Reducing turbulent transport in tokamaks by combining intrinsic rotation and the low momentum diffusivity regime

Haomin Sun1,∗, Justin Ball1, Stephan Brunner1, Anthony Field2, Bhavin Patel2, Daniel Kennedy2, Colin Roach2, Diego Jose Cruz-Zabala3, Fernando Puentes Del Pozo3, Eleonora Viezzer3, and Manuel Garcia Munoz3 1Ecole Polytechnique Fédérale de Lausanne (EPFL), Swiss Plasma Center (SPC), CH-1015 Lausanne, Switzerland
2UKAEA (United Kingdom Atomic Energy Authority), Culham Campus, Abingdon, Oxfordshire, OX14 3DB, UK
3Department of Atomic, Molecular and Nuclear Physics, University of Seville, Seville, Spain
* haomin.sun@epfl.ch
(October 14, 2024)
Abstract

Based on the analysis of a large number of high-fidelity nonlinear gyrokinetic simulations, we propose a novel strategy to improve confinement in tokamak plasmas by combining up-down asymmetric flux surface shaping with the Low Momentum Diffusivity (LMD) regime. We show that the intrinsic momentum flux driven by up-down asymmetry creates strong flow shear in the LMD regime that can significantly reduce energy transport, increasing the critical gradient by up to 25%percent2525\%25 %. In contrast to traditional methods for generating flow shear, such as neutral beam injection, this approach requires no external momentum source and is expected to scale well to large fusion devices. The experimental applicability of this strategy in spherical tokamaks is addressed via simulations by considering actual equilibria from MAST and a preliminary equilibrium from SMART.

preprint: APS/123-QED

Introduction–It is well known that strong toroidal rotation (specifically E×B𝐸𝐵E\times Bitalic_E × italic_B rotation shear) can reduce turbulent transport [1, 2, 3, 4, 4, 5, 6, 7, 6, 8, 9, 10, 11, 11, 12, 13, 14, 11, 15, 16, 17, 18, 19] and stabilize MagnetoHydroDynamic (MHD) instabilities [20, 21, 22, 23, 24, 21, 22], thereby improving the prospects for tokamak fusion power plants. Traditional methods of driving plasma rotation such as Neutral Beam Injection (NBI) [25, 26, 27] or Radio Frequency (RF) waves [28, 29, 30, 31, 32] are not expected to scale well to large devices [33]. Taking NBI as an example, we see that the ratio of the momentum pNBIsubscript𝑝NBIp_{\text{NBI}}italic_p start_POSTSUBSCRIPT NBI end_POSTSUBSCRIPT over the energy ENBIsubscript𝐸NBIE_{\text{NBI}}italic_E start_POSTSUBSCRIPT NBI end_POSTSUBSCRIPT injected by the beam scales as pNBI/ENBI1/ENBIsimilar-tosubscript𝑝NBIsubscript𝐸NBI1subscript𝐸NBIp_{\text{NBI}}/E_{\text{NBI}}\sim 1/\sqrt{E_{\text{NBI}}}italic_p start_POSTSUBSCRIPT NBI end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT NBI end_POSTSUBSCRIPT ∼ 1 / square-root start_ARG italic_E start_POSTSUBSCRIPT NBI end_POSTSUBSCRIPT end_ARG [34]. Thus, the more energetic beams needed to penetrate into larger and denser plasmas will be less effective at driving rotation.

An intriguing alternative is to utilize intrinsic rotation [35, 36, 37], spontaneously driven by the plasma turbulence itself under certain conditions. The symmetry properties of gyrokinetics [5, 38, 39, 40] show that the only way to create flow of the order of the sound speed in the tokamak core is to break the up-down symmetry of magnetic flux surfaces about the midplane [41, 42]. This approach is independent of external sources and is expected to scale well to large devices (as it persists in the limit ρ=ρi/a0superscript𝜌subscript𝜌𝑖𝑎0\rho^{*}=\rho_{i}/a\to 0italic_ρ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_a → 0, where ρisubscript𝜌𝑖\rho_{i}italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is ion gyroradius and a𝑎aitalic_a is tokamak minor radius). One can calculate the radial profile of rotation in the steady-state operation of a tokamak by requiring the total toroidal angular momentum flux to be zero, which is well approximated by considering only the ion contribution (Πi=0subscriptΠ𝑖0\Pi_{i}=0roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0) due to their large mass compared to electrons. Since the momentum flux depends on both the toroidal ion rotation ΩisubscriptΩ𝑖\Omega_{i}roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and its radial derivative dΩi/dx𝑑subscriptΩ𝑖𝑑𝑥d\Omega_{i}/dxitalic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_d italic_x, it is common to Taylor expand ΠisubscriptΠ𝑖\Pi_{i}roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT about Ωi=0subscriptΩ𝑖0\Omega_{i}=0roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 and dΩi/dx=0𝑑subscriptΩ𝑖𝑑𝑥0d\Omega_{i}/dx=0italic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_d italic_x = 0 to find [41]

Πi(Ωi,dΩidx)Πi,intnimiR02DΠidΩidxnimiR02PΠiΩi=0,similar-to-or-equalssubscriptΠ𝑖subscriptΩ𝑖𝑑subscriptΩ𝑖𝑑𝑥subscriptΠ𝑖𝑖𝑛𝑡subscript𝑛𝑖subscript𝑚𝑖subscriptsuperscript𝑅20subscript𝐷Π𝑖𝑑subscriptΩ𝑖𝑑𝑥subscript𝑛𝑖subscript𝑚𝑖subscriptsuperscript𝑅20subscript𝑃Π𝑖subscriptΩ𝑖0\Pi_{i}\left(\Omega_{i},\frac{d\Omega_{i}}{dx}\right)\simeq\Pi_{i,int}-n_{i}m_% {i}R^{2}_{0}D_{\Pi i}\frac{d\Omega_{i}}{dx}-n_{i}m_{i}R^{2}_{0}P_{\Pi i}\Omega% _{i}=0,roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , divide start_ARG italic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_x end_ARG ) ≃ roman_Π start_POSTSUBSCRIPT italic_i , italic_i italic_n italic_t end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_Π italic_i end_POSTSUBSCRIPT divide start_ARG italic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_x end_ARG - italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_Π italic_i end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 , (1)

where Πi,int=Πi(0,0)subscriptΠ𝑖𝑖𝑛𝑡subscriptΠ𝑖00\Pi_{i,int}=\Pi_{i}(0,0)roman_Π start_POSTSUBSCRIPT italic_i , italic_i italic_n italic_t end_POSTSUBSCRIPT = roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 , 0 ) is the intrinsic momentum flux generated by up-down asymmetry and the second and third terms are the diffusive and pinch terms, respectively. Here, nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is ion density, misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the ion mass, R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the major radius of the tokamak, x𝑥xitalic_x is the radial coordinate, DΠisubscript𝐷Π𝑖D_{\Pi i}italic_D start_POSTSUBSCRIPT roman_Π italic_i end_POSTSUBSCRIPT is the angular momentum diffusion coefficient and PΠisubscript𝑃Π𝑖P_{\Pi i}italic_P start_POSTSUBSCRIPT roman_Π italic_i end_POSTSUBSCRIPT is the angular momentum pinch coefficient. Since the pinch term is typically smaller than the diffusive term [43, 44, 45, 46], we neglect it for now, but show in Ref. [47] that this will only cause an underestimation of the flow shear created by up-down asymmetry. The strength of the ion momentum diffusivity is typically compared with the ion heat diffusivity DQisubscript𝐷𝑄𝑖D_{Qi}italic_D start_POSTSUBSCRIPT italic_Q italic_i end_POSTSUBSCRIPT through the ion Prandtl number

Pri=DΠiDQi.subscriptPr𝑖subscript𝐷Π𝑖subscript𝐷𝑄𝑖\text{Pr}_{i}=\frac{D_{\Pi i}}{D_{Qi}}.Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = divide start_ARG italic_D start_POSTSUBSCRIPT roman_Π italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_D start_POSTSUBSCRIPT italic_Q italic_i end_POSTSUBSCRIPT end_ARG . (2)

This allows one to estimate the rotation shear being driven by a given turbulence level. A low Prandtl number means that a given momentum source (external or intrinsic) drives larger rotation shear at a given turbulence level. Without considering the pinch term, Eq. (1) indicates that maximizing Πi,intsubscriptΠ𝑖𝑖𝑛𝑡\Pi_{i,int}roman_Π start_POSTSUBSCRIPT italic_i , italic_i italic_n italic_t end_POSTSUBSCRIPT or minimizing ion momentum diffusivity will strengthen velocity shear. Maximizing Πi,intsubscriptΠ𝑖𝑖𝑛𝑡\Pi_{i,int}roman_Π start_POSTSUBSCRIPT italic_i , italic_i italic_n italic_t end_POSTSUBSCRIPT has been the subject of previous work [42]. The focus of this letter is to minimize the Prandtl number so as to help stabilize turbulence. In conventional aspect ratio tokamaks, Pri1similar-to-or-equalssubscriptPr𝑖1\text{Pr}_{i}\simeq 1Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≃ 1 has been experimentally observed [48, 49], motivating many theoretical works to assume this [50, 51, 41, 42]. Recently, however, it has been shown by gyrokinetic simulations that a significantly lower Prandtl number can be achieved [52, 53], especially at tight aspect ratio and low safety factor [54, 55, 56]. However, the Prandtl number calculations in many works were simplified, since either parallel momentum flux [54] or the toroidal angular momentum flux generated by flow shear parallel to the magnetic field [55, 56] was calculated. A more complete calculation must consider all components of toroidal angular momentum flux [38, 13] generated by toroidal flow shear as actually present in tokamaks [57]. A thorough scan of the dependence of the Prandtl number on the various geometric properties of tight aspect ratio tokamaks has not been performed, nor has a self-consistent study of the stabilization effect of intrinsic flow shear by up-down asymmetry on plasma turbulence. Despite recent experimental tokamak data showing that one can obtain a Prandtl number lower than one [7, 58, 59, 44, 60, 45, 46], these results are often hard to interpret given the challenge of separating the three terms in Eq. (1), especially for tight aspect ratio spherical tokamaks [61, 45]. All these theoretical and experimental issues call for a comprehensive theoretical study of toroidal angular momentum transport and how this can be combined with rotation drive mechanisms to stabilize turbulence.

We perform a thorough study of the Low Momentum Diffusivity (LMD) regime using a large number of high-fidelity nonlinear (NL) local gyrokinetic simulations. We first use a circular geometry to determine the parameter regime that achieves LMD. We then consider a tilted elliptical geometry in order to drive rotation from up-down asymmetry. By scanning the flow shear dΩi/dx𝑑subscriptΩ𝑖𝑑𝑥d\Omega_{i}/dxitalic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_d italic_x to find the value at which momentum flux is zero, we show that turbulence can self-consistently drive an intrinsic flow shear that significantly reduces the heat flux. The experimental applicability of this approach is addressed via simulations of a MAST tokamak equilibrium and a preliminary SMART equilibrium. To our knowledge, this work not only provides the first comprehensive numerical study on toroidal angular momentum diffusivity in the LMD regime, but also demonstrates the first potentially practical method of driving strong rotation shear in future tokamak power plants.

Methods–We use the well-benchmarked code GENE [62, 63] to perform a large number of flux tube (local) gyrokinetic simulations and model core plasma turbulence. We consider the Miller representation [64] to parameterize the flux surfaces. To make our analysis as accurate as possible, we consider the actual toroidal flow shear (including its parallel and E×B𝐸𝐵E\times Bitalic_E × italic_B components) to drive momentum flux as well as calculate the exact toroidal angular momentum flux (according to Eq. (A.2) in Ref. [65]).

Analysis–We start with circular concentric flux surfaces, an inverse aspect ratio ϵ=r0/R0=0.36italic-ϵsubscript𝑟0subscript𝑅00.36\epsilon=r_{0}/R_{0}=0.36italic_ϵ = italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.36 (r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the minor radial location of the flux tube), adiabatic electrons, equal electron and ion temperatures, and a density gradient of R0/Ln=2.22subscript𝑅0subscript𝐿𝑛2.22R_{0}/L_{n}=2.22italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2.22. These parameters are the same as the Cyclone Base Case (CBC) [66] (except the tighter aspect ratio) and produces Ion Temperature Gradient (ITG) driven turbulence. We perform a parameter scan over safety factor 1.05q4.551.05𝑞4.551.05\leq q\leq 4.551.05 ≤ italic_q ≤ 4.55, magnetic shear 0.1s^1.60.1^𝑠1.60.1\leq\hat{s}\leq 1.60.1 ≤ over^ start_ARG italic_s end_ARG ≤ 1.6 and ion temperature gradient 4.96R0/LTi12.964.96subscript𝑅0subscript𝐿𝑇𝑖12.964.96\leq R_{0}/L_{Ti}\leq 12.964.96 ≤ italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT ≤ 12.96, comprising a 3D data set with 576 NL gyrokinetic simulations, where Lnsubscript𝐿𝑛L_{n}italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and LTsubscript𝐿𝑇L_{T}italic_L start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT are the characteristic gradient scale lengths of density and temperature, respectively. To drive non-zero toroidal angular momentum flux and compute the momentum diffusivity, we impose toroidal flow shear dΩi/dx𝑑subscriptΩ𝑖𝑑𝑥d\Omega_{i}/dxitalic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_d italic_x including a parallel and a perpendicular component of ω=(r0/q)dΩi/dx=0.12cs/R0subscript𝜔perpendicular-tosubscript𝑟0𝑞𝑑subscriptΩ𝑖𝑑𝑥0.12subscript𝑐𝑠subscript𝑅0\omega_{\perp}=-(r_{0}/q)d\Omega_{i}/dx=0.12c_{s}/R_{0}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - ( italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_q ) italic_d roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_d italic_x = 0.12 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where cs=Te/misubscript𝑐𝑠subscript𝑇𝑒subscript𝑚𝑖c_{s}=\sqrt{T_{e}/m_{i}}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = square-root start_ARG italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG is the sound speed, and Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is electron temperature. We use a simulation box with widths Lx×Ly=200ρi×200ρisubscript𝐿𝑥subscript𝐿𝑦200subscript𝜌𝑖200subscript𝜌𝑖L_{x}\times L_{y}=200\rho_{i}\times 200\rho_{i}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT × italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 200 italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT × 200 italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and number of grid points (nkx,nky,nz,nv||,nμ)=(192,64,64,32,9)(n_{k_{x}},n_{k_{y}},n_{z},n_{v_{||}},n_{\mu})=(192,64,64,32,9)( italic_n start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT | | end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) = ( 192 , 64 , 64 , 32 , 9 ), where ρisubscript𝜌𝑖\rho_{i}italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the ion gyroradius, (x,y,z,v||,μ)(x,y,z,v_{||},\mu)( italic_x , italic_y , italic_z , italic_v start_POSTSUBSCRIPT | | end_POSTSUBSCRIPT , italic_μ ) are the radial, binormal, parallel, parallel velocity, and magnetic moment directions, respectively, kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and kysubscript𝑘𝑦k_{y}italic_k start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT are the wavevectors in x𝑥xitalic_x and y𝑦yitalic_y directions. The heat flux and toroidal angular momentum flux, Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Π^isubscript^Π𝑖\hat{\Pi}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, are respectively normalized to gyroBohm units csniTi(ρi/R0)2subscript𝑐𝑠subscript𝑛𝑖subscript𝑇𝑖superscriptsubscript𝜌𝑖subscript𝑅02c_{s}n_{i}T_{i}(\rho_{i}/R_{0})^{2}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and cs2miniR0(ρi/R0)2subscriptsuperscript𝑐2𝑠subscript𝑚𝑖subscript𝑛𝑖subscript𝑅0superscriptsubscript𝜌𝑖subscript𝑅02c^{2}_{s}m_{i}n_{i}R_{0}(\rho_{i}/R_{0})^{2}italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where Tisubscript𝑇𝑖T_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the ion temperature.

Figure 1 (a) shows the constant Prandtl number manifold Pri=0.5subscriptPr𝑖0.5\text{Pr}_{i}=0.5Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.5 that we construct by linearly interpolating the PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT data from our 3D parameter scan. In simulations with up-down symmetric geometries, the Prandtl number is calculated by

Pri=Π^iQ^iR0LTiϵqωcsR0.subscriptPr𝑖subscript^Π𝑖subscript^𝑄𝑖subscript𝑅0subscript𝐿𝑇𝑖italic-ϵ𝑞subscript𝜔perpendicular-tosubscript𝑐𝑠subscript𝑅0\text{Pr}_{i}=\frac{\hat{\Pi}_{i}}{\hat{Q}_{i}}\frac{R_{0}}{L_{Ti}}\frac{% \epsilon}{q\,\omega_{\perp}}\frac{c_{s}}{R_{0}}.Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = divide start_ARG over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG divide start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT end_ARG divide start_ARG italic_ϵ end_ARG start_ARG italic_q italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG divide start_ARG italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG . (3)

Below the manifold, corresponding to lower values of q𝑞qitalic_q at fixed s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG and R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT, PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is smaller than 0.50.50.50.5. We choose to define this region as the LMD regime. Only the nonlinearly unstable region of the parameter space is presented in the figure. We see that s^1similar-to-or-equals^𝑠1\hat{s}\simeq 1over^ start_ARG italic_s end_ARG ≃ 1 and low R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT favor a low PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, reflected by a larger range in q𝑞qitalic_q with low PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The heat flux shown by the color of the manifold changes primarily with R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT, indicating that q𝑞qitalic_q and s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG do not significantly affect Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in the LMD regime. To show the parameter dependence of the Prandtl number more clearly, several contours of PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are shown in Fig. 1 (b)-(d) by holding either q𝑞qitalic_q, s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG or R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT constant. It is clear that PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT has a non-trivial dependence on these parameters. Figure 1 (c) and (d) show that PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT depends strongly on q𝑞qitalic_q, and Fig. 1 (b) shows that at low q𝑞qitalic_q, PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT becomes small over a large range in s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG and R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT. This feature, together with a small aspect ratio 1/ϵ=2.781italic-ϵ2.781/\epsilon=2.781 / italic_ϵ = 2.78 suggests that a more poloidally angled magnetic field reduces PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. While the dependence of PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT on R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT is weaker, being closer to marginal stability reduces PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The Prandtl number depends strongly on magnetic shear, with a factor of 2similar-toabsent2\sim 2∼ 2 decrease giong from s^=0.1^𝑠0.1\hat{s}=0.1over^ start_ARG italic_s end_ARG = 0.1 to s^1similar-to-or-equals^𝑠1\hat{s}\simeq 1over^ start_ARG italic_s end_ARG ≃ 1, as shown in Fig. 1 (d). Though not shown here, only considering the parallel component of Π^isubscript^Π𝑖\hat{\Pi}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT results in an overestimate of PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [47]. Adding kinetic electrons or changing to Trapped Electron Mode (TEM) turbulence retains the above features in PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, while increasing the aspect ratio 1/ϵ1italic-ϵ1/\epsilon1 / italic_ϵ strongly increases PrisubscriptPr𝑖\text{Pr}_{i}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [47].

Refer to caption
Figure 1: Visualizations of the Prandtl number. (a) shows the surface with Pri=0.5subscriptPr𝑖0.5\text{Pr}_{i}=0.5Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.5, below which we define as the LMD regime. The color map on the surface denotes the heat flux Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The contours in the bottom plane represent lines of constant q𝑞qitalic_q values on the surface. (b)-(d) provide contours of Prandtl number fixing one of our three parameters: (b) q=1.55𝑞1.55q=1.55italic_q = 1.55, (c) s^=1.0^𝑠1.0\hat{s}=1.0over^ start_ARG italic_s end_ARG = 1.0, and (d) R0/LTi=6.96subscript𝑅0subscript𝐿𝑇𝑖6.96R_{0}/L_{Ti}=6.96italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT = 6.96. The red contour lines denote Pri=0.5subscriptPr𝑖0.5\text{Pr}_{i}=0.5Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.5.
Refer to caption
Figure 2: Subplots (a) and (b) show Π^isubscript^Π𝑖\hat{\Pi}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as a function of ωsubscript𝜔perpendicular-to\omega_{\perp}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT obtained with NL simulations for s^=0.1^𝑠0.1\hat{s}=0.1over^ start_ARG italic_s end_ARG = 0.1 (red), s^=0.4^𝑠0.4\hat{s}=0.4over^ start_ARG italic_s end_ARG = 0.4 (green), s^=0.8^𝑠0.8\hat{s}=0.8over^ start_ARG italic_s end_ARG = 0.8 (blue) and s^=1.2^𝑠1.2\hat{s}=1.2over^ start_ARG italic_s end_ARG = 1.2 (black). Subplots (c) and (d) show a comparison of Π^i/Q^isubscript^Π𝑖subscript^𝑄𝑖\hat{\Pi}_{i}/\hat{Q}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as a function of ωsubscript𝜔perpendicular-to\omega_{\perp}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT between the QL model (black), NL simulations with four cases of ω={0,0.1,0.3,ωΠi=0}cs/R0\omega_{\perp}=\bigl{\{}0,-0.1,-0.3,\omega^{{\Pi}_{i}=0}_{\perp}\bigl{\}}c_{s}% /R_{0}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = { 0 , - 0.1 , - 0.3 , italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT } italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (red), and the full NL scans shown in (a) and (b) for s^=0.4^𝑠0.4\hat{s}=0.4over^ start_ARG italic_s end_ARG = 0.4 (green) and s^=0.8^𝑠0.8\hat{s}=0.8over^ start_ARG italic_s end_ARG = 0.8 (blue), respectively. All results are for q=2.05𝑞2.05q=2.05italic_q = 2.05, R0/LTi=10.96subscript𝑅0subscript𝐿𝑇𝑖10.96R_{0}/L_{Ti}=10.96italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT = 10.96. The error bars (calculated by a rolling average on the time traces) are shown for some example cases.

Next, we combine LMD with a source of intrinsic momentum flux to drive strong flow shear. To predict the self-consistent rotation gradient that could be achieved in experiments, we perform NL gyrokinetic simulations with adiabatic electrons for tilted elliptical magnetic flux surfaces with ϵ=0.36italic-ϵ0.36\epsilon=0.36italic_ϵ = 0.36, elongation κ=1.5𝜅1.5\kappa=1.5italic_κ = 1.5, and elongation tilt angle θκ=π/8subscript𝜃𝜅𝜋8\theta_{\kappa}=\pi/8italic_θ start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT = italic_π / 8. This tilt angle was identified to generate the strongest intrinsic momentum flux in a CBC-like equilibrium [42].

Without external momentum sources, steady state operation of a tokamak requires Π^i=0subscript^Π𝑖0\hat{\Pi}_{i}=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0. To enforce this for a given equilibrium in local gyrokinetic simulations, we need to scan over the flow shear ωsubscript𝜔perpendicular-to\omega_{\perp}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT to find the value at which Π^i=0subscript^Π𝑖0\hat{\Pi}_{i}=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0. This is illustrated in Fig. 2 (a), where we see that as we increase the magnitude of the (negatively valued) flow shear, Π^isubscript^Π𝑖\hat{\Pi}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT gradually drops to zero and then changes sign. The flow shear for which Π^i=0subscript^Π𝑖0\hat{\Pi}_{i}=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 is the expected steady-state value in experiment. We see that a higher s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG results in a higher magnitude of the steady-state flow shear |ω|subscript𝜔perpendicular-to|\omega_{\perp}|| italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT |. This is both because higher s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG lowers the Prandtl number (reflected in Fig. 1 (d)) and because higher s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG increases the intrinsic momentum flux (reflected by values of Π^isubscript^Π𝑖\hat{\Pi}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT at ω=0subscript𝜔perpendicular-to0\omega_{\perp}=0italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0 in Fig. 2 (a)) [40, 42]. As expected, the heat flux is reduced by a larger |ω|subscript𝜔perpendicular-to|\omega_{\perp}|| italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | [54] as shown in Fig. 2 (b).

Making a 4D NL scan of (q,s^,R0/LTi,ω)𝑞^𝑠subscript𝑅0subscript𝐿𝑇𝑖subscript𝜔perpendicular-to(q,\hat{s},R_{0}/L_{Ti},\omega_{\perp})( italic_q , over^ start_ARG italic_s end_ARG , italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) is computationally intensive. To reduce the numerical costs, we use a QuasiLinear (QL) model developed specifically to estimate Π^i/Q^isubscript^Π𝑖subscript^𝑄𝑖\hat{\Pi}_{i}/\hat{Q}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT [65]. This allows us to scan ωsubscript𝜔perpendicular-to\omega_{\perp}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT and find the steady-state flow shear ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{{\Pi}_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT such that Π^i(ωΠi=0)=0subscript^Π𝑖subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to0\hat{\Pi}_{i}(\omega^{\Pi_{i}=0}_{\perp})=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) = 0. We then compare two NL simulations, one with ω=ωΠi=0subscript𝜔perpendicular-tosubscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega_{\perp}=\omega^{\Pi_{i}=0}_{\perp}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT and one with ω=0subscript𝜔perpendicular-to0\omega_{\perp}=0italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0, to determine the heat flux reduction by intrinsic rotation. A benchmark of the flux ratio Π^i/Q^isubscript^Π𝑖subscript^𝑄𝑖\hat{\Pi}_{i}/\hat{Q}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT between NL simulations and our QL model is shown in Fig. 2 (c) and (d). We see that our QL model predicts ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{{\Pi}_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT well. Note that we use a linear interpolation between neighboring data points. To estimate the Prandtl number for up-down asymmetric cases, we perform two NL simulations at ω={0.1,0.3}cs/R0\omega_{\perp}=\bigl{\{}-0.1,-0.3\bigl{\}}c_{s}/R_{0}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = { - 0.1 , - 0.3 } italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. As explained in Ref. [47], we adjust Eq. (3) to be

Pr¯i=Δ(Π^iQ^i)R0LTiϵqΔωcsR0,subscript¯Pr𝑖Δsubscript^Π𝑖subscript^𝑄𝑖subscript𝑅0subscript𝐿𝑇𝑖italic-ϵ𝑞Δsubscript𝜔perpendicular-tosubscript𝑐𝑠subscript𝑅0\overline{\text{Pr}}_{i}=\Delta\left(\frac{\hat{\Pi}_{i}}{\hat{Q}_{i}}\right)% \frac{R_{0}}{L_{Ti}}\frac{\epsilon}{q\Delta\omega_{\perp}}\frac{c_{s}}{R_{0}},over¯ start_ARG Pr end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = roman_Δ ( divide start_ARG over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) divide start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT end_ARG divide start_ARG italic_ϵ end_ARG start_ARG italic_q roman_Δ italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG divide start_ARG italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , (4)

so that the Prandtl number is proportional to the slope in Fig. 2 (c) and (d), where ΔΔ\Deltaroman_Δ indicates the difference of quantities between their values at ω=0.1cs/R0subscript𝜔perpendicular-to0.1subscript𝑐𝑠subscript𝑅0\omega_{\perp}=-0.1c_{s}/R_{0}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - 0.1 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and 0.3cs/R00.3subscript𝑐𝑠subscript𝑅0-0.3c_{s}/R_{0}- 0.3 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Hence, four NL simulations are performed for each (q,s^,R0/LTi)𝑞^𝑠subscript𝑅0subscript𝐿𝑇𝑖(q,\hat{s},R_{0}/L_{Ti})( italic_q , over^ start_ARG italic_s end_ARG , italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT ), i.e. ω={0,0.1,0.3,ωΠi=0}cs/R0\omega_{\perp}=\bigl{\{}0,-0.1,-0.3,\omega^{{\Pi}_{i}=0}_{\perp}\bigl{\}}c_{s}% /R_{0}italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = { 0 , - 0.1 , - 0.3 , italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT } italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (shown in Fig. 2 (c) and (d) by the red lines). Note that the QL model is only used to obtain ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{{\Pi}_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT.

Figure 3 summarizes the effectiveness of up-down asymmetry combined with the LMD regime to create flow shear that stabilizes turbulence. When going from subplot (a), (b), (c) to (d), one gradually moves into the LMD regime. As we can see, the intrinsic flow shear strengthens, together with the decrease in Prandtl number, enabling a reduction in the heat flux by the intrinsic flow shear. Outside the LMD regime (Pr¯i0.5greater-than-or-equivalent-tosubscript¯Pr𝑖0.5\overline{\text{Pr}}_{i}\gtrsim 0.5over¯ start_ARG Pr end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≳ 0.5), illustrated by Figs. 3 (a) and (b), only weak flow shear (|ω|0.2cs/R0less-than-or-similar-tosubscript𝜔perpendicular-to0.2subscript𝑐𝑠subscript𝑅0|\omega_{\perp}|\lesssim 0.2c_{s}/R_{0}| italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | ≲ 0.2 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) is produced and the heat flux is barely affected. In the LMD regime (Pr¯i0.5less-than-or-similar-tosubscript¯Pr𝑖0.5\overline{\text{Pr}}_{i}\lesssim 0.5over¯ start_ARG Pr end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≲ 0.5), shown in Fig. 3 (c) and (d), the Prandtl number is low, enabling a strong intrinsic flow shear (|ω|0.3cs/R0greater-than-or-equivalent-tosubscript𝜔perpendicular-to0.3subscript𝑐𝑠subscript𝑅0|\omega_{\perp}|\gtrsim 0.3c_{s}/R_{0}| italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | ≳ 0.3 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) that significantly reduces the heat flux and can increase the critical gradient by up to 25%percent2525\%25 %. Therefore, a combination of LMD with up-down asymmetry can create strong stabilizing intrinsic flow shear in the tokamak core.

Refer to caption
Figure 3: The heat flux Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for the self-consistent value of flow shear ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{\Pi_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT (ensuring Π^i=0subscript^Π𝑖0\hat{\Pi}_{i}=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0) driven by up-down asymmetric flux surfaces (blue), compared to the same cases without any flow shear ω=0subscript𝜔perpendicular-to0\omega_{\perp}=0italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0 (red) as a function of R0/LTisubscript𝑅0subscript𝐿𝑇𝑖R_{0}/L_{Ti}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT for (a) q=2.05,s^=0.1formulae-sequence𝑞2.05^𝑠0.1q=2.05,\hat{s}=0.1italic_q = 2.05 , over^ start_ARG italic_s end_ARG = 0.1, (b) q=2.05,s^=0.4formulae-sequence𝑞2.05^𝑠0.4q=2.05,\hat{s}=0.4italic_q = 2.05 , over^ start_ARG italic_s end_ARG = 0.4, (c) q=2.05,s^=0.8formulae-sequence𝑞2.05^𝑠0.8q=2.05,\hat{s}=0.8italic_q = 2.05 , over^ start_ARG italic_s end_ARG = 0.8, and (d) q=1.05,s^=0.8formulae-sequence𝑞1.05^𝑠0.8q=1.05,\hat{s}=0.8italic_q = 1.05 , over^ start_ARG italic_s end_ARG = 0.8. The Prandtl numbers (black) are also shown for each equilibrium. The flow shear values ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{\Pi_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT (unit cs/R0subscript𝑐𝑠subscript𝑅0c_{s}/R_{0}italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) required to achieve Π^i=0subscript^Π𝑖0\hat{\Pi}_{i}=0over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 are indicated by the numbers neighboring each blue data point.
Refer to caption
Figure 4: Subplots (a) and (b) show the radial profiles of ion temperature and safety factor measured for MAST shot 24600 at t=0.28s𝑡0.28𝑠t=0.28sitalic_t = 0.28 italic_s. The profiles are obtained from TRANSP interpretative analysis, using the data processing code Pyrokinetics [67]. The three vertical dashed lines indicate the three considered radial locations ρpol={0.71,0.77,0.84}subscript𝜌𝑝𝑜𝑙0.710.770.84\rho_{pol}=\left\{0.71,0.77,0.84\right\}italic_ρ start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT = { 0.71 , 0.77 , 0.84 }. The inset in (b) shows the MAST experimental flux surface shape at these three radial locations as well as the artificially tilted elliptical geometry. (c) shows the experimentally measured toroidal flow shear |ω|subscript𝜔perpendicular-to|\omega_{\perp}|| italic_ω start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | produced by NBI (blue line) and the GENE predictions for intrinsic flow shear (stars) that could be created intrinsically if MAST could create the tilted elliptical geometries. (d) shows the heat flux with flow shear using the tilted elliptical geometry, where the vertical dashed lines indicate the predicted steady-state intrinsic flow shear.

To check the experimental feasibility of this strategy, we take an experimental equilibrium from the Mega Ampere Spherical Tokamak (MAST). We consider shot 24600 at time t=0.28s𝑡0.28𝑠t=0.28sitalic_t = 0.28 italic_s (with an on-axis NBI injection), which has the ion temperature profile shown in Fig. 4 (a) and the flux surfaces and q𝑞qitalic_q profile shown in (b). The experimentally measured kinetic profiles are averaged over 0.27st0.29s0.27𝑠𝑡0.29𝑠0.27s\leq t\leq 0.29s0.27 italic_s ≤ italic_t ≤ 0.29 italic_s to reduce the numerical uncertainties. We use ρpol=ψn1/2subscript𝜌𝑝𝑜𝑙subscriptsuperscript𝜓12𝑛\rho_{pol}=\psi^{1/2}_{n}italic_ρ start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT = italic_ψ start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT as the radial coordinate, where ψnsubscript𝜓𝑛\psi_{n}italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the normalized poloidal flux. As we can see, the q𝑞qitalic_q profile is low over a wide radial range, which, together with the tight aspect ratio of MAST, facilitates LMD. No significant MHD instabilities are observed at around t=0.28s𝑡0.28𝑠t=0.28sitalic_t = 0.28 italic_s [68] and the system is in a quasi-steady state. We perform gyrokinetic simulations at three radial locations ρpol={0.71,0.77,0.84}subscript𝜌𝑝𝑜𝑙0.710.770.84\rho_{pol}=\left\{0.71,0.77,0.84\right\}italic_ρ start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT = { 0.71 , 0.77 , 0.84 }, which are dominated by ITG driven turbulence. We tested other radial locations, but they are dominated by either Micro-Tearing Modes (MTM) or Electron Temperature Gradient (ETG) modes, which are computationally much more challenging for NL gyrokinetic simulations. Kinetic electrons are included in these simulations. Our simulations indicate that these three radial locations are close to or in the LMD regime, with Pri={0.58,0.40,0.41}subscriptPr𝑖0.580.400.41\text{Pr}_{i}=\left\{0.58,0.40,0.41\right\}Pr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { 0.58 , 0.40 , 0.41 }.

To verify that one can expect the LMD regime to reduce turbulence, we artificially modified the MAST flux surfaces to be tilted ellipses as shown in Fig. 4 (b) without considering the external torque from NBI, and performed NL flow shear scans to find ωΠi=0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to\omega^{\Pi_{i}=0}_{\perp}italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. These flow shear values are shown in Fig. 4 (c), alongside the experimental flow shear profile. This illustrates that experimentally significant flow shear levels driven by external NBI sources could actually be driven intrinsically if MAST could achieve tilted elliptical flux surfaces. Figure 4 (d) shows that this intrinsic flow shear almost entirely quenches the ITG turbulence and thus significantly reduces the heat flux.

While it is difficult to create strongly up-down asymmetric flux surfaces in MAST, a new spherical tokamak under construction, SMART [69, 70, 71, 72, 73], has stronger shaping capabilities. Figure 5 (a) shows a preliminary up-down asymmetric geometry that can potentially be achieved in SMART with the addition of its advanced coils. We consider one radial location at ρpol=0.79subscript𝜌𝑝𝑜𝑙0.79\rho_{pol}=0.79italic_ρ start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT = 0.79, dominated by ITG turbulence, and perform a similar NL flow shear scan as we did for MAST. The important parameters are R0/LTi=6.32subscript𝑅0subscript𝐿𝑇𝑖6.32R_{0}/L_{Ti}=6.32italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT italic_T italic_i end_POSTSUBSCRIPT = 6.32, q=1.34𝑞1.34q=1.34italic_q = 1.34, s^=1.25^𝑠1.25\hat{s}=1.25over^ start_ARG italic_s end_ARG = 1.25, ϵ=0.39italic-ϵ0.39\epsilon=0.39italic_ϵ = 0.39. A strong intrinsic flow shear of |ωΠi=0|=0.23cs/R0subscriptsuperscript𝜔subscriptΠ𝑖0perpendicular-to0.23subscript𝑐𝑠subscript𝑅0|\omega^{\Pi_{i}=0}_{\perp}|=0.23c_{s}/R_{0}| italic_ω start_POSTSUPERSCRIPT roman_Π start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT | = 0.23 italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, together with one order of magnitude heat flux reduction (shown in Fig. 5 (b) and (c), respectively) are observed, further demonstrating the experimental significance of our method.

Refer to caption
Figure 5: Subplot (a) shows an up-down asymmetric equilibrium that can be achieved by the SMART tokamak with its advanced coil set. Subplots (b) and (c) show the ratio Π^i/Q^isubscript^Π𝑖subscript^𝑄𝑖\hat{\Pi}_{i}/\hat{Q}_{i}over^ start_ARG roman_Π end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Q^isubscript^𝑄𝑖\hat{Q}_{i}over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as functions of flow shear from NL GENE simulations at ρpol=0.79subscript𝜌𝑝𝑜𝑙0.79\rho_{pol}=0.79italic_ρ start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT = 0.79.

Conclusions–In this letter, we propose and analyze a new method to drive strong intrinsic rotation shear in tokamaks: combining the LMD regime with up-down asymmetry. Using a large number of NL and QL gyrokinetic simulations, we self-consistently calculate the flow shear and show that it can significantly stabilize ITG turbulence. The experimental feasibility of this method is addressed by considering a MAST experimental equilibrium and a planned SMART equilibrium. Moreover, our results actually present a lower bound for the impact of the intrinsic flow shear as we neglected the pinch term in our analysis [47] and any external momentum drive. Based on our preliminary study, SMART should be able to test the main predictions in this Letter. This novel approach for generating strong flow shear does not rely on external momentum injection and therefore could be an attractive way of creating strong rotation in future large spherical tokamaks like STEP [74].

Acknowledgement–The simulations in this work are performed on CSCS Daint, Cineca Marconi and CSCS Lumi clusters, with a total computational cost of about 3×1073superscript1073\times 10^{7}3 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT CPU hours. The authors thank Prof. Ben McMillan, Dr. Antoine Hoffmann, Dr. Francis Casson, Arnas Volcokas and Alessandro Balestri for the fruitful discussions. This work has been carried out within the framework of the EUROfusion Consortium, partially funded by the European Union via the Euratom Research and Training Programme (Grant Agreement No. 101052200 - EUROfusion). The Swiss contribution to this work has been 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. This work was supported in part by the Swiss National Science Foundation and by the EPSRC Energy Programme (Grant Number EP/W006839/1).

References

  • Biglari et al. [1990] H. Biglari, P. H. Diamond, and P. W. Terry, Physics of Fluids B: Plasma Physics 2, 1 (1990).
  • Stambaugh et al. [1990] R. D. Stambaugh, S. M. Wolfe, R. J. Hawryluk, J. H. Harris, H. Biglari, S. C. Prager, R. J. Goldston, R. J. Fonck, T. Ohkawa, B. G. Logan, and E. Oktay, Physics of Fluids B: Plasma Physics 2, 2941 (1990).
  • Eriksson et al. [1997] L.-G. Eriksson, E. Righi, and K.-D. Zastrow, Plasma Physics and Controlled Fusion 39, 27 (1997).
  • Angioni et al. [2011] C. Angioni, R. M. McDermott, F. J. Casson, E. Fable, A. Bottino, R. Dux, R. Fischer, Y. Podoba, T. Pütterich, F. Ryter, and E. Viezzer (ASDEX Upgrade Team), Phys. Rev. Lett. 107, 215003 (2011).
  • Peeters et al. [2005] A. G. Peeters, C. Angioni, and the ASDEX Upgrade Team, Physics of Plasmas 12 (2005), 072515.
  • Schekochihin et al. [2008] A. A. Schekochihin, S. C. Cowley, and W. D. et al., Plasma Physics and Controlled Fusion 50, 124024 (2008).
  • de Vries et al. [2008] P. de Vries, M.-D. Hua, D. McDonald, C. Giroud, M. Janvier, M. Johnson, T. Tala, K.-D. Zastrow, and J. E. Contributors, Nuclear Fusion 48, 065006 (2008).
  • Mantica et al. [2009] P. Mantica, D. Strintzi, T. Tala, C. Giroud, T. Johnson, H. Leggate, E. Lerche, T. Loarer, A. G. Peeters, A. Salmi, S. Sharapov, D. Van Eester, P. C. de Vries, L. Zabeo, and K.-D. Zastrow, Phys. Rev. Lett. 102, 175002 (2009).
  • Ida et al. [1990] K. Ida, S. Hidekuma, Y. Miura, T. Fujita, M. Mori, K. Hoshino, N. Suzuki, and T. Yamauchi (JFT-2M Group), Phys. Rev. Lett. 65, 1364 (1990).
  • Newton et al. [2010] S. Newton, S. Cowley, and N. Loureiro, Plasma Physics and Controlled Fusion 52, 125001 (2010).
  • Schekochihin et al. [2012] A. A. Schekochihin, E. G. Highcock, and S. C. Cowley, Plasma Physics and Controlled Fusion 54, 055011 (2012).
  • Highcock et al. [2011] E. G. Highcock, M. Barnes, F. I. Parra, A. A. Schekochihin, C. M. Roach, and S. C. Cowley, Physics of Plasmas 18, 102304 (2011).
  • Barnes et al. [2011] M. Barnes, F. Parra, E. Highcock, A. Schekochihin, S. Cowley, and C. Roach, Phys. Rev. Lett. 106, 175004 (2011).
  • Highcock et al. [2012] E. G. Highcock, A. A. Schekochihin, and S. C. C. et al., Phys. Rev. Lett. 109, 265001 (2012).
  • Christen et al. [2021] N. Christen, M. Barnes, and F. I. Parra, Journal of Plasma Physics 87, 905870230 (2021).
  • McMillan et al. [2019] B. F. McMillan, J. Ball, and S. Brunner, Plasma Phys. Control. Fusion 61, 055006 (2019).
  • Ball et al. [2019] J. Ball, S. Brunner, and B. F. McMillan, Plasma Phys. Control. Fusion 61, 064004 (2019).
  • Noterdaeme et al. [2003] J.-M. Noterdaeme, E. Righi, V. Chan, J. deGrassie, K. Kirov, M. Mantsinen, M. Nave, D. Testa, K.-D. Zastrow, R. Budny, R. Cesario, A. Gondhalekar, N. Hawkes, T. Hellsten, P. Lamalle, F. Meo, F. Nguyen, and E.-J.-E. contributors, Nuclear Fusion 43, 274 (2003).
  • de Vries et al. [2006] P.��C. de Vries, K. M. Rantamäki, C. Giroud, E. Asp, G. Corrigan, A. Eriksson, M. de Greef, I. Jenkins, H. C. M. Knoops, P. Mantica, H. Nordman, P. Strand, T. Tala, J. Weiland, K.-D. Zastrow, and J. E. Contributors, Plasma Physics and Controlled Fusion 48, 1693 (2006).
  • Garofalo et al. [2002] A. M. Garofalo, E. J. Strait, L. C. Johnson, R. J. La Haye, E. A. Lazarus, G. A. Navratil, M. Okabayashi, J. T. Scoville, T. S. Taylor, and A. D. Turnbull, Phys. Rev. Lett. 89, 235001 (2002).
  • Aiba et al. [2009] N. Aiba, S. Tokuda, M. Furukawa, N. Oyama, and T. Ozeki, Nuclear Fusion 49, 065015 (2009).
  • Aiba et al. [2011] N. Aiba, M. Furukawa, M. Hirota, N. Oyama, A. Kojima, S. Tokuda, and M. Yagi, Nuclear Fusion 51, 073012 (2011).
  • Chu et al. [1999] M. Chu, L. Chen, L.-J. Zheng, C. Ren, and A. Bondeson, Nuclear Fusion 39, 2107 (1999).
  • Wahlberg and Bondeson [2000] C. Wahlberg and A. Bondeson, Physics of Plasmas 7, 923 (2000).
  • Groebner et al. [1990] R. J. Groebner, K. H. Burrell, and R. P. Seraydarian, Phys. Rev. Lett. 64, 3015 (1990).
  • Suckewer et al. [1981] S. Suckewer, H. Eubank, R. Goldston, J. McEnerney, N. Sauthoff, and H. Towner, Nuclear Fusion 21, 1301 (1981).
  • Goumiri et al. [2016] I. Goumiri, C. Rowley, S. Sabbagh, D. Gates, S. Gerhardt, M. Boyer, R. Andre, E. Kolemen, and K. Taira, Nuclear Fusion 56, 036023 (2016).
  • Hsuan et al. [1996] H. Hsuan, M. Bitter, C. K. Phillips, J. R. Wilson, C. Bush, H. H. Duong, D. Darrow, G. W. Hammett, K. W. Hill, R. P. Majeski, S. Medley, M. Petrov, E. Synakowski, M. Zarnstorff, and S. Zweben, AIP Conference Proceedings 355, 39 (1996).
  • Chang et al. [1999] C. S. Chang, C. K. Phillips, R. White, S. Zweben, P. T. Bonoli, J. E. Rice, M. J. Greenwald, and J. deGrassie, Physics of Plasmas 6, 1969 (1999).
  • Chan et al. [2002] V. S. Chan, S. C. Chiu, and Y. A. Omelchenko, Physics of Plasmas 9, 501 (2002).
  • Li and Wan [2011] J. Li and B. Wan, Nuclear Fusion 51, 094007 (2011).
  • Lyu et al. [2020] B. Lyu, F. D. Wang, J. Chen, R. J. Hu, Y. Y. Li, J. Fu, H. M. Zhang, M. Bitter, K. W. Hill, Y. J. Shi, M. Y. Ye, and B. N. Wan, Physics of Plasmas 27, 022511 (2020).
  • Liu et al. [2004] Y. Liu, A. Bondeson, Y. Gribov, and A. Polevoi, Nuclear Fusion 44, 232 (2004).
  • Parra et al. [2011a] F. I. Parra, M. Barnes, E. G. Highcock, A. A. Schekochihin, and S. C. Cowley, Phys. Rev. Lett. 106, 115004 (2011a).
  • Camenen et al. [2010] Y. Camenen, A. Bortolon, B. P. Duval, L. Federspiel, A. G. Peeters, F. J. Casson, W. A. Hornsby, A. N. Karpushov, F. Piras, O. Sauter, A. P. Snodin, and G. Szepesi, Phys. Rev. Lett. 105, 135003 (2010).
  • Hornsby et al. [2017] W. Hornsby, C. Angioni, E. Fable, P. Manas, R. McDermott, A. Peeters, M. Barnes, F. Parra, and T. A. U. Team, Nuclear Fusion 57, 046008 (2017).
  • Zhu et al. [2024] H. Zhu, T. Stoltzfus-Dueck, R. Hager, S. Ku, and C. S. Chang, Phys. Rev. Lett. 133, 025101 (2024).
  • Parra et al. [2011b] F. Parra, M. Barnes, and A. Peeters, Phys. Plasmas 18, 062501 (2011b).
  • Peeters et al. [2011] A. Peeters, C. Angioni, A. Bortolon, Y. Camenen, F. Casson, B. Duval, L. Fiederspiel, W. Hornsby, Y. Idomura, T. Hein, N. Kluy, P. Mantica, F. Parra, A. Snodin, G. Szepesi, D. Strintzi, T. Tala, G. Tardini, P. de Vries, and J. Weiland, Nuclear Fusion 51, 094027 (2011).
  • Parra and Barnes [2015] F. I. Parra and M. Barnes, Plasma Physics and Controlled Fusion 57, 045002 (2015).
  • Ball et al. [2014] J. Ball, F. I. Parra, M. Barnes, and W. D. et al., Plasma Physics and Controlled Fusion 56, 095014 (2014).
  • Ball et al. [2018] J. Ball, F. I. Parra, M. Landreman, and M. Barnes, Nuclear Fusion 58, 026003 (2018).
  • Peeters et al. [2007] A. G. Peeters, C. Angioni, and D. Strintzi, Phys. Rev. Lett. 98, 265003 (2007).
  • Guttenfelder et al. [2017] W. Guttenfelder, A. Field, I. Lupelli, T. Tala, S. Kaye, Y. Ren, and W. Solomon, Nuclear Fusion 57, 056022 (2017).
  • Zimmermann et al. [2022] C. F. B. Zimmermann, R. M. McDermott, E. Fable, C. Angioni, B. P. Duval, R. Dux, A. Salmi, U. Stroth, T. Tala, G. Tardini, T. Pütterich, the ASDEX Upgrade, and E. M. Teams, Plasma Physics and Controlled Fusion 64, 055020 (2022).
  • Zimmermann et al. [2023] C. Zimmermann, R. McDermott, C. Angioni, B. Duval, R. Dux, E. Fable, A. Salmi, U. Stroth, T. Tala, G. Tardini, T. Pütterich, and the ASDEX Upgrade Team, Nuclear Fusion 63, 126006 (2023).
  • Sun et al. [2024a] H. Sun, J. Ball, S. Brunner, A. Field, B. Patel, A. Balestri, D. Kennedy, C. Roach, D. J. Cruz-Zabala, F. P. D. Pozo, E. Viezzer, and M. G. Munoz, Physics of the low momentum diffusivity regime in tokamaks and its experimental applicability (2024a), arXiv:2408.12331 [physics.plasm-ph] .
  • Tala et al. [2011] T. Tala, A. Salmi, C. Angioni, F. Casson, G. Corrigan, J. Ferreira, C. Giroud, P. Mantica, V. Naulin, A. Peeters, W. Solomon, D. Strintzi, M. Tsalas, T. Versloot, P. de Vries, K.-D. Zastrow, and JET-EFDA contributors, Nuclear Fusion 51, 123002 (2011).
  • Weisen et al. [2012] H. Weisen, Y. Camenen, A. Salmi, T. Versloot, P. de Vries, M. Maslov, T. Tala, M. Beurskens, C. Giroud, and JET-EFDA contributors, Nuclear Fusion 52, 042001 (2012).
  • Diamond et al. [2008] P. H. Diamond, C. J. McDevitt, Ö. D. Gürcan, T. S. Hahm, and V. Naulin, Physics of Plasmas 15 (2008), 012303.
  • Diamond et al. [2009] P. Diamond, C. McDevitt, Ã. Gürcan, T. Hahm, W. X. Wang, E. Yoon, I. Holod, Z. Lin, V. Naulin, and R. Singh, Nuclear Fusion 49, 045002 (2009).
  • Holod and Lin [2008] I. Holod and Z. Lin, Physics of Plasmas 15 (2008), 092302.
  • Camenen et al. [2011] Y. Camenen, Y. Idomura, S. Jolliet, and A. Peeters, Nuclear Fusion 51, 073039 (2011).
  • Casson et al. [2009] F. J. Casson, A. G. Peeters, Y. Camenen, W. A. Hornsby, A. P. Snodin, D. Strintzi, and G. Szepesi, Physics of Plasmas 16, 092303 (2009).
  • McMillan [2015] B. F. McMillan, Physics of Plasmas 22, 020707 (2015).
  • McMillan and Dominski [2019] B. F. McMillan and J. Dominski, Journal of Plasma Physics 85, 175850301 (2019).
  • Abel et al. [2013] I. Abel, G. Plunk, E. Wang, M. Barnes, S. Cowley, W. Dorland, and A. Schekochihin, Rep. Prog. Phys 76, 116201 (2013).
  • McDermott et al. [2011] R. M. McDermott, C. Angioni, R. Dux, A. Gude, T. Pütterich, F. Ryter, G. Tardini, and the ASDEX Upgrade Team, Plasma Physics and Controlled Fusion 53, 035007 (2011).
  • Buchholz et al. [2015] R. Buchholz, S. Grosshauser, W. Guttenfelder, W. A. Hornsby, P. Migliano, A. G. Peeters, and D. Strintzi, Physics of Plasmas 22, 082307 (2015).
  • Hornsby et al. [2018] W. Hornsby, C. Angioni, Z. Lu, E. Fable, I. Erofeev, R. McDermott, A. Medvedeva, A. Lebschy, A. Peeters, and T. A. U. Team, Nuclear Fusion 58, 056008 (2018).
  • Meyer et al. [2009] H. Meyer, R. Akers, F. Alladio, L. Appel, K. Axon, N. B. Ayed, P. Boerner, R. Buttery, P. Carolan, D. Ciric, C. Challis, I. Chapman, G. Coyler, J. Connor, N. Conway, S. Cowley, M. Cox, G. Counsell, G. Cunningham, A. Darke, M. deBock, G. deTemmerman, R. Dendy, J. Dowling, A. Y. Dnestrovskij, Y. Dnestrovskij, B. Dudson, D. Dunai, M. Dunstan, A. Field, A. Foster, L. Garzotti, K. Gibson, M. Gryaznevich, W. Guttenfelder, N. Hawkes, J. Harrison, P. Helander, T. Hender, B. Hnat, M. Hole, D. Howell, M. D. Hua, A. Hubbard, M. Istenic, N. Joiner, D. Keeling, A. Kirk, H. Koslowski, Y. Liang, M. Lilley, S. Lisgo, B. Lloyd, G. Maddison, R. Maingi, A. Mancuso, S. Manhood, R. Martin, G. McArdle, J. McCone, C. Michael, P. Micozzi, T. Morgan, A. Morris, D. Muir, E. Nardon, G. Naylor, M. O’Brien, T. O’Gorman, A. Patel, S. Pinches, J. Preinhaelter, M. Price, E. Rachlew, D. Reiter, C. Roach, V. Rozhansky, S. Saarelma, A. Saveliev, R. Scannell, S. Sharapov, V. Shevchenko, S. Shibaev, H. Smith, G. Staebler, D. Stork, J. Storrs, A. Sykes, S. Tallents, P. Tamain, D. Taylor, D. Temple, N. Thomas-Davies, A. Thornton, A. Thyagaraja, M. Turnyanskiy, J. Urban, M. Valovic, R. Vann, F. Volpe, G. Voss, M. Walsh, S. Warder, R. Watkins, H. Wilson, M. Windridge, M. Wisse, A. Zabolotski, S. Zoletnik, O. Zolotukhin, and the MAST and NBI teams, Nuclear Fusion 49, 104017 (2009).
  • Jenko et al. [2000] F. Jenko, W. Dorland, M. Kotschenreuther, and B. Rogers, Physics of Plasmas 7, 1904 (2000).
  • Goerler et al. [2011] T. Goerler, X. Lapillonne, S. Brunner, T. Dannert, F. Jenko, F. Merz, and D. Told, Journal of Computational Physics 230, 7053 (2011).
  • Miller et al. [1998] R. L. Miller, M. S. Chu, J. M. Greene, Y. R. Lin-Liu, and R. E. Waltz, Physics of Plasmas 5, 973 (1998).
  • Sun et al. [2024b] H. Sun, J. Ball, S. Brunner, and A. Volčokas, Nuclear Fusion 64, 036026 (2024b).
  • Dimits et al. [2000] A. Dimits, G. Bateman, M. Beer, B. Cohen, W. Dorland, G. Hammett, C. Kim, J. Kinsey, M. Kotschenreuther, A. Kritz, et al., Physics of Plasmas 7, 969 (2000).
  • Patel et al. [2024] B. S. Patel, P. Hill, L. Pattinson, M. Giacomin, A. Bokshi, D. Kennedy, H. G. Dudding, J. F. Parisi, T. F. Neiser, A. C. Jayalekshmi, D. Dickinson, and J. R. Ruiz, Journal of Open Source Software 9, 5866 (2024).
  • Field et al. [2011] A. Field, C. Michael, R. Akers, J. Candy, G. Colyer, W. Guttenfelder, Y. c. Ghim, C. Roach, S. Saarelma, and the MAST Team, Nuclear Fusion 51, 063006 (2011).
  • Doyle et al. [2021] S. Doyle, D. Lopez-Aires, A. Mancini, M. Agredano-Torres, J. Garcia-Sanchez, J. Segado-Fernandez, J. Ayllon-Guerola, M. Garcia-Munoz, E. Viezzer, C. Soria-Hoyo, J. Garcia-Lopez, G. Cunningham, P. Buxton, M. Gryaznevich, Y. Hwang, and K. Chung, Fusion Engineering and Design 171, 112706 (2021).
  • Mancini et al. [2021] A. Mancini, J. Ayllon-Guerola, S. Doyle, M. Agredano-Torres, D. Lopez-Aires, J. Toledo-Garrido, E. Viezzer, M. Garcia-Munoz, P. Buxton, K. Chung, J. Garcia-Dominguez, J. Garcia-Lopez, M. Gryaznevich, J. Hidalgo-Salaverri, Y. Hwang, and J. Segado-Fernández, Fusion Engineering and Design 171, 112542 (2021).
  • Agredano-Torres et al. [2021] M. Agredano-Torres, J. Garcia-Sanchez, A. Mancini, S. Doyle, M. Garcia-Munoz, J. Ayllon-Guerola, M. Barragan-Villarejo, E. Viezzer, J. Segado-Fernandez, D. Lopez-Aires, J. Toledo-Garrido, P. Buxton, K. Chung, J. Garcia-Dominguez, L. Garcia-Franquelo, M. Gryaznevich, J. Hidalgo-Salaverri, Y. Hwang, J. Leon-Galvan, and J. Maza-Ortega, Fusion Engineering and Design 168, 112683 (2021).
  • Segado-Fernandez et al. [2023] J. Segado-Fernandez, A. Mancini, J. Garcia-Dominguez, J. Ayllon-Guerola, D. Cruz-Zabala, L. Velarde, M. Garcia-Munoz, E. Viezzer, C. Navarro, M. Agredano-Torres, and P. Vicente-Torres, Fusion Engineering and Design 193, 113832 (2023).
  • Podestá et al. [2024] M. Podestá, D. J. Cruz-Zabala, F. M. Poli, J. Dominguez-Palacios, J. W. Berkery, M. Garcia-Munoz, E. Viezzer, A. Mancini, J. Segado, L. Velarde, and S. M. Kaye, Plasma Physics and Controlled Fusion 66, 045021 (2024).
  • Anand et al. [2023] H. Anand, O. Bardsley, D. Humphreys, M. Lennholm, A. Welander, Z. Xing, J. Barr, M. Walker, J. Mitchell, and H. Meyer, Fusion Engineering and Design 194, 113724 (2023).