[1]\fnmAzarakhsh \surJalalvand [1,2]\fnmEgemen \surKolemen

[1]\orgdivDepartment of Mechanical and Aerospace Engineering, \orgnamePrinceton University, \orgaddress\cityPrinceton,\countryUSA

2]\orgnamePrinceton Plasma Physics Laboratory, \orgaddress\cityPrinceton, \countryUSA

3]\orgdivDepartment of Physics, \orgnameChung-Ang University, \orgaddress\citySeoul, \countrySouth Korea

4]\orgdivApplied Physics and Applied Mathematics, \orgnameColumbia University, \orgaddress\cityNew York, \countryUSA

5]\orgdivDepartment of Nuclear Engineering, \orgnameSeoul National University, \orgaddress\citySeoul, \countryRepublic of Korea

Discovering hidden physics using ML-based multimodal super-resolution measurement and its application to fusion plasmas

[email protected]    \fnmSangKyeun \surKim    \fnmJaemin \surSeo    \fnmQiming \surHu    \fnmMax \surCurie    \fnmPeter \surSteiner    \fnmAndrew Oakleigh \surNelson    \fnmYong-Su \surNa    [email protected] * [ [ [ [
Abstract

A non-linear complex system governed by multi-spatial and multi-temporal physics scales cannot be fully understood with a single diagnostic, as each provides only a partial view and much information is lost during data extraction. Combining multiple diagnostics also results in imperfect projections of the system’s physics. By identifying hidden inter-correlations between diagnostics, we can leverage mutual support to fill in these gaps, but uncovering these inter-correlations analytically is too complex. We introduce a groundbreaking machine learning methodology to address this issue. Our multimodal approach generates super-resolution data encompassing multiple physics phenomena, capturing detailed structural evolution and responses to perturbations previously unobservable. This methodology addresses a critical problem in fusion plasmas: the Edge Localized Mode (ELM), a plasma instability that can severely damage reactor walls. One method to stabilize ELM is using resonant magnetic perturbation to trigger magnetic islands. However, low spatial and temporal resolution of measurements limits the analysis of these magnetic islands due to their small size, rapid dynamics, and complex interactions within the plasma. With super-resolution diagnostics, we can experimentally verify theoretical models of magnetic islands for the first time, providing unprecedented insights into their role in ELM stabilization. This advancement aids in develo** effective ELM suppression strategies for future fusion reactors like ITER and has broader applications, potentially revolutionizing diagnostics in fields such as astronomy, astrophysics, and medical imaging.

keywords:
Fusion reactor, Machine learning, Synthetic diagnostics, Physics-preserving super-resolution

1 Introduction

In complex physical systems, diagnostic measurements are often intricately interconnected through fundamental physical principles. These underlying connections stem from the laws of nature that govern the behavior of matter and energy. For instance, electromagnetic events couple the measured signals, and equations of state relate variables such as pressure, volume, and temperature, providing a framework to infer one quantity from others. Similarly, coupled differential equations in fluid dynamics or plasma physics describe how multiple system parameters evolve interdependently over time. Such relationships are particularly evident in nuclear fusion, the focus field of this work, which is characterized by its intricate interplay of various physical phenomena.

The nuclear fusion technology aimed at producing eco-friendly energy is rapidly advancing through the synergy of academia and industry [1, 2, 3]. The success of nuclear fusion is fundamentally based on maintaining high-temperature, high-pressure hydrogen plasma without becoming unstable. It was recently shown that Artificial Intelligence (AI) can be a helpful tool to achieve that goal [4, 5, 6, 3]. Experimental fusion reactors like DIII-D [7] utilize various diagnostics for effective plasma monitoring necessary for this AI application [8]. For example, the Electron Cyclotron Emission (ECE) diagnostic system measures electron temperature [9], \ceCO2 interferometer (\ceInterferometer) measures electron density and its fluctuations [10], Motional Stark Effect (MSE) measures the magnetic field [11], and Thomson Scattering (TS) measures the electron temperature and density [12]. The different measurements each capture different physical properties, and form a complementary set for extracting as much information from the plasma as possible. Although it is likely that there exists some kind of correlation or coupling between the measurements of different diagnostics (see Section 6.1), our current scientific understanding is still not capable of specifying some of these relationships analytically. Machine Learning (ML) is a powerful tool for identifying hidden relationships in data [13]. Learning the hidden relationships among different diagnostics would be a great asset to enhance their measurements, and it also helps to find a minimal set of diagnostics for a future reactor in which the availability of diagnostics is limited due to the cost and hardware constraints.

One of the most critical issues for fusion reactors is the Edge Localized Mode (ELM), an instability that occurs at the plasma edge under high-confinement conditions. This edge instability delivers transient and intense heat flux outward, which can cause irreversible damage to the wall in a reactor-scale device. Therefore, understanding and controlling this phenomenon is a major challenge that must be resolved [5, 14]. However, the detailed physical mechanism of ELM s and structural response to the external field occurring within milliseconds are still subjects of ongoing debate. High-frequency diagnostics like ECE and \ceInterferometer possess sufficient time resolution to track these fast dynamics, but their limited spatial resolution and measurement conditions pose challenges in clearly observing the structural characteristics of ELM s. On the other hand, TS offers high spatial resolution capable of observing detailed structures, but its temporal resolution is too low to elucidate the exact mechanism of ELM s.

The current remedy to this issue is a specific operational method for TS, known as “bunch mode”, to increase the sampling rate of up to 10 kHztimes10kilohertz10\text{\,}\mathrm{kHz}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG [15, 16]. Despite its high pulse repetition, firing TS in “bunch mode” is limited by the heat capacity of the laser medium and limited measurement repetition. Therefore such an approach is typically reserved for very short periods of time or specific experiments where high-resolution temporal data is crucial [16].

Instead, we hypothesize that a data-driven ML model, so-called Diag2Diag, with multimodal inputs comprising the high-frequency diagnostics can effectively make use of internal correlations in order to estimate TS. This can enhance the temporal resolution of the existing TS diagnostics without upgrading hardwares, so-called Super-Resolution TS (SRTS) diagnostics, which enables deeper physical analysis of plasma behavior.

Various fields have developed ML-based spatial or temporal resolution enhancement techniques, but these mostly involve uni-modal resolution enhancement by learning linear or nonlinear interpolation of the given information [17, 18, 19, 20, 21]. These are applicable only to regularly sampled data and are challenging to generate finer-scale phenomena undetectable at the time resolution of the target sensor (more detail in Section 6.2). Our work goes beyond plausible interpolation; it is a physics-preserving super-resolution to reconstruct events missed by target diagnostics, by learning the correlation between different diagnostic measurements, which is, to our best knowledge, the first attempt of its kind.

Refer to caption
Figure 1: Main methodology. DiagA is essential to capture fast transient events near the edge of plasma. But due to its low temporal resolution and accuracy it fails to track the evolution of such events. Diag2Diag solves this problem by generating synthetic super resolution of DiagA by learning the correlation between DiagA data and other diagnostic measurements with higher resolutions and better accuracy.

Figure 1 summarizes the main methodology for this work. DIII-D utilizes hundreds of diagnostics for monitoring the plasma. These diagnostics measure various characteristics of plasma at different temporal resolutions. A potential ML model can learn the intrinsic correlations among diagnostics data and thus generate one from others. This works for both, time-series and spectrograms, although different variants of Artificial Neural Network (ANN) are used. The design choices and the optimization and training strategies are described in the following sections.

2 ML-based map** between different diagnostics

For develo** an ML-based SRTS diagnostic from other diagnostics, it is essential to verify the existence, strength, and robustness of correlation among them. We therefore approach this in several steps as described subsequently.

The aim of the first step is to show that we can reconstruct the spectrograms of one diagnostic based on another. As was discussed in Section 1, it is very likely that different diagnostics data have intrinsic correlations. Certain plasma instabilities and modes, such as ELM and Alfvén Eigenmode (AE), affect both electron temperature, which is measured by ECE, and density fluctuations measured by \ceInterferometer. We now show that an ANN is able to learn this relationship during AE modes by map** from ECE spectrograms to \ceInterferometer spectrograms as illustrated in Figure 2.

Figures 2(b) and (d) show example spectrograms obtained from the raw signals of ECE and \ceInterferometer, respectively, whose measurement positions and paths can be seen in Figure 2(a). We designed and trained a Convolutional Neural Network (CNN) that takes 40 ECE spectrograms as input and reconstructs 4 target \ceInterferometer spectrograms, as shown in Figure 2(c). The reconstructed synthetic \ceInterferometer spectrograms visually confirm the plausible reconstruction of features such as frequency chir** and harmonics as seen in Figure 2(d).

Refer to caption
Figure 2: Reconstructing \ceInterferometer spectrograms from ECE spectrograms of DIII-D shot 170669 using convolutional neural networks. (a) The configuration of 4 ECE and 40 ECE probes at DIII-D . (b) A tensor of (40×time×frequency)40𝑡𝑖𝑚𝑒𝑓𝑟𝑒𝑞𝑢𝑒𝑛𝑐𝑦(40\times time\times frequency)( 40 × italic_t italic_i italic_m italic_e × italic_f italic_r italic_e italic_q italic_u italic_e italic_n italic_c italic_y ) is supplied to CNN. (c) The configuration of CNN. (d) Visual comparison of measured and reconstructed spectrograms (e) Comparison of the Alfven Eigenmode detector output [22] supplied with the measured and reconstructed spectrograms.

Besides the visual comparison, we are also interested in how much the underlying physical information is preserved using this method. Therefore, we evaluated the preservation of physical information by performing a downstream task, AE instability detection [22], based on the measured and the reconstructed \ceInterferometer spectrograms. In Figure 2(e), we can see that the AE scores obtained from the reconstructed spectrograms (blue) closely match those from the measured spectrograms (red). This demonstrates that the results generated by ML contain sufficient hidden physical information, and thus it is supported that ML can extract the intrinsic correlation among diagnostic data. We can now take a step further and tackle the generation of TS signals from other diagnostics based on raw time-series, not a spectrogram.

3 Multimodal super-resolution diagnostic

In this section, we switch from spectrograms to time-series signals and show that the amplitude of a diagnostics can be reconstructed from other diagnostics, while preserving intrinsic physics. More importantly, we will show that if the input diagnostics are of much higher temporal resolution compared to the target one, such a model can be used to increase the time resolution of the target signals in a much more intelligent way compared to the conventional uni-modal interpolations. As a use case, we target TS, one of the most important diagnostics that measure the density and pressure profile of plasma. However as mentioned earlier, its low temporal resolution is a bottleneck in studying the plasma evolution in the rapidly changing events such as ELM.

We consider a suite of input diagnostics available at DIII-D including \ceInterferometer, ECE, Magnetic probes (Magnetics), Charge Exchange Recombination (CER), and MSE with typical sampling rates of 1.66 MHztimes1.66megahertz1.66\text{\,}\mathrm{MHz}start_ARG 1.66 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG, 500 kHztimes500kilohertz500\text{\,}\mathrm{kHz}start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, 2 MHztimes2megahertz2\text{\,}\mathrm{MHz}start_ARG 2 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG, 200 Hztimes200hertz200\text{\,}\mathrm{Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG, and 4 kHztimes4kilohertz4\text{\,}\mathrm{kHz}start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, respectively. To obtain a dataset suitable for this task, all the included diagnostics are aligned with the TS sampling time steps by matching their most recent measured sample. In this way, we create a dataset with which we train the Diag2Diag ANN for this task. Since the sampling steps of TS are not always uniform in time (See Figure 6), we opted for a memory-less neural network instead of the recurrent neural network commonly used in time-series analysis. However, we included the first and second derivatives of the high-resolution input diagnostics, ECE and \ceInterferometer, to include the temporal evolution information.

Refer to caption
Figure 3: (a) Comparison of the electron density by the measured TS and the synthetic SRTS, for the DIII-D shot 153761 [15] near the edge (Z=0.71m𝑍0.71𝑚Z=0.71mitalic_Z = 0.71 italic_m). Dαsubscript𝐷𝛼D_{\alpha}italic_D start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT with arbitrary units is plotted as an indicator of ELM s. An example of ELM event captured by both diagnostics, and another example only captured by SRTS are highlighted in green and purple, respectively. (b-c) Aggregating the measured TS density and temperature in three locations of plasma near the edge for several ELM cycles of the DIII-D shot 174832. The circle highlights the measures TS for one selected ELM cycle and the solid lines present the SRTS which agreeably match the measures TS. t=0𝑡0t=0italic_t = 0 represents the time when ELM is identified by Dαsubscript𝐷𝛼D_{\alpha}italic_D start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. (d-e) The evolution of SRTS between two consecutive measured TS near one ELM cycle across the plasma location.

Figure 3(a) shows, in blue, synthetic TS signals (or SRTS) reconstructed through the trained Diag2Diag from other high-frequency diagnostics, where the original TS signals are also shown with black dots. We can observe that the synthetic signals closely follow the original signals. Furthermore, while the original signals sometimes fail to capture ELM events (identifiable by spectral emission (Dαsubscript𝐷𝛼D_{\alpha}italic_D start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT)), the synthetic signals accurately capture the events missed between the original signals.

Validation: Investigating ELM cycles in DIII-D

When ELM instability occurs, a large amount of plasma quickly escapes from the boundary within milliseconds, and then the plasma gradually recovers. TS diagnostics can observe the density and temperature structure at this edge region, but are limited in capturing dynamics occurring over milliseconds. Recent research [23] overcame these resolution limits by aggregating the measurements from multiple repeated cycles of the fast activity under almost identical conditions to observe a complete evolution. In this subsection, over 20 highly reproduced cycles of ELM crash and recovery were aggregated from DIII-D shot 174823 to assume a ground truth of a complete evolution of an ELM cycle.

The aggregated density and temperature evolution measured by TS in three locations of plasma near the edge are shown in Figures 3(b-c) with transparent crosses, while measurements from a single cycle are shown as solid dots, with different colors for different measurement locations.

We used the Diag2Diag model to generate synthetic SRTS, shown with solid lines in Figure 3(b-c). The SRTS signal from a single cycle around time 3795 mstimes3795ms3795\text{\,}\mathrm{m}\mathrm{s}start_ARG 3795 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG not only follows the trend of the aggregated multiple TS measurements but also well overlies the TS measurements within that cycle. Figures 3(d-e) show the detailed evolution of plasma density and temperature across the plasma plasma location captured by SRTS in the same ELM cycle at 3795 mstimes3795ms3795\text{\,}\mathrm{m}\mathrm{s}start_ARG 3795 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG which is missed by TS between its two consecutive measurements at 3791 mstimes3791ms3791\text{\,}\mathrm{m}\mathrm{s}start_ARG 3791 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG and 3800 mstimes3800ms3800\text{\,}\mathrm{m}\mathrm{s}start_ARG 3800 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG.

In a more typical tokamak discharge, the plasma state continually changes, and ELM s occur more irregularly as shown in Figure 3(a). In such cases, it is not possible to reconstruct a single ELM cycle by aggregating multiple cycles, and our SRTS method will be highly beneficial.

4 Science discovery: Unveiling diagnostic evidence of RMP mechanism on the plasma boundary

In what follows, we investigate whether the synthetic super-resolution diagnostics can help to verify the hypotheses on the mechanism of plasma response to external field perturbations in fusion plasma physics that have been proposed theoretically or by simulations but have never been visualized with the experimental data due to the lack of diagnostic resolution.

One promising strategy to control ELM s is employing Resonant Magnetic Perturbation (RMP) s [24, 25, 26, 27, 28] generated by external 3D field coils depicted in Figure 4(a). These fields effectively reduce the temperature and density at the confinement pedestal, stabilizing the energy bursts in the edge region. Consequently, ITER will rely on RMP s to maintain a burst-free burning plasma in a tokamak, making it essential for the fusion community to understand and predict the its physics mechanism[29]. However, this issue has been remained a challenge for decades.

The leading theory [30, 31, 32, 33] for explaining the reduced pedestal by RMP s is the formation of magnetic islands by an external 3D field. The magnetic island is a ubiquitous feature in an electromagnetic system with plasmas [34] formed by field reconnection [35, 36]. This structure allows rapid heat (or temperature) and particle (or density) transport between adjacent magnetic field lines, strongly reducing the gradient of local heat and particle distribution or, in other words, profile flattening[37]. The existing theories explain that RMP forms static magnetic islands at the pedestal top and foot region, therefore reducing the pedestal by local profile flattening. As illustrated in Figure 4(a), the theory predicts that RMP s can create magnetic islands near the plasma boundary where pedestal sit in. This model has been successful in quantitatively explaining and predicting the RMP-induced pedestal degradation in real experiments [38, 39], reinforcing magnetic islands as a promising mechanism for RMP-induced pedestal degradation. Nevertheless, measuring the direct evidence of island or local profile flattening still remains a challenge. Extensive experimental efforts have been conducted for this reason and were able to capture the local flattening electron temperature profile [40] near the pedestal top, strongly supporting this theory. However, simultaneously measuring electron temperature and density both at the pedestal top and foot was not possible. In a previous study, rough evidence was observed in TS [33], but it was insufficient to derive a concrete conclusion, mainly due to a large uncertainty of measurement originating from narrow structure (expected from theory, see Fig.4(a)) and oscillatory nature of the plasma boundary. To address the diagnostic uncertainties caused by such system oscillation, one method is to increase the time sampling rate and use time averaging. However, in conventional TS, increasing the time resolution results in a trade-off with measured accuracy, eventually leading to observational limitations.

Refer to caption
Figure 4: Structure of 3D coils and islands by perturbed field (a), and the evidence in the simulation (b-d) and the SRTS diagnostic (e-g) for RMP-induced island mechanism on the plasma boundary in DIII-D shot 157545. Here, the Chebyshev filter is used to derive a statistically reliable time trace of the profile, leveraging the enhanced temporal resolution.

Interestingly, the SRTS has once again illuminated the profile evolution by RMP application, providing the first evidence of ”simultaneous” flattening of temperature and density profile at both top and the foot of the pedestal, strongly supporting the theoretical prediction of magnetic islands effect. This is possible by capturing the statically reliable time trace of profile with Chebyshev time filter, leveraging the enhanced temporal resolution by SRTS.

Figures 4(b-g) illustrate the recovery of temperature and density pedestals within 10 mstimes10millisecond10\text{\,}\mathrm{ms}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG after deactivating RMP, as captured through numerical modeling (Figure 4(b-d)) and SRTS (Figure 4(e-g)). The simulations reveal that the recovery of temperature and density pedestals begins at the top and foot, coinciding with the disappearance of islands. As depicted in Figures 4(d) and (g), the profile gradient recovers at these island locations, enhancing the overall profile. For instance, the measured temperature pedestal shows recovery at both the top and foot through an increasing gradient, displaying qualitative alignment with the simulation results. However, some discrepancies are noted, particularly in the density evolution at the pedestal foot in the SRTS, even though its gradient remains consistent with the modeling. These quantitative differences may stem from the TS’s limited spatial resolution at the boundary and the modeling assumptions such as fixed boundary conditions [41]. Nevertheless, the gradient evolution directly indicates a change in transport due to the RMP-induced islands during this perturbative profile evolution, highlighting that the SRTS successfully reveals the experimental island effect. This provides the ‘first‘ diagnostic evidence of profile flattening at magnetic islands, a key mechanism of RMP-induced pedestal degradation. This successful application of SRTS underscores its potential to reveal new physics beyond the limitations of conventional diagnostic techniques.

5 Conclusion

This study introduces a transformative approach in the field of signal processing and diagnostics through the development of a multimodal neural network, Diag2Diag, which significantly enhances temporal resolution. By leveraging the intrinsic correlations among various diagnostic measurements, we have demonstrated the potential to increase the temporal resolution of the Thomson Scattering diagnostics in fusion plasma from a standard 0.2 kHztimes0.2kilohertz0.2\text{\,}\mathrm{kHz}start_ARG 0.2 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG to an unprecedented 1 MHztimes1megahertz1\text{\,}\mathrm{MHz}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG. This improvement has unlocked new potentials in analyzing fast transient phenomena in plasma, such as the ELM s and the effects of RMP s on pedestal degradation, which were previously blurred or missed in lower resolution data. The ability to inspect these dynamics in greater detail provides new insights into plasma behavior, particularly in conditions where key physics is hidden in the milliseconds. This enhancement is not merely a technical improvement but a crucial enabler for deeper insights into plasma behaviors that are pivotal for advancing fusion reactors. Furthermore, the model’s ability to reconstruct and predict diagnostics from multimodal inputs opens new avenues for cost-effective and less hardware-dependent diagnostic systems. This is particularly beneficial for experimental setups where space and resources are limited, such as in smaller fusion test facilities or in environments where installing multiple high-resolution diagnostics is impractical.

The implications of this work extend well beyond the immediate application to magnetic fusion. The super-resolution capabilities developed here can significantly impact areas such as laser fusion data analysis, accelerator data analysis, and molecular dynamics research. In these fields, similar challenges exist where the time resolution of diagnostics is inadequate to capture fast phenomena effectively. By applying our method, researchers can potentially uncover new physical phenomena or confirm theoretical predictions that were previously unverifiable through experiments due to resolution constraints.

In conclusion, the Diag2Diag model not only addresses a critical need within the fusion community but also sets a precedent for the broader application of AI and machine learning in physical sciences. By pushing the boundaries of what can be observed and measured, this work contributes to the foundational technologies necessary for the realization of fusion energy and advances our understanding of complex physical systems across various scientific domains.

6 Methods

6.1 Underlying Physics in Coupling of Diagnostic Measurements in Plasma System

Diagnostics of electromagnetic systems involve measuring photons or waves to determine the physical quantities of these systems through post-processing. Due to the nature of the systems, these diagnostics are connected. Firstly, the measured signals are interconnected through electromagnetic interactions during system events. Additionally, the physical quantities obtained from signal processing are closely linked through momentum balances. Electromagnetic plasma quantities are governed by a series of momentum equations that encompass variables such as density, flow, temperature, and higher-order terms. Figure 5 illustrates the momentum equations for plasma density (n𝑛nitalic_n) and temperature (T𝑇Titalic_T), where D𝐷Ditalic_D represents particle diffusivity, v𝑣vitalic_v is plasma flow, Snsubscript𝑆𝑛S_{n}italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the particle source, q𝑞qitalic_q denotes heat flux, B𝐵Bitalic_B stands for magnetic field, j𝑗jitalic_j is plasma current, STsubscript𝑆𝑇S_{T}italic_S start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the heat source, and (α𝛼\alphaitalic_α, β𝛽\betaitalic_β) are constant coefficients determined by plasma properties [42]. These equations demonstrate how the measured plasma quantities are interrelated both spatially and temporally. For instance, the line-averaged density obtained from \ceInterferometer diagnostics is geometrically linked to the local density measured by TS by its definition. Simultaneously, temperatures measured by TS and ECE diagnostics, which are positioned differently, are spatially coupled through the gradient term in the momentum equations. Although the TS density and temperature do not directly interact in the equations, they are tightly linked via diffusive fluxes influenced by turbulence, flow, and sources in a self-consistent manner. This intricate physical coupling of various diagnostic measurements allows ML to identify and predict their interconnections effectively.

Refer to caption
Figure 5: Schematic of couplings between diagnostics of the plasma system. They are connected through electromagnetic interactions between signals. Simultaneously, derived quantities from these signals are coupled via geometric definitions, momentum balances, and high-order physics, including turbulence, flow, and source in the system.

6.2 Related works

In recent years different kinds of ANN have been used for upsampling visual data [43, 44, 45, 46, 47] and for radar data [48, 49, 50, 51]. These approaches are typically some kind of non-linear interpolation to add frames between existing video frames. More examples for ML-based upsampling were proposed for medical data [52] and for audio data [53, 54, 55, 56]. Similar to the video upsampling approaches, these approaches can be considered a subcategory of non-linear interpolation as well. In [57], an alternative to interpolation is suggested to estimate missing data in temporal data streams. It is to some extent a multimodal approach, because it fuses different kinds of information. However, the algorithm is limited towards estimating missing data or dealing with irregularly sampled data. Approaches like these work well for enhancing existing sequences, which are quasi-stationary in a way such that consecutive frames or samples do not change very fast.

However, in nuclear fusion, many spurious events like ELM can happen between two TS samples. By interpolating between consecutive TS samples, regardless of linearly or non-linearly, it is likely that we would miss such spurious events. In our work, we thus develop a novel method to generate additional TS samples based on other diagnostics. This is roughly inspired by other multimodal ML approaches, such as [58], where it was proposed to fuse Radar and camera data for an enhanced distance estimation. This is a multimodal approach and thus related to our approach, or [59], where machine learning was used to reveal the control mechanics of an insect wing hinge. This was also a multimodal approach in a way that the ML algorithm received different features recorded from flying insects. However, similar to the other approaches, no attempts to upsampling or estimating missing/in-between data are made. Also [60] presents an artificial neural network method to enhance historical electron temperature data from the decommissioned C-2U fusion device. The model significantly increases the effective sampling rate of TS temperature measurements, utilizing data from multiple diagnostics including the measured TS. The method’s effectiveness is demonstrated through comparisons with ensemble-averaged data for micro-burst instability study. The model’s main drawbacks include limited generalization to only temperature profile study for one specific plasma regime. Notably, the work does not explore the model’s potential for discovering new physics in fusion plasmas.

6.3 Diagnostic details for ELM

In order to let fusion energy be a viable energy source, it must achieve significant fusion gain through continuous fusion reactions. A prominent method to reach this objective is operating a tokamak in high-confinement mode (H-mode), which has a narrow edge transport barrier, also known as the pedestal. This feature significantly boosts plasma confinement within the reactor, enhancing fusion power and efficiency. However, operating in H-mode introduces a steep pressure gradient at the pedestal, leading to substantial operational risks. This gradient drives hazardous edge energy bursts due to a plasma instability known as ELM s. These bursts lead to sudden drops in the energy at the pedestal, causing severe, transient heat fluxes on the reactor walls. This results in damaging material, potential surface erosion and melting, with heat energy reaching approximately 20 MJ m2times20timesmegajoulemeter220\text{\,}\mathrm{MJ}\text{\,}{\mathrm{m}}^{-2}start_ARG 20 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_MJ end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_m end_ARG start_ARG - 2 end_ARG end_ARG end_ARG, which is an unacceptable level for fusion reactors. From ITER, future machines will not allow even the first ELM. Therefore, to advance tokamak designs toward practical application in fusion energy, it is crucial to develop dependable methods to consistently suppress these edge burst events.

A limitation of some diagnostics, such as TS is the low temporal resolution of only 200 Hztimes200hertz200\text{\,}\mathrm{Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG, which does not allow for detecting and tracking fast events like ELM (1 msabsenttimes1millisecond\leq$1\text{\,}\mathrm{ms}$≤ start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG). Figure 6 shows an example of missing ELM in a discharge, due to the low temporal resolution of TS. Nevertheless, it is still important to detect such events reliably, as they can have a strong impact on plasma behavior.

Refer to caption
Figure 6: (a) Configuration of some diagnostics at DIII-D . (b) Example of TS signal for DIII-D discharge 190736, in red, along with the Dαsubscript𝐷𝛼D_{\alpha}italic_D start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT measurements, as an indicator of ELM, and a collection of other diagnostics that is used to increase the resolution of TS. (c) The same data as (b) but zoomed in to show the sampling points of the diagnostics around two examples of ELM event. Due to the low sampling rate, this ELM is not observed by TS. However, thanks to their high temporal resolution, diagnostics including ECE, \ceInterferometer, and Magnetics capture that.

On the other hand, diagnostics like \ceInterferometer, ECE, have much higher temporal resolution with sampling frequencies around MHzmegahertz\mathrm{MHz}roman_MHz, which allows for a much more detailed analysis of the plasma. However, these diagnostics have different characteristics compared to TS. While TS offers detailed insights into both electron density and temperature with high accuracy, it requires complex setups and is usually more resource-intensive. A \ceInterferometer provides a more straightforward approach to measuring electron density, excelling in situations that require rapid response and continuous monitoring. Furthermore, ECE and TS are both pivotal diagnostic tools used in tokamaks for measuring electron temperature, yet they operate on distinctly different principles and offer unique advantages. ECE utilizes the natural microwave emissions from electrons gyrating around magnetic field lines to provide excellent temporal resolution, allowing for the monitoring of rapid plasma changes and instabilities, though its effectiveness can be limited by variations in magnetic field strength. On the other hand, TS involves firing a laser into the plasma and analyzing the scattered light, which provides robust, absolute measurements of both electron temperature and density with less susceptibility to magnetic influences. While ECE excels in continuous data collection and fine temporal analysis, TS offers superior spatial resolution and is less dependent on external conditions, making it invaluable for comprehensive, though typically less frequent, plasma evaluations. If it would be possible to find a correlation between the measurements of those high-resolution diagnostics and TS, this would be useful for develo** new physical analyses.

6.4 Data acquisition

For this experiment, we used discharges from the DIII-D tokamak that include all data from the key diagnostics of interest (CER, \ceInterferometer, ECE, MSE, and TS). We randomly selected 4000400040004000 discharges recorded between the years 2017201720172017 and 2022202220222022 to ensure a diverse and representative dataset. The diagnostic data was collected using the DIII-D MDSplus [61] and PTDATA [62] systems. These diagnostics are generally provided as time-series data streams with varying sampling frequencies, ranging from 200 Hztimes200hertz200\text{\,}\mathrm{Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG for TS up to 1.66 MHztimes1.66megahertz1.66\text{\,}\mathrm{MHz}start_ARG 1.66 end_ARG start_ARG times end_ARG start_ARG roman_MHz end_ARG for \ceInterferometer. The specific pre-processing steps applied to the data for the different experiments conducted in this study are detailed in the following sections.

6.5 Feature extraction

For the spectrogram experiments, we consider the \ceInterferometer and ECE diagnostics. We compute logarithmic magnitude spectrogram from time-series of the raw diagnostics. For each channel (40404040 ECE channels and 4444 \ceInterferometer channels), we therefore used hamming windows of 1 mstimes1millisecond1\text{\,}\mathrm{ms}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG with 0.5 mstimes0.5millisecond0.5\text{\,}\mathrm{ms}start_ARG 0.5 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG overlap. In this way, it was ensured that the different magnitude spectrogrames are aligned in time. The spectrograms were afterwards converted to a logarithmic scale, clipped and rescaled to the range of [0, 1]. Given the noisy nature of the ECE signals and after rescaling the spectrograms to the range of [0,1], the spectrograms are enhanced using a pipeline of image processing filters that includes

  • Quantile Filtering with a threshold of 0.9,

  • Gaussian Blur Filtering on patches of size 31x3,

  • Subtracting average per frequency bin

We used the ECE spectrograms as inputs to our model. Since we treated every ECE channel independently during feature extraction, we obtained one spectrogram per channel, resulting in 40404040 input spectrograms (one per ECE channel). Since our model is designed to estimate the \ceInterferometer spectrograms, it predicts four output spectrogram channels corresponding to the four \ceInterferometer interferometer channels.

For the time-series models, the different diagnostic measurements have varying sampling rates, and some are even non-uniformly sampled in time. Since the aim of time-series data analysis was to increase the resolution of TS, we used its timestamps as a reference and aligned all diagnostic modalities to TS by matching their most recent measured samples in time. This resulted in an amount of 135 233135233135\,233135 233 training, 22 0842208422\,08422 084 validation, and 18 7211872118\,72118 721 test samples.

For \ceInterferometer and ECE, we also included the first and second temporal derivatives. Therefore, we smoothed the signals with a moving average window of 1 mstimes1millisecond1\text{\,}\mathrm{ms}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG (1660166016601660 \ceInterferometer samples and 500500500500 ECE samples), and then computed the first and second temporal derivatives of the smoothed signal also with a window of 1 mstimes1millisecond1\text{\,}\mathrm{ms}start_ARG 1 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG. In this way, we can consider a temporal context of 4 mstimes4millisecond4\text{\,}\mathrm{ms}start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_ms end_ARG.

The diagnostics CER and MSE have a low temporal resolution, i.e., sampling frequencies of 200 Hztimes200hertz200\text{\,}\mathrm{Hz}start_ARG 200 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG and 4 kHztimes4kilohertz4\text{\,}\mathrm{kHz}start_ARG 4 end_ARG start_ARG times end_ARG start_ARG roman_kHz end_ARG, respectively. In this paper, we assume that they evolve only slowly in time. For the upsampling experiments, we thus pad these diagnostics after a measured sample with constant values until the next measured sample arrived.

The diagnostics (CER, \ceInterferometer, ECE, and MSE) together with the derivatives of \ceInterferometer (4444 channels \rightarrow 12121212 dimensions including derivatives) and ECE (42424242 channels \rightarrow 126126126126 dimensions) lead to an input size of 192192192192. From there, we map to TS with 288288288288 dimensions for plasma density and temperature.

6.6 Spectrogram model development

The resulting multi-channel ECE spectrograms were used as the input to a CNN, and the multi-channel \ceInterferometer spectrograms were used as the target outputs. We optimized all important hyper-parameters based on the 11\mathcal{L}1caligraphic_L 1 loss to minimize the difference between the ground truth and the estimated outputs on the validation set.

The optimization process of the model involved several key steps:

  • The model underwent training for up to 500500500500 epochs.

  • We implemented early stop** with a patience threshold of 20202020 epochs, during which we monitored the validation loss for any improvements.

  • The AdamW optimizer [63], known for decoupling weight decay from the learning rate, was utilized to minimize the 11\mathcal{L}1caligraphic_L 1 loss function.

  • We conducted a comprehensive hyper-parameter optimization through a randomized search across 1000100010001000 iterations for all hyper-parameters listed in Table 1.

The exact search space of the hyper-parameters and their optimized values obtained from the randomized search are summarized in Table 1.

Table 1: Optimized hyperparameters for the spectrogram prediction CNN model.
Hyper-parameter Search space Optimized value
Batch size 1 to 8range1818start_ARG 1 end_ARG to start_ARG 8 end_ARG, random integers 2222
Kernel size 3 to 15range315315start_ARG 3 end_ARG to start_ARG 15 end_ARG odd integers 7777
Learning rate 1×105 to 1range1E-511\text{\times}{10}^{-5}1start_ARG start_ARG 1 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 5 end_ARG end_ARG end_ARG to start_ARG 1 end_ARG log uniform 0.482×1030.482E-30.482\text{\times}{10}^{-3}start_ARG 0.482 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 3 end_ARG end_ARG
Final 11\mathcal{L}1caligraphic_L 1 loss 1.2×1031.2E-31.2\text{\times}{10}^{-3}start_ARG 1.2 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 3 end_ARG end_ARG

To reduce the amount of training time, we randomly selected 518518518518 discharges from the entire dataset to conduct the hyperparameter optimization. The model with the best performing hyperparameter setting (achieving an 11\mathcal{L}1caligraphic_L 1 loss of 1.2×1031.2E-31.2\text{\times}{10}^{-3}start_ARG 1.2 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 3 end_ARG end_ARG on the validation set) was then re-trained on all available discharges.

The best-performing model is a CNN that transforms the ECE spectrograms with 40404040 channels subsequently to 32 , 16 and 83216832168start_ARG 32 end_ARG , start_ARG 16 end_ARG and start_ARG 8 end_ARG feature maps and finally to the \ceInterferometer spectrograms with 4444 channels. For each feature map, 2D filter kernels with a size of 7×7777\times 77 × 7 are used. Batch normalization was used separately for each channel, and parametric ReLUReLU\mathrm{ReLU}roman_ReLU activation functions were used after each batch normalization layer. The model had in total 95 8239582395\,82395 823 trainable parameters (i.e., filter kernels for each feature map, batch normalization parameters, and negative slope of the parametric ReLUReLU\mathrm{ReLU}roman_ReLU activation function).

6.7 Time-series model development

For the time-series prediction task, we employed a Multilayer Perceptron (MLP) model. The input data to the MLP comprised the CER, \ceInterferometer, ECE, MSE, and magnetic diagnostics, along with the first and second temporal derivatives of the \ceInterferometer and ECE signals, resulting in a total input size of 236236236236 dimensions. The target output was the TS diagnostic data, which had 80808080 dimensions representing electron temperature and density across various spatial locations. The target data were augmented by factor 2222 by using the upper and lower intervals of each sample as additional targets.

The MLP model was trained for a maximum of 500500500500 epochs, with an early stop** mechanism implemented to halt the training process if the validation loss did not improve for 20202020 consecutive epochs. The AdamW optimizer [63] was employed to minimize the 11\mathcal{L}1caligraphic_L 1 loss function during training.

As for the spectrogram model, a comprehensive hyperparameter optimization was undertaken using a randomized search approach spanning 2000200020002000 iterations. The hyper-parameters jointly optimized included the batch size, hidden layer size, dropout rate, and learning rate.

Table 2 summarizes the optimized hyperparameter values obtained from the randomized search process.

Table 2: Optimized hyperparameters for the time-series MLP model.
Hyper-parameter Search space Optimized value
Batch size 1 to 2048range1204812048start_ARG 1 end_ARG to start_ARG 2048 end_ARG, powers of 2 1024102410241024
Hidden layer size 192 to 2048range19220481922048start_ARG 192 end_ARG to start_ARG 2048 end_ARG integers 952952952952
Dropout 0 to 1range0101start_ARG 0 end_ARG to start_ARG 1 end_ARG uniform 0.0760.0760.0760.076
Learning rate 1×105 to 1range1E-511\text{\times}{10}^{-5}1start_ARG start_ARG 1 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 5 end_ARG end_ARG end_ARG to start_ARG 1 end_ARG log uniform 1.998×1031.998E-31.998\text{\times}{10}^{-3}start_ARG 1.998 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG - 3 end_ARG end_ARG
Final R2superscript𝑅2R^{2}italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT score 0.920.920.920.92

6.8 Uncertainty quantification

To estimate the uncertainty in super-resolution TS, we employed a Bayesian Neural Network (BNN) with the architecture described in Section 6.7. For each TS channel, we calculated the standard deviation of the BNN outputs, particularly focusing on the pedestal area.

Figure 7 illustrates the average standard deviations of the neural network outputs per channel over the validation set of discharges, depicted as red error bars. The TS channels are represented by their relative distance from the core of the plasma (ψn=0subscript𝜓𝑛0\psi_{n}=0italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 0) to the edge (ψn=1subscript𝜓𝑛1\psi_{n}=1italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 1). For comparison, we also show the empirically measured diagnostic errors, which include contributions from background light, dark noise, and pulse error.

Although the sources of diagnostic uncertainty differ from those of model uncertainty, our results indicate that the model uncertainty falls in the range of the diagnostic errors that are generally accepted by physicists. This suggests that proposed model provides a reliable estimate of uncertainty that enhances the confidence in the super-resolution TS measurements.

Refer to caption
Figure 7: Comparison of the SRTS and the measured TS diagnostic uncertainty for both electron density nesubscript𝑛𝑒n_{e}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT and temperature Tesubscript𝑇𝑒T_{e}italic_T start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. The dashed line shows the average measured TS.

6.9 Thomson Scattering in “bunch mode”

Thomson Scattering, as a popular and reliable diagnostic technique, has successfully measured electron temperatures and electron number densities of plasmas for many years. However, conventional TS techniques operate only at tens of hertz. To accurately resolve the fast transient dynamics, the Thomson scattering lasers can be fired in a bunch mode, which enabled temporal resolution of up to 10 µstimes10microsecond10\text{\,}\mathrm{\SIUnitSymbolMicro s}start_ARG 10 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_s end_ARG. This increase in temporal resolution is achieved by using multiple lasers in the same path with pulses interleaved closely in time. Normally, the lasers are phased to produce pulses at fairly regular intervals (exact regularity is not possible with the specific combination of 20 Hztimes20hertz20\text{\,}\mathrm{Hz}start_ARG 20 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG and 50 Hztimes50hertz50\text{\,}\mathrm{Hz}start_ARG 50 end_ARG start_ARG times end_ARG start_ARG roman_Hz end_ARG lasers being used at DIII-D ). In bunch mode, the phase shifts are adjusted so that all lasers fire in rapid succession, followed by a cool down. This bunch mode encompasses between 3 and 7 laser pulses depending on the time in the discharges.

Figure 8(a) presents the comparison of SRTS with the measured TS fired at bunch mode for the DIII-D shot 153761. Figure 8(b-e) are zoomed in of different window times. Figure 8 (f-g) are further zoomed in frames up to one TS fire of the TS lasers. The matches between the measured TS and SRTS confirms the reliability of SRTS. Also we observe the SRTS superiority in capturing ELM s in comparison to the measured TS even in “bunch mode”.

Refer to caption
Figure 8: Comparison of the SRTS and the TS fired in “bunch mode” for measuring the electron density in DIII-D shot 153761 at the pedestal (Z=0.71m).

6.10 Research method

To avoid any bias during model development and evaluation, each of the following steps in this research was conducted independently by separate researchers in a feed-forward manner as presented in Figure 9:

Refer to caption
Figure 9: Research steps
  1. 1.

    The data scientists developed diagnostic dataset for training the neural network aiming for generating synthetic super resolution Thomson Scattering (SRTS). In this phase, the evaluation metric was simply the similarity between the model’s output and the measured TS, whenever the measurement was available.

  2. 2.

    For physics validation we generated the super resolution diagnostic for a known ELMy discharge and asked an ELM-expert physicist to validate the behavior of the super resolution diagnostic.

  3. 3.

    The data scientist then delivered the generated super-resolution diagnostic for the target plasma discharge to an experimental physicist to extract the plasma profile from that.

  4. 4.

    We then asked another physicist with expertise in simulation to obtain the simulation results for the target plasma discharge.

  5. 5.

    In the final phase, we compared the plasma profiles extracted from our generated diagnostics and the simulation results. They matched nicely!

This indicates that our results are not biased based on prior physics knowledge, and we also did not rework our ML model to match our results with the simulation.

\bmhead

Acknowledgements

This material is based upon work supported by the U.S. Department of Energy, Office of Science, Office of Fusion Energy Sciences, using the National Fusion Facility, a DOE Office of Science user facility, under Award DE-FC02-04ER54698. In addition this research was supported by the U.S. Department of Energy, under Awards DE-SC0024527, DE-SC0015480, DE-SC0022270 and DE-SC0022272, as well as the National Research Foundation of Korea (NRF) Award RS-2024-00346024 funded by the Korea government (MSIT).

Disclaimer: This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

Author contributions

A.J. is the main author of the manuscript and contributed to develo** the multimodal model and data science analyses. S.K. contributed to the physics analysis of Diag2Diag for the RMP mechanism on the plasma boundary. J.S contributed to the general physics analysis and writing the manuscript. Q.H. contributed to the physics simulation of the RMP mechanism on the plasma boundary. M.C. and P.S. contributed to the DIII-D data collection and data preprocessing for the multimodal model development and writing the manuscript. A.O.N contributed the physics analysis of Diag2Diag for ELM cycles. Y.S.N and E.K. contributed to the conception of this work, analyses, and writing the manuscript.

Competing interests

The authors declare no competing interests.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request. The source code for the models developed in this work can be found in [64].

References

  • \bibcommenthead
  • Waldrop [2014] Waldrop, M.M.: Plasma physics: The fusion upstarts. Nature 511(7510), 398–400 (2014) https://doi.org/10.1038/511398a
  • [2] Start-up fusion. Nature Physics 17(10), 1069–1069 (2021) https://doi.org/10.1038/s41567-021-01390-1
  • Degrave et al. [2022] Degrave, J., Felici, F., Buchli, J., Neunert, M., Tracey, B., Carpanese, F., Ewalds, T., Hafner, R., Abdolmaleki, A., Casas, D., Donner, C., Fritz, L., Galperti, C., Huber, A., Keeling, J., Tsimpoukelli, M., Kay, J., Merle, A., Moret, J.-M., Noury, S., Pesamosca, F., Pfau, D., Sauter, O., Sommariva, C., Coda, S., Duval, B., Fasoli, A., Kohli, P., Kavukcuoglu, K., Hassabis, D., Riedmiller, M.: Magnetic control of tokamak plasmas through deep reinforcement learning. Nature 602(7897), 414–419 (2022) https://doi.org/10.1038/s41586-021-04301-9
  • Seo et al. [2024] Seo, J., Kim, S., Jalalvand, A., Conlin, R., Rothstein, A., Abbate, J., Erickson, K., Wai, J., Shousha, R., Kolemen, E.: Avoiding fusion plasma tearing instability with deep reinforcement learning. Nature 626(8000), 746–751 (2024) https://doi.org/10.1038/s41586-024-07024-9
  • Kim et al. [2024] Kim, S.K., Shousha, R., Yang, S.M., Hu, Q., Hahn, S.H., Jalalvand, A., Park, J.-K., Logan, N.C., Nelson, A.O., Na, Y.-S., Nazikian, R., Wilcox, R., Hong, R., Rhodes, T., Paz-Soldan, C., Jeon, Y.M., Kim, M.W., Ko, W.H., Lee, J.H., Battey, A., Yu, G., Bortolon, A., Snipes, J., Kolemen, E.: Highest fusion performance without harmful edge energy bursts in tokamak. Nature Communications 15(1), 3990 (2024) https://doi.org/10.1038/s41467-024-48415-w
  • Chouchene et al. [2024] Chouchene, S., Brochard, F., Lemoine, N., Cavalier, J., Desecures, M., Weinzettl, V.: Mutual interactions between plasma filaments in a tokamak evidenced by fast imaging and machine learning. Phys. Rev. E 109, 045201 (2024) https://doi.org/10.1103/PhysRevE.109.045201
  • Ding et al. [2024] Ding, S., Garofalo, A.M., Wang, H.Q., Weisberg, D.B., Li, Z.Y., Jian, X., Eldon, D., Victor, B.S., Marinoni, A., Hu, Q.M., Carvalho, I.S., Odstrčil, T., Wang, L., Hyatt, A.W., Osborne, T.H., Gong, X.Z., Qian, J.P., Huang, J., McClenaghan, J., Holcomb, C.T., Hanson, J.M.: A high-density and high-confinement tokamak plasma regime for fusion energy. Nature (2024) https://doi.org/10.1038/s41586-024-07313-3
  • Boivin et al. [2005] Boivin, R.L., Luxon, J.L., Austin, M.E., Brooks, N.H., Burrell, K.H., Doyle, E.J., Fenstermacher, M.E., Gray, D.S., Groth, M., Hsieh, C.-L., Jayakumar, R.J., Lasnier, C.J., Leonard, A.W., McKee, G.R., Moyer, R.A., Rhodes, T.L., Rost, J.C., Rudakov, D.L., Schaffer, M.J., Strait, E.J., Thomas, D.M., Van Zeeland, M., Watkins, J.G., Watson, G.W., Wong, C.P.C.: Diii-d diagnostic systems. Fusion Science and Technology 48(2), 834–851 (2005) https://doi.org/10.13182/FST05-A1043
  • Haskey et al. [2022] Haskey, S.R., Grierson, B.A., Stagner, L., Chrystal, C., Bortolon, A., Laggner, F.M.: Details of the neutral energy distribution and ionization source using spectrally resolved Balmer-alpha measurements on DIII-D. Review of Scientific Instruments 93(10), 103504 (2022) https://doi.org/10.1063/5.0101854 https://pubs.aip.org/aip/rsi/article-pdf/doi/10.1063/5.0101854/16594766/103504_1_online.pdf
  • Strait [2006] Strait, E.J.: Magnetic diagnostic system of the DIII-D tokamak. Review of Scientific Instruments 77(2), 023502 (2006) https://doi.org/10.1063/1.2166493 https://pubs.aip.org/aip/rsi/article-pdf/doi/10.1063/1.2166493/16053320/023502_1_online.pdf
  • Holcomb et al. [2006] Holcomb, C.T., Makowski, M.A., Jayakumar, R.J., Allen, S.A., Ellis, R.M., Geer, R., Behne, D., Morris, K.L., Seppala, L.G., Moller, J.M.: Motional Stark effect diagnostic expansion on DIII-D for enhanced current and Er profile measurements. Review of Scientific Instruments 77(10), 10–506 (2006) https://doi.org/10.1063/1.2235812 https://pubs.aip.org/aip/rsi/article-pdf/doi/10.1063/1.2235812/11170995/10e506_1_online.pdf
  • Ponce-Marquez et al. [2010] Ponce-Marquez, D.M., Bray, B.D., Deterly, T.M., Liu, C., Eldon, D.: Thomson scattering diagnostic upgrade on DIII-Da). Review of Scientific Instruments 81(10), 10–525 (2010) https://doi.org/10.1063/1.3495759 https://pubs.aip.org/aip/rsi/article-pdf/doi/10.1063/1.3495759/13981451/10d525_1_online.pdf
  • Shousha et al. [2023] Shousha, R., Seo, J., Erickson, K., Xing, Z., Kim, S., Abbate, J., Kolemen, E.: Machine learning-based real-time kinetic profile reconstruction in diii-d. Nuclear Fusion 64(2), 026006 (2023) https://doi.org/10.1088/1741-4326/ad142f
  • Joung et al. [2024] Joung, S., Smith, D.R., McKee, G., Yan, Z., Gill, K., Zimmerman, J., Geiger, B., Coffee, R., O’Shea, F.H., Jalalvand, A., Kolemen, E.: Tokamak edge localized mode onset prediction with deep neural network and pedestal turbulence. Nuclear Fusion 64(6), 066038 (2024) https://doi.org/10.1088/1741-4326/ad43fb
  • Diallo et al. [2015] Diallo, A., Groebner, R.J., Rhodes, T.L., Battaglia, D.J., Smith, D.R., Osborne, T.H., Canik, J.M., Guttenfelder, W., Snyder, P.B.: Correlations between quasi-coherent fluctuations and the pedestal evolution during the inter-edge localized modes phase on DIII-D). Physics of Plasmas 22(5), 056111 (2015) https://doi.org/10.1063/1.4921148 https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.4921148/13785237/056111_1_online.pdf
  • HE et al. [2019] HE, Z., SMITH, C., ZHANG, Z., BIEWER, T.M., JIANG, N., HSU, P.S., ROY, S.: Pulse-burst laser-based 10 khz thomson scattering measurements. Plasma Science and Technology 21(10), 105603 (2019) https://doi.org/10.1088/2058-6272/ab2e30
  • Fasseaux et al. [2024] Fasseaux, H., Loyez, M., Caucheteur, C.: Machine learning unveils surface refractive index dynamics in comb-like plasmonic optical fiber biosensors. Communications Engineering 3(1), 34 (2024) https://doi.org/10.1038/s44172-024-00181-9
  • Ward et al. [2022] Ward, E.N., Hecker, L., Christensen, C.N., Lamb, J.R., Lu, M., Mascheroni, L., Chung, C.W., Wang, A., Rowlands, C.J., Schierle, G.S.K., Kaminski, C.F.: Machine learning assisted interferometric structured illumination microscopy for dynamic biological imaging. Nature Communications 13(1), 7836 (2022) https://doi.org/10.1038/s41467-022-35307-0
  • Balestriero et al. [2022] Balestriero, R., Glotin, H., Baranuik, R.: Interpretable and learnable super-resolution time-frequency representation. In: Bruna, J., Hesthaven, J., Zdeborova, L. (eds.) Proceedings of the 2nd Mathematical and Scientific Machine Learning Conference. Proceedings of Machine Learning Research, vol. 145, pp. 118–152. PMLR, ??? (2022). https://proceedings.mlr.press/v145/balestriero22a.html
  • Fukami et al. [2023] Fukami, K., Fukagata, K., Taira, K.: Super-resolution analysis via machine learning: a survey for fluid flows. Theoretical and Computational Fluid Dynamics 37(4), 421–444 (2023) https://doi.org/10.1007/s00162-023-00663-0
  • Ren et al. [2023] Ren, P., Rao, C., Liu, Y., Ma, Z., Wang, Q., Wang, J.-X., Sun, H.: Physr: Physics-informed deep super-resolution for spatiotemporal data. Journal of Computational Physics 492, 112438 (2023) https://doi.org/10.1016/j.jcp.2023.112438
  • Garcia et al. [2023] Garcia, A.V., Jalalvand, A., Steiner, P., Rothstein, A., Van Zeeland, M., Heidbrink, W.W., Kolemen, E.: Comparison of machine learning systems trained to detect alfvén eigenmodes using the co2 interferometer on diii-d. Nuclear Fusion 63(12), 126039 (2023) https://doi.org/10.1088/1741-4326/acfe8b
  • Nelson et al. [2021] Nelson, A.O., Xing, Z.A., Izacard, O., Laggner, F.M., Kolemen, E.: Interpretative sol modeling throughout multiple elm cycles in diii-d. Nuclear Materials and Energy 26, 100883 (2021) https://doi.org/10.1016/j.nme.2020.100883
  • Liang et al. [2007] Liang, Y., Koslowski, H.R., Thomas, P.R., Nardon, E., Alper, B., Andrew, P., Andrew, Y., Arnoux, G., Baranov, Y., Bécoulet, M., Beurskens, M., Biewer, T., Bigi, M., Crombe, K., De La Luna, E., Vries, P., Fundamenski, W., Gerasimov, S., Giroud, C., Gryaznevich, M.P., Hawkes, N., Hotchin, S., Howell, D., Jachmich, S., Kiptily, V., Moreira, L., Parail, V., Pinches, S.D., Rachlew, E., Zimmermann, O.: Active control of type-i edge-localized modes with n=1𝑛1n=1italic_n = 1 perturbation fields in the jet tokamak. Phys. Rev. Lett. 98, 265004 (2007) https://doi.org/10.1103/PhysRevLett.98.265004
  • Kirk et al. [2010] Kirk, A., Nardon, E., Akers, R., Bécoulet, M., Temmerman, G.D., Dudson, B., Hnat, B., Liu, Y.Q., Martin, R., Tamain, P., Taylor, D., MAST team: Resonant magnetic perturbation experiments on mast using external and internal coils for elm control. Nuclear Fusion 50(3), 034008 (2010) https://doi.org/10.1088/0029-5515/50/3/034008
  • Suttrop et al. [2011] Suttrop, W., Eich, T., Fuchs, J.C., Günter, S., Janzer, A., Herrmann, A., Kallenbach, A., Lang, P.T., Lunt, T., Maraschek, M., McDermott, R.M., Mlynek, A., Pütterich, T., Rott, M., Vierle, T., Wolfrum, E., Yu, Q., Zammuto, I., Zohm, H.: First observation of edge localized modes mitigation with resonant and nonresonant magnetic perturbations in asdex upgrade. Phys. Rev. Lett. 106, 225004 (2011) https://doi.org/10.1103/PhysRevLett.106.225004
  • Sun et al. [2016] Sun, Y., Liang, Y., Liu, Y.Q., Gu, S., Yang, X., Guo, W., Shi, T., Jia, M., Wang, L., Lyu, B., Zhou, C., Liu, A., Zang, Q., Liu, H., Chu, N., Wang, H.H., Zhang, T., Qian, J., Xu, L., He, K., Chen, D., Shen, B., Gong, X., Ji, X., Wang, S., Qi, M., Song, Y., Yuan, Q., Sheng, Z., Gao, G., Fu, P., Wan, B.: Nonlinear transition from mitigation to suppression of the edge localized mode with resonant magnetic perturbations in the east tokamak. Phys. Rev. Lett. 117, 115001 (2016) https://doi.org/10.1103/PhysRevLett.117.115001
  • Park et al. [2018] Park, J.-K., Jeon, Y., In, Y., Ahn, J.-W., Nazikian, R., Park, G., Kim, J., Lee, H., Ko, W., Kim, H.-S., Logan, N.C., Wang, Z., Feibush, E.A., Menard, J.E., Zarnstroff, M.C.: 3D field phase-space control in tokamak plasmas. Nature Physics 14(12), 1223–1228 (2018) https://doi.org/10.1038/s41567-018-0268-8
  • Loarte et al. [2014] Loarte, A., Huijsmans, G., Futatani, S., Baylor, L.R., Evans, T.E., Orlov, D.M., Schmitz, O., Becoulet, M., Cahyna, P., Gribov, Y., Kavin, A., Naik, A.S., Campbell, D.J., Casper, T., Daly, E., Frerichs, H., Kischner, A., Laengner, R., Lisgo, S., Pitts, R.A., Saibene, G., Wingen, A.: Progress on the application of elm control schemes to iter scenarios from the non-active phase to dt operation. Nuclear Fusion 54(3), 033007 (2014) https://doi.org/10.1088/0029-5515/54/3/033007
  • Snyder et al. [2012] Snyder, P.B., Osborne, T.H., Burrell, K.H., Groebner, R.J., Leonard, A.W., Nazikian, R., Orlov, D.M., Schmitz, O., Wade, M.R., Wilson, H.R.: The EPED pedestal model and edge localized mode-suppressed regimes: Studies of quiescent H-mode and development of a model for edge localized mode suppression via resonant magnetic perturbationsa). Physics of Plasmas 19(5), 056115 (2012) https://doi.org/10.1063/1.3699623 https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.3699623/13989482/056115_1_online.pdf
  • Nazikian et al. [2016] Nazikian, R., Kirk, A., Suttrop, W., Cavedon, M., Grierson, B., Evans, T.E.: First observation of elm suppression by magnetic perturbations in asdex upgrade in a shape-matching identity experiment with diii-d. In: 26th IAEA Fusion Energy Conference (FEC 2016). IAEA, ??? (2016). https://hdl.handle.net/11858/00-001M-0000-002C-A0F9-B
  • Orain et al. [2019] Orain, F., Hoelzl, M., Mink, F., Willensdorfer, M., Bécoulet, M., Dunne, M., Günter, S., Huijsmans, G., Lackner, K., Pamela, S., Suttrop, W., Viezzer, E., Team, A.U., Team, E.M.: Non-linear modeling of the threshold between ELM mitigation and ELM suppression by resonant magnetic perturbations in ASDEX upgrade. Physics of Plasmas 26(4), 042503 (2019) https://doi.org/10.1063/1.5091843 https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5091843/14020497/042503_1_online.pdf
  • Hu et al. [2019] Hu, Q.M., Nazikian, R., Grierson, B.A., Logan, N.C., Park, J.-K., Paz-Soldan, C., Yu, Q.: The density dependence of edge-localized-mode suppression and pump-out by resonant magnetic perturbations in the DIII-D tokamak. Physics of Plasmas 26(12), 120702 (2019) https://doi.org/10.1063/1.5134767 https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5134767/13931866/120702_1_online.pdf
  • Furth et al. [1963] Furth, H.P., Killeen, J., Rosenbluth, M.N.: Finite‐Resistivity Instabilities of a Sheet Pinch. The Physics of Fluids 6(4), 459–484 (1963) https://doi.org/10.1063/1.1706761 https://pubs.aip.org/aip/pfl/article-pdf/6/4/459/12485401/459_1_online.pdf
  • Parker [1957] Parker, E.N.: Sweet’s mechanism for merging magnetic fields in conducting fluids. Journal of Geophysical Research (1896-1977) 62(4), 509–520 (1957) https://doi.org/10.1029/JZ062i004p00509 https://agupubs.onlinelibrary.wiley.com/doi/pdf/10.1029/JZ062i004p00509
  • Sweet [1958] Sweet, P.A.: 14. the neutral point theory of solar flares. Symposium - International Astronomical Union 6, 123–134 (1958) https://doi.org/10.1017/S0074180900237704
  • Fitzpatrick [1995] Fitzpatrick, R.: Helical temperature perturbations associated with tearing modes in tokamak plasmas. Physics of Plasmas 2(3), 825–838 (1995) https://doi.org/10.1063/1.871434 https://pubs.aip.org/aip/pop/article-pdf/2/3/825/19326142/825_1_online.pdf
  • Orain et al. [2019] Orain, F., Hoelzl, M., Mink, F., Willensdorfer, M., Bécoulet, M., Dunne, M., Günter, S., Huijsmans, G., Lackner, K., Pamela, S., Suttrop, W., Viezzer, E., Team, A.U., Team, E.M.: Non-linear modeling of the threshold between ELM mitigation and ELM suppression by resonant magnetic perturbations in ASDEX upgrade. Physics of Plasmas 26(4), 042503 (2019) https://doi.org/10.1063/1.5091843 https://pubs.aip.org/aip/pop/article-pdf/doi/10.1063/1.5091843/14020497/042503_1_online.pdf
  • Markl et al. [2023] Markl, M., Ulbl, P., Albert, C.G., Angioni, C., Buchholz, R., Heyn, M.F., Kasilov, S.V., Kernbichler, W., Suttrop, W., Willensdorfer, M., ASDEX Upgrade Team: Kinetic study of the bifurcation of resonant magnetic perturbations for edge localized mode suppression in asdex upgrade. Nuclear Fusion 63(12), 126007 (2023) https://doi.org/10.1088/1741-4326/acf20c
  • Hu et al. [2020] Hu, Q.M., Nazikian, R., Grierson, B.A., Logan, N.C., Orlov, D.M., Paz-Soldan, C., Yu, Q.: Wide operational windows of edge-localized mode suppression by resonant magnetic perturbations in the diii-d tokamak. Phys. Rev. Lett. 125, 045001 (2020) https://doi.org/10.1103/PhysRevLett.125.045001
  • Yu et al. [2020] Yu, Q., Strumberger, E., Igochine, V., Lackner, K., Laqua, H.P., Zanini, M., Braune, H., Hirsch, M., Höfel, U., Marsen, S., Stange, T., Wolf, R.C., Günter, S., Wendelstein 7-X Team: Numerical modeling of the electron temperature crashes observed in wendelstein 7-x stellarator experiments. Nuclear Fusion 60(7), 076024 (2020) https://doi.org/10.1088/1741-4326/ab9258
  • Hutchinson et al. [2001] Hutchinson, I., Freidberg, J., et al.: Introduction to plasma physics. Electronic book available from: http://silas. psfc. mit. edu/introplasma/index. html (2001)
  • Wang et al. [2024] Wang, Z., Lin, S., Teng-Levy, M., Chu, P., Wolfe, B.T., Wong, C.-S., Campbell, C.S., Yue, X., Zhang, L., Aberle, D., Alvarez, M.A., Broughton, D., Chen, R.T., Cheng, B., Chu, F., Fossum, E.R., Foster, M.A., Huang, C., Kilic, V., Krushelnick, K., Li, W., Loomis, E., au2, T.S.J., Sjue, S.K., Tomkins, C., Yarotski, D.A., Zhu, R.: Physics-informed Meta-instrument for eXperiments (PiMiX) with applications to fusion energy (2024)
  • Jiang et al. [2018] Jiang, H., Sun, D., Jampani, V., Yang, M.-H., Learned-Miller, E., Kautz, J.: Super slomo: High quality estimation of multiple intermediate frames for video interpolation. In: Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR) (2018)
  • Niklaus and Liu [2018] Niklaus, S., Liu, F.: Context-aware synthesis for video frame interpolation. In: Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR) (2018)
  • Bao et al. [2019] Bao, W., Lai, W.-S., Ma, C., Zhang, X., Gao, Z., Yang, M.-H.: Depth-aware video frame interpolation. In: Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR) (2019)
  • Xue et al. [2019] Xue, T., Chen, B., Wu, J., Wei, D., Freeman, W.T.: Video enhancement with task-oriented flow. International Journal of Computer Vision 127(8), 1106–1125 (2019) https://doi.org/10.1007/s11263-018-01144-2
  • Che et al. [2018] Che, Z., Purushotham, S., Li, G., Jiang, B., Liu, Y.: Hierarchical deep generative models for multi-rate multivariate time series. In: Dy, J., Krause, A. (eds.) Proceedings of the 35th International Conference on Machine Learning. Proceedings of Machine Learning Research, vol. 80, pp. 784–793. PMLR, ??? (2018). https://proceedings.mlr.press/v80/che18a.html
  • Recla and Schmitt [2022] Recla, M., Schmitt, M.: Deep-learning-based single-image height reconstruction from very-high-resolution sar intensity data. ISPRS Journal of Photogrammetry and Remote Sensing 183, 496–509 (2022) https://doi.org/10.1016/j.isprsjprs.2021.11.012
  • Recla and Schmitt [2023] Recla, M., Schmitt, M.: Improving deep learning-based height estimation from single sar images by injecting sensor parameters. In: IGARSS 2023 - 2023 IEEE International Geoscience and Remote Sensing Symposium, pp. 1806–1809 (2023). https://doi.org/10.1109/IGARSS52108.2023.10282228
  • Roßberg and Schmitt [2023] Roßberg, T., Schmitt, M.: Temporal upsampling of ndvi time series by rnn-based fusion of sparse optical and dense sar-derived ndvi data. In: IGARSS 2023 - 2023 IEEE International Geoscience and Remote Sensing Symposium, pp. 5990–5993 (2023). https://doi.org/10.1109/IGARSS52108.2023.10282861
  • Bellos et al. [2019] Bellos, D., Basham, M., Pridmore, T., French, A.P.: A convolutional neural network for fast upsampling of undersampled tomograms in x-ray ct time-series using a representative highly sampled tomogram. Journal of Synchrotron Radiation 26(3), 839–853 (2019) https://doi.org/10.1107/S1600577519003448
  • Pascual et al. [2017] Pascual, S., Bonafonte, A., Serrà, J.: Segan: Speech enhancement generative adversarial network. In: Proceedings of the Interspeech 2017, pp. 3642–3646 (2017). https://doi.org/10.21437/Interspeech.2017-1428
  • Donahue et al. [2019] Donahue, C., McAuley, J., Puckette, M.: Adversarial audio synthesis. In: International Conference on Learning Representations (2019). https://openreview.net/forum?id=ByMVTsR5KQ
  • Pandey and Wang [2020] Pandey, A., Wang, D.: Densely connected neural network with dilated convolutions for real-time speech enhancement in the time domain. In: ICASSP 2020 - 2020 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), pp. 6629–6633 (2020). https://doi.org/10.1109/ICASSP40776.2020.9054536
  • Kumar et al. [2019] Kumar, K., Kumar, R., Boissiere, T., Gestin, L., Teoh, W.Z., Sotelo, J., Brébisson, A., Bengio, Y., Courville, A.C.: Melgan: Generative adversarial networks for conditional waveform synthesis. In: Wallach, H., Larochelle, H., Beygelzimer, A., Alché-Buc, F., Fox, E., Garnett, R. (eds.) Advances in Neural Information Processing Systems, vol. 32. Curran Associates, Inc., ??? (2019). https://proceedings.neurips.cc/paper_files/paper/2019/file/6804c9bca0a615bdb9374d00a9fcba59-Paper.pdf
  • Yoon et al. [2019] Yoon, J., Zame, W.R., Schaar, M.: Estimating missing data in temporal data streams using multi-directional recurrent neural networks. IEEE Transactions on Biomedical Engineering 66(5), 1477–1490 (2019) https://doi.org/10.1109/TBME.2018.2874712
  • Li et al. [2024] Li, Y., Wang, Y., Meng, C., Duan, Y., Ji, J., Zhang, Y., Zhang, Y.: Farfusion: A practical roadside radar-camera fusion system for far-range perception. IEEE Robotics and Automation Letters, 1–8 (2024) https://doi.org/10.1109/LRA.2024.3387700
  • Melis et al. [2024] Melis, J.M., Siwanowicz, I., Dickinson, M.H.: Machine learning reveals the control mechanics of an insect wing hinge. Nature 628(8009), 795–803 (2024) https://doi.org/10.1038/s41586-024-07293-4
  • Player et al. [2022] Player, G., Magee, R.M., Tajima, T., Trask, E., Zhai, K., TAE Team: Enhancing historical electron temperature data with an artificial neural network in the c-2u frc. Nuclear Fusion 62(12), 126019 (2022) https://doi.org/10.1088/1741-4326/ac8fa3
  • Fredian and Stillerman [2002] Fredian, T.W., Stillerman, J.A.: Mdsplus. current developments and future directions. Fusion Engineering and Design 60(3), 229–233 (2002) https://doi.org/10.1016/S0920-3796(02)00013-3
  • Schissel and McHarg Jr [2000] Schissel, D., McHarg Jr, B.: Data analysis infrastructure at the diii-d national fusion facility. General Atomics report No. GA-A23474 (2000)
  • Loshchilov and Hutter [2019] Loshchilov, I., Hutter, F.: Decoupled weight decay regularization. In: International Conference on Learning Representations (2019). https://openreview.net/forum?id=Bkg6RiCqY7
  • Jalalvand et al. [2024] Jalalvand, A., Curie, M., Steiner, P., Kolemen, E.: Diag2Diag: Discovering hidden physics using multimodal super-resolution. GitHub (2024). https://github.com/PlasmaControl/diag2diag.git