion coefficients ηensubscript𝜂𝑒𝑛\eta_{en}italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT and ηeisubscript𝜂𝑒𝑖\eta_{ei}italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT at the main X-point in the newly observed explosive reconnection stage (b) and the profiles of the logarithm of the total magnetic diffusion (ηensubscript𝜂𝑒𝑛\eta_{en}italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT+ηeisubscript𝜂𝑒𝑖\eta_{ei}italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT) at various times along the y-direction through the main X-point in the case with β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.05 and bpertsubscript𝑏𝑝𝑒𝑟𝑡b_{pert}italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT = 0.005 at Z=1250 km above the solar surface.
Refer to caption

(a)(i)(ii)(iii)(iv)

Refer to caption

(b)

Refer to caption

(c)

Refer to caption

(d)

Refer to caption

(e)

Refer to caption

(f)

Refer to caption

(g)

Figure 3: The distributions of magnetic field (solid lines) and out-of-plane electric current density jz at four different time points during the magnetic reconnection process in the case with β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.05 and bpert = 0.005 at Z=1250 km above the solar surface (a), Distribution of the primary variables along the vertical line in the third panel of jz which crosses a pair of shocks (b)-(g). Bn (and Vn) is the normal component of the magnetic field (and velocity), and Bt (and Vt) is transverse component of the shock front, respectively. The black vertical lines in (b)-(g) represent the positions of the shock front. The red boxes in (a) are the regions for calculating the reconnection rates.

3 Simulation results

Figure 1(a) shows time evolution of the reconnection rate, γ𝛾\gammaitalic_γ = Vy-aver/VA-aver (Zafar et al., 2023), for different initial perturbations (bpert) at middle chromospheric altitudes. Vy-aver and VA-aver are the average inflow and Alfven velocities within the red boxes in Figure 3(a). For the reconnection rate calculation, we used boxes of different sizes near the principal reconnection X-point. We found that the box size affects the amplitudes of reconnection rates slightly but has no effect on the evolution trend. The main conclusion does not depend on the choices of the boxes used to calculate the reconnection rate. The solid and dash-dotted lines correspond to Z = 1250 km and 1400 km, respectively. Like a typical reconnection process, an initial elongated Sweet-Parker phase with a reconnection rate of the order of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT is followed by the faster plasmoid-mediated phase, where the reconnection rate reaches to around 0.02 (Ni et al., 2015). Then another abrupt explosive reconnection process occurs, the reconnection rate is additionally increased to a higher value. The maximum value reaches to around 0.06 or above in the explosive Petschek-like stage and then drop back to similar-to\sim0.02. The explosive Petschek-like phase is followed by a plateau phase, where the reconnection rate remains at its maximum value for a period of time. The plateau phase persists longer for small bpert, but its duration decreases with increasing initial magnetic perturbation. The onset time of explosive reconnection is also affected by the strength of initial magnetic perturbation, which increases by decreasing bpert. The case with a high initial perturbation bpert = 0.03 (red line in Figure 1(a)) shows several sharp spikes in the reconnection rate. The initial spike in the reconnection rate right after the start of the simulation (red solid line in Figure 1(a)), where the reconnection rate reaches around 0.02, is due to the strong initial perturbation in this simulation case. However, the other tall short-lived spikes in the reconnection rate at t similar-to\sim 14 sec and 16 sec correspond to explosive Petschek-like reconnection, indicating that this is an intermittent process.

Next, the effects of initial plasma β𝛽\betaitalic_β (β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) and plasma parameters at different altitudes (Table 1) on the dynamics of this newly observed explosive fast intermittent reconnection events are explored. Figure 1(b) shows that an explosive Petschek-like reconnection phase occurs at both middle chromospheric altitudes, regardless of the initial plasma β𝛽\betaitalic_β value. Simulations with higher β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT have a longer plateau phase, whereas those with lower β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT have a shorter one. The explosive reconnection phase is triggered earlier when strength of the initial magnetic field is stronger, and the initial plasma β𝛽\betaitalic_β is lower. The time evolution of the reconnection rate at various altitudes, from the photosphere to the top of the chromosphere with the same initial plasma β𝛽\betaitalic_β (β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5) and perturbation (bpert = 0.009) shows that the reconnection rate peaks-up violently and attains the values of about 0.06 or above (Figure 1(c)).

The fast reconnection phase is observed in all simulation cases, suggesting that the transition from plasmoid-dominated to the explosive faster Petschek-like reconnection phase are independent of magnetic perturbations (bpert), initial plasma β𝛽\betaitalic_β (β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) and physical parameters (T0,ρ0subscript𝑇0subscript𝜌0T_{0},\rho_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) of various altitudes in the lower solar atmosphere. However, the difference in bpert and β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT effect the duration of the plateau phase and the timing of the explosive reconnection regime onset. Here we want to mention that no explosive reconnection phase has been observed at Z = 2000 km in the presence of ambipolar diffusion with β0=1.33subscript𝛽01.33\beta_{0}=1.33italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.33 and b=pert0.005{}_{pert}=0.005start_FLOATSUBSCRIPT italic_p italic_e italic_r italic_t end_FLOATSUBSCRIPT = 0.005 in our previous study (Zafar et al., 2023). To clarify the dynamics of explosive faster reconnection phase at Z = 2000 km above the solar surface, we performed simulations with different initial plasma β𝛽\betaitalic_β (β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.3, 3.0 and 5.0) and initial magnetic perturbations. Additionally, we carried out a simulation with the same initial plasma β𝛽\betaitalic_β (β0=1.33subscript𝛽01.33\beta_{0}=1.33italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.33) as in Case-VI from the previous study, but with a perturbation of b=pert0.001{}_{pert}=0.001start_FLOATSUBSCRIPT italic_p italic_e italic_r italic_t end_FLOATSUBSCRIPT = 0.001. An explosive Petschek-like reconnection phase is observed in all these simulations with different β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and bpert at Z = 2000 km (not shown here). We reviewed Case-VI (Z = 2000 km, β0subscript𝛽0\beta_{0}italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.33, b=pert0.005{}_{pert}=0.005start_FLOATSUBSCRIPT italic_p italic_e italic_r italic_t end_FLOATSUBSCRIPT = 0.005) from our previous study  (Zafar et al., 2023) and found that this simulation case terminated earlier, which could reveal why an explosive faster reconnection phase was not observed. Therefore, we expect that the plasmoid-mediated phase in Case-VI of the previous study would have transformed into an explosive Petschek-like reconnection phase, if the simulations had been run for a longer duration.

To explore the underlying trigger mechanism of explosive fast reconnection, the temporal evolution of various plasma parameters are analyzed during the explosive fast reconnection stage in each case, Figure 2(a) shows the results in the case with β0=0.05subscript𝛽00.05\beta_{0}=0.05italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.05, bpert=0.005subscript𝑏𝑝𝑒𝑟𝑡0.005b_{pert}=0.005italic_b start_POSTSUBSCRIPT italic_p italic_e italic_r italic_t end_POSTSUBSCRIPT = 0.005 at Z=1250 km (green solid line in Figure 1(a)) above the solar surface. It is quite evident that the decrease in plasma temperature, density and pressure at the main X-point occurs just before the onset of explosive fast reconnection. Rapid drop in plasma temperature and density cause a reduction in gas pressure at the X-point, an increase in the plasma inflow velocity and leads to a spontaneous fast magnetic reconnection. After careful analysis of the simulation results, we identified several reasons for the decrease in plasma density and temperature at the primary X-point. These include radiative cooling, motion of the primary X-point along the current sheet and ejection of hot plasma from the primary X-point. Radiative cooling restricts the increase in plasma temperature during the reconnection process and has a significant impact on the distribution of plasma density and temperature. In some cases, such as photospheric and lower chromospheric reconnection events, it has a substantial contribution to decrease plasma temperature. In the non-linear stage, the distributions of plasma temperature and density along the current sheet is non-uniform and there is a possibility that the X-point moves from the hot and dense region to the cool and less dense one, leading to faster decrease in temperature and density. Another possibility is the ejection of hot plasma from the main X-point and its surroundings, resulting in a sudden decrease of plasma temperature, density and hence pressure. Two drastic spikes in the reconnection rate at Z = 2000 km (see the cyan line in Figure 1(c)) at t = 85 sec and 105 sec corresponds to temperature drop caused by the motion of the X-point and the ejection of hot plasma from the surrounding of X-point, respectively.

The magnetic diffusion caused by electron-ion collision, ηeisubscript𝜂𝑒𝑖\eta_{ei}italic_η start_POSTSUBSCRIPT italic_e italic_i end_POSTSUBSCRIPT and electron-neutral collision, ηensubscript𝜂𝑒𝑛\eta_{en}italic_η start_POSTSUBSCRIPT italic_e italic_n end_POSTSUBSCRIPT both increase significantly when the temperature at the X-point drops during t = 28.5 sec to t = 31.5 sec in the case with β0=0.05subscript𝛽00.05\beta_{0}=0.05italic_β start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.05 and bpert = 0.005 at Z = 1250 km above the solar surface (Figure 2(b)). The distributions of the logarithm of the total magnetic diffusion through the main X-point along the y-direction at different times are shown in Figure 2(c). It is observed that the magnetic diffusion at the main X-point is actually much smaller than that of outside the current sheet. But, the local Petschek-like reconnection still happens.

Figure 3(a) shows the 2D distributions of the current density Jz (background color) and field lines (black solid lines) at various stages of the reconnection process. After the Sweet-Parker reconnection stage, the current sheet become unstable and leads to multiple plasmoids. With the occurrence of the plasmoid instability the reconnection rate peaks up and reaches to a value of about 0.02. One such stage having a chain of magnetic islands is shown in Figure 3a(i), at the same time the reconnection rate is about similar-to\sim 0.02 (green line in Figure 1(a)). Subsequently, all magnetic islands wiped out with time and the current sheet becomes bifurcated (Figures 3a(ii) and (iii)). In the proximity of X-point, the magnetic field lines have V-shaped structure while it possesses W-shaped structure when far away from the X-point. These are important characteristic of Petschek like scenario. The peak of magnetic reconnection (t = 28.5-31.5 s, green line in Figure 1(a)) corresponds to the phase where the current sheet dynamics is Petschek like. At about t = 31.5 s, the current sheet again becomes unstable and breakup into plasmoids (Figure 3a(iv)). The reconnection rate drops again to around 0.02 (t>>>31.5 s, green line in Figure 1(a)) with the generation of these new plasmoids. The simulation with initial perturbation of 0.03 (red line in Figure 1(a)) clearly shows that this is a cyclic process, and the reconnection rate fluctuates between 0.07 and 0.02.

To examine Petschek shocks, the distributions of temperature (T)𝑇(T)( italic_T ), mass density (ρ)𝜌(\rho)( italic_ρ ), gas pressure (P)𝑃(P)( italic_P ), out of plane current density (Jz)subscript𝐽𝑧(J_{z})( italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ), magnetic fields (B)𝐵(B)( italic_B ) and velocity (V)𝑉(V)( italic_V ) along the dotted black sampling line in Figure 3a(iii) are illustrated in Figures 3(b)-3(g). It is evident that most variables changes drastically as it crosses the shock fronts marked by dotted lines in the Figures 3(b)-3(g). The absolute values of Bt and Vn decreases, Bn does not vary much, and the other physical quantities increase from the upstream to the downstream regions. Such discontinuous structure have been investigated in many previous studies (Sato, 1979; Ugai & Tsuda, 1979; Mei et al., 2012), where the slow mode shocks are confirmed to be formed. This analysis reveals that the system changes from plasmoid-mediated phase to the Petschek-like reconnection (t similar-to\sim 28.5s-31.5 green solid line in Figure 1(a)) and hence leading to explosive faster reconnection with the rate \geq 0.06 becomes established.

4 Discussion and Conclusion

We have investigated the magnetic reconnection phenomena at various altitudes in the partially ionized low solar atmosphere with different initial conditions based on the single-fluid MHD model. The time-dependent ionizations of the partially ionized helium-hydrogen plasmas and suitable radiative cooling models are considered in these numerical experiments. The results presented in this work reported for the first time that the plasmoid-dominated fast reconnection phase make transition to an explosive faster Petschek-like reconnection phase in partially ionized plasmas.

The careful analysis has revealed that the sudden decrease of plasma temperature and density at the main X-point are responsible for sudden drop of plasma pressure, leading to an explosive fast reconnection rate. The bifurcated shape of the local current sheet and existence of pair of slow mode shocks during the newly observed faster reconnection phase clearly proved the occurrence of the Petschek-like reconnection. The reconnection rate reached to 0.06 or above during the explosive phase in all these simulation experiments, which is almost three times higher than that during the plasmoid-dominated reconnection phase. The reconnection rate drops back to 0.02similar-toabsent0.02\sim 0.02∼ 0.02 when the local current sheet with the primary X-point gets elongated and new plasmoids are generated. The occurrence of such a faster reconnection phase is independent of initial plasma β𝛽\betaitalic_β, magnetic perturbation and plasma parameters at different altitudes. However, the timing of the explosive reconnection regime onset, and the duration of the plateau phase is much smaller in the lower β𝛽\betaitalic_β and higher perturbation cases.

The Petschek-like reconnection observed in our simulations differs from the classical Petschek reconnection model, that is derived based on the static state MHD equations. The classical one is a steady reconnection model, and the derived reconnection rate scales with Lundquist number S as γsimilar-to𝛾absent\gamma\simitalic_γ ∼ 1/logS. Therefore, the reconnection rate predicated by the classical Petschek model is 0.1 for a typical Lundquist number S 1010similar-toabsentsuperscript1010\sim 10^{10}∼ 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT of a flare current sheet in the corona, which is sufficiently fast to explain the timescale of solar flares. However, MHD simulations failed to obtain a Petschek solution for a uniform resistivity profile. The Petschek solution can be physically realized only when there is a localized anomalous resistivity at the X-point. It is worth nothing that in order to produce anomalous resistivity, the length scale of the current sheet is comparable to the ion Larmor radius or the ion inertial length, hence non-MHD (collisionless effects) are likely to become important before steady-state Petschek reconnection is realized (Zweibel & Yamada, 2009). In addition, such a steady reconnection process cannot well match the explosive reconnection process with bursty radiation enhancements. However, the Petschek-like reconnection in our simulations is non-steady process, highly dynamic, and intermittent in the whole reconnection process.

We should also mention that the unstable local Petschek-like reconnection has been identified during the plasmoid instability stage in the previous simulations with full ionized plasmas (Baty, 2012; Shibayama et al., 2015, 2019), but the reconnection rates in those works only reached about 0.01-0.02, the second faster reconnection stage was not presented. Our results revealed a fast reconnection mechanism with a reconnection rate of order 0.1 in the partially ionized low solar atmosphere in the MHD scale. Apart from partially ionized solar plasma, our analysis can also be useful to better understand fast magnetic reconnection in other partially ionized environments such as interstellar medium, protoplanetary discs and magnetic reconnection studies in laboratory experiments.

Acknowledgments

This research was supported by the National Key R&\&&D Program of China Nos. 2022YFF0503804 (2022YFF0503800) and 2022YFF0503003 (2022YFF0503000); the NSFC Grants 12373060 and 11933009; the Strategic Priority Research Program of CAS with grants XDB0560000 and XDA17040507; the outstanding member of the Youth Innovation Promotion Association CAS (No. Y2021024); the Yunling Talent Project for the Youth; the Basic Research of Yunnan Province in China Grant 202401AS070044; the Yunling Scholar Project of the Yunnan Province and the Yunnan Province ScientistWorkshop of Solar Physics; Yunnan Key Laboratory of Solar Physics and Space Science under the number 202205AG070009; The numerical calculations and data analysis have been done on Hefei advanced computing center and on the Computational Solar Physics Laboratory of Yunnan Observatories. We benefit from the discussions of the ISSI-BJ Team ”Solar eruptions: preparing for the next generation multi-waveband coronagraph.”

References

  • Ali & Zhu (2019) Ali, A., & Zhu, P. 2019, Physics of Plasmas, 26
  • Avrett & Loeser (2008) Avrett, E. H., & Loeser, R. 2008, The Astrophysical Journal Supplement Series, 175, 229
  • Barata & Conde (2010) Barata, J., & Conde, C. 2010, Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 619, 21
  • Baty (2012) Baty, H. 2012, Physics of Plasmas, 19
  • Bhattacharjee et al. (2009) Bhattacharjee, A., Huang, Y.-M., Yang, H., & Rogers, B. 2009, Physics of Plasmas, 16
  • Biskamp (1986) Biskamp, D. 1986, The Physics of fluids, 29, 1520
  • Brandenburg & Zweibel (1994) Brandenburg, A., & Zweibel, E. G. 1994, Astrophysical Journal, Part 2-Letters (ISSN 0004-637X), vol. 427, no. 2, p. L91-L94, 427, L91
  • Brandenburg & Zweibel (1995) —. 1995, Effects of pressure and resistivity on the ambipolar diffusion singularity: too little, too late, Tech. rep., SCAN-9504110
  • Carlsson & Leenaarts (2012) Carlsson, M., & Leenaarts, J. 2012, Astronomy & Astrophysics, 539, A39
  • Carlsson & Stein (1997) Carlsson, M., & Stein, R. F. 1997, The Astrophysical Journal, 481, 500
  • Carlsson & Stein (2002) —. 2002, The Astrophysical Journal, 572, 626
  • Chen et al. (2023) Chen, F., Rempel, M., & Fan, Y. 2023, The Astrophysical Journal Letters, 950, L3
  • Daughton et al. (2009) Daughton, W., Roytershteyn, V., Albright, B., et al. 2009, Physical review letters, 103, 065004
  • De Pontieu et al. (2007) De Pontieu, B., Hansteen, V., van der Voort, L. R., van Noort, M., & Carlsson, M. 2007, The Astrophysical Journal, 655, 624
  • Ellerman (1917) Ellerman, F. 1917, Astrophysical Journal, vol. 46, p. 298, 46, 298
  • Fang et al. (2002) Fang, C., Chen, P., & Ding, M. 2002, in COSPAR Colloquia Series, Vol. 14, Elsevier, 3–8
  • Fermo et al. (2012) Fermo, R., Drake, J., & Swisdak, M. 2012, Physical review letters, 108, 255005
  • Furth et al. (1973) Furth, H., Rutherford, P., & Selberg, H. 1973, The Physics of Fluids, 16, 1054
  • Gan & Fang (1990) Gan, W., & Fang, C. 1990, The Astrophysical Journal, 358, 328
  • Georgoulis et al. (2002) Georgoulis, M. K., Rust, D. M., Bernasconi, P. N., & Schmieder, B. 2002, The Astrophysical Journal, 575, 506
  • Huang & Bhattacharjee (2012) Huang, Y.-M., & Bhattacharjee, A. 2012, Physical review letters, 109, 265002
  • Huang et al. (2011) Huang, Y.-M., Bhattacharjee, A., & Sullivan, B. P. 2011, Physics of Plasmas, 18
  • Khomenko & Collados (2012) Khomenko, E., & Collados, M. 2012, The Astrophysical Journal, 747, 87
  • Kulsrud (1998) Kulsrud, R. M. 1998, Physics of Plasmas, 5, 1599
  • Leake et al. (2013) Leake, J. E., Lukin, V. S., & Linton, M. G. 2013, Physics of Plasmas, 20, 061202
  • Leake et al. (2012) Leake, J. E., Lukin, V. S., Linton, M. G., & Meier, E. T. 2012, The Astrophysical Journal, 760, 109
  • Leenaarts & Carlsson (2009) Leenaarts, J., & Carlsson, M. 2009, in The Second Hinode Science Meeting: Beyond Discovery-Toward Understanding, Vol. 415, 87
  • Li et al. (2016) Li, L., Zhang, J., Peter, H., et al. 2016, Nature Physics, 12, 847
  • Lin et al. (2005) Lin, J., Ko, Y.-K., Sui, L., et al. 2005, The Astrophysical Journal, 622, 1251
  • Liu et al. (2002) Liu, B., Mineshige, S., & Shibata, K. 2002, The Astrophysical Journal, 572, L173
  • Liu et al. (2023) Liu, M., Ni, L., Cheng, G., Ziegler, U., & Lin, J. 2023, Research in Astronomy and Astrophysics
  • Mei et al. (2012) Mei, Z., Shen, C., Wu, N., et al. 2012, Monthly Notices of the Royal Astronomical Society, 425, 2824
  • Moore et al. (2011) Moore, R. L., Sterling, A. C., Cirtain, J. W., & Falconer, D. A. 2011, The Astrophysical Journal Letters, 731, L18
  • Murtas et al. (2021) Murtas, G., Hillier, A., & Snow, B. 2021, Physics of Plasmas, 28, 032901
  • Murtas et al. (2022) —. 2022, Physics of Plasmas, 29
  • Ni et al. (2021) Ni, L., Chen, Y., Peter, H., Tian, H., & Lin, J. 2021, Astronomy & Astrophysics, 646, A88
  • Ni et al. (2022) Ni, L., Cheng, G., & Lin, J. 2022, Astronomy & Astrophysics, 665, A116
  • Ni et al. (2020) Ni, L., Ji, H., Murphy, N. A., & Jara-Almonte, J. 2020, Proceedings of the Royal Society A, 476, 20190867
  • Ni et al. (2015) Ni, L., Kliem, B., Lin, J., & Wu, N. 2015, The Astrophysical Journal, 799, 79
  • Ni et al. (2018) Ni, L., Lukin, V. S., Murphy, N. A., & Lin, J. 2018, Physics of Plasmas, 25
  • Parker (1957) Parker, E. N. 1957, Journal of Geophysical Research, 62, 509
  • Paschmann et al. (1979) Paschmann, G., Sonnerup, B. Ö., Papamastorakis, I., et al. 1979, Nature, 282, 243
  • Peter et al. (2014) Peter, H., Tian, H., Curdt, W., et al. 2014, Science, 346, 1255726
  • Petschek (1964) Petschek, H. E. 1964, in Proceedings of a Symposium Held at the Goddard Space Flight Center, Greenbelt, Maryland, October 28-30, 1963, Vol. 50, Scientific and Technical Information Division, National Aeronautics and …, 425
  • Sato (1979) Sato, T. 1979, Journal of Geophysical Research: Space Physics, 84, 7177
  • Shepherd & Cassak (2010) Shepherd, L. S., & Cassak, P. 2010, Physical review letters, 105, 015004
  • Shibata & Tanuma (2001) Shibata, K., & Tanuma, S. 2001, Earth, Planets and Space, 53, 473
  • Shibata et al. (2007) Shibata, K., Nakamura, T., Matsumoto, T., et al. 2007, Science, 318, 1591
  • Shibayama et al. (2019) Shibayama, T., Kusano, K., Miyoshi, T., & Bhattacharjee, A. 2019, Physics of Plasmas, 26
  • Shibayama et al. (2015) Shibayama, T., Kusano, K., Miyoshi, T., Nakabou, T., & Vekstein, G. 2015, Physics of Plasmas, 22
  • Sweet (1958) Sweet, P. A. 1958, in Electromagnetic Phenomena in Cosmical Physics, ed. B. Lehnert, Vol. 6, 123
  • Takasao et al. (2011) Takasao, S., Asai, A., Isobe, H., & Shibata, K. 2011, The Astrophysical Journal Letters, 745, L6
  • Tian et al. (2016) Tian, H., Xu, Z., He, J., & Madsen, C. 2016, The Astrophysical Journal, 824, 96
  • Ugai & Tsuda (1979) Ugai, M., & Tsuda, T. 1979, Journal of Plasma Physics, 22, 1
  • Uzdensky & Kulsrud (2000) Uzdensky, D., & Kulsrud, R. 2000, Physics of Plasmas, 7, 4018
  • Vranjes & Krstic (2013) Vranjes, J., & Krstic, P. 2013, Astronomy & Astrophysics, 554, A22
  • Wargnier et al. (2023) Wargnier, Q., Martinez-Sykora, J., Hansteen, V., & De Pontieu, B. 2023, The Astrophysical Journal, 946, 115
  • Yokoyama & Shibata (1995) Yokoyama, T., & Shibata, K. 1995, Nature, 375, 42
  • Yokoyama & Shibata (2001) —. 2001, The Astrophysical Journal, 549, 1160
  • Zafar et al. (2023) Zafar, A., Ni, L., Lin, J., & Ziegler, U. 2023, Physics of Plasmas, 30
  • Zhu et al. (2020a) Zhu, B., Yan, H., Zhong, Y., et al. 2020a, Applied Mathematical Modelling, 78, 932
  • Zhu et al. (2020b) —. 2020b, Applied Mathematical Modelling, 78, 968
  • Ziegler (2011) Ziegler, U. 2011, Journal of Computational Physics, 230, 1035
  • Zweibel & Yamada (2009) Zweibel, E. G., & Yamada, M. 2009, Annual review of astronomy and astrophysics, 47, 291