Microtearing modes (MTMs) are suggested as a candidate for anomalous thermal transport in tokamak H-mode discharges. This study investigates MTMs in tokamak plasmas, employing simulations in the BOUT++ framework. It simplifies and linearizes the governing equations in detailed linear simulations. The study meticulously evaluates various conductivity models under diverse plasma conditions and collision regimes. The research thoroughly assesses different conductivity models across a range of plasma conditions and collision regimes. A unified dispersion relation that includes both MTM and Drift-Alfvén Wave (DAW) instabilities is derived, showing that DAW and MTM instabilities occur at varying distances from the rational surface. Specifically, MTMs become unstable near the rational surface but stabilize farther away, while drift-Alfvén instability appears farther from the rational surface. The study also re-derives MTM dispersion relations using Ohm's law and the vorticity equation, providing a thorough analysis of electromagnetic and electrostatic interactions in tokamaks. Global simulations demonstrate an inverse correlation between MTM growth rates and collisionality, and a direct correlation with temperature gradients. The nonalignment of the rational surface with the peak stabilizes the MTMs. Nonlinear simulations highlight electron temperature relaxation as the primary saturation mechanism for MTMs, with magnetic flutter identified as the dominant mode of electron thermal transport.
I. INTRODUCTION
The microtearing mode (MTM) is an electromagnetic micro-instability that significantly affects thermal transport and energy confinement in the tokamak plasma. In standard tokamak H-mode discharges, such as those observed in DIII-D,1 JET,2 and others, MTMs are frequently observed. MTM can dominate electron thermal transport in the JET ITER-like wall pedestal under certain conditions,3 plays a role in determining the behavior of electron temperature during an ELM cycle,4 and in double barrier operations, it generates a substantial electron thermal transport at the Inner Transport Barrier (ITB).5 In the high-β regime, MTMs may become more prominent as KBMs enter their second stable regime, thereby playing a significant role alongside the behavior of KBMs. Therefore, MTMs can become the dominant low- modes as β increase in spherical tokamak devices like NSTX.6,7
MTM was initially recognized as a temperature-gradient-driven tearing mode in the ST Tokamak in the 1970s. Using a variant method, Hazeltine et al.8 first proposed a linear kinetic conductivity model for this type of mode. This kinetic theory was further developed by Drake et al.,9 with the model being extended to a semi-collisional regime. More recently, Larakers et al.10,11 have successfully expanded the theory to include the global radial variation effect, providing a comprehensive explanation for the separate bands observed in the MTM frequency spectrum.
BOUT++ is a C++ framework specifically developed for solving initial value partial differential equations12 related to magnetized plasmas. It ensures a clear delineation between numerical and physics development coding, thereby enhancing the efficiency of the simulation code development process. Successfully applied in tokamak boundary plasma simulations,13,14 including edge-localized modes, boundary turbulence, and blobs. Plasma models in BOUT++ have been validated with experimental data and results from other codes.
In this work, we have extended the capabilities of the BOUT++ code to include the simulation of the MTM. This paper presents our global simulation of the collisional MTM in a shifted-circle tokamak geometry. We have implemented Hassam's fluid model for the MTM,15 incorporating electron inertia and a time-dependent thermal force into Ohm's law equation. Both global linear and nonlinear simulations are performed to investigate the MTM characteristics.
This paper is organized as follows. In Sec. II, a brief survey of existing theoretical models on MTM is provided. Section III describes the simulation model and parameters used in this study in detail. Linear simulation results are summarized in Sec. IV as we explore the mode structures and the relation between frequency and growth rate spectrum and collisionality and temperature gradient. The effect of alignment on the growth rate is investigated as well. Nonlinear MTM simulations are presented in Sec. V. We find the saturation mechanism of MTMs due to temperature profile relaxation and compare the heat transport resulting from E × B convection and magnetic flutter. Finally, Sec. VI summarizes our findings.
II. THEORY OF MICROTEARING MODE
In this section, we delve into the theory of MTMs, examining its dispersion relation and exploring its connection to classical tearing modes as well as DAW instabilities.
A. Review and comparison of MTM models
In summary, for local assumption around the rational surface, Larakers's conductivity model reduces to Drake's model in both and limits, and reduces to Hassam's model in limit. Larakers's model thus bridges the two limits and also includes effects.
B. Dispersion relations of MTM
Diagram of the normalized distribution of absolute parallel magnetic vector potential, , calculated from Eq. (18) using a shooting method. The horizontal axis denotes the distance with respect to the rational surface. The parameters are, , δ = 10, ky = 0.
Diagram of the normalized distribution of absolute parallel magnetic vector potential, , calculated from Eq. (18) using a shooting method. The horizontal axis denotes the distance with respect to the rational surface. The parameters are, , δ = 10, ky = 0.
However, across the current layer, there is a discontinuity of due to the current sheet, and the prime denotes derivative in the radial direction. This discontinuity can be represented by a tearing parameter, calculated under the constant- assumption, . The is solved in the outer region. Approximately , and .19 Then the solution of Eq. (18) can be obtained from an integration within the current layer width, Δj.18
Numerical solution of dispersion relation Eq. (21). Panels (a) and (b) show the frequency and growth rate vs normalized .
Numerical solution of dispersion relation Eq. (21). Panels (a) and (b) show the frequency and growth rate vs normalized .
III. FLUID MODEL AND SIMULATION SETUP
In this section, we introduce the control equations, equilibrium profiles, and parameters used in this study.
A. Microtearing mode model for simulation
If the diamagnetic drift effect is considered (i.e., finite ), then the second term in the right-hand side of vorticity Eq. (24) must be included to maintain equilibrium.26 Note in our simulations, cold ion model was adapted so that the perturbation of ion density and temperature vanish, making the second term in the right-hand side of Eq. (27) always zero. At the same time, the Er shear flow, , is included to balance equilibrium ion diamagnetic drift for particle conservation.27 Also, Er shear flow introduces the Doppler shift effect, . Consequently, when referring to the equilibrium ion diamagnetic drift effect, we consider the equilibrium flow induced by equilibrium ion diamagnetic drift, balanced by the Er shear flow-essentially investigating the impact of Er shear flow.
B. Simulation setup
In our simulation, we utilize a shifted circular geometry equilibrium generated by CORSICA.28 The equilibrium profile is displayed in Fig. 3. Our simulation domain extends from the flux surface to . To simplify the physics and focus on the impact of temperature gradients, we maintain a constant density of throughout the domain, thereby avoiding the influence of density gradients.
(a) Electron temperature equilibrium profile, (b) safety factor profile, and (c) profile normalized to its toroidal mode number.
(a) Electron temperature equilibrium profile, (b) safety factor profile, and (c) profile normalized to its toroidal mode number.
The safety factor profile with for the simulation domain is intentionally set to be flatter than usual to distinguish rational surfaces clearly for discussions. In contrast to the standard CORSICA-generated equilibrium, our equilibrium features a lower current profile and a stronger magnetic field. It's worth mentioning that Seto et al.29 have made significant developments in BOUT++ to enable accurate low-n mode simulations, although this version may come with slower code execution speed. So, we do not perform very low-n modes simulation.
A field-aligned coordinate is implemented in BOUT++,30 where , with ν being the local field-line pitch. Therefore, is along the cross flux direction, and are along the parallel and bi-normal directions. For linear simulation, the grid points are in x, y, and z directions respectively, and for nonlinear simulation, . It is a twist-shift boundary condition in the y direction and periodic in the Z direction. For , the boundary condition is the Neumann condition in the inner boundary and the Dirichlet in the outer boundary; for U, it is Dirichlet in both boundaries. Other parameters are: Alfvén time , normalization length , Alfvén speed , normalization magnetic field , and normalization electron temperature .
IV. GLOBAL LINEAR SIMULATION OF MTM
Electrostatic potential and normalized parallel magnetic vector potential are solved from Eqs. (36) and (37) by the invert-Laplace Laplacian equation solver of BOUT++. This solver is introduced in detail in the Appendix. In our simulations, we exclude the current driven term by default, , to focus on the temperature gradient-driven MTM.
A. Mode structure
A linear simulation of n = 8 is performed. Figure 4 displays the poloidal and radial distributions of and at in the linear simulation. These mode structures exhibit typical tearing parities, with being even parity and odd parity to the rational surface q = 2.75. In Fig. 5, a Poincaré plot reveals the formation of magnetic island chains around rational surfaces, providing compelling evidence for the identification of the MTM.
Poloidal distributions of (a) normalized parallel magnetic vector (b) normalized electrostatic potential , and (c) normalized electron temperature perturbation for n = 8 MTM. Radial distributions of (d) (e) , and (f) at the inner mid-plane. is normalized to is normalized to , and is normalized to .
Poloidal distributions of (a) normalized parallel magnetic vector (b) normalized electrostatic potential , and (c) normalized electron temperature perturbation for n = 8 MTM. Radial distributions of (d) (e) , and (f) at the inner mid-plane. is normalized to is normalized to , and is normalized to .
Poicare plot in a poloidal slice presents magnetic island chains in the linear simulation at . The horizontal axis denotes the poloidal angle, where is the out mid-plane. The vertical axis denotes the normalized flux surface. The dashed red line indicates the location of rational surface.
Poicare plot in a poloidal slice presents magnetic island chains in the linear simulation at . The horizontal axis denotes the poloidal angle, where is the out mid-plane. The vertical axis denotes the normalized flux surface. The dashed red line indicates the location of rational surface.
B. Dependency on collisionality and temperature gradient
We perform a scan in with for the n = 8 mode to investigate the effects of collisionality. As shown in Fig. 6(a), the growth rate decreases as collisionality increases. In comparison with the simulation, the local theory tends to overestimate the linear growth rate. This overestimation is attributed to the theory being local and not considering magnetic shear effects. The declining growth rate in the collisional regime with increasing collisionality is a distinct characteristic that sets it apart from the classical tearing mode. In Fig. 6(b), it is evident that collisionality has a relatively minor impact on mode frequency in the collisional regime. The scan in indicates that collisionality plays a stabilizing effect for MTM. However, this finding is a result of our theoretical model, which is only valid in the strongly collisional regime. In general, MTMs are destabilized by collisionality in the presence of a temperature gradient. In the weak collision regime, the growth rate is proportional to collisionality, indicating that collisionality destabilizes MTM,5 which distinguishes it from drift-wave instabilities.
(a) Linear growth rate γ and (b) mode frequency vs for . The blue solid is calculated by Eq. (38). The scatter plots are the simulation results, where the red solid dots denote the simulations without curvature and diamagnetic drift effects, the magenta stars denote the simulations with curvature effect included, and the green squares denote the simulations with both curvature and diamagnetic drift effects included. (c) Linear growth rate and (d) mode frequency vs CT for .
(a) Linear growth rate γ and (b) mode frequency vs for . The blue solid is calculated by Eq. (38). The scatter plots are the simulation results, where the red solid dots denote the simulations without curvature and diamagnetic drift effects, the magenta stars denote the simulations with curvature effect included, and the green squares denote the simulations with both curvature and diamagnetic drift effects included. (c) Linear growth rate and (d) mode frequency vs CT for .
Another scan is performed in CT using the same approach, as presented in Fig. 6(c). As CT increases, the growth rate shows a proportional increase. This is because, with the higher CT, effectively both the thermal force and the time-dependent thermal force increase. These two forces drive the unstable MTM. In other words, when CT is higher, there is more effective free energy from the electron temperature gradient for electron motion and leads to greater instability in the MTM. This provides evidence that the MTM is primarily driven by the temperature gradient, as opposed to the current gradient characteristic for the classical tearing mode. The mode frequency is noticeably influenced by CT, as shown in Fig. 6(d). An increase in CT leads to a corresponding increase in the MTM frequency. This observation suggests that the MTM frequency is directly proportional to the electron diamagnetic frequency.
The effects of curvature and ion diamagnetic drift (correspondingly with the Er shear flow induced) are investigated. As illustrated in Fig. 6, the magenta stars denote the cases with curvature drive [denoted by the third term on the RHS of Eq. (33)] included and the green squares denote the cases with both curvature and ion diamagnetic drift effect included. The curvature almost has no influence on the linear growth rate and mode frequency indicating that the MTM is not a curvature-driven mode.
The reduction in growth rate is attributed to the non-local shear of E × B flow. As illustrated in Fig. 7, RMS distributions at outer mid-plane of different collisionality are presented. Following the definition of Yagyu and Numata,31 the current layer width δc is defined by the full width at half maximum height of where the height is measured from the top to the first local minimal value. The current layer widths of are . MTM is localized around the bottom of the Er well where the local shear is nearly zeros. In such instances, the Er shear flow has minimal impact on the linear growth rate. However, as collisionality increases, the current layer width broadens. Consequently, the Er shear flow could suppress the MTM leading to a decrease in the linear growth rate.
Diagram of Er profile and distributions of RMS at outer mid-plane of n = 8 modes for , and with .
Diagram of Er profile and distributions of RMS at outer mid-plane of n = 8 modes for , and with .
C. Alignment of rational surfaces with the profile
MTM is unstable only when the rational surface is in good alignment with the peak of the electron diamagnetic frequency, , profile. As pointed out by Larakers and Curie,32 MTM is unstable with , where Δs is the distance that one rational surface shifts away from the peak of the location of and is the characteristic width of profile. Once the rational surface is too far away from the peak , MTM cannot survive. The competition between the nonalignment damping effect and gradient-driven effect determines the unstable regime of MTM. The nonalignment is a stabilizing effect. Our conclusion is consistent with Larakers's global calculation,11 Curie's SLiM model,32,33 and also Hassan's pedestal simulation.34
A linear simulation is conducted to confirm this prediction. The equilibrium temperature and safety factor profiles align with those shown in Fig. 3. Toroidal modes are scanned from n = 5 to 11. In Fig. 8(a), the locations of rational surfaces for different toroidal modes are labeled with various colored vertical lines. Correspondingly, Fig. 8(b) displays the growth rate spectrum. The red dotted line denotes the cases without curvature effect and peeling drive effect [related to and denoted by the second term on the RHS of Eq. (33)], the blue dotted line denotes the cases with peeling drive effect included, and the magenta line denotes the cases with both curvature and peeling drive effect included.
(a) The diagram of profile and location of rational surfaces of different toroidal mode numbers; (b) the growth rate of different toroidal mode number cases. (c)–(e) presents the mode structure of different poloidal number modes corresponding to n = 8, 9, 11 cases. The dashed vertical red line denotes the location of peak and the dashed vertical gray lines denote the location of rational surfaces. The solid black curve represents the safety factor profile.
(a) The diagram of profile and location of rational surfaces of different toroidal mode numbers; (b) the growth rate of different toroidal mode number cases. (c)–(e) presents the mode structure of different poloidal number modes corresponding to n = 8, 9, 11 cases. The dashed vertical red line denotes the location of peak and the dashed vertical gray lines denote the location of rational surfaces. The solid black curve represents the safety factor profile.
The rational surface of n = 8 almost perfectly aligns with the peak of , while the rational surface of n = 9 is slightly shifted. Under the discussion in the previous section, when the drive increases, the growth rate is correspondingly higher. A higher toroidal mode number leads to a more pronounced profile since , and then increases the drive. The mode structures of the most unstable mode, n = 8, 9, 11, in Fig. 8(b) are presented in Figs. 8(c)–8(e). Our global simulations show that the perturbations peak at the location of peak but not always at the rational surface. For n = 9, 11 modes, the most unstable m modes offset the rational surface and the symmetry of the mode structure breaks down. As predicted by the local theory, the n = 9 mode is expected to have a larger growth rate than that of n = 8. However, n = 9 has a smaller growth rate despite having a larger . This is due to the non-local nonalignment effect, the linear growth rate of the n = 9 mode decreases.
V. GLOBAL NONLINEAR SIMULATION OF MTM
In this section, we conduct an initial nonlinear simulation to explore the nonlinear saturation mechanism of MTM. We will compare radial thermal transport resulting from convection and magnetic flutter. The nonlinear simulation retains all terms except the peeling drive in Eqs. (24)–(26) and utilizes the same equilibrium as in the previous section. Also, the Er shear flow is included by default in the following simulations.
A. Saturation of collisional MTM due to profile relaxation
The surface-averaged perturbations of the absolute values of electrostatic potential and normalized parallel magnetic vector potential are presented in Fig. 9(a). After approximately time, the MTM enters a saturation state. Notably, the temperature profile changes the evolution. We depict the initial temperature profile with a dashed black curve, and the profile after 3000 Alfvén times with a solid blue curve, as shown in Fig. 9(b). The corresponding normalized gradient profiles are shown as dotted lines with the same colors.
(a) The time evolution of the root mean square (RMS) of magnetic perturbations is depicted. The solid blue and orange curves represent the evolution of flux surface-averaged magnetic vector potential and electrostatic potential perturbations at the q = 2.75 rational surface, respectively. The dashed curves indicate the perturbation evolution with equilibrium Er shear flow. (b) The electron temperature profiles at different time are illustrated. The dashed black curve represents the initial profile, the blue curve represents the profile after , and the red curve represents the profile after with equilibrium Er shear flow. The corresponding normalized temperature gradient profiles are plotted in dotted lines with the same color.
(a) The time evolution of the root mean square (RMS) of magnetic perturbations is depicted. The solid blue and orange curves represent the evolution of flux surface-averaged magnetic vector potential and electrostatic potential perturbations at the q = 2.75 rational surface, respectively. The dashed curves indicate the perturbation evolution with equilibrium Er shear flow. (b) The electron temperature profiles at different time are illustrated. The dashed black curve represents the initial profile, the blue curve represents the profile after , and the red curve represents the profile after with equilibrium Er shear flow. The corresponding normalized temperature gradient profiles are plotted in dotted lines with the same color.
It is evident that the temperature profile changes, with the peak gradient decreasing occurring around the q = 2.75 rational surface. Consequently, the linear-driven force decreases, leading to a gradual cessation of perturbation growth, ultimately resulting in mode saturation. In our nonlinear simulations, Only the zonal component of temperature perturbation is kept, the zonal component of and are removed, and correspondingly, zonal flow and zonal field are not included. We are not going to discuss how they affect the saturation process and will discuss it in another paper. In comparison with the scenario without the equilibrium Er shear flow, perturbations exhibit a reduced linear growth rate but maintain a similar amplitude in the nonlinear phase. Without the equilibrium Er shear flow, the temperature gradient exhibits a more significant decrease.
B. Electron radial thermal transport
To illustrate the MTM thermal transport, we use the case with Er shear flow as an example. The normalized χe of the two components are shown in Fig. 10. The normalized factor represents the gyro-Bohm diffusion coefficient, and is the ion gyroradius normalized to minor radius. Figures 10(a) and 10(c) present the temporal evolution of the distributions of the two components of χe. In comparison with the magnetic flutter component, the electron thermal transport by E × B convection is negligible. This is attributed to the changes in magnetic field topology induced by MTM. It can be illustrated in Fig. 11. In the nonlinear phase, such as , the original closed flux surfaces break down due to the strong magnetic perturbation compared to that in the linear phase at . There forms a magnetic island chain around the q = 2.75 rational surface and a stochastic magnetic field in the major domain between q = 2.67 and 2.833 rational surfaces. For the electrons, the parallel conductivity is large, and the parallel thermal transport has a radial component due to the magnetic perturbation. Therefore, the electron thermal transport is enhanced, and the magnetic flutter component dominates the electron thermal transport.
The surface averaged thermal transport coefficient is normalized by the Bohm diffusion coefficient. (a) and (c) represent the temporal evolution of radial distributions of magnetic flutter component and E × B convection respectively. The horizontal dashed white line denotes the evolution of χe at rational surface. Correspondingly, they are plotted in solid lines in panels (b) and (d).
The surface averaged thermal transport coefficient is normalized by the Bohm diffusion coefficient. (a) and (c) represent the temporal evolution of radial distributions of magnetic flutter component and E × B convection respectively. The horizontal dashed white line denotes the evolution of χe at rational surface. Correspondingly, they are plotted in solid lines in panels (b) and (d).
Poincare plots in the nonlinear simulation at (top), (middle), and (bottom), corresponding to the simulation case with Er shear flow. The horizontal dashed red lines show the location of different rational surfaces.
Poincare plots in the nonlinear simulation at (top), (middle), and (bottom), corresponding to the simulation case with Er shear flow. The horizontal dashed red lines show the location of different rational surfaces.
In our simulation, we observe an asymmetric spreading of MTM turbulence. There is an evident inward spreading as indicated in Fig. 10(a). Such an inward spreading intensifies the thermal transport near the pedestal top. As shown in Fig. 10(b), the thermal transport coefficient at increases after a temporal delay and eventually reaches the same level as that at .
C. The influence of the free streaming parameter αH on electron radial thermal transport
In this section, we discuss how the free streaming parameter αH in the flux limit heat flux model affects the electron thermal transport in the global nonlinear simulations.
(a)Total effective correlation length , parallel correlation length , collision mean free path , and parallel effective correlation length vs αH; (b) temporal evolution of flux surface averaged at q = 2.75 rational surface. The dashed horizontal lines denote the period-averaged within . The averaged are for . (c) The electron thermal transport coefficient of the three cases within the nonlinear phase . The averaged values are denoted by the dashed horizontal lines and they are 9.45, 11.53, and 2.77 normalized to DgB for respectively.
(a)Total effective correlation length , parallel correlation length , collision mean free path , and parallel effective correlation length vs αH; (b) temporal evolution of flux surface averaged at q = 2.75 rational surface. The dashed horizontal lines denote the period-averaged within . The averaged are for . (c) The electron thermal transport coefficient of the three cases within the nonlinear phase . The averaged values are denoted by the dashed horizontal lines and they are 9.45, 11.53, and 2.77 normalized to DgB for respectively.
The electron parallel heat flux plays two roles. It does not only contribute to a large thermal transport but also dissipates the perturbations. As αH increases, increases correspondingly. However, a larger χfl makes perturbation, like , better damped as shown in Fig. 12(b). An increase in αH from 0.03 to 0.15 makes the perturbation decrease from to . Therefore, these two effects compete to determine electron thermal transport. As shown in Fig. 12(c), when , the averaged electron thermal transport coefficient is the largest in comparison with that of and .
At the q = 2.75 rational surface, (a) temporal evolution of total magnetic flutter electron thermal transport coefficient (blue), , for ; (b) the dots in different colors denote the calculated at each time step of different αH cases within . Black curves in different styles are calculated from Eq. (50). is the electron gyro-radius normalized to minor radius; (c) time averaged thermal transport coefficient (red dots) vs the free streaming parameter αH.
At the q = 2.75 rational surface, (a) temporal evolution of total magnetic flutter electron thermal transport coefficient (blue), , for ; (b) the dots in different colors denote the calculated at each time step of different αH cases within . Black curves in different styles are calculated from Eq. (50). is the electron gyro-radius normalized to minor radius; (c) time averaged thermal transport coefficient (red dots) vs the free streaming parameter αH.
A thermal transport coefficient in this form indicates that thermal transport can be conducted by a stochastic magnetic field.36,37 It is consistent with the discussion in Sec. V B.
However, from the discussion above, we learn that electron thermal transport is strongly dependent on the model of parallel heat flux. Our flux-limit heat flux model can capture some instinct characteristics of microtearing turbulence like the transport scaling law Eq. (51), but quantitatively the results are determined by the αH we chose.
VI. SUMMARY
In our research, we have developed a fluid simulation code for studying the collisional MTM within the BOUT++ framework. We should note that BOUT++ has been updated to include the capability for simulating this mode and make it possible to investigate how MTM interacts with other ion modes like the kinetic ballooning mode.
In our linear simulations, we observe that the MTM exhibits a tearing parity in its mode structure. We explore the dependencies of growth rate and mode frequency on collisionality and temperature gradient. Within the collisional regime, the linear growth rate of the MTM decreases as collisionality increases, while it remains directly proportional to the electron temperature gradient. The Er shear flow causes a reduction in the linear growth rate especially when the collision is strong. Additionally, we investigate the non-local alignment effect, revealing that the MTM becomes unstable only when the rational surface aligns with the peak of the profile; and the nonalignment effect serves as a stabilizing factor.
In our nonlinear simulations, we find that the saturation of the MTM is primarily due to electron temperature profile relaxation with decreasing temperature gradient around the rational surface. Furthermore, we observe that electron thermal transport from magnetic flutter significantly exceeds that generated by convection due to the change in the topology of the magnetic field by MTM. Evident inward turbulence spreading is observed to intensify the thermal transport near the pedestal top. In our simulations, the electron thermal transport is quantitatively influenced by the heat flux closure or the free streaming parameter αH in the flux-limit heat flux model.
In the H-mode discharge, a substantial electron temperature gradient at the edge can supply the necessary free energy to destabilize microtearing modes (MTMs). These modes are distinguished by their narrow radial width while spanning an ion-mode length scale in the poloidal direction. Each m, n mode localizes around its respective rational surface, yet it may become unstable across multiple surfaces, influenced by the q profile and the electron diamagnetic frequency profile. Specifically, in the pedestal region, where the q profile can steepen, rational surfaces tend to cluster more closely, allowing MTMs to extend over a wider radial range. Consequently, MTMs can induce anomalous thermal transport in the H-mode pedestal, thereby affecting the confinement time. In NSTX H-mode plasmas, the confinement time is observed to be inversely proportional to the electron collisionality,38 , with MTM-driven transport being a potential contributor to this scaling, notably in the weakly collisional regime. Contrary to this, our simulations of collisional MTMs show a different trend where higher collisionality stabilizes MTMs, while lower collisionality induces unstable MTMs, leading to increased thermal turbulence transport and decreased confinement. The different results in comparison with the experiment are due to the collision regime. Either very low or high collisionality stabilizes MTMs, and our present modeling is limited to the high collisionality regime. We plan to investigate MTMs in the weakly collisional regime in future research.
Hassam's model is valid in the strong collision regime, and it limits our simulation to reach out to the weak collision regime, where . Unfortunately, in the edge pedestal of modern Tokamak devices, the electron diamagnetic frequency is close to the collision frequency, which means for the MTM . In the gyro-kinetic simulations of today's Tokamak devices, the most unstable mode is found to be .39 We also recognize additional mechanisms for MTM saturation, such as turbulence regulation by zonal flows (ZFs) and energy cascading due to nonlinear toroidal mode couplings.40 Our current BOUT++ simulations primarily focus on the relaxation of the electron temperature profile, partly because ZF calculations are still in development, resulting in their exclusion. Prioritizing the integration of ZFs in future research is essential to fully understand their regulatory role. We also plan to delve into the effects of energy cascading on MTM saturation, acknowledging its potentially significant impact. Our simulations are currently limited by the use of Hassam's model, which assumes , confining our analysis to lower-n modes. Our objective is to advance a fluid model of MTM that can handle arbitrary frequency responses, enabling the inclusion of higher-n modes to extensively assess energy cascading.
In our simulation, we are focused on investigating the unstable microtearing modes. To simplify the analysis, we have considered a limiting case with infinite ηe, thereby omitting the complexities associated with the density gradient and its stabilizing effects.31 This method enables us to pinpoint the maximum growth rate, indicative of the most unstable limit. Moving forward, we plan to integrate the density gradient effects into our research to comprehensively understand their impact on MTM stability.
From the discussion above, we find that the electron thermal transport of tearing mode turbulence in fluid simulations is strongly influenced by the heat flux model/closure. Simulations with the Landau-fluid closure21,41 and perhaps more sophisticated parallel heat flux models42–44 may provide quantitatively more accurate calculation of electron thermal transport. Other nonlinear effects like energy cascading due to toroidal toroidal mode coupling are believed to relieve the issue. Coupling both ion modes and MTM in one simulation to investigate the thermal transport for both ions and electrons is planned as a future study.
ACKNOWLEDGMENTS
The authors would like to thank Dr. N. Li, Dr. C. Xiao, Dr. Z. Guo, Dr. Y. Wang, Dr. C. Li, Dr. Y. Chen, and Dr. M. Curie for helpful discussions. K. Fan and C. Dong are supported by the National MCF Energy R&D Program (Grant No. 2018YFE0311300). B. Zhu and X. Q. Xu are supported by of the U.S. Department of Energy by Lawrence Livermore National Laboratory (Contract No. DE-AC52-07NA27344). T. Xia is supported by the National Natural Science Foundation of China (12175275) and the Youth Innovation Promotion Association Chinese Academy of Sciences (Y2021114). This research used resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the U.S. Department of Energy (Contract No. DE-AC02-05CH11231) using NERSC awards FES-ERCAP0024033, FES-ERCAP0023237, and FES-ERCAP0024539.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
K. Fan: Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). X. Q. Xu: Funding acquisition (equal); Investigation (equal); Methodology (equal); Software (equal); Supervision (equal); Writing – review & editing (equal). B. Zhu: Investigation (supporting); Methodology (equal); Software (equal); Writing – review & editing (equal). C. Dong: Investigation (supporting); Methodology (equal); Writing – review & editing (equal). T. Xia: Methodology (supporting); Software (supporting); Writing – review & editing (equal). Z. Li: Methodology (supporting); Writing – review & editing (supporting).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.