Direct prediction of saturated neoclassical tearing modes in slab using an equilibrium approach

E. Balkovic\aff1    J. Loizu\aff1    J. P. Graves\aff1    Y.-M. Huang\aff2    C. Smiet\aff1 \aff1 École Polytechnique Fédérale de Lausanne, Swiss Plasma Center, CH-1015 Lausanne, Switzerland
\aff2 Princeton University, Princeton NJ 08543, USA
Abstract

We demonstrate for the first time that the nonlinear saturation of neoclassical tearing modes (NTMs) can be found directly using a variational principle based on Taylor relaxation, without needing to simulate the intermediate, resistivity-dependent dynamics. As in previous investigations of classical tearing mode saturation (Loizu et al., 2020; Loizu & Bonfiglio, 2023), we make use of SPEC (Hudson et al., 2012), an equilibrium solver based on the variational principle of the Multi-Region relaxed MHD, featuring stepped pressure profiles and arbitrary magnetic topology. We work in slab geometry and employ a simple bootstrap current model Jbs=C\bnablapsubscript𝐽𝑏𝑠𝐶\bnabla𝑝J_{bs}=C\>\bnabla pitalic_J start_POSTSUBSCRIPT italic_b italic_s end_POSTSUBSCRIPT = italic_C italic_p to study the bootstrap-driven tearing modes, scanning over the asymptotic matching parameter ΔsuperscriptΔ\Delta^{\prime}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and bootstrap current strength. Saturated island widths produced by SPEC agree well with the predictions of an initial value resistive MHD code (Huang & Bhattacharjee, 2016) while being orders of magnitude faster to calculate. Additionally, we observe good agreement with a simple analytical Modified Rutherford Equation, without requiring any fitting coefficients. The match is obtained for both linearly unstable classical tearing modes in the presence of bootstrap current, and neoclassical tearing modes, which are linearly stable but nonlinear-unstable due to the effects of the bootstrap current.

1 Introduction

Tearing modes convert magnetic field energy into kinetic energy of the plasma through spontaneous magnetic reconnection (Goldston, 2020). Generally present in toroidally-confined magnetic fusion plasmas, these internal modes violate the frozen-in constraint of ideal MHD, modifying the flux surface topology and creating magnetic islands. In many cases, tearing modes do not lead to the complete collapse of the plasma, but rather grow into non-linearly saturated states (Kong, 2020) that can strongly increase global transport and lower performance (Günter et al., 1999). These instabilities often appear in the form of neoclassical tearing modes (NTM), the behavior of which is driven by the bootstrap current. The latter is generated by the collisional interaction of trapped and passing particles in a toroidal plasma and is typically proportional to pressure gradients (Helander & Sigmar, 2005). As the transport of particles and heat along the magnetic field dominates that across the field, islands tend to locally flatten pressure (Fitzpatrick, 1995), which leads to the removal of bootstrap current inside the island, driving the tearing mode unstable. Effects of bootstrap current on tearing modes include an enhanced growth of linearly unstable tearing modes and, more importantly, a nonlinear destabilization of configurations that are linearly stable to tearing modes.

As a first approximation, tearing modes are studied using linear theory which involves tearing-like perturbations of an initial equilibrium. These are constructed by solving the linearized resistive MHD equations in a narrow layer around the resonant surface, where the safety factor q𝑞qitalic_q is rational, and performing an asymptotic matching with a solution of the linearized ideal MHD equations outside the layer. If a mode can be found that lowers the MHD potential energy, 𝒲=𝑑V[B2/2μ0+p/(γ1)]𝒲differential-d𝑉delimited-[]superscript𝐵22subscript𝜇0𝑝𝛾1\mathcal{W}=\int dV\>[B^{2}/2\mu_{0}+p/(\gamma-1)]caligraphic_W = ∫ italic_d italic_V [ italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_p / ( italic_γ - 1 ) ], then the equilibrium is deemed unstable to this mode. In the large aspect ratio limit, an important quantity is ΔsuperscriptΔ\Delta^{\prime}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT which comes from the discontinuity of the derivative of the perturbed flux across the resistive layer for a given perturbation. Neglecting the stabilizing effects of pressure (Glasser et al., 1976), if Δ>0superscriptΔ0\Delta^{\prime}>0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0, then a particular tearing mode is unstable and will lower 𝒲𝒲\mathcal{W}caligraphic_W, and if Δ<0superscriptΔ0\Delta^{\prime}<0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < 0 the mode will increase 𝒲𝒲\mathcal{W}caligraphic_W as it takes more energy to bend the fieldlines than what is released by reconnection. The linear approach provides a good start for uncovering the physics of tearing modes, but it only yields information on linear growth and stability. Therefore, it is of limited use for understanding nonlinear growth and saturation (Goldston, 2020), as is measured in experiments (Kong, 2020).

To study the nonlinear dynamics, Rutherford (1973) started from the linear theory and extended it into a nonlinear regime, deriving an ODE for the evolution of the island width w𝑤witalic_w, which is a good measure of the mode amplitude and also relates to the level of confinement degradation of the underlying equilibrium. Fitzpatrick (1995) extended Rutherford’s model by including the effects of the bootstrap current. An example of such ODE is:

μ0ηwt=Δ0+Δstab(w)+Δbs(w)subscript𝜇0𝜂𝑤𝑡subscriptsuperscriptΔ0subscriptsuperscriptΔstab𝑤subscriptsuperscriptΔbs𝑤\frac{\mu_{0}}{\eta}\frac{\partial w}{\partial t}=\Delta^{\prime}_{0}+\Delta^{% \prime}_{\textrm{stab}}(w)+\Delta^{\prime}_{\textrm{bs}}(w)divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_η end_ARG divide start_ARG ∂ italic_w end_ARG start_ARG ∂ italic_t end_ARG = roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT stab end_POSTSUBSCRIPT ( italic_w ) + roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ( italic_w ) (1)

where Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the value of ΔsuperscriptΔ\Delta^{\prime}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT for perturbations of an initial equilibrium with w=0𝑤0w=0italic_w = 0, Δstab(w)0subscriptsuperscriptΔstab𝑤0\Delta^{\prime}_{\textrm{stab}}(w)\leq 0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT stab end_POSTSUBSCRIPT ( italic_w ) ≤ 0 includes stabilizing terms and ensures saturation of the island, and Δbs(w)subscriptsuperscriptΔbs𝑤\Delta^{\prime}_{\textrm{bs}}(w)roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ( italic_w ) accounts for the effects of the bootstrap current. The bootstrap term is destabilizing and can be found analytically in the large aspect ratio limit, Δbs(w)1/wsimilar-tosubscriptsuperscriptΔbs𝑤1𝑤\Delta^{\prime}_{\textrm{bs}}(w)\sim 1/wroman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ( italic_w ) ∼ 1 / italic_w (Fitzpatrick, 1995). This term diverges for small islands and so is modified to Δbs(w)w2/(w2+w02)similar-tosubscriptsuperscriptΔbs𝑤superscript𝑤2superscript𝑤2superscriptsubscript𝑤02\Delta^{\prime}_{\textrm{bs}}(w)\sim w^{2}/(w^{2}+w_{0}^{2})roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ( italic_w ) ∼ italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), thus introducing a regularizing scale, w0(χ/χ)1/4similar-tosubscript𝑤0superscriptsubscript𝜒perpendicular-tosubscript𝜒parallel-to14w_{0}\sim({\chi_{\perp}}/{\chi_{\parallel}})^{1/4}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ ( italic_χ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_χ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT, which describes the effects of incomplete pressure flattening in an island. Here χsubscript𝜒perpendicular-to\chi_{\perp}italic_χ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT and χsubscript𝜒parallel-to\chi_{\parallel}italic_χ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT are the heat diffusion coefficients across and along the magnetic field, respectively. The typical evolution of an island for a linearly stable equilibrium (Δ0<0subscriptsuperscriptΔ00\Delta^{\prime}_{0}<0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0) using either of the two bootstrap terms is shown in Figure 1. When w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT isn’t included, the mode is always unstable and will grow regardless of the initial island size. However, when w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is included and incomplete flattening is accounted for, then the mode will not grow unless it has been perturbed with a sufficient seed, i.e. if the initial island size is sufficiently large. Seeding dynamics are complicated (Muraglia et al., 2021), but it is important to note that, in the large island limit ww0much-greater-than𝑤subscript𝑤0w\gg w_{0}italic_w ≫ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , the saturated island is essentially the same, regardless of whether the seeding mechanism is included or not. For the purposes of this paper, we study the ww0much-greater-than𝑤subscript𝑤0w\gg w_{0}italic_w ≫ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT regime, which is one of greater relevance to magnetic fusion experiments (in a typical tokamak w01cmsubscript𝑤01cmw_{0}\approx 1\textrm{cm}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 1 cm while wsat4cmsubscript𝑤sat4cmw_{\textrm{sat}}\approx 4\textrm{cm}italic_w start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT ≈ 4 cm) (Sauter et al., 2002). Non-linear equations such as Eq. 1 are known as Modified Rutherford equations (MRE) and are often used in tokamak experiments (Kong, 2020) as they are quick to evaluate and give good estimates for the saturated island widths. However, MRE often includes other Δ(w)superscriptΔ𝑤\Delta^{\prime}(w)roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_w ) ad-hoc terms (Kong et al., 2022) which are device-specific and can feature many fitting coefficients.

Refer to caption
Figure 1: Sketch of the width growth rate as a function of width, as described by Equation 1. The black curve shows the behavior with the diverging 1/w1𝑤1/w1 / italic_w term, while the red curve demonstrates the effect of the w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT correction to ΔbssubscriptsuperscriptΔbs\Delta^{\prime}_{\textrm{bs}}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT.

A more accurate approach to studying nonlinear tearing modes is the time integration of resistive MHD equations, evolving a perturbed equilibrium into a steady state featuring a saturated island. Such a method gives a detailed description of tearing mode dynamics and can be expanded to add various non-MHD effects, but it is numerically very expensive (Hoelzl et al., 2021; Teng et al., 2018). The cost comes from the large separation between ideal and resistive timescales (measured by Lundquist number S=τres/τAlfven1𝑆subscript𝜏ressubscript𝜏Alfvenmuch-greater-than1S=\tau_{\textrm{res}}/\tau_{\textrm{Alfven}}\gg 1italic_S = italic_τ start_POSTSUBSCRIPT res end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT Alfven end_POSTSUBSCRIPT ≫ 1), which requires long simulation times, and the separation between system size and tearing spatial scales, which requires fine meshes. In summary, given existing approaches, it is difficult to predict the saturation of tearing modes in a manner that is simultaneously fast, geometry-agnostic, and rooted in solid theory.

However, there may exist another approach based on energy principles that can remedy some of these stated drawbacks. The method was pioneered by Cooper et al. (2010), who used the ideal MHD equilibrium solver VMEC (Hirshman & Whitson, 1983) to predict the saturated state of internal kink modes in a tokamak. The idea is that, in addition to the initial MHD equilibrium, one can think of the saturated state of an instability as yet another (lower energy) MHD equilibrium. Thus, if an equilibrium code exists where both the initial and saturated states are in the space of possible solutions, then by clever perturbations in the parameter space it may be possible to directly jump from the initial to the final state. In addition to ideal modes such as the internal kink, this approach proved valuable for finding nonlinear states of non-resonant modes (Brunetti et al., 2014), where q=m/n𝑞𝑚𝑛q=m/nitalic_q = italic_m / italic_n surface is not present. Here, nonlinear states from VMEC (Hirshman & Whitson, 1983) agreed well with a resistive MHD code when a mode was nonresonant, but led to a poor agreement otherwise.

Loizu extended the method for resonant resistive instabilities. His work showed that such an approach can be used for the prediction of the saturated state of classical tearing modes, both in the slab (Loizu et al., 2020) and cylindrical geometry (Loizu & Bonfiglio, 2023). Tearing mode prediction required using an equilibrium code such as the Stepped Pressure Equilibrium Code (SPEC) (Hudson et al., 2012), which does not constrain the magnetic topology around the resistive layer. In this paper, we present an extension of the work done in slab (Loizu et al., 2020) where the effects of bootstrap current are included and lead to the growth and saturation of neoclassical tearing modes. The paper is organized as follows. In section 2 we introduce the set of tearing-unstable initial equilibria, including the geometry and starting profiles. Section 3 describes how the saturation of tearing modes is predicted using resistive MHD initial value simulations. Finally, section 4 presents the study of nonlinear saturation using SPEC, along with a comparison of the variational approach to the resistive MHD and MRE.

2 Family of tearing unstable equilibria

To demonstrate the new approach for the prediction of tearing mode saturation with bootstrap current, we work in slab geometry. This simplifies the analysis since the slab does not have curvature or a coordinate singularity. In addition, we can set up an equilibrium that is stable to ideal modes and that has only one resonant surface where a tearing mode can grow unstable. Furthermore, analytical linear and nonlinear theories are simpler in slab geometry. We use a cartesian coordinate system where x𝑥xitalic_x is a radial-like coordinate, y𝑦yitalic_y is the poloidal-like direction, z𝑧zitalic_z is the toroidal-like direction, and the slab is doubly periodic in y𝑦yitalic_y and z𝑧zitalic_z. The size of the slab is fixed to 2π2𝜋2\pi2 italic_π along x𝑥xitalic_x and z𝑧zitalic_z, while we consider different values L𝐿Litalic_L of the system size along y𝑦yitalic_y, thus allowing us to scan different stability regimes, with effects similar to varying the safety factor q𝑞qitalic_q in toroidal geometry. An initial equilibrium magnetic field 𝑩0=Bz0z^+z^×\bnablaψ0subscript𝑩0subscript𝐵𝑧0^𝑧^𝑧\bnablasubscript𝜓0\boldsymbol{B}_{0}=B_{z0}\hat{z}+\hat{z}\times\bnabla\psi_{0}bold_italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT over^ start_ARG italic_z end_ARG + over^ start_ARG italic_z end_ARG × italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is constructed with a flux function ψ0(x)=1/cosh2(x)+αx2subscript𝜓0𝑥1superscript2𝑥𝛼superscript𝑥2\psi_{0}(x)=1/\cosh^{2}{(x)}+\alpha x^{2}italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = 1 / roman_cosh start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_x ) + italic_α italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, similar to that in Loureiro et al. (2005). Associated to this flux function is the poloidal field, By0(x)=ψ0(x)subscript𝐵𝑦0𝑥superscriptsubscript𝜓0𝑥B_{y0}(x)=\psi_{0}^{\prime}(x)italic_B start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT ( italic_x ) = italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ), and the toroidal current, Jz0(x)=ψ0′′(x)subscript𝐽𝑧0𝑥superscriptsubscript𝜓0′′𝑥J_{z0}(x)=\psi_{0}^{\prime\prime}(x)italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT ( italic_x ) = italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_x ), which we call the ohmic current. The initial pressure p0(x)subscript𝑝0𝑥p_{0}(x)italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) is simply a linear function of x𝑥xitalic_x, with the scale being given by the condition on the average plasma beta β=2μ0p/B23%𝛽delimited-⟨⟩2subscript𝜇0𝑝superscript𝐵2percent3\beta=\langle 2\mu_{0}p/B^{2}\rangle\approx 3\%italic_β = ⟨ 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_p / italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ ≈ 3 %. The equilibrium, profiles of which are shown in Fig. 2, is ideally stable and has a single resonant surface at x=0𝑥0x=0italic_x = 0, and Bz0subscript𝐵𝑧0B_{z0}italic_B start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT is an essentially constant strong guide field, such that the mode is approximately incompressible (Goldston, 2020), with minor variations to set up an initial force balance, Bz02+|\bnablaψ0|2+p0=constsuperscriptsubscript𝐵𝑧02superscript\bnablasubscript𝜓02subscript𝑝0constB_{z0}^{2}+|\bnabla\psi_{0}|^{2}+p_{0}=\textrm{const}italic_B start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = const. The field at the resonant surface is solely in the z𝑧zitalic_z direction, which means that the mode, in order to have the resonance condition 𝒌𝑩=0𝒌𝑩0\boldsymbol{k}\cdot\boldsymbol{B}=0bold_italic_k ⋅ bold_italic_B = 0, is symmetric in the z𝑧zitalic_z direction. The bootstrap current, which in a torus originates from the field curvature, is not inherently present in a slab equilibrium. We overcome this by employing a simple model for the bootstrap current 𝑱bs=C^C0(p/x)z^subscript𝑱bs^𝐶subscript𝐶0𝑝𝑥^𝑧\boldsymbol{J}_{\textrm{bs}}=\hat{C}\>C_{0}(\partial p/\partial x)\>\hat{z}bold_italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT = over^ start_ARG italic_C end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( ∂ italic_p / ∂ italic_x ) over^ start_ARG italic_z end_ARG, where C0=Jz0(0)/(p0(0)/x)subscript𝐶0subscript𝐽𝑧00subscript𝑝00𝑥C_{0}=J_{z0}(0)/(\partial p_{0}(0)/\partial x)italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT ( 0 ) / ( ∂ italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 ) / ∂ italic_x ) and C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG is a free constant which specifies the relative strength of the bootstrap and ohmic current in the initial equilibrium. This term has been successfully used in previous studies of slab NTMs (Muraglia et al., 2021). Thanks to the POEM theory (Escande & Ottaviani, 2004; Militello & Porcelli, 2004), the stabilization term of MRE, Eq. 1, can be expressed analytically as Δstab(w)=(1/2.44d2)wsubscriptsuperscriptΔstab𝑤12.44superscript𝑑2𝑤\Delta^{\prime}_{\textrm{stab}}(w)=-(1/2.44d^{2})wroman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT stab end_POSTSUBSCRIPT ( italic_w ) = - ( 1 / 2.44 italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_w, where d𝑑ditalic_d is the length scale of the initial equilibrium current, d=Jz0(0)/Jz0′′(0)0.35𝑑subscript𝐽𝑧00superscriptsubscript𝐽𝑧0′′00.35d=\sqrt{J_{z0}(0)/J_{z0}^{\prime\prime}(0)}\approx 0.35italic_d = square-root start_ARG italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT ( 0 ) / italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( 0 ) end_ARG ≈ 0.35. Using the POEM stabilization term and a simplified bootstrap contribution Δbs=C^C0prb(1/w)subscriptsuperscriptΔ𝑏𝑠^𝐶subscript𝐶0subscriptsuperscript𝑝𝑟𝑏1𝑤\Delta^{\prime}_{bs}=\hat{C}\>C_{0}\>p^{\prime}_{r}\>b\>(1/w)roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b italic_s end_POSTSUBSCRIPT = over^ start_ARG italic_C end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_b ( 1 / italic_w ), where prsubscriptsuperscript𝑝𝑟p^{\prime}_{r}italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the pressure gradient across the resonant surface and b𝑏bitalic_b is a fitting coefficient (which can be estimated analytically, see Appendix B), we arrive at the following expression

wsat=1.22d2(Δ0+Δ02+1.64a2C^C0prb)subscript𝑤𝑠𝑎𝑡1.22superscript𝑑2subscriptsuperscriptΔ0superscriptsubscriptΔ021.64superscript𝑎2^𝐶subscript𝐶0subscriptsuperscript𝑝𝑟𝑏w_{sat}=1.22d^{2}\left(\Delta^{\prime}_{0}+\sqrt{\Delta_{0}^{\prime 2}+1.64a^{% -2}\>\hat{C}\>C_{0}\>p^{\prime}_{r}\>b}\right)italic_w start_POSTSUBSCRIPT italic_s italic_a italic_t end_POSTSUBSCRIPT = 1.22 italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + square-root start_ARG roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ 2 end_POSTSUPERSCRIPT + 1.64 italic_a start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT over^ start_ARG italic_C end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_b end_ARG ) (2)

This equation provides an estimate of the island width for tearing modes with small constant Jbssubscript𝐽bsJ_{\textrm{bs}}italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT and is valid only in the large island limit, ww0much-greater-than𝑤subscript𝑤0w\gg w_{0}italic_w ≫ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

3 Resistive MHD

The canonical model equations describing the nonlinear evolution of tearing modes are the resistive MHD equations:

ρt+\bnabla(ρ𝒖)=0𝜌𝑡bold-⋅\bnabla𝜌𝒖0\displaystyle\frac{\partial\rho}{\partial t}+\bnabla\boldsymbol{\cdot}\left(% \rho\boldsymbol{u}\right)=0divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG + bold_⋅ ( italic_ρ bold_italic_u ) = 0 (3)
(ρ𝒖)t+\bnabla(ρ𝒖𝒖)=\bnabla(p+12B2)+\bnabla(𝑩𝑩)+\bnabla(μρ\mathsfbiϵ\displaystyle\frac{\partial(\rho\boldsymbol{u})}{\partial t}+\bnabla% \boldsymbol{\cdot}\left(\rho\boldsymbol{u}\boldsymbol{u}\right)=-\bnabla\left(% p+\frac{1}{2}B^{2}\right)+\bnabla\boldsymbol{\cdot}\left(\boldsymbol{B}% \boldsymbol{B}\right)+\bnabla\boldsymbol{\cdot}\left(\mu\rho\boldsymbol{% \mathsfbi{\epsilon}}divide start_ARG ∂ ( italic_ρ bold_italic_u ) end_ARG start_ARG ∂ italic_t end_ARG + bold_⋅ ( italic_ρ bold_italic_u bold_italic_u ) = - ( italic_p + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + bold_⋅ ( bold_italic_B bold_italic_B ) + bold_⋅ ( italic_μ italic_ρ bold_italic_ϵ
pt+\bnabla(p𝒖)=(1γ)(p\bnabla𝒖χ2\bnabla(ρ𝒃𝒃\bnabla(p/ρ)))𝑝𝑡bold-⋅\bnabla𝑝𝒖1𝛾bold-⋅𝑝\bnabla𝒖bold-⋅subscript𝜒parallel-to2\bnablabold-⋅𝜌𝒃𝒃\bnabla𝑝𝜌\displaystyle\frac{\partial p}{\partial t}+\bnabla\boldsymbol{\cdot}\left(p% \boldsymbol{u}\right)=(1-\gamma)\left(p\bnabla\boldsymbol{\cdot}\boldsymbol{u}% -\frac{\chi_{\parallel}}{2}\bnabla\boldsymbol{\cdot}\left(\rho\boldsymbol{b}% \boldsymbol{b}\boldsymbol{\cdot}\bnabla(p/\rho)\right)\right)divide start_ARG ∂ italic_p end_ARG start_ARG ∂ italic_t end_ARG + bold_⋅ ( italic_p bold_italic_u ) = ( 1 - italic_γ ) ( italic_p bold_⋅ bold_italic_u - divide start_ARG italic_χ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG bold_⋅ ( italic_ρ bold_italic_b bold_italic_b bold_⋅ ( italic_p / italic_ρ ) ) ) (4)
ψt=(𝒖×𝑩)z+η(Jz0Jz)+ηJbs𝜓𝑡subscript𝒖𝑩𝑧𝜂subscript𝐽𝑧0subscript𝐽𝑧𝜂subscript𝐽bs\displaystyle\frac{\partial\psi}{\partial t}=(\boldsymbol{u}\times\boldsymbol{% B})_{z}+\eta(J_{z0}-J_{z})+\eta J_{\textrm{bs}}divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_t end_ARG = ( bold_italic_u × bold_italic_B ) start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_η ( italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) + italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT (5)

These equations are solved, in a dimensionless form, using the code HMHD (Huang & Bhattacharjee, 2016). Viscous forces are introduced in the momentum balance, Eq. 3.2, through the strain tensor \mathsfbiϵ=(\bnabla𝒖+\bnablaT𝒖)/2\mathsfbibold-italic-ϵ\bnabla𝒖superscript\bnabla𝑇𝒖2\boldsymbol{\mathsfbi{\epsilon}}=(\bnabla\boldsymbol{u}+\bnabla^{T}\boldsymbol% {u})/2bold_italic_ϵ = ( bold_italic_u + start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_italic_u ) / 2. In Ohm’s law, Eq. 3.4, ηJz0𝜂subscript𝐽𝑧0\eta J_{z0}italic_η italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT is the loop voltage which maintains a non-zero steady state, and ηJbs𝜂subscript𝐽bs\eta J_{\textrm{bs}}italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT acts like a secondary driving term, allowing for the inclusion of an effective bootstrap current, as used by Muraglia et al. (2021). The variables are normalized as follows (real \rightarrow normalized): length x,y,zx,y,z/L0formulae-sequence𝑥𝑦𝑧𝑥𝑦𝑧subscript𝐿0x,y,z\rightarrow x,y,z/L_{0}italic_x , italic_y , italic_z → italic_x , italic_y , italic_z / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, density ρρ/ρ0𝜌𝜌subscript𝜌0\rho\rightarrow\rho/\rho_{0}italic_ρ → italic_ρ / italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, magnetic field 𝑩𝑩/B0𝑩𝑩subscript𝐵0\boldsymbol{B}\rightarrow\boldsymbol{B}/B_{0}bold_italic_B → bold_italic_B / italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, current 𝑱𝑱/(B0/L0μ0)𝑱𝑱subscript𝐵0subscript𝐿0subscript𝜇0\boldsymbol{J}\rightarrow\boldsymbol{J}/(B_{0}/L_{0}\mu_{0})bold_italic_J → bold_italic_J / ( italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), velocities 𝒖𝒖/uA𝒖𝒖subscript𝑢A\boldsymbol{u}\rightarrow\boldsymbol{u}/u_{\textrm{A}}bold_italic_u → bold_italic_u / italic_u start_POSTSUBSCRIPT A end_POSTSUBSCRIPT (with Alfven speed uA=B0/μ0ρ0subscript𝑢Asubscript𝐵0subscript𝜇0subscript𝜌0u_{\textrm{A}}=B_{0}/\sqrt{\mu_{0}\rho_{0}}italic_u start_POSTSUBSCRIPT A end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / square-root start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG), time tt/τA𝑡𝑡subscript𝜏At\rightarrow t/\tau_{\textrm{A}}italic_t → italic_t / italic_τ start_POSTSUBSCRIPT A end_POSTSUBSCRIPT (with Alfven time τA=L0/uAsubscript𝜏Asubscript𝐿0subscript𝑢A\tau_{\textrm{A}}=L_{0}/u_{\textrm{A}}italic_τ start_POSTSUBSCRIPT A end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT A end_POSTSUBSCRIPT), and pressure pp/(B02μ0)𝑝𝑝superscriptsubscript𝐵02subscript𝜇0p\rightarrow p/(B_{0}^{2}\mu_{0})italic_p → italic_p / ( italic_B start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). The parameters of the model are the adiabatic coefficient γ𝛾\gammaitalic_γ, which we fix to that of the ideal gas γ=5/3𝛾53\gamma=5/3italic_γ = 5 / 3, parallel diffusion coefficient, χ10subscript𝜒parallel-to10\chi_{\parallel}\approx 10italic_χ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ 10, used for numerical stability, dynamic viscosity μ𝜇\muitalic_μ, and resistivity η𝜂\etaitalic_η. The latter two are specified in terms of their respective dimensionless numbers, the Lundquist number S=L0uA/η𝑆subscript𝐿0subscript𝑢A𝜂S=L_{0}u_{\textrm{A}}/\etaitalic_S = italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT A end_POSTSUBSCRIPT / italic_η, and the magnetic Prandtl number Prm=μ/η𝑃subscript𝑟𝑚𝜇𝜂Pr_{m}=\mu/\etaitalic_P italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_μ / italic_η. For the simulations shown below, the dimensionless numbers are S[103,104]𝑆superscript103superscript104S\in[10^{3},10^{4}]italic_S ∈ [ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ] and Prm=[101,101]𝑃subscript𝑟𝑚superscript101superscript101Pr_{m}=[10^{-1},10^{1}]italic_P italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = [ 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ]. The code relies on the 5-point finite difference scheme for the spatial discretization, and the trapezoidal leapfrog method for the temporal discretization. The simulations are performed using dynamic time steps based on the CFL condition (Courant et al., 1967) for Alfvén dynamics. The mesh comprises approximately 400x400 points, periodic boundary conditions are applied in the y𝑦yitalic_y direction, and wall boundary conditions are applied in the x𝑥xitalic_x direction.

Simulating a tearing mode entails setting up the initial equilibrium, Figure 2, according to the profiles derived from the initial flux function ψ0subscript𝜓0\psi_{0}italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and pressure p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The initial density is constant and set to ρ[2,8]𝜌28\rho\in[2,8]italic_ρ ∈ [ 2 , 8 ] in normalized units. Note that the initial equilibrium does not include bootstrap current. The next step involves perturbing the initial flux with ψp=ψ^pcos(y/L)cos(x/2π)subscript𝜓𝑝subscript^𝜓𝑝𝑦𝐿𝑥2𝜋\psi_{p}=\hat{\psi}_{p}\cos{(y/L)\cos{(x/2\pi)}}italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_cos ( italic_y / italic_L ) roman_cos ( italic_x / 2 italic_π ), where ψ^psubscript^𝜓𝑝\hat{\psi}_{p}over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the perturbation magnitude (the exact value does not affect saturation, see Fig. 3). The mode then grows spontaneously and reaches, on an approximately resistive time scale, τR=μ0L2/η=SτAlfvensubscript𝜏𝑅subscript𝜇0superscript𝐿2𝜂𝑆subscript𝜏Alfven\tau_{R}={\mu_{0}L^{2}}/{\eta}=S\tau_{\textrm{Alfven}}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_η = italic_S italic_τ start_POSTSUBSCRIPT Alfven end_POSTSUBSCRIPT, a state where the profiles and island size stabilize. The time trace of the width of a typical island is depicted in Figure 3. The final step is to fit a decaying exponential of the form w(t)=c0+c1exp(c2t)𝑤𝑡subscript𝑐0subscript𝑐1subscript𝑐2𝑡w(t)=c_{0}+c_{1}\exp(c_{2}t)italic_w ( italic_t ) = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_exp ( italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_t ) to the final stage of the island evolution. The best fit for the constant coefficient gives the saturated island width, wsat:=w(t)=c0assignsubscript𝑤sat𝑤𝑡subscript𝑐0w_{\textrm{sat}}:=w(t\rightarrow\infty)=c_{0}italic_w start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT := italic_w ( italic_t → ∞ ) = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Fig. 3 also shows that the value of wsatsubscript𝑤satw_{\textrm{sat}}italic_w start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT is independent of η𝜂\etaitalic_η, μ𝜇\muitalic_μ, as well as the mesh resolution. The structure of the saturated island is shown in Figure 4a, where an m=1𝑚1m=1italic_m = 1 mode can be seen to dominate. This supports an underlying assumption in MRE, which is that the nonlinear profiles have the same spatial dependence as the linear ones and are of single harmonicity.

Refer to caption
Figure 2: Initial profiles (α=0.002𝛼0.002\alpha=-0.002italic_α = - 0.002) of the poloidal field, toroidal current, and pressure, as used in HMHD (red) and SPEC (blue). SPEC equilibrium consists of Nvol=151subscript𝑁𝑣𝑜𝑙151N_{vol}=151italic_N start_POSTSUBSCRIPT italic_v italic_o italic_l end_POSTSUBSCRIPT = 151 volumes, the largest of which is the central volume, located around x=0𝑥0x=0italic_x = 0.
Refer to caption
Figure 3: Time evolution of the island width from a seed to a saturated island, as a function of the resistive time. Here the equilibrium has L=10.7d𝐿10.7𝑑L=10.7ditalic_L = 10.7 italic_d and Δ0d=0.97subscriptsuperscriptΔ0𝑑0.97\Delta^{\prime}_{0}\>d=0.97roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = 0.97. Different curves correspond to (1) a reference case (solid black), (2) a smaller seed (blue triangles), (3) smaller resistivity (orange circles), (4) finer mesh (green crosses), and (5) smaller viscosity (red stars).
Refer to caption
Figure 4: Separatrices (red) and flux contours (black) of saturated islands in HMHD (left) and SPEC (right) for an equilibrium with L=7.8d𝐿7.8𝑑L=7.8ditalic_L = 7.8 italic_d, which corresponds to Δ0d=0.23subscriptsuperscriptΔ0𝑑0.23\Delta^{\prime}_{0}\>d=-0.23roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = - 0.23. The SPEC plot also depicts a selection of the ideal interfaces (green), the shape of which resembles the flux contours in HMHD.

The islands shown in Figs. 3,4 are for a slab of a certain size L𝐿Litalic_L, to which corresponds a value of Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. It is interesting to study the dependence of wsatsubscript𝑤satw_{\textrm{sat}}italic_w start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT on Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Such a study is shown in Figure 5, where the island width is depicted as a function of the linear stability parameter Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The black triangles reproduce the original study by Loizu et al. (2020), showing that classical modes (Jbs=0subscript𝐽bs0J_{\textrm{bs}}=0italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT = 0) do not have a saturated state if linearly stable (Δ0<0subscriptsuperscriptΔ00\Delta^{\prime}_{0}<0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0) and do have a saturated state, with wsatsubscript𝑤satw_{\textrm{sat}}italic_w start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT approximately linear in Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, if linearly unstable (Δ0>0subscriptsuperscriptΔ00\Delta^{\prime}_{0}>0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0). However, this is no longer true for tearing modes with bootstrap current, simulations of which are shown as black crosses (for C^=0.048^𝐶0.048\hat{C}=0.048over^ start_ARG italic_C end_ARG = 0.048). In this case, linearly stable modes (Δ0��<0subscriptsuperscriptΔ00\Delta^{\prime}_{0}<0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0) can in fact grow into saturated islands. These are usually referred to as neoclassical tearing modes (NTM) and are commonly observed in tokamaks (Kong, 2020). As depicted in Figure 1, NTMs need to be initiated with a sufficiently large seed, otherwise islands collapse back into the initial equilibrium. The precise size of this seed, however, does not impact the saturated island (see Fig. 3). For linearly unstable equilibria (Δ0>0subscriptsuperscriptΔ00\Delta^{\prime}_{0}>0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0), the bootstrap current again acts in a destabilizing manner, resulting in an enhanced island width as compared to the cases with Jbs=0subscript𝐽bs0J_{\textrm{bs}}=0italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT = 0.

Refer to caption
Figure 5: Scan of island width as a function of the linear stability Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Classical modes are shown from SPEC (green circle), HMHD (black triangle). Bootstrap-enhanced classical modes and linearly stable neoclassical tearing modes are shown from SPEC (blue diamond) and HMHD (black cross). The analytical prediction from MRE is also displayed (red dash). The bootstrap coefficient used is C^=0.048^𝐶0.048\hat{C}=0.048over^ start_ARG italic_C end_ARG = 0.048.

4 Variational approach

For directly predicting the saturation of tearing modes, we use SPEC (Hudson et al., 2012), a code that finds Multi Region Relaxed MHD (MRxMHD) equilibrium states using a variational principle. Taylor (1974) hypothesized that resistive plasmas often evolve in a way that conserves global magnetic helicity =Vp𝑑V𝑨𝑩subscriptsubscript𝑉𝑝bold-⋅differential-d𝑉𝑨𝑩\mathcal{H}=\int_{V_{p}}dV\boldsymbol{A}\boldsymbol{\cdot}\boldsymbol{B}caligraphic_H = ∫ start_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_d italic_V bold_italic_A bold_⋅ bold_italic_B while minimizing plasma energy 𝒲𝒲\mathcal{W}caligraphic_W. Here 𝑨𝑨\boldsymbol{A}bold_italic_A is the vector potential and Vpsubscript𝑉𝑝V_{p}italic_V start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the plasma volume. Hole et al. (2006) extended this idea to build the MRxMHD model that subdivides a plasma into Nvolsubscript𝑁volN_{\textrm{vol}}italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT constant-pressure Taylor-relaxed volumes 𝒱lsubscript𝒱𝑙\mathcal{V}_{l}caligraphic_V start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT which are separated by ideal interfaces lsubscript𝑙\mathcal{I}_{l}caligraphic_I start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT. Thus SPEC solves MRxMHD by finding the extrema of the energy functional

=𝒲+μ(0)=l𝒱l𝑑V(B22μ0+plγ1)+μl(𝒱l𝑑V𝑨𝑩l0)𝒲𝜇superscript0subscript𝑙subscriptsubscript𝒱𝑙differential-d𝑉superscript𝐵22subscript𝜇0subscript𝑝𝑙𝛾1subscript𝜇𝑙subscriptsubscript𝒱𝑙bold-⋅differential-d𝑉𝑨𝑩subscriptsuperscript0𝑙\mathcal{F}=\mathcal{W}+\mu(\mathcal{H}-\mathcal{H}^{0})=\sum_{l}\int_{% \mathcal{V}_{l}}dV\>\left(\frac{B^{2}}{2\mu_{0}}+\frac{p_{l}}{\gamma-1}\right)% +\mu_{l}\left(\int_{\mathcal{V}_{l}}dV\>\boldsymbol{A}\boldsymbol{\cdot}% \boldsymbol{B}-\mathcal{H}^{0}_{l}\right)caligraphic_F = caligraphic_W + italic_μ ( caligraphic_H - caligraphic_H start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT caligraphic_V start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_d italic_V ( divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG start_ARG italic_γ - 1 end_ARG ) + italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( ∫ start_POSTSUBSCRIPT caligraphic_V start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_d italic_V bold_italic_A bold_⋅ bold_italic_B - caligraphic_H start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT )

where l0subscriptsuperscript0𝑙\mathcal{H}^{0}_{l}caligraphic_H start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT are the target helicities and μlsubscript𝜇𝑙\mu_{l}italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT are the Lagrange multipliers. In slab geometry, the interfaces lsubscript𝑙\mathcal{I}_{l}caligraphic_I start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT are represented as

𝒙l(θ,ζ)=mxl,mcos(mθ)x^+Lθ2πy^+ζz^subscript𝒙𝑙𝜃𝜁subscript𝑚subscript𝑥𝑙𝑚𝑚𝜃^𝑥𝐿𝜃2𝜋^𝑦𝜁^𝑧\boldsymbol{x}_{l}(\theta,\zeta)=\sum_{m}x_{l,m}\cos{(m\theta)}\>\hat{x}+L% \frac{\theta}{2\pi}\>\hat{y}+\zeta\>\hat{z}bold_italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_θ , italic_ζ ) = ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT roman_cos ( italic_m italic_θ ) over^ start_ARG italic_x end_ARG + italic_L divide start_ARG italic_θ end_ARG start_ARG 2 italic_π end_ARG over^ start_ARG italic_y end_ARG + italic_ζ over^ start_ARG italic_z end_ARG

where xl,msubscript𝑥𝑙𝑚x_{l,m}italic_x start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT are the coefficients of a Fourier expansion that is numerically truncated at m=mpol𝑚subscript𝑚polm=m_{\textrm{pol}}italic_m = italic_m start_POSTSUBSCRIPT pol end_POSTSUBSCRIPT. The field within each subvolume is found through minimization with respect to the vector potential 𝑨𝑨\boldsymbol{A}bold_italic_A, and force balance across the interfaces is achieved by varying the interface harmonics xl,msubscript𝑥𝑙𝑚x_{l,m}italic_x start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT. SPEC finds the states that extremize \mathcal{F}caligraphic_F, which satisfy (Hudson et al., 2012)

\bnabla×𝑩=μl𝑩\bnabla𝑩subscript𝜇𝑙𝑩\displaystyle\bnabla\times\boldsymbol{B}=\mu_{l}\boldsymbol{B}× bold_italic_B = italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT bold_italic_B (6)
𝑭l=[p+B2/2μ0]l=0subscript𝑭𝑙subscriptdelimited-[]𝑝superscript𝐵22subscript𝜇0𝑙0\displaystyle\boldsymbol{F}_{l}=[p+B^{2}/2\mu_{0}]_{l}=0bold_italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = [ italic_p + italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = 0 (7)

where the first equation describes linear force-free fields inside each 𝒱lsubscript𝒱𝑙\mathcal{V}_{l}caligraphic_V start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, and the second equation represents the force balance across the two sides of each lsubscript𝑙\mathcal{I}_{l}caligraphic_I start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT. The position and shape of each interface is in fact determined by Eq. 4.2, which is the local equivalent of the MHD force balance equation 𝑱×𝑩=\bnablap𝑱𝑩\bnabla𝑝\boldsymbol{J}\times\boldsymbol{B}=\bnabla pbold_italic_J × bold_italic_B = italic_p. The standard approach for finding minimal MRxMHD states uses a Newton method for finding the zero of 𝑭lsubscript𝑭𝑙\boldsymbol{F}_{l}bold_italic_F start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and the alternative is a gradient descent algorithm for minimizing the functional \mathcal{F}caligraphic_F. We found that the Newton method alone is not robust enough for equilibria with islands. Instead, a more robust approach involves first minimizing the energy functional using a descent method, which is slower but more stable, and then using the Newton method to make fine adjustments to interface harmonics, resulting in an equilibrium with good force balance. The setup for solving MRxMHD in slab geometry with SPEC for a case with Nvol=3subscript𝑁vol3N_{\textrm{vol}}=3italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT = 3 is shown in Figure 6.

Refer to caption
Figure 6: The setup for SPEC in slab geometry for a case of Nvol=3subscript𝑁vol3N_{\textrm{vol}}=3italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT = 3. SPEC solves for Beltrami fields within each volume and seeks force balance across ideal interfaces by perturbing the shapes of these interfaces.

Inputs to SPEC include four physical constraints per volume that are fixed during functional minimization. These are the pressure plsubscript𝑝𝑙p_{l}italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, the integrated toroidal and poloidal fluxes, ΔΨltΔsubscriptsuperscriptΨ𝑡𝑙\Delta\Psi^{t}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and ΔΨlpΔsubscriptsuperscriptΨ𝑝𝑙\Delta\Psi^{p}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, and the helicities lsubscript𝑙\mathcal{H}_{l}caligraphic_H start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT inside each volume. Alternatively, SPEC can directly solve Eqs. 4.1 and 4.2 provided a different set of four constraints, (plsubscript𝑝𝑙p_{l}italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, ΔΨltΔsubscriptsuperscriptΨ𝑡𝑙\Delta\Psi^{t}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, ΔΨlpΔsubscriptsuperscriptΨ𝑝𝑙\Delta\Psi^{p}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, μlsubscript𝜇𝑙\mu_{l}italic_μ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT) or (plsubscript𝑝𝑙p_{l}italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, ΔΨltΔsubscriptsuperscriptΨ𝑡𝑙\Delta\Psi^{t}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, Ilvolsubscriptsuperscript𝐼vol𝑙I^{\textrm{vol}}_{l}italic_I start_POSTSUPERSCRIPT vol end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, Ilsurfsubscriptsuperscript𝐼surf𝑙I^{\textrm{surf}}_{l}italic_I start_POSTSUPERSCRIPT surf end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT), where Ilvolsubscriptsuperscript𝐼vol𝑙I^{\textrm{vol}}_{l}italic_I start_POSTSUPERSCRIPT vol end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT is the integrated toroidal current in each volume and Ilsurfsubscriptsuperscript𝐼surf𝑙I^{\textrm{surf}}_{l}italic_I start_POSTSUPERSCRIPT surf end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT is the sheet current on each ideal interface (Baillod et al., 2021).

While the linear growth of tearing modes occurs on a time scale that is intermediate between Alfvenic, τAlfvensubscript𝜏Alfven\tau_{\textrm{Alfven}}italic_τ start_POSTSUBSCRIPT Alfven end_POSTSUBSCRIPT, and resistive, τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT (for slab τ=τAlfven2/5τR3/5𝜏superscriptsubscript𝜏Alfven25superscriptsubscript𝜏𝑅35\tau=\tau_{\textrm{Alfven}}^{2/5}\tau_{R}^{3/5}italic_τ = italic_τ start_POSTSUBSCRIPT Alfven end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 / 5 end_POSTSUPERSCRIPT italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 / 5 end_POSTSUPERSCRIPT (Goldston, 2020)), the nonlinear saturation occurs on an approximately resistive time scale, as is demonstrated in Figure 3. Therefore, Taylor’s hypothesis of global helicity conservation, which works well for faster relaxation events (Berger, 1999), is not valid for slow tearing modes. Instead, the POEM theory (Militello & Porcelli, 2004) reveals that a good invariant is the flux surface average of the toroidal current Jz(ψ)=Jz0ψsubscript𝐽𝑧𝜓subscriptdelimited-⟨⟩subscript𝐽𝑧0𝜓J_{z}(\psi)=\langle J_{z0}\rangle_{\psi}italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_ψ ) = ⟨ italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT, with ψ𝜓\psiitalic_ψ being the nonlinear flux function. See Loizu et al. (2020) for a numerical demonstration of this invariant and its comparison with helicity. To construct equilibria where an integrated version of Jz(ψ)=Jz0ψsubscript𝐽𝑧𝜓subscriptdelimited-⟨⟩subscript𝐽𝑧0𝜓J_{z}(\psi)=\langle J_{z0}\rangle_{\psi}italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_ψ ) = ⟨ italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT is fixed, we run SPEC with the constraints (plsubscript𝑝𝑙p_{l}italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, ΔΨltΔsubscriptsuperscriptΨ𝑡𝑙\Delta\Psi^{t}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, Ilvolsubscriptsuperscript𝐼vol𝑙I^{\textrm{vol}}_{l}italic_I start_POSTSUPERSCRIPT vol end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, Ilsurfsubscriptsuperscript𝐼surf𝑙I^{\textrm{surf}}_{l}italic_I start_POSTSUPERSCRIPT surf end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT).

As with HMHD, finding tearing modes in SPEC involves first setting up the equilibrium. The initial domain is covered along the x𝑥xitalic_x direction with ideal interfaces, with a number Nvolsubscript𝑁volN_{\textrm{vol}}italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT that is sufficient to reproduce the equilibrium profiles. Special care needs to be given to the placement of the two interfaces bounding the resonant surface x=0𝑥0x=0italic_x = 0. If these are placed too closely to the resonant surface then they might limit the growth of the island. Conversely, if the spacing between them is too large then the analytical profiles are not well represented around x=0𝑥0x=0italic_x = 0, and can be artificially stabilized. The arbitrariness in interface placement is removed by placing the two interfaces symmetrically around x=0𝑥0x=0italic_x = 0 with a spacing of ΨwsubscriptΨ𝑤\Psi_{w}roman_Ψ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT (defined as the normalized toroidal flux enclosed by these two interfaces) which maximizes the width of the resulting saturated island. For more details on interface placement, see Loizu et al. (2020). From the interface positions and the analytical equilibrium profiles, the SPEC input profiles including ΔΨltΔsubscriptsuperscriptΨ𝑡𝑙\Delta\Psi^{t}_{l}roman_Δ roman_Ψ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, Ilvolsuperscriptsubscript𝐼𝑙volI_{l}^{\textrm{vol}}italic_I start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT vol end_POSTSUPERSCRIPT, and plsubscript𝑝𝑙p_{l}italic_p start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT are calculated. Since the pressure in SPEC is stepped, the bootstrap model Jbs=C^C0p/xsubscript𝐽bs^𝐶subscript𝐶0𝑝𝑥J_{\textrm{bs}}=\hat{C}\>C_{0}\>\partial p/\partial xitalic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT = over^ start_ARG italic_C end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∂ italic_p / ∂ italic_x is implemented as a set of sheet currents on the interfaces Ilsurf=C^C0Δpisuperscriptsubscript𝐼𝑙surf^𝐶subscript𝐶0Δsubscript𝑝𝑖I_{l}^{\textrm{surf}}=\hat{C}\>C_{0}\>\Delta p_{i}italic_I start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT surf end_POSTSUPERSCRIPT = over^ start_ARG italic_C end_ARG italic_C start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_Δ italic_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The integrated bootstrap current in SPEC is the same as in HMHD, and, as we will show, equivalent saturated islands can be obtained despite the discreteness of the current sheets. Note that the bootstrap current in SPEC, unlike in HMHD, exists already in the initial equilibrium, which explains why the initial toroidal volume currents, shown in Fig. 2, are not exactly the same between the two codes. SPEC is run with these inputs to find the initial MRxMHD state, profiles of which are shown in Fig. 2.

The next step is to evaluate linear stability and determine the unstable eigenmode with which the initial equilibrium can be perturbed, before a secondary equilibrium with an island is sought. This is accomplished by analyzing the Hessian matrix (Kumar, 2022) defined as

\mathsfbiHij=𝑿i𝑿j\mathsfbisubscript𝐻𝑖𝑗subscript𝑿𝑖subscript𝑿𝑗\mathsfbi{H}_{ij}=\frac{\partial\mathcal{F}}{\partial\boldsymbol{X}_{i}\>% \partial\boldsymbol{X}_{j}}italic_H start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG ∂ caligraphic_F end_ARG start_ARG ∂ bold_italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∂ bold_italic_X start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG

where i,j𝑖𝑗i,jitalic_i , italic_j index the packed degrees of freedom of the interfaces 𝑿={xl,m|l=1,,Nvol;m=0,,mpol}𝑿conditional-setsubscript𝑥𝑙𝑚formulae-sequence𝑙1subscript𝑁vol𝑚0subscript𝑚pol\boldsymbol{X}=\{x_{l,m}\>|\>l\!=\!1,...,N_{\textrm{vol}};\>m\!=\!0,...,m_{% \textrm{pol}}\}bold_italic_X = { italic_x start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT | italic_l = 1 , … , italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT ; italic_m = 0 , … , italic_m start_POSTSUBSCRIPT pol end_POSTSUBSCRIPT }. Eigendecomposition of \mathsfbiH\mathsfbi𝐻\mathsfbi{H}italic_H gives information on linear stability, where a negative eigenvalue implies instability and the associated eigenmode describes the shape of the mode in the space of interface Fourier coefficients. Since the initial equilibrium is ideally stable, depending on L𝐿Litalic_L, there is either zero or one negative eigenvalue, the eigenmode of which has the shape of an odd-parity tearing-like mode. Linear stability in SPEC has been verified for classical tearing modes (Loizu et al., 2020), while the stability of modes with bootstrap current is more subtle. In fact, as discussed in Sec. 1 and illustrated in Fig. 1, a system for which w00subscript𝑤00w_{0}\rightarrow 0italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → 0 (which corresponds to perfect pressure flattening in the island) is unconditionally unstable. Since the pressure profile in SPEC is, by definition, flat in the relaxation region, we expect that tearing modes will always be unstable when bootstrap effects are included. This is in fact what we observe when looking at the Hessian (see Appendix A). Nevertheless, as discussed in Sec. 1 and seen in Fig. 1, the saturation of an island is essentially independent of w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in the limit ww0much-greater-than𝑤subscript𝑤0w\gg w_{0}italic_w ≫ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT studied here. After finding the eigendecomposition of the Hessian, the next step involves perturbing the interfaces of the SPEC equilibrium, creating the seed island. The perturbation is constructed from the unstable eigenmode which is scaled to give the largest possible seed that does not overlap the ideal interfaces. Finally, SPEC is run for the second time, iterating on the interface harmonics to reach a secondary MRxMHD equilibrium with a lower energy and a topology of an island.

The results of running SPEC for a range of equilibria can be seen in Figure 5. The green circles correspond to classical tearing modes without bootstrap current, as reported in Loizu et al. (2020). In this case, SPEC recovers the island width obtained using HMHD as well as the prediction by the POEM theory (Militello & Porcelli, 2004). As is shown, SPEC correctly predicts the linear stability threshold such that stable equilibria for which Δ0<0subscriptsuperscriptΔ00\Delta^{\prime}_{0}<0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < 0 are also nonlinearly stable. The blue diamonds in Fig. 5 show the results from running SPEC with the inclusion of bootstrap current (C^=0.048^𝐶0.048\hat{C}=0.048over^ start_ARG italic_C end_ARG = 0.048). SPEC again manages to predict islands both in the case of classical modes (enhanced with bootstrap) and neoclassical tearing modes (which are only destabilized by the bootstrap current effect). A slight difference exists in the width between SPEC and HMHD for the larger islands, which we believe may be due to a violation of the current invariant with the increasing island size. Also shown in the figure is a prediction of an MRE, Eq. 2.1, which includes the bootstrap contribution. The results related to the MRE rely on a constant that we fit numerically, b=4.25𝑏4.25b=4.25italic_b = 4.25, as is standard for these models in general (Kong, 2020), instead of using the less-accurate, analytical value of b=8𝑏8b=8italic_b = 8, as derived in Appendix B. The fit is obtained by matching the island width for the equilibrium with Δ0=0subscriptsuperscriptΔ00\Delta^{\prime}_{0}=0roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.

The analysis in Figure 5 was done using a fixed bootstrap strength, C^=0.048^𝐶0.048\hat{C}=0.048over^ start_ARG italic_C end_ARG = 0.048. To verify that the prediction from SPEC is independent of the exact value of C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG, we chose two equilibria, one linearly stable, Δ0d=0.26subscriptsuperscriptΔ0𝑑0.26\Delta^{\prime}_{0}\>d=-0.26roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = - 0.26, and one linearly unstable, Δ0d=0.26subscriptsuperscriptΔ0𝑑0.26\Delta^{\prime}_{0}\>d=0.26roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = 0.26, and we vary the constant C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG from 00 to 0.30.30.30.3. The resulting island widths are shown in Figure 7, where it can be seen that the value of C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG does not significantly affect the agreement between SPEC, HMHD, and the analytical MRE. In addition to the island width, one can look at the structure of the island in SPEC and compare it with the island generated in HMHD. As shown in Figure 4, the island from SPEC matches that of HMHD closely. Its shape is dominantly given by the m=1𝑚1m=1italic_m = 1 harmonic, but it seems to have a slightly larger content of m>1𝑚1m>1italic_m > 1 harmonics. It is interesting to note that the two interfaces around the island, shown in green in Fig. 4b, are effectively coincident on the island separatrix, such that the entire volume between them is filled by the island. This is a consequence of the choice of ΨwsubscriptΨ𝑤\Psi_{w}roman_Ψ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT that maximizes the island width.

The saturated profiles in SPEC also agree well with the profiles from HMHD, as can be seen in Figure 8. The pressure in HMHD flattens across the island, which in this case happens primarily due to the generated parallel plasma flow, and not the parallel diffusion. In SPEC, the flattening is a consequence of the MRxMHD model. The radial profile of Bysubscript𝐵𝑦B_{y}italic_B start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT over the entire domain is not much affected by the tearing mode, meaning that the initial agreement between SPEC and HMHD is sustained in the saturated state.

Two important parameters in SPEC are mpolsubscript𝑚polm_{\textrm{pol}}italic_m start_POSTSUBSCRIPT pol end_POSTSUBSCRIPT and Nvolsubscript𝑁volN_{\textrm{vol}}italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT. The first one determines the number of poloidal harmonics used for representing the interfaces and force imbalances, while the second one is the number of relaxed volumes in SPEC and it determines how well the outer ideal solution is represented. In Figure 9, the saturation width of a classical tearing mode obtained with SPEC is shown as a function of both mpolsubscript𝑚polm_{\textrm{pol}}italic_m start_POSTSUBSCRIPT pol end_POSTSUBSCRIPT and Nvolsubscript𝑁volN_{\textrm{vol}}italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT. It can be seen that the saturated island width is converged, in the sense that increasing the two parameters beyond mpol=3subscript𝑚pol3m_{\textrm{pol}}=3italic_m start_POSTSUBSCRIPT pol end_POSTSUBSCRIPT = 3 and Nvol=151subscript𝑁vol151N_{\textrm{vol}}=151italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT = 151, which are used for the simulations presented in this paper, leads to less than 1%percent11\%1 % difference in the calculated widths.

Refer to caption
Figure 7: Scan of island width as a function of the bootstrap current strength for SPEC (green and blue) and HMHD (black), in the case of a classical tearing mode (Δ0d=0.26subscriptsuperscriptΔ0𝑑0.26\Delta^{\prime}_{0}\>d=0.26roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = 0.26) and an NTM (Δ0d=0.26subscriptsuperscriptΔ0𝑑0.26\Delta^{\prime}_{0}\>d=-0.26roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d = - 0.26).
Refer to caption
Figure 8: Profiles of a) pressure and b) poloidal field for a saturated tearing mode from SPEC and HMHD. Flattened pressure in HMHD matches well with SPEC where the pressure is, ipso facto, flat inside the island. The poloidal field is mostly unchanged from the initial equilibrium.
Refer to caption
Figure 9: Convergence of the width of a classical tearing mode with the bootstrap current in SPEC with respect to the number of a) poloidal harmonics and b) relaxed volumes. Reaching an equilibrium with a saturated island requires at least 3 harmonics.

5 Discussion

We have demonstrated that an "equilibrium approach", based on the solver SPEC (Hudson et al., 2012), can be used to directly predict the saturated state of tearing modes that evolve under the influence of the neoclassical bootstrap current. Good agreement with the resistive MHD equations, solved using the code HMHD (Huang & Bhattacharjee, 2016), was found for linearly unstable classical tearing modes, where the bootstrap current enhances the width of the saturated island. More importantly, we have shown that the approach can be applied to neoclassical tearing modes, which are linearly stable and for which bootstrap current is the primary driving mechanism. In addition to demonstrating the agreement in the width of saturated islands, a good match was also shown for the saturated island shapes and profiles of the pressure and the magnetic field. The independence of the SPEC results with respect to resolution parameters was established, as shown by the scans of the island width as a function of the number of poloidal harmonics and relaxed volumes. Having demonstrated that SPEC can correctly reproduce the saturated state of tearing modes, it is important to stress the relative benefits of this approach. First and foremost, the new method is significantly faster than resistive MHD simulations. For the scan of tearing modes with the bootstrap current in Figure 5, it took approximately 300 CPU hours to simulate with HMHD the saturation of just a single tearing mode. Conversely, it took a mere 2 CPU hours to simulate the entire dataset for SPEC, which includes 30 different saturated islands. An additional benefit is that the SPEC-based approach does not involve fitting coefficients. Such is the case with the analytical MRE model which, despite being fast, has several coefficients that need to be fit to a specific device (10similar-toabsent10\sim 10∼ 10 coefficients for the TCV tokamak) (Kong, 2020). The last notable benefit of the approach is that it is general, both in terms of geometry and particular instability. While initial work (Loizu et al., 2020) and its extension presented here were done in a slab, this approach could be applied to other configurations. The method has already been applied to classical tearing modes in a cylindrical tokamak (Loizu & Bonfiglio, 2023), and work has started in toroidal geometry. We are interested in predicting 2-1 NTMs observed in the TCV tokamak (Kong, 2020) and comparison with experimental measurements and MRE models. More broadly, this work promotes the development of a fast and simple tool for the general prediction of saturated instabilities in arbitrary geometries. In particular, this method may eventually allow for quick prediction of nonlinear saturation in stellarators, which have been observed (de Aguilera et al., 2015; Watanabe et al., 2005; Geiger J. E., 2004) to operate stably in regimes deemed unstable by linear theory.

6 Acknowledgments

This work was supported in part by the Swiss National Science Foundation. And by a grant from the Simons Foundation (1013657, JL). This work has been carried out within the framework of the EUROfusion Consortium, via the Euratom Research and Training Programme (Grant Agreement No 101052200 — EUROfusion) and funded by the Swiss State Secretariat for Education, Research and Innovation (SERI). Views and opinions expressed are however those of the author(s) only and do not necessarily reflect those of the European Union, the European Commission, or SERI. Neither the European Union nor the European Commission nor SERI can be held responsible for them. We thank Antoine Baillod for his help with the codes and insightful discussions.

Appendix A

Linear stability in SPEC has been verified against numerical calculations of Δ0subscriptsuperscriptΔ0\Delta^{\prime}_{0}roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for classical tearing modes (Loizu et al., 2020). The stability for equilibria with bootstrap current is more involved. In Figure 10, we measure the size of a marginally stable slab as a function of the number of relaxed volumes. Each set of points in the plot represents marginal stability, such that points above are unstable and below stable. The top curve in blue represents equilibria without bootstrap current, and it can be seen that it has good convergence as Nvolsubscript𝑁volN_{\textrm{vol}}italic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT is increased, in this case to a marginal size of 3.0similar-toabsent3.0\sim 3.0∼ 3.0. However, the same is not true for the equilibria with bootstrap current shown in red, for which the limit Nvolsubscript𝑁volN_{\textrm{vol}}\rightarrow\inftyitalic_N start_POSTSUBSCRIPT vol end_POSTSUBSCRIPT → ∞ does not seem to exist. This leads us to believe that such equilibria are linearly unconditionally unstable in SPEC, where stability in any particular equilibria is due to limited resolution. Such a conclusion is supported by the bootstrap term from the MRE (see Figure 1). In SPEC, the ratio of parallel to perpendicular heat conductivity is infinite, which means that the term that accounts for heat transport in the island becomes w0=0subscript𝑤00w_{0}=0italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0. This leads to the bootstrap contribution becoming Δbs1/wsimilar-tosubscriptsuperscriptΔbs1𝑤\Delta^{\prime}_{\textrm{bs}}\sim 1/wroman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ∼ 1 / italic_w, which also leads to unconditional linear stability in the MRE model.

Refer to caption
Figure 10: Size of a marginally stable slab in SPEC as a function of the number of volumes. The top curve (blue) is for the case with Jbs=0subscript𝐽bs0J_{\textrm{bs}}=0italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT = 0, which converges to a well-defined value. The bottom curve (red) has Jbs0subscript𝐽bs0J_{\textrm{bs}}\neq 0italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT ≠ 0 and it does not have a limit.

Appendix B

Here we derive the MRE for the width of tearing modes in slab with bootstrap current. Neglecting flows which are assumed to be weak, we start from Ohm’s law

ψt=η(Jz0Jz)+ηJbs𝜓𝑡𝜂subscript𝐽𝑧0subscript𝐽𝑧𝜂subscript𝐽bs\frac{\partial\psi}{\partial t}=\eta(J_{z0}-J_{z})+\eta J_{\textrm{bs}}divide start_ARG ∂ italic_ψ end_ARG start_ARG ∂ italic_t end_ARG = italic_η ( italic_J start_POSTSUBSCRIPT italic_z 0 end_POSTSUBSCRIPT - italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) + italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT (8)

Expressing the flux function as a sum of the equilibrium and perturbed part, ψ(x,y,t)=ψ0(x)+ψ1(x,y,t)𝜓𝑥𝑦𝑡subscript𝜓0𝑥subscript𝜓1𝑥𝑦𝑡\psi(x,y,t)=\psi_{0}(x)+\psi_{1}(x,y,t)italic_ψ ( italic_x , italic_y , italic_t ) = italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) + italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ), and assuming gradients in x𝑥xitalic_x dominate over those in y𝑦yitalic_y, /x/ymuch-greater-than𝑥𝑦\partial/\partial x\gg\partial/\partial y∂ / ∂ italic_x ≫ ∂ / ∂ italic_y,

ψ1t=η2ψ1x2+ηJbssubscript𝜓1𝑡𝜂superscript2subscript𝜓1superscript𝑥2𝜂subscript𝐽bs\frac{\partial\psi_{1}}{\partial t}=\eta\frac{\partial^{2}\psi_{1}}{\partial x% ^{2}}+\eta J_{\textrm{bs}}divide start_ARG ∂ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = italic_η divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT (9)

Assuming all quantities are of a single harmonicity in the poloidal y𝑦yitalic_y direction, k=2π/L𝑘2𝜋𝐿k=2\pi/Litalic_k = 2 italic_π / italic_L, the functions are expressed as f(x,y,t)f^(x,t)cos(ky)similar-to𝑓𝑥𝑦𝑡^𝑓𝑥𝑡𝑘𝑦f(x,y,t)\sim\hat{f}(x,t)\cos(ky)italic_f ( italic_x , italic_y , italic_t ) ∼ over^ start_ARG italic_f end_ARG ( italic_x , italic_t ) roman_cos ( italic_k italic_y ). Next, we integrate both sides from w/2𝑤2-w/2- italic_w / 2 to w/2𝑤2w/2italic_w / 2, where w𝑤witalic_w is the island width at time t𝑡titalic_t

w/2w/2𝑑xψ1t=w/2w/2𝑑xη2ψ1x2+w/2w/2𝑑xηJbssuperscriptsubscript𝑤2𝑤2differential-d𝑥subscript𝜓1𝑡superscriptsubscript𝑤2𝑤2differential-d𝑥𝜂superscript2subscript𝜓1superscript𝑥2superscriptsubscript𝑤2𝑤2differential-d𝑥𝜂subscript𝐽bs\int_{-w/2}^{w/2}dx\>\frac{\partial\psi_{1}}{\partial t}=\int_{-w/2}^{w/2}dx\>% \eta\frac{\partial^{2}\psi_{1}}{\partial x^{2}}+\int_{-w/2}^{w/2}dx\>\eta J_{% \textrm{bs}}∫ start_POSTSUBSCRIPT - italic_w / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_w / 2 end_POSTSUPERSCRIPT italic_d italic_x divide start_ARG ∂ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = ∫ start_POSTSUBSCRIPT - italic_w / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_w / 2 end_POSTSUPERSCRIPT italic_d italic_x italic_η divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ∫ start_POSTSUBSCRIPT - italic_w / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_w / 2 end_POSTSUPERSCRIPT italic_d italic_x italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT (10)

For the term on the left hand side, we apply the constant-ψ𝜓\psiitalic_ψ approximation to lift ψ𝜓\psiitalic_ψ outside the integral, and for the bootstrap term, we assume that the pressure gradient, and hence Jbssubscript𝐽bsJ_{\textrm{bs}}italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT, is slowly varying across the island. We obtain

ψ1tw=ηψ1x|w/2w/2+ηJbswsubscript𝜓1𝑡𝑤evaluated-at𝜂subscript𝜓1𝑥𝑤2𝑤2𝜂subscript𝐽bs𝑤\frac{\partial\psi_{1}}{\partial t}w=\eta\frac{\partial\psi_{1}}{\partial x}% \Bigr{|}_{-w/2}^{w/2}+\eta J_{\textrm{bs}}wdivide start_ARG ∂ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG italic_w = italic_η divide start_ARG ∂ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x end_ARG | start_POSTSUBSCRIPT - italic_w / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_w / 2 end_POSTSUPERSCRIPT + italic_η italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT italic_w (11)

To proceed further, we need an additional equation that relates the perturbed flux and instantaneous island width. Following an approach similar to Goldston (2020), the total flux is again composed of equilibrium and perturbed contributions. Expanding around the resonant surface, we use first-order expansion for the field By0(x)=By0xsubscript𝐵𝑦0𝑥subscriptsuperscript𝐵𝑦0𝑥B_{y0}(x)=B^{\prime}_{y0}xitalic_B start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT ( italic_x ) = italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT italic_x, which gives ψ0=1/2By0x2subscript𝜓012subscriptsuperscript𝐵𝑦0superscript𝑥2\psi_{0}=-1/2\>B^{\prime}_{y0}x^{2}italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 1 / 2 italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, since we know that By0(x)=ψ0/xsubscript𝐵𝑦0𝑥subscript𝜓0𝑥B_{y0}(x)=-{\partial\psi_{0}}/{\partial x}italic_B start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT ( italic_x ) = - ∂ italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ∂ italic_x. For the perturbed flux, we use the zeroth-order expansion in x𝑥xitalic_x and the fact that the island is of a single harmonicity ψ1(x,y,t)=ψ^cos(ky)subscript𝜓1𝑥𝑦𝑡^𝜓𝑘𝑦\psi_{1}(x,y,t)=-\hat{\psi}\cos(ky)italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ) = - over^ start_ARG italic_ψ end_ARG roman_cos ( italic_k italic_y ).

Evaluating the flux function at the X-point, we get ψ(x=0,y=0)=ψ^𝜓formulae-sequence𝑥0𝑦0^𝜓\psi(x=0,y=0)=-\hat{\psi}italic_ψ ( italic_x = 0 , italic_y = 0 ) = - over^ start_ARG italic_ψ end_ARG, and the following expression traces the field line (xlsubscript𝑥𝑙x_{l}italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, ylsubscript𝑦𝑙y_{l}italic_y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT) corresponding to that flux

ψ^(1cos(kyl))=1/2By0xl2^𝜓1𝑘subscript𝑦𝑙12subscriptsuperscript𝐵𝑦0subscriptsuperscript𝑥2𝑙\hat{\psi}(1-\cos(ky_{l}))=1/2\>B^{\prime}_{y0}x^{2}_{l}over^ start_ARG italic_ψ end_ARG ( 1 - roman_cos ( italic_k italic_y start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ) = 1 / 2 italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT (12)

The island width is then obtained from the maximum excursion of the field line in x𝑥xitalic_x

w(t)=2max(xl)=[16ψ^By0]1/2𝑤𝑡2subscript𝑥𝑙superscriptdelimited-[]16^𝜓subscriptsuperscript𝐵𝑦012w(t)=2\max(x_{l})=\left[\frac{16\hat{\psi}}{B^{\prime}_{y0}}\right]^{1/2}italic_w ( italic_t ) = 2 roman_max ( italic_x start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = [ divide start_ARG 16 over^ start_ARG italic_ψ end_ARG end_ARG start_ARG italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT end_ARG ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT (13)

Thus we obtain the relation between the reconnected flux and the island width

ψ=bw2 with b=By016𝜓𝑏superscript𝑤2 with 𝑏subscriptsuperscript𝐵𝑦016\psi=bw^{2}\>\>\textrm{ with }b=\frac{B^{\prime}_{y0}}{16}italic_ψ = italic_b italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with italic_b = divide start_ARG italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y 0 end_POSTSUBSCRIPT end_ARG start_ARG 16 end_ARG (14)

Finally, this expression is inserted into the relation Eq. B4, leading to the MRE model

2ηwt=Δ+1bJbs1w2𝜂𝑤𝑡superscriptΔ1𝑏subscript𝐽bs1𝑤\frac{2}{\eta}\frac{\partial w}{\partial t}=\Delta^{\prime}+\frac{1}{b}J_{% \textrm{bs}}\frac{1}{w}divide start_ARG 2 end_ARG start_ARG italic_η end_ARG divide start_ARG ∂ italic_w end_ARG start_ARG ∂ italic_t end_ARG = roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG italic_b end_ARG italic_J start_POSTSUBSCRIPT bs end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_w end_ARG (15)
with Δ=1ψ1ψ1x|w/2w/2with superscriptΔevaluated-at1subscript𝜓1subscript𝜓1𝑥𝑤2𝑤2\textrm{with }\Delta^{\prime}=\frac{1}{\psi_{1}}\frac{\partial\psi_{1}}{% \partial x}\Bigr{|}_{-w/2}^{w/2}with roman_Δ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x end_ARG | start_POSTSUBSCRIPT - italic_w / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_w / 2 end_POSTSUPERSCRIPT (16)

References

  • de Aguilera et al. (2015) de Aguilera, Adriana M., Castejón, Francisco, Ascasíbar, Enrique, Blanco, Emilio, la Cal, Eduardo De, Hidalgo, Carlos, Liu, Bing, López-Fraguas, Antonio, Medina, Francisco, Ochando, María Antonia, Pastor, Ignacio, Ángeles Pedrosa, María, Milligen, Boudewijn Van, Velasco, José Luis & the TJ-II Team 2015 55 (11), 113014.
  • Baillod et al. (2021) Baillod, A., Loizu, J., Qu, Z.S., Kumar, A. & Graves, J.P. 2021 Computation of multi-region, relaxed magnetohydrodynamic equilibria with prescribed toroidal current profile. Journal of Plasma Physics 87 (4), 905870403.
  • Berger (1999) Berger, Mitchell A 1999 Introduction to magnetic helicity. Plasma Physics and Controlled Fusion 41 (12B), B167.
  • Brunetti et al. (2014) Brunetti, D., Graves, J.P., Cooper, W.A. & Terranova, D. 2014 Ideal saturated mhd helical structures in axisymmetric hybrid plasmas. Nuclear Fusion 54 (6), 064017.
  • Cooper et al. (2010) Cooper, W. A., Graves, J. P., Pochelon, A., Sauter, O. & Villard, L. 2010 Tokamak magnetohydrodynamic equilibrium states with axisymmetric boundary and a 3d helical core. Phys. Rev. Lett. 105, 035003.
  • Courant et al. (1967) Courant, R., Friedrichs, K. & Lewy, H. 1967 On the partial difference equations of mathematical physics 11 (2), 215–234.
  • Escande & Ottaviani (2004) Escande, D.F. & Ottaviani, M. 2004 Simple and rigorous solution for the nonlinear tearing mode. Physics Letters A 323 (3), 278–284.
  • Fitzpatrick (1995) Fitzpatrick, Richard 1995 Helical temperature perturbations associated with tearing modes in tokamak plasmas. Physics of Plasmas 2, 825–838.
  • Geiger J. E. (2004) Geiger J. E., Weller, A. Zarnstorff M. C. Nührenberg C. Werner A. H. F. Kolesnichenko Y. I. W7-AS Team and Neutral Beam Injection Group 2004 Equilibrium and stability of high-b plasmas in wendelstein 7-as. Fusion Science and Technology 46 (1), 13–23.
  • Glasser et al. (1976) Glasser, A. H., Greene, J. M. & Johnson, J. L. 1976 Resistive instabilities in a tokamak. The Physics of Fluids 19 (4), 567–574.
  • Goldston (2020) Goldston, R.J 2020 Introduction to Plasma Physics. CRC Press.
  • Günter et al. (1999) Günter, S, Gude, A, Maraschek, M, Yu, Q & the ASDEX Upgrade Team 1999 Influence of neoclassical tearing modes on energy confinement. Plasma Physics and Controlled Fusion 41 (6), 767.
  • Helander & Sigmar (2005) Helander, Per & Sigmar, D. J. 2005 Collisional transport in magnetized plasmas. Cambridge University Press.
  • Hirshman & Whitson (1983) Hirshman, S. P. & Whitson, J. C. 1983 Steepest-descent moment method for three-dimensional magnetohydrodynamic equilibria. The Physics of Fluids 26 (12), 3553–3568.
  • Hoelzl et al. (2021) Hoelzl, M., Huijsmans, G.T.A., Pamela, S.J.P., Bécoulet, M., Nardon, E., Artola, F.J., Nkonga, B., Atanasiu, C.V., Bandaru, V., Bhole, A., Bonfiglio, D., Cathey, A., Czarny, O., Dvornova, A., Fehér, T., Fil, A., Franck, E., Futatani, S., Gruca, M., Guillard, H., Haverkort, J.W., Holod, I., Hu, D., Kim, S.K., Korving, S.Q., Kos, L., Krebs, I., Kripner, L., Latu, G., Liu, F., Merkel, P., Meshcheriakov, D., Mitterauer, V., Mochalskyy, S., Morales, J.A., Nies, R., Nikulsin, N., Orain, F., Pratt, J., Ramasamy, R., Ramet, P., Reux, C., Särkimäki, K., Schwarz, N., Verma, P. Singh, Smith, S.F., Sommariva, C., Strumberger, E., van Vugt, D.C., Verbeek, M., Westerhof, E., Wieschollek, F. & Zielinski, J. 2021 The jorek non-linear extended mhd code and applications to large-scale instabilities and their control in magnetically confined fusion plasmas. Nuclear Fusion 61 (6), 065001.
  • Hole et al. (2006) Hole, M. J., Hudson, S. R. & Dewar, R. L. 2006 Stepped pressure profile equilibria in cylindrical plasmas via partial taylor relaxation. Journal of Plasma Physics 72 (6), 1167–1171.
  • Huang & Bhattacharjee (2016) Huang, Yi-Min & Bhattacharjee, A. 2016 Turbulent magnetohydrodynamic reconnection mediated by the plasmoid instability. The Astrophysical Journal 818 (1), 20.
  • Hudson et al. (2012) Hudson, S. R., Dewar, R. L., Dennis, G., Hole, M. J., McGann, M., von Nessi, G. & Lazerson, S. 2012 Computation of multi-region relaxed magnetohydrodynamic equilibria. Physics of Plasmas 19.
  • Kong (2020) Kong, Mengdi 2020 Towards integrated control of tokamak plasmas: physics-based control of neoclassical tearing modes in the tcv tokamak p. 228.
  • Kong et al. (2022) Kong, M, Felici, F, Sauter, O, Galperti, C, Vu, T, Ham, C J, Hender, T C, Maraschek, M, Reich, M, the TCV team, the ASDEX Upgrade team, the MAST team & the EUROfusion MST1 Team 2022 Physics-based control of neoclassical tearing modes on tcv. Plasma Physics and Controlled Fusion 64 (4), 044008.
  • Kumar (2022) Kumar, A. 2022 Plasma Physics and Controlled Fusion 64 (6), 065001.
  • Loizu & Bonfiglio (2023) Loizu, J. & Bonfiglio, D. 2023 Nonlinear saturation of resistive tearing modes in a cylindrical tokamak with and without solving the dynamics. Journal of Plasma Physics 89 (5), 905890507.
  • Loizu et al. (2020) Loizu, J., Huang, Y. M., Hudson, S. R., Baillod, A., Kumar, A. & Qu, Z. S. 2020 Direct prediction of nonlinear tearing mode saturation using a variational principle. Physics of Plasmas 27.
  • Loureiro et al. (2005) Loureiro, N. F., Cowley, S. C., Dorland, W. D., Haines, M. G. & Schekochihin, A. A. 2005 x𝑥xitalic_x-point collapse and saturation in the nonlinear tearing mode reconnection. Phys. Rev. Lett. 95, 235003.
  • Militello & Porcelli (2004) Militello, Fulvio & Porcelli, F. 2004 Simple analysis of the nonlinear saturation of the tearing mode. Physics of Plasmas 11, L13.
  • Muraglia et al. (2021) Muraglia, M., Poyé, A., Agullo, O., Dubuit, N. & Garbet, X. 2021 Nonlinear dynamics of ntm seeding by turbulence. Plasma Physics and Controlled Fusion 63.
  • Rutherford (1973) Rutherford, P. H. 1973 Nonlinear growth of the tearing mode. The Physics of Fluids 16 (11), 1903–1908.
  • Sauter et al. (2002) Sauter, O, Buttery, R J, Felton, R, Hender, T C, Howell, D F & contributors to the EFDA-JET Workprogramme 2002 Marginal beta-limit for neoclassical tearing modes in jet h-mode discharges. Plasma Physics and Controlled Fusion 44 (9), 1999.
  • Taylor (1974) Taylor, J. B. 1974 Relaxation of toroidal plasma and generation of reverse magnetic fields. Phys. Rev. Lett. 33, 1139–1141.
  • Teng et al. (2018) Teng, Q., Ferraro, N., Gates, D.A. & White, R.B. 2018 Nonlinear simulations of thermo-resistive tearing mode formalism of the density limit. Nuclear Fusion 58 (10), 106024.
  • Watanabe et al. (2005) Watanabe, K.Y., Sakakibara, S., Narushima, Y., Funaba, H., Narihara, K., Tanaka, K., Yamaguchi, T., Toi, K., Ohdachi, S., Kaneko, O., Yamada, H., Suzuki, Y., Cooper, W.A., Murakami, S., Nakajima, N., Yamada, I., Kawahata, K., Tokuzawa, T., Komori, A. & experimental group, LHD 2005 Effects of global mhd instability on operational high beta-regime in lhd. Nuclear Fusion 45 (11), 1247.