Cross-beam energy transfer (CBET) results from two-beam energy exchange via seeded stimulated Brillouin scattering, which detrimentally reduces laser-energy absorption for direct-drive inertial confinement fusion. Consequently, ablation pressure and implosion velocity suffer from the decreased absorption, reducing target performance in both symmetric and polar direct drive. Additionally, CBET alters the time-resolved scattered-light spectra and redistributes absorbed and scattered-light–changing shell morphology and low-mode drive symmetry. Mitigating CBET is demonstrated in inertial confinement implosions at the National Ignition Facility by detuning the laser-source wavelengths (±2.3 Å UV) of the interacting beams. In polar direct drive, wavelength detuning was shown to increase the equatorial region velocity experimentally by 16% and to alter the in-flight shell morphology. These experimental observations are consistent with design predictions of radiation–hydrodynamic simulations that indicate a 10% increase in the average ablation pressure. These results indicate that wavelength detuning successfully mitigates CBET. Simulations predict that optimized phase plates and wavelength-detuning CBET mitigation utilizing the three-legged beam layout of the OMEGA Laser System significantly increase absorption and achieve >100-Gbar hot-spot pressures in symmetric direct drive.

In direct-drive inertial confinement fusion (ICF), laser beams irradiate a low-Z shell (plastic or Be) containing a layer of frozen deuterium–tritium (DT) and ablatively drive an implosion. The ultimate goal of ICF is ignition and energy gain; the minimum shell kinetic energy required for ignition (defined as when the energy from DT fusion reactions exceeds the laser energy incident on the target) is given by (Ref. 1)

(1)

where the three implosion parameters—α,vimp, and Pabl [adiabat (the ratio of the fuel pressure to the Fermi-degenerate pressure at peak implosion velocity), implosion velocity, and ablation pressure, respectively]—are determined by the deposition of the laser energy into the coronal plasma of the target, the subsequent heat conduction to the ablation surface, and the resulting equation of state (EOS) of the shell material. Cross-beam energy transfer (CBET)2 has been identified in direct-drive experiments on the OMEGA3 and National Ignition Facility (NIF)4 Laser Systems to reduce absorption and thereby ablation pressure and implosion velocity. The CBET effect reduces the laser absorption by 20% to 30%, the implosion velocity by 10%–15% and the ablation pressure by up to 50% in ignition-scale designs. Reducing the target mass compensates for CBET losses, but the thinner shells become compromised as a result of hydrodynamic instability growth.5 As Eq. (1) indicates, the significant loss caused by CBET precludes ignition when Emin exceeds the laser system's capabilities, motivating CBET mitigation as a critical challenge facing direct drive.

The role of CBET in direct drive was identified in early research6,7 but only recently identified as the leading cause of decreased energy coupling. The critical role of CBET became apparent when attempts were made to match multiple calculated observables (shell morphology, trajectory, scattered-light spectra and power, and shock timing) with experiments.8,9 Historically, the role of CBET was masked by using a flux-limited electron-transport model,45 reproducing either scattered-light power or shell trajectory but ultimately unable to match other observables. Good agreement with all of the multiple experimental observables occurred8,9 with the inclusion of both the CBET and nonlocal electron transport10 models in 1-D LILAC11 and 2-D DRACO12 simulations.

CBET laser–plasma interaction (LPI) results from two-beam energy exchange via stimulated Brillouin scattering (SBS),2 which reduces absorbed light and consequently reduces ablation pressure and implosion velocity. The dominant CBET loss mechanism in direct drive occurs when rays counter-propagate (backscatter mode), thereby increasing scattered light eliding absorption, as illustrated in Fig. 1(a). For the ignition-relevant overlapped peak-integrated beam intensities of ∼4–8 × 1014 W/cm2 for the NIF experiments described in this paper, CBET is calculated to reduce laser absorption by 22%, the average implosion speed by ∼9%, and the average ablation pressure by 35%. While these numbers are lower than expected for ignition-scale designs with similar overlapped intensity, the coronal plasma volume is smaller, corresponding to the appropriately scaled target radius and shot energy available for these experiments that was determined by using the existing far-field spot sizes currently equipped on the NIF beams. These drive-related results are consistent with other ongoing OMEGA-8 and NIF-scale9 experiments.

FIG. 1.

(a) The effect of cross-beam energy transfer (CBET) in polar direct drive (PDD) predominantly affects the equatorial region due to the requisite repointing and because the beam port angles that would have induced significant polar CBET are absent in the PDD configuration; (b) successful CBET mitigation benefits the same region. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

FIG. 1.

(a) The effect of cross-beam energy transfer (CBET) in polar direct drive (PDD) predominantly affects the equatorial region due to the requisite repointing and because the beam port angles that would have induced significant polar CBET are absent in the PDD configuration; (b) successful CBET mitigation benefits the same region. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

Close modal

Mitigation strategies for the deleterious CBET effects invoke combinations of spatial (by reducing the interacting beam size and potentially on-target uniformity, beam overlap, and speckle reduction),13,14 temporal (via altering the timing of interacting beams),15 and wavelength domains.14 Wavelength detuning, the focus of this paper, works by altering the resonance condition between interacting beams.2 Utilizing wavelength separation and continuous, wide bandwidth to defeat laser–plasma interactions (LPI's) dates back many decades with a rich history.16–18 Distributing the laser energy into separated or continuous-wavelength bands decreases the intensity of the individual interactions, shifts resonance attributes, and potentially disrupts the coherent buildup within the plasma with a sufficiently wide bandwidth for each LPI process. The seeded SBS process of CBET does not result from noise like other LPI processes, but rather from another coherent beam, and its suppression requires larger bandwidths since the growth time-scales are shorter. In addition, even with a low gain coefficient, CBET levels can be large because of the large initial amplitude of the “seed.” Wavelength detuning was first examined for indirect drive,19 and subsequently for direct drive, but prematurely dismissed as a viable option.20 

After a more-detailed discussion of the CBET modeling process used in DRACO, the wavelength-detuning mitigation scheme will be presented using experimental evidence from the NIF Laser System. The amount of wavelength detuning currently available on the NIF (±2.3 Å UV) sufficiently demonstrates CBET mitigation for these initial proof-of-principle experiments. Wavelength detuning combined with a unique spatial approach will be discussed in the section on the OMEGA Laser System that applies to any direct-drive ICF laser system. The spatial approach employs a technique applied to a distributed phase plate (DPP) referred to as spot-masking apodization (SMA),21 which enhances CBET mitigation while having low impact on target illumination uniformity, beam overlap, or speckle reduction.

In direct drive, many overlapping beams interact with each other in a complicated tangle of intensity, directions, and wavelengths depending on the beam port, laser, and pulse-shape configurations surrounding the imploding target. In addition, each beam strongly refracts and frequency chirps in the expanding, evolving plasma atmosphere during propagation and then scatters energy spectra in a wide spread of exiting paths. The radiation–hydrodynamics code DRACO is an arbitrary Lagrangian–Eulerian (ALE) simulation code capable of modeling ICF implosions, including the effects of CBET in a 2-D polar-cylindrical mesh. The DRACO CBET package (Adaawam22) is an integral part of the highly scalable parallel 3-D ray-trace package (Mazinisin23),24 which models each beam as a set of adaptively chosen rays to minimize noise; see  Appendix A for more details. The Mazinisin package propagates refracting rays in 3-D within a 2-D plasma atmosphere and accounts for arbitrary multiple beam orientations, arbitrary far-field spot-shapes, independent pulse shapes, beam repointing, target offset, laser absorption, frequency chirp, speckle imprint, and smoothing by spectral dispersion (SSD)26 and for many nonuniform and random beam-to-beam variations such as pulse power imbalances, mistiming, or mispointing, as well as the 3-D CBET interaction effects (using Adaawam) with or without speckled beams. DRACO is able to emulate the effect of 3D modal power induced by laser imbalances by imposing any non-zero m-mode spectral power into a corresponding m = 0,ℓ-mode perturbation. Since the radiation-hydrodynamics code DRACO is only 2-D, the 3-D spatial feedback into the corona cannot be modeled; however, some useful 3-D scattering and deposition effects of refraction and CBET are captured within diagnostic output and can be correlated to experiment. This section concentrates on the modeling of CBET.

An extension to the plane-wave CBET model2 (see Fig. 2) adapts the steady-state and strong-damping limit fluid model to 3-D interacting rays in Adaawam by generalizing the wave vector description, accounting for the probe (kprobe) and pump (kpump) rays' direction and wavelength and their connection to the ion-acoustic wave vector description, viz., kakpumpkprobe. The extension expressed as an exponential factor, dτCBET, describes the amount of energy either gained or lost via an interaction with a background pump field over a small path length, viz.,

(2)

where

(3)
(4)
(5)

and where the spatially dependent plasma variables, Z,Te, Ti, vfluid, and ca are the average atomic number, the electron and ion temperatures, the expanding plasma fluid velocity, and the sound speed, respectively; νa is the dimensionless ion-acoustic damping coefficient [set to 0.2 using the 1-D code LILAC as guidance (Ref. 20)]; the ray propagation–dependent variables ds, ωprobe, and ωpump are the probe ray's path length and the probe and pump angular frequencies, respectively; in addition, the physical constants e, me are the electron charge and mass and the basic light constants c and λ0 are the vacuum speed of light and vacuum central laser wavelength, respectively. The normalized electron number density is defined as nene/necrit, where ne is the spatially dependent electron number density and the critical electron number density is given by necritmeω02/(4πe2),44 where ω02πc/λ0 is the central laser angular frequency. The resonance function P(η), defined in Eq. (3), determines the relative strength of the CBET interaction and gain or loss depending on its sign because all other parameters in Eq. (2) are strictly positive. The value and sign of the resonance function depend directly on the wave-vector–matching condition η given in Eq. (4), where gain occurs when η>0 and loss for η<0. Perfect wave-vector matching occurs when η=±1, which directly corresponds to the maximal amplitude P|η=±1=±1/νa at resonance. Random polarization is accounted for using Eq. (5), but this relationship is not exactly known and is a potential reason for the need of the experimental adjustment parameter fexp that is lumped into that expression. Further research into the polarization dependence is expected to help determine the proper formulation of ςpol and possibly reduce the required value of fexp. A tunable beam with variable polarization is presently being added to the OMEGA laser for fundamental research on CBET.25 In addition, since the factor νa proportionally changes the resonance function as P|η=±1=±1/νa the value used could account for a portion of the experimental adjustment factor fexp. Future investigation of this constant will be pursued to assess the correct value given the potential of wavelength or hydrodynamic dependence of the value of νa.

FIG. 2.

Diagram illustrating the dominant backscatter mode in direct drive that occurs for interacting, counter-propagating beams. The ion-acoustic wave ka results from the interference of the pump and probe beams, kpump and kprobe, respectively. The expanding plasma fluid velocity vfluid resonantly interacts with ka when the phase-matching condition, Eq. (4), is achieved and mediates energy transfer between kpump and kprobe. In the backscatter CBET mode, the pump and probe rays counter-propagate, where one beam is inbound and the other has refracted through the plasma atmosphere and is outbound. The resulting ion-acoustic wave vector, kakpumpkprobe, has a large magnitude (as much as twice a single component) and its direction oppositely oriented to the fluid velocity, vfluid. Consequently, the CBET resonance condition Eq. (4) becomes dominated by the magnitude of kavfluid and the wavelength or color separation of the pump and probe no longer controls the CBET gain or loss. In fact, the outbound probe ray always gains energy because of CBET regardless of its color (red- or blue-shifted) relative to the pump field under normal circumstances.

FIG. 2.

Diagram illustrating the dominant backscatter mode in direct drive that occurs for interacting, counter-propagating beams. The ion-acoustic wave ka results from the interference of the pump and probe beams, kpump and kprobe, respectively. The expanding plasma fluid velocity vfluid resonantly interacts with ka when the phase-matching condition, Eq. (4), is achieved and mediates energy transfer between kpump and kprobe. In the backscatter CBET mode, the pump and probe rays counter-propagate, where one beam is inbound and the other has refracted through the plasma atmosphere and is outbound. The resulting ion-acoustic wave vector, kakpumpkprobe, has a large magnitude (as much as twice a single component) and its direction oppositely oriented to the fluid velocity, vfluid. Consequently, the CBET resonance condition Eq. (4) becomes dominated by the magnitude of kavfluid and the wavelength or color separation of the pump and probe no longer controls the CBET gain or loss. In fact, the outbound probe ray always gains energy because of CBET regardless of its color (red- or blue-shifted) relative to the pump field under normal circumstances.

Close modal

Rays transport energy and not intensity; for this reason, modeling the CBET interaction [as well as inverse bremsstrahlung (IBS)] involves differential changes of the probe ray's energy. The only two LPI effects simulated in DRACO are IBS and CBET. The resultant probe-ray energy, Eprobe1 for any single CBET interaction with a particular pump-intensity–field, wave-vector pair {Ipump,kpump} over a small path length ds is given by Eprobe1=Eprobe0exp(dτCBET), where Eprobe0 is the probe ray's energy prior to the interaction. The probe ray's energy gain (when dτCBET > 0) or loss (when dτCBET < 0) resulting from the single interaction is calculated by ΔEprobe=Eprobe0[exp(dτCBET)1]. Notice that the exponential gain/loss factor dτCBET is dependent only on the pump intensity in this formulation and not on the probe ray's intensity or energy. Therefore, any swelling of the probe-ray intensity that might occur during refraction in small volumes after a turning point cannot affect the outcome of the probe ray's energy gain/loss. However, a significant increase in intensity directly corresponds to a proportionally increased ray-density which results in more individual CBET interactions which may result in increased net CBET gain and loss if the interaction volume is significant.

The CBET model given by Eqs. (2)–(5) includes relevant SBS physics and results in gain/loss for a probe ray interacting with the total background pump field as the probe refracts and traverses the computational mesh. This modeling assumes that the CBET response is steady-state based on the strong damping limit which is qualified by the 10ps stabilization time of detailed full-field simulations47 which is faster than typical 100ps hydrodynamic responses. The CBET modeling presented in this paper assumes there is a linear response in that there is no dependence on the probe intensity which is valid for the intensity ranges presented here. DRACO does have the facility to model the non-linear response for cases of high intensity fields which allows for the saturation effect of the electron density in relation to the IAW where the CBET response will then depend on both the intensity of the pump and probe. It is curious to note that including this dependence on the probe intensity actually helps stabilize the CBET interactions by making the formulation symmetric; however, this modeling then adds another adjustable parameter that dictates the saturation limit. DRACO also has options (with and without saturation) to run using a kinetic based model. These kinetic based options were not fully developed based upon the observations discussed in Ref. 20 given the intensities and scale-lengths of current experiments; the 1-D modeling was used to guide the development effort of DRACO for the sake of time. If the fluid model falls short of emulating experiments DRACO is prepared for the task of simulating with the kinetic models.

The background pump-intensity field Ipump represents an accumulation of all rays from any port traversing each computational zone while accounting for all their intensities, wave vectors, and chirped wavelengths in an angular spectrum representation (ASR)26,27 that divides the accumulated ray energies (weighted as multiplying by path length and dividing by volume, viz., the volume-weighted intensity) into angular and wavelength spectral components (see Fig. 3). In a typical simulation, 50–75k rays commonly traverse cells of dimension 20–50um (imprint simulation uses finer resolution but similar ray density) and accumulate to form the ASR in the active CBET region. The volume weighted average is the appropriate quantity corresponding to the inherent hydrodynamic scale of interest in the simulation; if smaller spatial scales are important then the mesh is designed to have higher resolution. An irregular geometry of a cell is properly handled using this volume weighted intensity construction. For example, a wide cell-area accepts proportionally higher ray density while the shorter cell-depth along the propagation direction governs the pathlength. The opposite aspect ratio yields a smaller ray density with longer pathlengths. In addition, rays from the same port average in intensity whereas rays from differing ports add in intensity; this could not be handled correctly if rays tracked intensity in lieu of energy.

FIG. 3.

Illustration of the angular spectrum representation (ASR). The electron density gradient causes a ray to refract through the expanding plasma atmosphere, which induces a Doppler shift because of the time-dependent change in the electron density. The ASR, indicated by the blue sphere located at the position vector r0 at the cell center of each computational cell, accumulates the weighted ray energy as a function of direction and color, viz., (θ,ϕ,λ) from all the beams that propagate through any cell as Ipump, where the weighting scales as the ray path length per cell volume. A ray that significantly refracts or reenters a cell causes two entries into the ASR: in and outbound.

FIG. 3.

Illustration of the angular spectrum representation (ASR). The electron density gradient causes a ray to refract through the expanding plasma atmosphere, which induces a Doppler shift because of the time-dependent change in the electron density. The ASR, indicated by the blue sphere located at the position vector r0 at the cell center of each computational cell, accumulates the weighted ray energy as a function of direction and color, viz., (θ,ϕ,λ) from all the beams that propagate through any cell as Ipump, where the weighting scales as the ray path length per cell volume. A ray that significantly refracts or reenters a cell causes two entries into the ASR: in and outbound.

Close modal

Each computational zone in the corona contains a unique ASR divided over all propagation angles and wavelengths that defines the background pump field illuminating the zone. For example, the total overlapped inbound intensity of the pump field may be ∼1 × 1015 W/cm2 for all beam ports incident on a cell but this summation represents an ASR that is distributed over many different directions and wavelengths, effectively lowering the intensity of each pump-field interaction and where the net CBET response is obviously the sum over all the individual interactions which are individually dependent on intensity, angles and wavelength. The calculated pump-field intensity accounts for all relevant changes caused by IBS absorption, gain/loss as a result of CBET, and any effective increase or decrease caused by refraction. In fact, the potential intensity increase from refraction located after the turning points (defined as the minimum radius reached where the direction vector becomes normal to the position vector) are inherently included and define a refractive shadow boundary surface with a sharp interface. The intensity increase from refraction can become significant (but is never infinite because rays never cross each other at the same optical pathlength, i.e., always non-zero ray-area, under relevant ICF conditions which only generates a weak form of volume focusing in stark contrast to important distributed focusing like from lens aberrations that occupy significant volume and families of rays converging into the same region); this increase tends to be balanced by occurring within vanishingly small volumes where the net effect is properly handled via the volume weighted intensity. This volume tends to be small because, away from the refractive shadow boundary, refraction in a negative electron density gradient profile is highly divergent which restricts the intensity peaks to small volumes. Any corresponding increase in CBET gain or loss is inherently handled in DRACO since an increase in intensity directly relates to a proportional increase in ray density which increases the density of individual CBET interactions; the net result may increase CBET gain and loss only if the interaction volume is significant.

Any probe ray that traverses a computational zone experiences many instances of CBET gain and loss via summations over all interactions with the entire ASR as well as losses via collisional IBS. All instances of CBET and IBS are accumulated into a total exponential gain/loss factor dτ=dτCBETgain+dτCBETloss+dτIBS over the ray's trajectory through the computational zone. The ray's trajectory is not directly affected by CBET or IBS, but solely determined by the coronal electron density profile of the current time step. However, the trajectory eventually becomes altered by the combined effects of CBET and IBS indirectly since the increased gain and/or scatter and collisional losses eventually change the coronal temperature gradients, which ultimately feed back into the electron density gradients during the hydrodynamic evolution. The total exponential factor dτ represents an accumulation of all gains and losses from the entire ASR and IBS along the trajectory within a computational cell. The total change of a probe ray's energy is given by ΔEprobe=Eprobe0[exp(dτ)1], which asymptotically goes to ΔEprobeEprobe0dτ(1+0.5dτ) as dτ0. Additionally, the ray's average energy over the path length, ds, is exactly given by Eprobe=ΔEprobe/dτ, which asymptotically goes to EprobeEprobe0(1+0.5dτ) as dτ0. However, after all the gain/loss effects accumulate into the total dτ, subsequent separation into different gain and loss energy components become intractable, in general, because of the nonlinear nature of the of the exponential operator. It is tempting to assert that, for example, the total CBET gain, ΔEprobegain is of the form Eprobe0[exp(dτCBETgain)1], where dτCBETgain is the total accumulated CBET gain portion. This is mathematically incorrect because the exponential operators are not separable, resulting in different answers depending on the order of operations taken to separate gain from loss; attempting to balance this disparity using an average of gain/loss and loss/gain is equally unsound and computationally inefficient. If the ray's path length, ds, is small enough such that each elemental contribution to the total dτ is roughly constant or the energy changes are evenly distributed over the whole path, it can be shown that gains and losses separate via a simple ratio scheme. This assumption is consistent with the usage of an ASR that represents the entire computational cell. Assuming constant values or an even distribution of the dτ components over the path length ds total gain and loss components are exactly separable within a sufficiently small volume, e.g., the computational cell, using a straightforward ratio scheme. The total CBET gain, CBET loss, and IBS loss are given by ΔEprobegain=dτCBETgainEprobe,ΔEprobeloss=dτCBETlossEprobe, and ΔEprobeIBS=dτIBSEprobe, respectively, which are exact and computationally efficient under the constant cell-wise or evenly distributed ASR/IBS assumption.

Adaawam calculates the CBET interaction self-consistently in conjunction with the hydrodynamic evolution of the ICF target (via a split-step technique). In addition, Adaawam captures all the necessary coupled interactions between the dynamic electron density profile, temperature, and plasma-flow velocity that dictates the behavior of CBET and vice versa resulting from the strongly coupled CBET and hydrodynamics, which significantly affect each other. Averaging out coronal angular spatial variations of the plasma before calculating CBET or employing a corona not derived using CBET (or from a lesser dimensional code) and its induced spatial variations completely misses the coupling of this essential spatial effect. Adaawam uses advanced iterative feedback control [proportional-integral-differential (PID) control loops28] to stabilize the CBET coupled many-beam interactions while maintaining energy conservation (see Fig. 4). The DRACO CBET model compares with many experimental observables across a range of implosions on OMEGA29 and the NIF.7 An experimentally determined CBET gain multiplier of fexp = 1.5 (from unrelated OMEGA shots29) that uses the first-principles EOS tables was applied to all pre- and post-shot simulations without attempting to fit the NIF shots having similar intensity but different scale lengths and pulse shapes. The CBET gain multiplier of fexp = 1.5 that applies across laser systems indicates a predictive ability in regards to shell morphology and trajectory on the initial wavelength-detuning shot campaign at the tested ∼4–8 × 1014 W/cm2 intensity range. The nonlocal electron transport package10 is used in modeling experiments on both OMEGA and the NIF.

FIG. 4.

The advanced proportional-integral-differential (PID) feedback controller employed in Adaawam. Feedback through a PID-controller loop provides vital control over CBET energy balance. Feedback minimizes energy imbalance through a PID loop by adjusting the ASR until the adjustment returns to zero. Energy is conserved by bringing all neighboring cells into equilibrium.

FIG. 4.

The advanced proportional-integral-differential (PID) feedback controller employed in Adaawam. Feedback through a PID-controller loop provides vital control over CBET energy balance. Feedback minimizes energy imbalance through a PID loop by adjusting the ASR until the adjustment returns to zero. Energy is conserved by bringing all neighboring cells into equilibrium.

Close modal

Maximal CBET occurs in the rapidly expanding coronal plasma where two interacting rays satisfy the ion-acoustic-wave–matching conditions [i.e., Eq. (4), when η= ±1] that account for propagation direction, wavelength, and fluid flow; e.g., a CBET resonance occurs at the Mach-1 surface given a radial plasma flow for directly opposed radially propagating rays of equal angular frequency or wavelength. The significant CBET-interaction region for a directly driven imploding ICF target tends to form at supersonic speeds above the Mach-1 surface for laser beams without detuning (e.g., 1 ≲ Mach ≲ 1.5). The ray's angular frequency is given by its initial value and the integrated temporal derivative of the refractive index (a generalization of the conventional Doppler shift30). The expanding plasma dynamically alters the instantaneous refractive index in space via the temporal dependence electron density, and thereby the angular frequency and wavelength, and is independent of ray propagation direction14 (see the  Appendix B for more details). Consequently, the CBET resonance features adapt as the coronal plasma evolves along with the interacting ray wavelengths. These resultant interactions directly map onto a frequency-chirped scattered-light measurement that can be employed to analyze the implosions and laser–plasma interaction physics. The 2-D code DRACO produces chirped scattered-light diagnostics that have azimuthally varying features resulting from the variation in the modeled 3-D beam-port geometry coupled with the polar and azimuthal angular dependence of CBET (see Figs. 5 and 6). However, as self-emission measurements (see Fig. 7) indicate, there are additional azimuthal variations in the ablation surface (not included in DRACO since the necessary azimuthal feedback cannot be modeled in a 2-D code) that are imprinted in the corona; these azimuthal gradients consequently alter the energy deposition and scattered light. Many closely spaced experimental diagnostic ports are necessary to correlate them on average to a 2-D simulation, which does not include the necessary 3-D feedback into the electron density to reproduce an experiment without averaging (as previously addressed). To resolve these expected variations, future 3-D simulations will attempt to verify the measured azimuthal variations and determine the magnitude of this effect, the impact on the implosions, and possibly alter the pointings to minimize the azimuthal modulations. Alternatively, increasing the number and distribution of scattered-light diagnostic parts could help resolve any discrepancies observed.

FIG. 5.

Spatial scattered-light diagnostic from 2-D DRACO simulations of (a) N160406–001 and (b) N160821–002 showing the angular variations induced from beam-port geometry, repointing, spot shape the angular interaction dependence of CBET. (c) The corresponding azimuthally integrated results of each simulation are plotted showing the locations of NIF scattered-light diagnostic ports and the sensitive polar angles expected to detect effects of wavelength detuning. The predicted azimuthal variations are expected to change when coupled to a 3-D hydrodynamic code, so the experimental results therefore are not expected to be an exact match to a 2-D code unless many more diagnostic ports are made available to take the appropriate averages.

FIG. 5.

Spatial scattered-light diagnostic from 2-D DRACO simulations of (a) N160406–001 and (b) N160821–002 showing the angular variations induced from beam-port geometry, repointing, spot shape the angular interaction dependence of CBET. (c) The corresponding azimuthally integrated results of each simulation are plotted showing the locations of NIF scattered-light diagnostic ports and the sensitive polar angles expected to detect effects of wavelength detuning. The predicted azimuthal variations are expected to change when coupled to a 3-D hydrodynamic code, so the experimental results therefore are not expected to be an exact match to a 2-D code unless many more diagnostic ports are made available to take the appropriate averages.

Close modal
FIG. 6.

Chirped scattered-light diagnostic from experiments and corresponding 2-D DRACO simulations of (a) N160406–001 and (b) N160821–002. The color contour maps of the experimental results are shown along with 10% of peak contours (solid lines) of both the experiments (black) and simulations (magenta).

FIG. 6.

Chirped scattered-light diagnostic from experiments and corresponding 2-D DRACO simulations of (a) N160406–001 and (b) N160821–002. The color contour maps of the experimental results are shown along with 10% of peak contours (solid lines) of both the experiments (black) and simulations (magenta).

Close modal
FIG. 7.

[(a) and (d)] Simulation and [(b), (c), and (e)] experimental self-emission images (of shots N160821-002, N170103-001, and N170716-003) from the [(a), (b), and (e)] polar and [(c) and (d)] equatorial views showing the correlation of the experimentally observed features to the features predicted by the 2-D code DRACO including the long-duration (∼400ps) self-emitting core feature. The experimental images show m= 4 and m= 8 features on the inner rings of the polar view; both features are prominent in the northern-hemisphere cone-swapping shot N170103-001. Future simulations of these experiments with a 3-D code are expected to match the experimental observations.

FIG. 7.

[(a) and (d)] Simulation and [(b), (c), and (e)] experimental self-emission images (of shots N160821-002, N170103-001, and N170716-003) from the [(a), (b), and (e)] polar and [(c) and (d)] equatorial views showing the correlation of the experimentally observed features to the features predicted by the 2-D code DRACO including the long-duration (∼400ps) self-emitting core feature. The experimental images show m= 4 and m= 8 features on the inner rings of the polar view; both features are prominent in the northern-hemisphere cone-swapping shot N170103-001. Future simulations of these experiments with a 3-D code are expected to match the experimental observations.

Close modal

Wavelength detuning between crossing beams responds differently in indirect- versus direct-drive ICF implosions, depending on the dominant CBET mode. In indirect drive, the sign of small wavelength detuning (<2 Å UV) is used to control the direction of energy transfer between co-propagating interacting beams by leveraging the CBET resonance for the forward-scatter (or sidescatter) mode where the pump and probe beams are either both inbound or outbound,14 i.e., |kavfluid||ωpumpωprobe| in Eq. (4) (see Fig. 8 for an illustrated example), implying that the frequency difference controls the sign of η. The magnitude of the kavfluid term becomes insignificant for two reasons: (1) the pump- and probe-vectors co-propagate within a small angle, leading to a small-magnitude ion-acoustic wave vector, kakpumpkprobe; and (2) its direction is nearly orthogonal to the fluid velocity, vfluid. The main CBET interaction occurs near the laser entrance hole of the indirect-drive hohlraum. The forward-scatter mode effect on spatial distribution of laser-energy deposition is more dramatic in indirect drive because there is significant propagation (or throw) after the interaction. The propagation distance separates the far-field spots before deposition on the inside of the hohlraum wall. Any forward-scatter mode interactions that occur change the balance on the hohlraum wall because of the large spot separation. While this mode occurs in direct drive, it negligibly alters scattered-light loss because any energy exchanged is merely deposited in slightly different regions (because of the short propagation distance before deposition and the significant beam overlap) contributing to distortion; some modest additional distortion at small wavelength separations can arise.14 In contrast, an outbound ray in the dominant backscatter mode in direct drive always experiences CBET gain regardless of the wavelength-difference sign or magnitude (for nominal levels) because the ion-acoustic wave's contribution dominates the CBET resonance function, i.e., |kavfluid||ωpumpωprobe| in Eq. (4) (see Fig. 2 for an illustrated example), implying that the frequency difference cannot control the sign of η under normal circumstances.14 Under atypical conditions, the outbound ray may experience a loss resonance for extreme wavelength separation (e.g., >20 Å UV) but insignificantly impacts scattered light because the outbound rays typically transport little energy.

FIG. 8.

In the side- of forward-scatter CBET mode the pump and probe rays are both inbound or outbound and co-propagate within a small angle. The resulting ion-acoustic wave vector, kakpumpkprobe, has a small magnitude and its direction is nearly orthogonal to the fluid velocity, vfluid. Consequently, the CBET resonance condition [Eq. (4)] becomes controlled by the wavelength or color separation of the pump and probe.

FIG. 8.

In the side- of forward-scatter CBET mode the pump and probe rays are both inbound or outbound and co-propagate within a small angle. The resulting ion-acoustic wave vector, kakpumpkprobe, has a small magnitude and its direction is nearly orthogonal to the fluid velocity, vfluid. Consequently, the CBET resonance condition [Eq. (4)] becomes controlled by the wavelength or color separation of the pump and probe.

Close modal

The ensemble CBET exchange is best described as an interaction volume (a weighted volume that determines the interaction strength, which depends on path length, intensity, wavelength, electron density, coronal temperature, fluid velocity, etc.) because any high-gain region is equally matched by loss and significant CBET occurs only when the ensemble interaction volume is large. For example, there might be high intensity after a turning point over insignificant path lengths and a diminishing area that form an ineffective and vanishingly small interaction volume with minimal resulting CBET. In addition, the physical diffraction-limited transverse speckle sizes formed from realistic focused beams are significantly larger in volume, which disrupt and diffuse any small-scale (∼micron or less) diffraction effects from self-interfering idealized smooth beams near the refractive shadow boundary. For speckle or the diffraction pattern to cause a CBET response to deviate significantly from this linear view the conditions untypical or irrelevant to ICF implosions need to exist to induce nonlinear behavior which also imparts saturation effects. As examples, this can occur using a construction of speckle not adhering to the known Goodman spectrum21 resulting in incorrect amplitudes or widths, not modeling refraction in spherical geometry causing speckle to interact over long distances by minimizing transverse movement of speckle and artificially enhancing the CBET response or combining all the power into a single simulated beam (or too few beams) instead of from many beams (>10 for OMEGA or >40 for NIF) normally overlapping from a wide distribution of beam angles resulting in non-ICF relevant conditions. There is little evidence to suggest that nonlinear conditions occur in typical direct-drive ICF implosions but they may become important to some degree at ignition scale. To be hydrodynamically significant, the average energy of rays over the entire computational cell must be significant, i.e., the intensity defined as energy times the path length divided by cell volume should be substantial. This definition is consistent with the CBET interaction via the ASR that represents the entire computational cell. A method that views the CBET interaction over the entire cell but employs high intensities that arise from refraction without compensating for the proportionally diminishing cross-sectional area and very small path lengths will inadvertently overpredict the CBET response. In contrast, attempting to model field-swelling by applying a maxima function actually dampens or saturates the CBET response rather than modeling what the name suggests, i.e., swelling implies that enhancement not dampening of CBET occurs. The preferred method to handle large intensities over significant interaction volumes is modeling pump-depletion and is currently being added to DRACO which doesn't require auxiliary saturation. Currently a modest form of saturation can be employed in DRACO to prevent excessive gain during convergence but typically is not required.

The resonant CBET gain region of the outbound rays in the backscatter mode never disappears but rather shifts into a smaller interaction volume because the relative instantaneous wavelength difference changes the ion-acoustic-wave–matching conditions of the interacting rays. The resonance region bifurcates and shifts both farther out in the corona (where the outbound rays have lower intensities and experience higher expanding fluid velocity and lower electron density) and closer inside the corona (where the interaction becomes shielded by the refractive shadow-boundary surface and/or outbound rays that have negligible intensity)14 [see Fig. 1(a)]. The net effect of wavelength detuning shifts the significant CBET-interaction region to lower Mach numbers covering both sub- and supersonic regions (e.g., 0.7 ≲ Mach ≲ 1.3.), i.e., moves the region closer toward the ablation surface. A sufficiently large wavelength separation (detuning) significantly reduces CBET exchange for direct drive by decreasing the interaction volume (e.g., ±6 Å UV). In contrast, an insufficient wavelength i.e., <2 Å UV, separation can lead to deposition and shell distortion via the forward-scatter mode.14 The efficacy of wavelength-detuning CBET mitigation diminishes as the plasma expands and the target implodes, which causes the CBET resonance regions to drift gradually into larger interaction volumes during the drive pulse.14 Selecting large wavelength-detuning values delay the inevitable onset of diminished mitigation and, for this reason, using the largest wavelength separation available yields the most favorable results.14 The decay rate of detuning efficacy differs according to the relative wavelength shift and causes asymmetries which are remedied using balanced designs. Simulations predict that wavelength-detuning CBET mitigation is effective for both symmetric direct drive (SDD) on OMEGA and polar direct drive (PDD), the present illumination configuration on the NIF, since the same mechanisms occur in both configurations, although the positive impact is more pronounced for PDD.14 Note that because the Randall-based CBET-model assumes a steady-state response, only the effects of wavelength separation can be properly treated and the dampening due to the coherence-disruption effects of bandwidth on LPI-process growth cannot be modeled. This also applies to the future modeling of smoothing by spectral dispersion (SSD);26 see  Appendix B for a brief discussion.

The efficacy of wavelength-detuning CBET mitigation has been studied on the NIF.31 The target designed for these wavelength-detuning shots on the NIF was adapted from existing 600-kJ designs,9 where the trajectories and the shape of the imploding shell and scattered light were well described by the CBET model in DRACO. The platform target design is shown in the inset of Fig. 9, where the laser beam powers (shown in red) produce a peak overlapped intensity of ∼4–8 × 1014 W/cm2 at the initial target radius.

FIG. 9.

The National Ignition Facility (NIF) PDD target design for wavelength detuning with cone swapping to induce a wavelength difference across the equator. Inset: the warm plastic (CH), 1160-μm-radius, 100-μm-thick shell with a 20-atm D2 gas fill; (red) the total 590-kJ design pulse; (blue) the 45-kJ backlighter pulse. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

FIG. 9.

The National Ignition Facility (NIF) PDD target design for wavelength detuning with cone swapping to induce a wavelength difference across the equator. Inset: the warm plastic (CH), 1160-μm-radius, 100-μm-thick shell with a 20-atm D2 gas fill; (red) the total 590-kJ design pulse; (blue) the 45-kJ backlighter pulse. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

Close modal

The indirect-drive NIF beam geometry distributes 192 beam ports [grouped into 48 quads, shown as projected circles in Fig. 10(a)] toward the poles of the NIF target chamber, forming cones of quads that share a common polar angle.32 Repointing higher-intensity beams from lower latitudes toward the equator partially compensates for the NIF port geometry and higher incident angles when illuminating direct-drive targets. In this modified configuration, referred to as PDD,24,33 CBET predictably dominates in the equatorial region where most of the crossing-beam interactions occur due to the requisite repointing,14,34 as shown in Fig. 1(b). The beam port angles that would have induced significant polar CBET are absent in the PDD configuration. As a result, PDD implosions tend to become oblate because CBET reduces the laser drive preferentially in the equatorial region. With this motivation, a basic wavelength-detuning strategy exploits the PDD configuration, where each hemisphere has a different wavelength or color. However, the nominal symmetric wavelength mapping [see Fig. 10(a)] developed for indirect-drive targets precludes achieving true hemispheric wavelength detuning using typical PDD repointing configurations.24 

FIG. 10.

NIF Quad-Port Hammer projections for the wavelength-detuning CBET mitigation scheme. (a) Indirect-drive mapping where the colored symbols indicate relative wavelength; (b) PDD repoint mapping that achieves hemispheric detuning, typical northern-hemisphere repointing, and southern-hemisphere cone swapping. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

FIG. 10.

NIF Quad-Port Hammer projections for the wavelength-detuning CBET mitigation scheme. (a) Indirect-drive mapping where the colored symbols indicate relative wavelength; (b) PDD repoint mapping that achieves hemispheric detuning, typical northern-hemisphere repointing, and southern-hemisphere cone swapping. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

Close modal

The NIF fiber front end32 supports three separate initial colors or wavelength shifts Δλ0 = {λ1,λ2,λ3} detuned from a central wavelength λ0 ∼ 351 nm. Currently, the three-color {λ1,λ2,λ3} mapping onto the NIF indirect-drive ports is symmetric about the equator [see Fig. 10(a)]. The symmetric color mapping precludes achieving a wavelength difference about the equator using typical PDD repointing schemes. A dramatic repointing (referred to as “cone swapping”) induces a wavelength difference within the equatorial region and when applied in either the northern or southern hemisphere [see Fig. 1(b) for the southern case]. For cone swapping, in one hemisphere the higher-latitude ports (“inner cones”: {λ1, λ2}) are repointed to the equator and the lower-latitude ports (“outer cones”: {λ3}) are repointed to the mid- and high latitudes. For the first set of wavelength-detuning experiments described here, two different colors were specified such that λ1 = λ2λ3. The current NIF configuration, while not optimal, is capable of achieving a modest wavelength-detuning level Δλ0 = {+2.3,+2.3,–2.3} Å UV, which is adequate for these proof-of-principle experiments. Cone swapping plus wavelength detuning induces the desired partial hemispheric wavelength difference between beams crossing the equatorial region. However, cone swapping is not intended for high-performance implosions because it induces a north/south asymmetry because of the combination of large repointing in only one hemisphere and the consequential illumination by nonideal spot shapes in the swapped locations (since DPP reconfiguration takes considerable operations time).

The far-field spot envelope (induced from DPP's21 and small-divergence smoothing) quad mapping is given by the current indirect-drive configuration on the NIF: the inner and outer cones. The inner cones (λ1, λ2; red/green projected circles in Fig. 10) use a wide elliptical spot shape not well suited for the equatorial region but appropriate for the polar and mid-latitude illumination; the wide spot has a super-Gaussian ∼3 profile with ∼2-mm-wide major axis and ∼4:3 ellipticity. The outer cones (λ3; blue projected circles in Fig. 10) use a narrow elliptical spot shape adequate for the equatorial region but not well suited for polar or mid-latitude illumination; the narrow spot has a super-Gaussian ∼3 profile with ∼1.5-mm-wide major axis and ∼2:1 ellipticity. The values of the beam energy and repointing were additionally adjusted in the cone-swapping hemisphere to compensate for the swapped spot shapes and the higher incident angles using established PDD design principles.24 The cone-swapping repointing scheme and the fixed DPP quad-mapping result in nonoptimal implosion symmetry [see Fig. 10(b)] because the illumination pattern in the cone-swapping hemisphere is nonideal and the incident angles are increased (decreasing hydroefficiency). As previously mentioned, however, cone swapping does provide the desirable equatorial wavelength separation of opposing beams from each hemisphere. For this reason, fusion yield and areal density are not metrics for these experiments, which concentrate instead on observables directly related to laser-energy absorption: implosion trajectory, shell morphology, and scattered light. Future reconfigurations (optimal DPP's for PDD,21,24 flexible color mapping,35,36 and larger wavelength separation35,36) can relieve these constraints, and simulations predict improved overall fusion performance.

With this motivation, for the first time in direct-drive ICF, wavelength-detuning CBET mitigation was demonstrated and shown to improve energy coupling. The NIF PDD wavelength-detuning CBET mitigation campaign shots were performed in three pairs. Each pair consisted of one implosion backlit with ∼6.7-keV x rays produced from a planar Fe foil target energized by two quads of NIF beams with 45 kJ (see the blue curve in Fig. 9) of UV laser energy per beam with an equatorial view of the compressing shell. The second implosion of the pair examined self-emission images (800eVhν10keV) of the compressing target from equatorial and polar views using a 25-μm Be filter. Additional diagnostics measured both hard x rays produced by energetic electrons arising from the stimulated Raman scattering (SRS) and possibly two-plasmon–decay (TPD) instabilities. The inferred levels of these instabilities contain at most only a few percent of the incident energy and do not affect the analysis of the laser–target coupling and CBET.8 The mitigation of CBET is expected to increase the observed levels of SRS and TPD because the resulting increased incident intensity could rise above the various LPI instability activation thresholds. Discussion of the observed increased TPD and SRS signals is planned for a future publication. It should also be noted that a larger wavelength separation than used in these experiments is expected to reduce TPD efficiency which implies that the TPD response should dampen rather than increase as inferred in these experiments given the larger bandwidth.47 

The first pair of control experiments (shots N160405 and N160406) with the same wavelength for all the beams (zero detuning) was performed to establish the baseline experimental observables. Next, two pairs of experiments with a detuning mapping of Δλ0 = {+2.3,+2.3,–2.3} Å UV were performed to evaluate the efficacy of wavelength-detuning CBET mitigation. The zero-detuning and first-detuning shots (N160821–001 and N160821–002) employed southern-hemisphere cone swapping, as illustrated in Fig. 10(b). The second-detuning shots (N170102 and N170103) applied northern-hemisphere cone swapping, primarily to effectively image the self-emission from the opposite/antipodal pole and to observe the expected image inversion. The repointing (accounting for mirror-image cone swapping) and pulse shapes were nominally identical for all shots where the only intended difference was the wavelength configuration.

The simulated and measured backlit gated x-ray radiographs are analyzed to show shell morphology evolution and in-flight shell trajectory, which are both used to infer energy coupling. The gated images (gate time ∼100 ps) shown in Fig. 11 compare the shell morphology for the three backlit shots. The experimental framing-camera images are a composite of several images close in time for this slowly moving target that were cross-correlated and adjusted for magnification to enhance the signal-to-noise ratio; the measurements used a 30–μm pinhole. The simulated images are post-processed with the x-ray imaging code Spect3D37 with pinhole and gate values matching the experimental setup. The first two rows are radiographs of matched post-shot simulations and experimental results for the baseline zero-detuning and wavelength-detuning shots with southern-hemisphere cone swapping. The last row shows radiographs for detuning shots with northern-hemisphere cone swapping. All the backlit radiograph data show remarkable agreement between simulation and experiment, especially the expected trend for the detuning shots. A mere ∼2% to 3% additional laser energy is absorbed with detuning, but since this energy is localized to the equatorial coronal volume fraction (∼25%), and the deposition is redistributed closer to the ablation surface increasing hydrodynamic efficiency, the result is dramatic as observed with the gated x-ray radiographs.

FIG. 11.

Comparison of backlit radiographs from post-shot DRACO simulations and NIF experimental results near the end of the laser pulse at t = 8.5 ns. The dashed lines indicate the outer shell surface extracted from each image defined by the steepest gradient in the inward radial direction. Reprinted with permission from Phys. Rev. Lett. 120, 085001–4 (2018). Copyright (2018) American Physical Society.

FIG. 11.

Comparison of backlit radiographs from post-shot DRACO simulations and NIF experimental results near the end of the laser pulse at t = 8.5 ns. The dashed lines indicate the outer shell surface extracted from each image defined by the steepest gradient in the inward radial direction. Reprinted with permission from Phys. Rev. Lett. 120, 085001–4 (2018). Copyright (2018) American Physical Society.

Close modal

Most notable was the design prediction and the subsequent measurement of the equatorial mass accumulation with active wavelength detuning (bottom two rows in Fig. 11). As predicted, the mass accumulation flipped orientation when cone swapping was applied to the opposite hemisphere. The wavelength-detuning design attempted to minimize the ℓ = 2 Legendre mode while accounting for the spot shapes, pointing, and energies in conjunction with the expected increased drive in the equatorial region caused by CBET mitigation. The equatorial mass accumulation is a common feature in PDD implosions with small ℓ = 2 contributions and not directly related to CBET mitigation. The accumulation is caused by unbalanced lateral mass flow toward the equator (from primarily oblique incidence) when sufficient equatorial drive is available (e.g., from CBET mitigation) while achieving a small ℓ = 2 component. Corrections for the equatorial mass accumulation consist of design attributes that specifically address the equatorial region, e.g., using optimal DPP's or shell-mass contouring.

The shell trajectory is inferred from the simulated and experimental backlit radiographs by first extracting the outer steepest gradient surface or radii [see Figs. 11 and 12 (inset)]. The majority of the CBET gain occurs in the equatorial region [Fig. 1(b)] and consequently the region expected to benefit from wavelength detuning. Both the surface-area–weighted average of the entire extracted surface and a range restricted to the equatorial region (shown here) demonstrate the benefit. When the extracted shell surface is restricted to the equatorial region (±30° region about the equator) and plotted as a function of time (see Fig. 12), the inferred implosion speed increases as a result of wavelength-detuning CBET mitigation. The equatorial shell speed increases 9% from 144 to 157 μm/ns based on simulation (experimentally a 16% increase from 133 to 154 μm/ns) because wavelength-detuning CBET mitigation deposits ∼3% additional energy within the small volume over the equator with enhanced hydroefficiency due to coronal deposition closer to the ablation surface. The enhanced equatorial velocity consistently results when comparing the extracted outer shell contours taken from zero detuning and detuning shots in Fig. 12 (inset), where the entire surface-area–weighted average implosion speed increases experimentally by 13%. This observation in simulation and experiment suggests that CBET mitigation via wavelength detuning promises to alleviate the excessive losses in directly driven ICF targets. Simulations predict that as the wavelength separation increases so does the efficacy of the mitigation.35 Simulations also predict that large wavelength separations (±6 Å UV) also significantly benefit implosions with ignition-relevant implosion speeds (∼350 to 400 μm/ns) (Ref. 38).

FIG. 12.

Equatorial shell trajectories from post-processed simulated (solid lines) and experimental (symbols) backlit radiographs. The red lines/symbols represent the baseline zero-detuning experiment (N160405). The blue lines/symbols represent the average of the two detuning experiments (N160821 and N170102). The inset shows superimposed extracted surfaces from the experimental radiographs of Fig. 11, exemplifying the equatorial mitigation. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

FIG. 12.

Equatorial shell trajectories from post-processed simulated (solid lines) and experimental (symbols) backlit radiographs. The red lines/symbols represent the baseline zero-detuning experiment (N160405). The blue lines/symbols represent the average of the two detuning experiments (N160821 and N170102). The inset shows superimposed extracted surfaces from the experimental radiographs of Fig. 11, exemplifying the equatorial mitigation. Reprinted with permission from Marozas et al., Phys. Rev. Lett. 120, 085001 (2018). Copyright 2018 American Physical Society.

Close modal

The simulated self-emission images in both the polar and azimuthal views predict the presence of a core self-emission feature; see Fig. 7. The corresponding experimental self-emission measurements confirm the presence of this self-emitting core feature. In fact, only this series of PDD shots using wavelength detuning consistently measure this long-duration (∼400ps) feature on every self-emission measurement shot. The self-emitting core signature for wavelength detuning shots help qualify current 2-D and future 3-D simulations as well as provides an excepted diagnostic marker of the wavelength detuning shots.

Additional NIF capabilities improve PDD target–energy coupling according to simulations. The proposed capabilities include flexible color-to-port mapping, custom DPP's, and an increase in wavelength separation. Flexible color-to-port mapping eliminates cone swapping and consequently improves the north/south asymmetry. Providing a full set of optimized SMA-DPP's mounted to the ideal port locations minimizes repointing (assuming flexible color remapping) and further improves the north/south symmetry. SMA-DPP's also improve energy efficiency (by eliminating over-the-horizon energy dumping caused by repointing), improve uniformity, and increase CBET mitigation (by tailoring the SMA to be slightly interior to the initial target radius) (see Fig. 13). Increased wavelength separation provides improved CBET mitigation and the flexible color remapping permits a balanced tri-color configuration (see Fig. 14).35,36 Simulations of intermediate-energy (∼700-kJ) designs on the NIF illustrate the expected improvements over these proof-of-principle experiments with additional NIF capabilities (see Fig. 15). The proposed additional NIF capabilities provide dramatic symmetry control and higher convergence and demonstrate control over the equatorial mass accumulation. The progress toward providing these additional NIF capabilities can be taken in a staged approach where, for example, one-third of the total SMA-DPP's are used to populate the beam ports repointed to the equatorial region as an initial step toward the ultimate goal.

FIG. 13.

Optimized spot-masking apodization–distributed phase plate (SMA-DPP) for the equatorial repointed beams using NIF PDD. The tailored shape improves both energy efficiency and uniformity. The SMA aspect also can limit the over-the-horizon energy that provides CBET mitigation without significantly impacting uniformity.

FIG. 13.

Optimized spot-masking apodization–distributed phase plate (SMA-DPP) for the equatorial repointed beams using NIF PDD. The tailored shape improves both energy efficiency and uniformity. The SMA aspect also can limit the over-the-horizon energy that provides CBET mitigation without significantly impacting uniformity.

Close modal
FIG. 14.

The balanced tri-color scheme proposed for the NIF provides a large amount of CBET mitigation while the color mixture balances the mitigation between the northern and southern hemispheres. A flexible remapping of NIF's fiber front end would be required to achieve this color distribution. The retroreflections from the final optics would be necessary in order for the larger wavelengths to separate.

FIG. 14.

The balanced tri-color scheme proposed for the NIF provides a large amount of CBET mitigation while the color mixture balances the mitigation between the northern and southern hemispheres. A flexible remapping of NIF's fiber front end would be required to achieve this color distribution. The retroreflections from the final optics would be necessary in order for the larger wavelengths to separate.

Close modal
FIG. 15.

The results of a highly optimized NIF PDD simulation at a similar distance traveled as in Fig. 11. The simulation includes CBET, detuning specified in Fig. 14, and optimal SMA-DPP's. The simulation achieved high convergence and maintained adequate symmetry down to a convergence ratio of ∼20. This figure illustrates the level of improvement expected with additional NIF capabilities.

FIG. 15.

The results of a highly optimized NIF PDD simulation at a similar distance traveled as in Fig. 11. The simulation includes CBET, detuning specified in Fig. 14, and optimal SMA-DPP's. The simulation achieved high convergence and maintained adequate symmetry down to a convergence ratio of ∼20. This figure illustrates the level of improvement expected with additional NIF capabilities.

Close modal

The OMEGA Laser System employs a three-way equal-energy split initiating the eventual separation into 60 large-area beams. Each leg of the three-way split contains 20 beams distributed around the OMEGA target chamber in a manner that minimizes the nonuniformity for each leg separately (see Fig. 16). Replacing OMEGA's large-aperture ring amplifier (LARA) and three-way split with three independent vertical amplifier sections (similar to the NIF) would outfit OMEGA with three independently controllable wavelengths in sets of 20 beams [originally proposed internally at the Laboratory for Laser Energetics (LLE) circa 2013]. Phase modulation generates multiple coexisting spectral lines without the severe coherent interference expected by co-propagating beams with different wavelengths. Phase modulating each beam to create widely separated lines would be superior to three sets of 20 beams since it effectively multiplies the total number of interacting sources and lowers each source line's intensity, which also lowers CBET interaction strength in concert with wavelength separation. However, this option requires substantial research to achieve the required line separation and to manage any issues amplifying this bandwidth (e.g., large-scale optical parametric amplification), whereas the option proposed here is feasible with currently available optics. Note that depending on the wavelength separation and desired shot energy, this potential upgrade could require wide-bandwidth optical coatings, new amplifier glass, and frequency-conversion crystals. While the original intention of the spatial distribution of the legs was not to support multiple colors, the spatial distribution of the legs is advantageous for CBET wavelength-detuning mitigation in the symmetric direct-drive configuration because it creates a distribution of wavelengths around the target that reduces the CBET interactions that occur using dissimilar wavelengths.

FIG. 16.

Diagram of the OMEGA target chamber showing the distribution of the three legs of 20 beams each, where each color (red, green, and blue) represents each leg and not necessarily its wavelength.

FIG. 16.

Diagram of the OMEGA target chamber showing the distribution of the three legs of 20 beams each, where each color (red, green, and blue) represents each leg and not necessarily its wavelength.

Close modal

A cryogenic DT target matched with a triple-picketed 21-kJ pulse provides a platform to test multiple CBET mitigation schemes via simulation (see Fig. 17), where the target's nominal initial outer radius is rt = 415 μm, which includes the 80-μm DT ice layer and plastic (CH) overcoat. The energy was lowered for wavelength detuning, which was projected to be limited to 21 kJ with the current amplifier-gain window, although it is expected to have full system energy with an upgrade to wider-bandwidth laser glass.

FIG. 17.

(a) The OMEGA cryogenic target and (b) 21-kJ pulse shape used to compare CBET mitigation schemes. The target has a 3.3-atm fill near the triple point with a 327-μm radius, surrounded by an 80-μm cryogenic DT ice layer and a plastic (CH) overcoat, totaling 415-μm outer radius. The pulse shape was limited to 21 kJ to satisfy the projected energy limitations of using zooming DPP's46 and wide wavelength separations. It is expected that laser glass upgrades would allow a larger system energy (>26 kJ) with Δλ0=±6ÅUV.

FIG. 17.

(a) The OMEGA cryogenic target and (b) 21-kJ pulse shape used to compare CBET mitigation schemes. The target has a 3.3-atm fill near the triple point with a 327-μm radius, surrounded by an 80-μm cryogenic DT ice layer and a plastic (CH) overcoat, totaling 415-μm outer radius. The pulse shape was limited to 21 kJ to satisfy the projected energy limitations of using zooming DPP's46 and wide wavelength separations. It is expected that laser glass upgrades would allow a larger system energy (>26 kJ) with Δλ0=±6ÅUV.

Close modal

The test platform uses a pulse and cryogenic DT target designed to produce high compression and hot-spot pressures exceeding 100 Gbar in reference simulations without including CBET. The hot-spot pressure is defined as the peak neutron-weighted total pressure attained interior to the compressed shell during the implosion. The value of 100 Gbar qualifies the implosion as ignition-scale–relevant conditions and is currently a major goal of the Omega Laser Facility. The non-CBET reference simulation uses a spot shape labeled SG5, which is defined as a super-Gaussian (SG) order equal to 5.0 with 95% enclosed energy set to rt (equivalent to a 5% intensity point set to 466.9 μm). The non-CBET simulation becomes the reference and goal that effectively measures success of any CBET-mitigation scheme because any successful CBET-mitigation scheme should converge toward this non-CBET simulation, i.e., the maximum expected laser absorption given that DRACO only simulates IBS and CBET. Another important reference simulation enables the CBET effect without employing any mitigation scheme that measures the overall impact of CBET, i.e., the minimum expected laser absorption, and also employs the SG5 spot shape. The significant CBET-interaction region for this reference case tends to be supersonic (1.0 ≲ Mach ≲ 1.5) during the main pulse. These two simulations without and with CBET set the upper and lower laser-absorption bounds, where their difference sets the maximum recoverable energy that can be expected and this difference can be used to measure the success of a particular mitigation scheme.

The two CBET mitigation schemes compared in this paper are a wavelength-detuning scheme with customized SMA-DPP's and a spatial scheme that underfills the target by 25% by decreasing the diameter of a typical far-field spot shape which is typically denoted as R75. Note that both of these mitigation schemes employ the spatial domain mitigation scheme to varying degrees. The degree of spatial mitigation directly impacts target performance, which can undermine any resulting benefits of CBET mitigation.

The first CBET mitigation scheme employs wavelength detuning with a customized SMA-DPP. The wavelength-detuning configuration is distributed as shown in Fig. 16, and the wavelength-detuning configuration is Δλ0={6,0,+6}ÅUV. The wavelength-detuning case also employs SMA-DPP's labeled SMA90SG3.3 to optimize the on-target uniformity while gently removing some over-the-horizon energy to spatially mitigate CBET. The SMA-DPP's were designed for a slightly overfilled SG = 3.3 with 95% enclosed energy set to rt (equivalent to a 5% intensity point set to 457.4 μm), which tends to minimize on-target uniformity, and using another super-Gaussian modulation that apertures the spot tails to trim the energy at 0.90 * rt while maintaining the underlying optimal super-Gaussian 3.3. The significant CBET-interaction region for this detuning case is shifted to lower sonic speeds and is approximately centered on the Mach-1 surface (0.7 ≲ Mach ≲ 1.3) during the main pulse.

The second CBET mitiagation scheme to be compared employs a spot shape that underfills the target by 25% (R75). The reduced spot shape specifically examined in this paper is denoted as SG4-R75 (Ref. 13) because the spot's SG = 4.0 with 95% enclosed energy set to 315 μm (∼0.75 rt; equivalent to a 5% intensity point set to 352.0 μm). There are a few detrimental side effects on target performance that result from underfilling the target by too much. The SG4-R75 spot has a deleterious impact on target performance by increasing the target's sensitivity to imbalances (such as mistiming, mispointing, and offset) and increasing the imprint susceptibility (short-wavelength modes imprint on the ablation surface caused by the speckled far-field pattern of a DPP) because of the lower effective number of overlapped beams; also the CBET activity interior to the beam increases because the beam intensity increases for the same delivered energy. The SG4-R75 case is more sensitive to laser imbalances (typical OMEGA values: 10% rms constant power imbalance, 5-ps rms mistiming, and 10-μm rms mispointing), as illustrated in Fig. 18 for power imbalance, where the induced modal structure is larger for ℓ < 10 Legendre modes.

FIG. 18.

Hard-sphere projection spectra comparing a target fully illuminated by a super-Gaussian (SG) = 3.4% and 5% intensity spot radius of 1.05 * rt (blue lines), cf. an underfilled target using a SG = 4.0 with 95% enclosed energy set to 315 μm (red lines), where rt is the initial target radius. The thick lines are projections without any imbalances. The thin lines result from hard-sphere projections that include 10% constant power imbalance. The plots illustrate how underfilling the target compromises uniformity (cf. ℓ = 10) and makes the projection more sensitive to imbalances (higher power in modes ℓ < 10). Although not plotted, the underfilled R75 spot projected onto target is also more-sensitive target offset mistiming and mispointing as well as suffering from increased imprint susceptibility.

FIG. 18.

Hard-sphere projection spectra comparing a target fully illuminated by a super-Gaussian (SG) = 3.4% and 5% intensity spot radius of 1.05 * rt (blue lines), cf. an underfilled target using a SG = 4.0 with 95% enclosed energy set to 315 μm (red lines), where rt is the initial target radius. The thick lines are projections without any imbalances. The thin lines result from hard-sphere projections that include 10% constant power imbalance. The plots illustrate how underfilling the target compromises uniformity (cf. ℓ = 10) and makes the projection more sensitive to imbalances (higher power in modes ℓ < 10). Although not plotted, the underfilled R75 spot projected onto target is also more-sensitive target offset mistiming and mispointing as well as suffering from increased imprint susceptibility.

Close modal

The results from all four runs are tabulated in Table I, where the first two rows are the reference runs that bracket the CBET effect. The non-CBET run absorbs 78% of the laser energy and achieves the desired goal of >100-Gbar hot-spot pressure, where the achieved pressure is 169 Gbar by design to account for laser system imbalances and because CBET mitigation schemes are not 100% effective. The CBET effect results in a decreased absorption fraction of 22.9% energy difference that degrades the hot-spot pressure ∼5×, which is 100-Gbar goal. The 22.9% energy difference between CBET and non-CBET runs can be used to measure the success of a mitigation scheme using the same target design. The SG4-R75 DPP helps to mitigate CBET, increases the absorption fraction to 65.2%, and more than doubles the hot-spot pressure to 72 Gbar, which is still under the design goal of 100 Gbar using a simulation under ideal conditions. The CBET mitigation of SG4-R75 yields a (65.2–55.1)/22.9 = 44.1% mitigation success factor. The SMA90SG3.3 simulation further improves the CBET mitigation, bringing the absorption fraction to 73.6% (nearly equivalent to the non-CBET run) and the hot-spot pressure above the design goal at 119 Gbar. The CBET mitigation of SMA90SG3.3 yields a (73.6 – 55.1)/22.9 = 80.8% mitigation success factor.

TABLE I.

The tabulated data from the four OMEGA-scale DRACO simulations evaluating cross-beam energy transfer (CBET) mitigation schemes under ideal conditions. The first two rows are the reference runs (shaded areas), without and with CBET, which define the CBET bracket. The third row shows data from the SG4-R75 simulation and the fourth row from the SMA90SG3.3 simulation. The values in the parentheses are from 1-D LILAC simulations showing good agreement with the DRACO results. The values in the brackets are simulations under nonideal conditions with simulated laser imblances using the format of [ImbA,ImbB] where the value in position ImbA does not apply non-zero m-modes but the value in position ImbB does apply non-zero m-modes. Highlight denotes CBET bracket.

CBETΔλ0,UVFar-field spotEabs (%)Phs (Gbar)YDT (×1014)Note
No SG5 78.0 (78.7) 169 (166) 2.29 (2.67) Reference 
Yes SG5 55.1 32 0.09 Worst case 
Yes R75SG4 65.2 72 [69,60] 0.36 [0.32,0.12] R75 
Yes {±6,0}Å SMA 90% SG3.3 73.6 119 [111,0.86] 1.17 [0.97,0.36] Δλ0 
CBETΔλ0,UVFar-field spotEabs (%)Phs (Gbar)YDT (×1014)Note
No SG5 78.0 (78.7) 169 (166) 2.29 (2.67) Reference 
Yes SG5 55.1 32 0.09 Worst case 
Yes R75SG4 65.2 72 [69,60] 0.36 [0.32,0.12] R75 
Yes {±6,0}Å SMA 90% SG3.3 73.6 119 [111,0.86] 1.17 [0.97,0.36] Δλ0 

The resultant shell morphology is illustrated by plotting the mass density near peak compression (see Fig. 19). Peak compression occurs at t ∼2.9 ns for the non-CBET reference run and the SMA90SG3.3 because the absorption fractions are similar, whereas it occurs at t ∼3.05 ns for the SG4-R75 run because of its lower energy absorption. Since the CBET mitigation for SMA90SG3.3 was nearly complete, the shell morphology is similar to the reference case with the exception that it is slightly delayed and the plotting resolution was not fine enough to capture it closer to its true peak; therefore, the lower mass density shown in Fig. 19(b). The ℓ = 10 Legendre mode is more pronounced in the SMA90SG3.3 case because of the use of SMA-DPP's with 90% sub-aperturing, which was expected but deemed acceptable for the benefit of improved CBET mitigation. However, the ℓ = 10 Legendre mode is far more severe for the SG4-R75 case and is the primary reason for its poor performance. The CBET mitigation benefit of the SG4-R75 spot does not outweigh the damage done to the shell. While it may be possible to fine-tune the pulse and target for higher hot-spot pressures using an R75 type of DPP just as SMA90SG3.3 can be retuned for the improved performance, this was not the point of this comparison. The point was to ascertain the mitigation success factor for different mitigation schemes by varying nothing except the mitigation scheme since success is determined relative to the non-CBET reference run. The fact remains that the SG4-R75 mitigation success factor is ∼0.5× that of SMA90SG3.3. This testing platform technique can be used to evaluate a variety of CBET mitigation schemes and is not limited to the two cases performed here.

FIG. 19.

Shell mass-density contour plots from DRACO simulations near peak compression plotted at similar distance traveled under ideal conditions. (a) The non-CBET SG5 reference run at t= 2.9 ns, (b) the SG4-R75 run at t= 3.05 ns [delayed relative to (a) because of the lower absorption fraction] and (c) the SMA-DPP SG3.3 with wavelength detuning run at t= 2.9 ns. Note the shell for (c) has not quite achieved maximum compression but was plotted for the distance traveled closest to (a). The results from these runs are tabulated in Table I.

FIG. 19.

Shell mass-density contour plots from DRACO simulations near peak compression plotted at similar distance traveled under ideal conditions. (a) The non-CBET SG5 reference run at t= 2.9 ns, (b) the SG4-R75 run at t= 3.05 ns [delayed relative to (a) because of the lower absorption fraction] and (c) the SMA-DPP SG3.3 with wavelength detuning run at t= 2.9 ns. Note the shell for (c) has not quite achieved maximum compression but was plotted for the distance traveled closest to (a). The results from these runs are tabulated in Table I.

Close modal

To estimate the impact of laser system imbalances on each CBET mitigation scheme, more simulations were run that include mistiming = 5-ps rms, mispointing = 10-μm rms, and a constant power imbalance = 10% relative to the nominal values, where random numbers are drawn from a normal (Gaussian) distribution. The random numbers drawn for each simulation are repeatably random; therefore the same set is used for the sake of a fair comparison. The resultant shell morphology is illustrated by plotting the mass density near peak compression (see Fig. 20). The results from these runs are tabulated in Table I; the values of which are within the brackets in the third and fourth rows. The SG4-R75 case has lower hot-spot pressure and yield, while the SMA90SG3.3 still achieves the >100-Gbar goal despite being lowered by the imposed laser imbalances.

FIG. 20.

Shell mass-density contour plots from DRACO simulations near peak compression plotted at similar distance traveled under nonideal conditions that include mistiming = 5-ps rms, mispointing = 10-μm rms, and a constant power imbalance = 10% relative to the nominal values, where random numbers are drawn from a normal (Gaussian) distribution. (a) The SG4-R75 run at t = 3.05 ns [delayed relative to (b) because of the lower absorption fraction] and (b) the SMA-DPP SG3.3 with wavelength detuning run at t = 2.9 ns. The results from these runs are tabulated in Table I; the values are in the brackets.

FIG. 20.

Shell mass-density contour plots from DRACO simulations near peak compression plotted at similar distance traveled under nonideal conditions that include mistiming = 5-ps rms, mispointing = 10-μm rms, and a constant power imbalance = 10% relative to the nominal values, where random numbers are drawn from a normal (Gaussian) distribution. (a) The SG4-R75 run at t = 3.05 ns [delayed relative to (b) because of the lower absorption fraction] and (b) the SMA-DPP SG3.3 with wavelength detuning run at t = 2.9 ns. The results from these runs are tabulated in Table I; the values are in the brackets.

Close modal

Draco has the ability to emulate the effect of the non-zero m-modes by projecting the effects of the various laser imbalances onto a hardsphere while next performing a spherical harmonic decomposition on-the-fly during the simulation, extracting the power of the non-zero m-modes and then subsequently imposing this power via an m = 0,ℓ-mode laser perturbation corresponding to every ℓ-mode supported by the mesh resolution. This is typically considered an over-estimation of the true non-zero m-mode effect on target performance because all the power is presented as a single mode instead of spread over many m-modes. The effect of non-zero m-modes on the target performance using this method is tabulated in Table I although not illustrated in the plots. An inspection of the tabular data suggests, at least as far as modeled in a 2-D code, that the detrimental effects on target performance due to laser imbalances are of the same level as the degradation due to the SG4-R75 spot-shape. For example, the neutron yield degrades by about a factor of 3× going from SMA90SG3.3 to SG4-R75 for each of the three simulations examples and the degradation is also ∼3× when comparing either design and applying laser imbalance with non-zero m-modes. The effects of the various laser imbalances may be larger due to the smaller surface area of true non-zero m-modes and their potential faster growth if the power was properly accounted for running a 3-D simulation; this comparison is planned for future research. The dominant detrimental effect of the small spot shape alone was sufficient to demonstrate the clear advantages of using SMA spot-shapes and wavelength detuning while attempting to account for laser imbalances. This is not to argue that laser imbalances are unimportant but rather to illustrate the effects of the smaller spot shape relative to laser imbalances. The lower performance using the non-zero m-modes suggests that an improved target and pulse shape design would be prudent, e.g., using a large pulse energy.

In conclusion, the first direct-drive wavelength-detuning CBET mitigation experiments on the NIF with a modest wavelength difference between crossing beams confirmed improved coupling as predicted by multidimensional hydrodynamic simulations. The improved coupling was inferred from the faster shell trajectory. A increased array of scattered-light detectors could be used on the NIF to measure improved absorption since the current number of detectors are too few for an accurate inference considering the amount of azimuthal variations. These direct-drive proof-of-principle experiments are the first such experiments and provide a path forward to recovering the energy loss caused by CBET. Simulations predict that as the wavelength separation increases (e.g., the ±6-Å UV predicted NIF full-energy limit14), the equatorial drive continues to improve and requires rebalancing to minimize ℓ = 2. Simulations also indicate that judicious use of all three colors with flexible color mapping in the fiber front end on the NIF produces better-balanced CBET mitigation designs in PDD.14,35,36 Simulations predict that symmetric direct drive on OMEGA will benefit from wavelength detuning since its three main driver legs already distribute adequately over the target. Additional CBET mitigation domains may be combined with wavelength detuning, e.g., optimized spot shapes that reduce the laser energy refracting over the horizon while maintaining optimal shape (SMA).21 Future experiments are planned to scope out the capabilities of wavelength-detuning CBET mitigation to further improve coupling and to address the asymmetry by proposing system changes to both OMEGA and the NIF: adding multiple wavelength sources to OMEGA, expanding the NIF's wavelength-detuning range, using SMA-DPP's, different wavelengths within NIF's quads, and remapping the NIF fiber front end to obviate cone swapping.

This material is based upon work supported by the Department of Energy National Nuclear Security Administration under Award No. DE-NA0001944, the University of Rochester, and the New York State Energy Research and Development Authority.

This report was prepared as an account of work sponsored by an agency of the U.S. Government. Neither the U.S. 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 U.S. Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the U.S. Government or any agency thereof.

The Mazinisin 3-D ray trace is part of the multiphysics radiation–hydrodynamics ALE 2-D code DRACO. Each physics package solves its own differential equations by employing the common split-step technique, where other physics effects are considered frozen or constant, while the current physics package is active during the discrete time step of the simulation. Typically, the result of one physics module builds a source term that drives a subsequent module, e.g., the ray trace stores the absorbed laser energy, which is then used to drive the thermal motion of the electrons and ions. Once all physics modules are completed, the time step is advanced and the process is repeated. The ray-trace package is highly scalable to run in parallel (with tested scalability to ∼1200 cores without any noticeable saturation) via the domain decomposition package Daashkaa,39 which is integrated into Mazinisin; this combination affords the CBET package Adaawam the necessary scalability and reduced memory footprint.

The Mazinisin package propagates refracting rays in 3-D within a 2-D plasma atmosphere and accounts for arbitrary multiple beam orientations, arbitrary far-field spot-shapes, independent pulse shapes, beam repointing, target offset, laser absorption, frequency chirp, speckle imprint, and smoothing by spectral dispersion (SSD)26 and for many nonuniform and random beam-to-beam variations such as pulse power imbalances, mistiming, or mispointing, as well as the 3-D CBET interaction effects (using Adaawam) with or without speckled beams. DRACO is able to emulate the effect of 3D modal power induced by laser imbalances by imposing any non-zero m-mode spectral power into a corresponding m = 0,ℓ-mode perturbation. Since the radiation-hydrodynamic code DRACO is only 2-D, the 3-D spatial feedback into the corona cannot be modeled; however, some useful 3-D scattering and deposition effects of refraction and CBET are captured within diagnostic output and can be correlated to experiment.

DRACO simulates ICF implosions of spherical targets while being a 2-D code. The logical mesh for a spherical object should reflect the symmetry of the object. The type of mesh that DRACO employs is an rz polar cylindrical mesh divided into quadrilateral computational cells or zones, where the orthogonal cylindrical coordinate system maps as Rx2+y2 and Z = z, where {x,y,z}3 define the 3-D Cartesian coordinates. The polar mesh in the R–Z plane is considered rotated 2π azimuthally, Φ, about the symmetry axis ẑ to form the complete initially spherical object, where the 3-D cylindrical coordinates are {R,Z,Φ}. Both the Cartesian and cylindrical 3-D coordinate systems are related to the orthogonal spherical coordinate system {r,θ,ϕ}, where angles θ and ϕ are referred to as the polar and azimuthal angles, respectively. The equivalences or transformations between the three 3-D coordinate systems are given by

(A1)
(A2)

and where the first two equations in Eq. (A2) imply

(A3)

where ϕ = tan−1(y/x) and θ = sin−1(R/r), but the ratios given in Eq. (A3) are more efficient if angles are not required.

The Lagrangian aspect of DRACO means that the underlying mesh flows with the fluid which, in general, alters the mesh from an orthogonal grid. This implies that ray tracing must be performed within a nonuniform grid precluding any simplifications or high-efficiency methods provided by regular orthogonal meshes. Regardless, a successful ray trace must be efficient and low noise because the laser is the energy source that ultimately drives all phenomenon in an ICF implosion, which is inherently an unstable hydrodynamic problem that amplifies perturbations whether they are intentional or not.

A ray trace can be divided into three broad classes of problems to solve that are equally and deceptively complex because of the myriad of issues that can occur on an irregular mesh; (1) ray initialization and entrance-cell ray entry, (2) refractive ray propagation with IBS and CBET, and (3) cell exit and cell-to-cell transfer. Each problem class contributes appreciably to defining a high-quality algorithm that generates low noise and high accuracy deposition. The second problem class differentiates itself as being the most time consuming but it is not the most important. Ignoring the importance of each class will lead to unsatisfactory results like noise-dominated shell morphology.

The first problem class involves selecting the initial ray position, direction, and energy as well as determining the cell-entry position. The choices made here significantly affect the noise induced in the simulation. Judiciously selecting the ray positions, for example, overwhelmingly impacts subsequent binning noise in the simulation. Before any results are discussed, the other two problem classes must be addressed because the results of the first problem class are inherently coupled to the next two problem classes.

The second problem class involves 3-D refractive ray tracing with CBET and IBS effects through an arbitrary quadrilateral mesh; currently, these are the only two LPI processes simulated in DRACO. The CBET effect does not alter the ray trajectory directly within a time step and does not come into play until after the ray traverses a computational cell because of its high computational costs. After the ray trajectory is formed the CBET coefficient and frequency chirp integrated along the path are determined, CBET is calculated as described in the second section [Eqs. (2)–(5)] using either a single or multiple path segments depending on the curvature. The calculation of IBS will be discussed alongside ray tracing. The discussion of frequency chirp is in  Appendix B. Mazinisin propagates rays in 3-D through the azimuthally symmetric 2-D atmosphere. Since the underlying physical quantities that govern the ray trace are defined on a 2-D RZ polar cylindrical mesh and the rays propagate in 3-D Cartesian coordinates a mapping to and from the cylindrical coordinates, is required to interpolate the physical quantities as a function of any ray's position vector r.

At the heart of any ray trace is the nonlinear partial-differential equation called the eikonal equation |S|2=nrefr2 (Ref. 40), from which one can derive the ray equation, viz.,

(A4)
(A5)

where r is the position vector and the differentials are taken with respect to the elemental propagation path length ds. The ray equation may be rewritten as

(A6)

where the two vectors are identified as the unit tangential vector

(A7)

and the curvature vector of the ray Kd2r/ds2, while the directional derivative of the refractive index is given by dnrefr/dst̂nrefr (Ref. 41). The unit tangential vector t̂ is also known as the propagation vector since it defines the propagation direction. The curvature vector is related to the propagation vector as

(A8)

and consequently obeys Kt̂=0. Solving Eq. (A6) for the curvature vector yields the exact expression

(A9)
(A10)

where nrefr1ne (Ref. 42), nrefr, and ne are all functions of the position vector r. In general, the solution to Eq. (A6) is analytically intractable, but numerical solutions are possible.

For example, through linearization of the position vector and tangential vector via a truncated Taylor series expansions about the current position and propagation vectors r0 and t̂0 over a small step size ds, viz.,

(A11)
(A12)

where the unit tangential vector t̂ defines the propagation direction. A first-order (in error) Euler quadrature algorithm (Refs. 41 and 43) is defined by evaluating Eq. (A10) at the current position and propagation vectors r0 and t̂0 to obtain an updated curvature vector K, which are then substituted into Eqs. (A11) and (A12) for a small step size ds to obtain the next position and propagation vectors. The error is controlled by setting the step size ds using |K|, viz.,

(A13)

where ϑcomp is a user-defined compression factor that controls the accuracy, Lchar defines the characteristic size of the computational cell, and ε is the machine precision. The first-order method is adequate for regions of low curvature, i.e., low |K| out in the corona ne0.2, because to get reasonable accuracy in high-curvature regions, the compression parameter ϑcomp must be set so small that it is computationally more efficient to use adaptive integration schemes. In addition, the first-order method can have difficulty turning near critical, i.e., regions of very high curvature, which leads to deposition noise and rays hitting the critical surface even when using small ϑcomp values.

Higher-order (in error) adaptive quadrature methods resolve many issues relating to accuracy, efficiency, and high curvature. In addition, it becomes straightforward to include adaptive IBS integration, further increasing accuracy and efficiency consistently with ray trajectory integration. Any ordinary differential equation (ODE) can be reduced to a set of coupled first-order differential equations and written generically as dyn/ds=fn(s,y1,,yN),n{1..N}, where the functions fn(s,y1,,yN) define the known derivatives to be evaluated along the integration path. The coupled set of first-order differential equations defined for the ray trace are given by the seven-element variable vector

(A14)

and the corresponding seven-element derivatives vector

(A15)

where ξ is the optional IBS absorption kernel and is defined as42,44

(A16)

and arranged with all constants toward the left and spatial variables toward the right, where kB is the Boltzmann constant and lnΛe is the spatially dependent electron Coulomb logarithm. Equations (A14)–(A16) form the basis of high-order adaptive quadrature methods.

In an adaptive quadrature method, an ODE driver routine given in Ref. 43 interfaces with an adaptive ODE stepper routine to solve Eq. (A14) by evaluating the derivatives from Eq. (A15) from a fourth- or fifth-order algorithm (such as the Cash–Karp, Runge–Kutta method43). The adaptive stepper adjusts the steps over an integration interval [s1,s2]={sn;s1sns2}, where snsn1+ds defines a variably spaced sequence taking large ds steps over low gradient terrain and iteratively refined small ds steps as the gradients increase in order to maintain a specified level of accuracy. The result is an efficient algorithm that takes the smallest number of steps and, consequently, function evaluations while maintaining a specified accuracy. This is especially important in high curvature regions where the ray trajectory is making the tightest turns and in this region depositing the highest energy fractions. Low-order methods, e.g., parabolic trajectories or the Euler quadrature, cannot prevent hitting the critical surface or making deposition errors because the gradients are not constant over a computational cell, and in high-gradient regions, a low-order method requires exceedingly small step sizes, making it impractical. However, using adaptive quadrature succeeds in this regard because this method adapts to spatially varying gradients, which leads to more-accurate trajectories and deposition while ensuring that rays never hit the critical surface (except for rare degenerate normal incidence cases, which occur only for one ray per beam in a radial plasma); see Fig. 21.

FIG. 21.

Radially integrated deposition patterns at 450 ps comparing ray-trace numerical integration models using the basic inverse projection. (a) The first-order Euler method described in Eqs. (A11) and (A12), (b) the adaptive fifth-order Runge–Kutta described in Eqs. (A14)–(A16). Both methods used Eq. (A13) to select the major step-size through the cells. The plots have the same vertical range but are centered on different values because of the overall deposition errors incurred in (a). This error could be reduced by requesting even finer step size resolution, whereas option (b) does not have these issues because the step-size is adaptively chosen and the integrator is of higher order. More importantly, the binning noise at the beam centers is more severe. The residual binning errors are remedied by using adaptive surrogate surface selection to account for refraction.

FIG. 21.

Radially integrated deposition patterns at 450 ps comparing ray-trace numerical integration models using the basic inverse projection. (a) The first-order Euler method described in Eqs. (A11) and (A12), (b) the adaptive fifth-order Runge–Kutta described in Eqs. (A14)–(A16). Both methods used Eq. (A13) to select the major step-size through the cells. The plots have the same vertical range but are centered on different values because of the overall deposition errors incurred in (a). This error could be reduced by requesting even finer step size resolution, whereas option (b) does not have these issues because the step-size is adaptively chosen and the integrator is of higher order. More importantly, the binning noise at the beam centers is more severe. The residual binning errors are remedied by using adaptive surrogate surface selection to account for refraction.

Close modal

The ODE driver routine requires modification to handle situations found in ICF implosions and the distorted meshes resulting in ALE hydrocodes. The driver must be able to call different types of stepper routines such as the Runge–Kutta or Bulirsch–Stoer methods43 to take advantage of their different strengths. (DRACO offers both user control as well as an automatic selection capability.) The routine also needs flexibility to be called for many types of integration, i.e., ray trajectory, absorption, or any other ODE equations appropriate to the ray trace. The driver also must be able to temporarily guide a ray past critical while it converges to the correct trajectory that avoids critical. Going past the critical density is allowed by using appropriate asymptotic limits that avoid triggering a stiff-equation response, i.e., using the machine precision ε in refractive index formula, which smoothly goes to this limit. Another adaptation allows the algorithm to dynamically decide the most-efficient integration order to apply, which boosts ray-tracing efficiency while not sacrificing accuracy. The most-important adaptation is limiting the overall path length when crossing the cell-edge boundary or up to any singularities such as the critical surface.

Numerical integration of Eq. (A6) requires repeated interpolation of the electron density and its gradient along the ray trajectory when the underlying functional form is not known. Likewise integrating Eq. (A15) requires continuous interpolation of the kernel Eq. (A16) and not its individual components because this leads to scalloping when different interpolated functions are multiplied. Cell-centered quantities are first interpolated onto the cell vertices using volumetric mass-weighted averaging. The plasmas being modeled are naturally continuous and likewise the interpolator should mimic this continuity and not lead to unphysical jump conditions across cell edges. Isoparametric mapping within the quadrilateral cells satisfies this desirable and essential condition and is employed in Mazinisin. The basic isoparametric mapping can be extended to include radial dependence other than linear, e.g., an exponential or 1/r2 radial decay of the electron density can be employed to provide larger cells with an improved accuracy that reflects the assumed radial functional form.

Higher-order (referring to the expansions and not the quadrature type) methods can be used to solve Eq. (A6), but linear methods tend to be superior because they involve many simple calculations that can achieve similar accuracy but in less time because of the computational overhead of the higher-order methods. The seemingly mathematical elegance of higher-order methods is outweighed by the computational complexity. Higher-order trajectories intersecting with nonlinear boundaries imply high-order roots, which increases solution complexity and computational overhead for no real gain when lower-order [i.e., Eq. (A15)] suffice and yield high accuracy and high efficiency. In addition, using a paucity of rays in lieu of high density of rays may seem like an attractive option at first, however, it suffers by trading off efficiency of propagating of many simple ray trajectories with the consequential inaccurate volume interpolation and complicated intersection calculations for deposition especially where it matters the most in ICF implosions, i.e., near critical.

The third problem class involves cell exit and cell-to-cell transfer. Cell-edge crossing detection uses a metric or “ruler” to measure the inside/outside condition at a cell's edge; the same metric is applied from any cell and during propagation and this consistency rules out machine or metric precision errors. Since neighboring cells share the same cell edge, the same detection response for a given cell face is guaranteed regardless of the cell sharing the cell face. Even if the metric has slight numerical errors, e.g., because of low accuracy or machine precision, it will not matter with regard to cell-edge crossing detection since the answer is absolute and invariant of which cell queries the inside/outside condition. This holds true because each pair of neighboring cells share the same interface and its accompanying functional form and therefore the same answer is absolute and guaranteed regardless of any small errors inherent in the cell-edge detection metric. The only possible side effect of numerical errors in the cell-edge detection is slight deposition errors but this is controllable and vanishingly small to machine precision by appropriately setting the detection tolerance. Cell-edge detection uses a root-polishing/root-finding algorithm (Brent's method43) that has been modified to bias the resulting position to be just outside the current cell; this technique is powerful and efficient. Use of bracketed roots guarantees the solution of finding a ray's exit; the side benefit is that rays are never lost or fail to propagate. There does exist, however, the possible trivial root where the chosen ds [Eq. (A13)] lands exactly on the cell's edge, which can be biased to prevent this situation at an additional computational cost. The trivial root eventually happens (albeit rarely) when propagating trillions of rays over thousands of time steps in thousands of simulations; the rare trivial root is solved by nudging the ray by a small step to force a positive cell-edge crossing condition. The adaptive root finder allows the hyper refinement of the step size to bring the ray arbitrary close to the cell's edge while ensuring cell crossing; the side effect is reduced energy binning noise. Another anomalous situation that can occur for unphysical meshes causes a ray to be infinitely guided internally between two density peaks without losing energy; unphysical mesh detection is necessary to prevent this unphysical situation. Cell-to-cell transfer is somewhat trivial since the mesh topology is known and each neighbor is known across the cell edge. The possible transfer across a cell corner is robustly handled automatically without the need for complicated corner detection logic. For instance, if the transfer into the edge-based neighbor is done and the corner crossing has occurred, this next cell would immediately trigger a cell crossing, which would immediately transfer the ray into its cell-edge neighbor, effectively transferring the ray across a cell corner. Removing unnecessary and complicated logic like the corner-crossing detection makes Mazinisin's algorithm robust and efficient. This also implies that rays are never lost since cell exit is guaranteed.

Now that the three problem classes have been outlined, the results of the adaptive inverse projection algorithm can be discussed. The adaptive inverse projection algorithm is used to initialize ray starting points, which encourage maximal deposition into the intended cells, greatly reducing binning noise see (Fig. 22). The method adapts to shell distortions as well as refraction to minimize binning noise. Compared to a random distribution with the same number of rays chosen randomly each ray-trace step, the different approaches have very different noise backgrounds. The random distribution technique relies on the incoherent addition of the noise patterns from different time steps that ideally dampens the noise response below the applied signal. In practice, however, to achieve the same level of noise reduction, a random method would have to increase the ray density many fold per time step (or equivalently increasing the number of ray-trace steps) because the insufficient smoothing obtained by averaging random distributions has diminishing returns. The use of an inverse projection algorithm achieves low noise through the judicious choice of launch positions that tend to deposit the bulk of energy in the intended cells, thereby minimizing binning noise.

FIG. 22.

Comparison of two DRACO simulations of the same target past cold start at 50 ps both using 1-M rays per time step showing the radial sum of the deposition pattern at 50 ps. (a) Two-dimensional random distribution of rays within each beam with a new realization every time step; and (b) using the inverse projection algorithm. Both methods use the same ray-path and absorption integrators. The only difference is the choice of ray distribution.

FIG. 22.

Comparison of two DRACO simulations of the same target past cold start at 50 ps both using 1-M rays per time step showing the radial sum of the deposition pattern at 50 ps. (a) Two-dimensional random distribution of rays within each beam with a new realization every time step; and (b) using the inverse projection algorithm. Both methods use the same ray-path and absorption integrators. The only difference is the choice of ray distribution.

Close modal

The basic inverse projection algorithm determines the ray launch positions for each beam port starting at cell positions specified within the 3-D plasma mesh, as position vectors rp3. The beam port is defined as a position/direction vector P=RpP̂, where Rp is a radius just outside the maximum extent of the plasma and the beam-port unit vector

(A17)

which points toward the beam port's central axis and is defined with the beam port's spherical coordinate angles {θPrt,ϕPrt}. A straight-line back projection along the trajectory given by the beam-port direction vector maps these points into the far-field plane of each beam port [see Fig. 23(a)], viz.,

(A18)

where rff3 under the condition that 0<N<Rp, where the range restrictions N > 0 define the visibility of rp from the beam port P and N<Rp results from the far-field plane's definition. A far-field transformation maps the position vector rff into the local 2-D Cartesian coordinates of the beam port's far-field plane, viz.,

(A19)

where the far-field transformation matrix is given by

(A20)

and rFF2, by construction, since the transformation maps into a plane. The mapped far-field coordinate determines the potential ray's intensity by interpolating this position into the beam's intensity pattern. The launch positions in the far-field plane form concentric ellipses [see Fig. 23(b)]. Each starting cell position in the plasma has an associated elemental surface area tiled together with all the neighboring points. The dot product of the beam-port unit vector and the surface area outward normal determines the points' visibility (when positive definite), the projected area onto the far-field plane, and thereby the power to be given to the ray by multiplying the visible ray's area and intensity. The starting positions in the 3-D plasma form a closed and connected surface subtending 4π steradian that projects a continuous and connected surface area onto the far-field plane; e.g., if the surface is spherical with radius rs, this maps onto a πrs2 area in the far-field plane. The first inverse-projection surface is typically chosen near critical since this region requires the lowest noise because it gives rise to the highest hydroefficiency, i.e., noise near critical adversely impacts target imprint more than noise far out in the corona. Once the plasma has expanded, the first surface provides insufficient coverage of the plasma back-projected onto the far-field plane. Other inverse projection surfaces are then selected to provide complete coverage of the full plasma extent. When the next subsequent surface area Anext is back-projected into the far-field plane, only the relative complement set (Anext/Aprev, where Aprev is the previous back-projected area) is allowed to propagate.

FIG. 23.

The basic inverse projection algorithm determines the ray launch positions. Starting with cell positions within (a) the 3-D target, these points are back-projected into (b) the far-field plane along the trajectory defined by the beam-port angle. When the beam-port is not aligned with ẑ, the launch positions in the far-field plane form concentric ellipses.

FIG. 23.

The basic inverse projection algorithm determines the ray launch positions. Starting with cell positions within (a) the 3-D target, these points are back-projected into (b) the far-field plane along the trajectory defined by the beam-port angle. When the beam-port is not aligned with ẑ, the launch positions in the far-field plane form concentric ellipses.

Close modal

As the plasma expands, the importance of refraction begins to play a role. Refraction causes the ray intended to deposit in one cell to deposit in a neighboring cell, leading to binning noise. An algorithm that adapts to the increased refraction effect becomes necessary to reduce noise. If the projected ray intensity (onto the inverse projection surface) is accumulated and plotted by impact parameter (dot product of the ray direction and a sphere's outward normal), the distribution peaks for a particular impact parameter. These rays become the most-significant players that determine the extent of binning noise. Minimizing the binning noise for these rays will minimize the noise for the total deposition. With this in mind, an adaptive algorithm optimizes a search for a surrogate inverse-projection surface that will minimize binning noise. The optimal surface is found when these rays are back-projected from this surrogate inverse projection surface and are then turned in the intended cell when propagated with refraction (see Fig. 24). Since the greatest energy is deposited where a ray turns, binning noise is minimized for the rays with the impact parameter corresponding to the peak of the distribution. The remaining rays surrounding the peak also have their binning noise reduced when using the surrogate inverse projection surface. The surrogate surface lies interior to the intended inverse projection surface. This adaptive algorithm works well for 2-D hydrocodes and optimizes the surrogate surfaces as the plasma evolves. The results of the adaptive inverse projection algorithm are compared to the basic algorithm (see Fig. 25). The binning noise is noticeable in Fig. 25(a) early in time, cf. to Fig. 25(b), and this excessive noise imprints in the shell and develops mass perturbations that persist late in time. These mass-density perturbations are remedied using the adaptive algorithm. Improved algorithms are being developed that account for refractive back-propagation that will further reduce binning noise and apply to 3-D hydrocodes.

FIG. 24.

The adaptive inverse projection algorithm attempts to compensate for refraction by finding a surrogate surface that is interior to the intended surface (in this example, the critical surface). The surrogate surface is found by optimizing the deposition for a particular impact parameter that defines the maximum distribution for rays. The rays are allowed to refract, and when the optimal surface is found, the ray will turn in the intended cell and deposit the greatest energy. When all rays, regardless of impact parameter, are back-projected from the surrogate surface, the overall binning noise is reduced.

FIG. 24.

The adaptive inverse projection algorithm attempts to compensate for refraction by finding a surrogate surface that is interior to the intended surface (in this example, the critical surface). The surrogate surface is found by optimizing the deposition for a particular impact parameter that defines the maximum distribution for rays. The rays are allowed to refract, and when the optimal surface is found, the ray will turn in the intended cell and deposit the greatest energy. When all rays, regardless of impact parameter, are back-projected from the surrogate surface, the overall binning noise is reduced.

Close modal
FIG. 25.

Radially integrated deposition patterns at 50 ps comparing binning noise where the primary inverse projection surface is the critical surface. (a) The basic inverse projection algorithm is used without locating the surrogate surface; (b) the surrogate surface is adaptively chosen to account for refraction of the rays. Both methods use the same number of rays and adaptive fifth-order ray-trace integration. Binning noise is minimized, especially near the beam port centers, when the inverse projection surface is adaptively selected.

FIG. 25.

Radially integrated deposition patterns at 50 ps comparing binning noise where the primary inverse projection surface is the critical surface. (a) The basic inverse projection algorithm is used without locating the surrogate surface; (b) the surrogate surface is adaptively chosen to account for refraction of the rays. Both methods use the same number of rays and adaptive fifth-order ray-trace integration. Binning noise is minimized, especially near the beam port centers, when the inverse projection surface is adaptively selected.

Close modal

The severity of residual noise induced from a random ray distribution versus the adaptive inverse projection algorithm is demonstrated using imprint simulations. Imprint is modeled in a NIF-scale 1.5-MJ ignition target covering ℓ modes 2:2:200 on a half-sphere using 750-k rays per ray-trace step. Both simulations use the same adaptive fifth-order Runge–Kutta quadrature. The simulations show dramatically different noise characteristics (see Fig. 26). The 2-D random distribution (purple) noise floor overwhelms the modal cutoff at ℓ = 200. A low-noise simulation should show a distinct modal cutoff past the applied signal modes (ℓ = 200 in this case). Noise absolutely corrupts any simulation or other deposition code that cannot show the distinct modal cutoff past the applied signal, and the results cannot be trusted to represent mode growth of the applied signal. In addition, simulations or other deposition codes that apply imprint signal past that supported by the mesh resolution (e.g., at least 10–12 zones per smallest wavelength) tend to compound the numerical noise by confusing noise with signal that cannot be properly supported by the mesh resolution and mask a noisy algorithm by inaccurately interpreting it as true signal. Only the adaptive inverse projection algorithm shows high fidelity where the noise floor is below the applied mode signal. The noise from the 2-D random distribution also affects applied ℓ modes 2:2:200 as compared to the adaptive inverse projection algorithm (red). The random ray distribution has insufficient noise reduction ability, cf. the adaptive inverse projection algorithm.

FIG. 26.

Shell-perturbation spectra of an NIF-scale 1.5-MJ ignition target comparison for imprint simulations including ℓ modes= 2:200 at 3.7 ns using different ray initializations and 750-k rays per time step. The 2-D random distribution (purple) noise floor overwhelms the modal cutoff at ℓ= 200. The noise from the 2-D random distribution also affects applied modes ℓ =2:2:200 as compared to the adaptive inverse projection algorithm (red). Both simulations used the same number of rays; 750 k rays per ray-trace step and same adaptive fifth-order ray-trace integrator.

FIG. 26.

Shell-perturbation spectra of an NIF-scale 1.5-MJ ignition target comparison for imprint simulations including ℓ modes= 2:200 at 3.7 ns using different ray initializations and 750-k rays per time step. The 2-D random distribution (purple) noise floor overwhelms the modal cutoff at ℓ= 200. The noise from the 2-D random distribution also affects applied modes ℓ =2:2:200 as compared to the adaptive inverse projection algorithm (red). Both simulations used the same number of rays; 750 k rays per ray-trace step and same adaptive fifth-order ray-trace integrator.

Close modal

Another side benefit of using the adaptive inverse projection algorithm is that the expected pristine deposition patterns can be used to diagnose ray-trace problems. Ray-trace codes that use a random distribution induce large of peak-to-peak noise on each time step, making it impossible to determine ray-tracing artifacts resulting from algorithmic problems or assumptions; any potential problems remain hidden, masked and overwhelmed by the noisy deposition. Ray-trace problems are noticed only when a smooth deposition is expected, e.g., via inverse projection, or after it is too late and noise has imprinted in the shell.

The CBET effect is altered by the relative wavelength or angular frequency of the interacting rays, as specified by the wave-vector–matching condition η given in Eq. (4). Knowing the instantaneous wavelength or angular frequency of the interacting rays is an important aspect of CBET simulations. In addition, understanding the scattered-light spectral features aids diagnoses of the experiments with CBET mitigation schemes. Streaked scattered-light measurements record the spectral history of the CBET interaction, and various spectral features signify its activity (see Fig. 6). This appendix describes the Doppler shift in expanding plasmas and explains some observed spectral features.

A light-based ray's vacuum wavelength relates to the angular frequency ω in the usual way, viz.,

(B1)

The refractive index within a plasma, given by42 

(B2)

nonlinearly depends on the normalized electron density ne, both of which are functions of space and time. A field propagating through a medium with a time-varying refractive index becomes temporally phase modulated over the path length L, which alters the characteristic angular frequency, viz.,

(B3)

(a generalization of the conventional Doppler shift30). The expanding plasma dynamically alters the instantaneous refractive index in space via the electron density Eq. (B2), and thereby the angular frequency Eq. (B3) and wavelength Eq. (B1). This phase-modulation process is independent of the ray-propagation direction.

Taking the temporal derivative of Eq. (B2), a differential equation relating the temporal derivatives of the refractive index nrefr and the electron density ne is found, viz.,

(B4)

where the sign determines the shift, e.g., negative is a blue shift and positive is a red shift. An instantaneous wavelength change is commonly referred to as a “chirp” because the angular frequency and wavelength vary as functions of time similar to a bird's song. The frequency chirp is independent of propagation direction, i.e., an inbound ray suffers the same chirp as an outbound ray traversing the same spatial region. The frequency chirp's sign depends not only on an apparent receding or advancing surface as in the common Doppler shift but depends additionally on the sign of the time-varying ne. For example, a receding negative electron density gradient generates a red shift, while a receding positive electron density gradient generates a blue shift; in this way the Doppler shift given in Eq. (B4) is not analogous to the common Doppler shift but they both alter the apparent wavelength. The effect given in Eq. (B4) is best thought of as optical phase modulation to avoid misconception.

Invoking the equation of continuity forms the solution to the temporal derivative of the electron density. In a simulation, the cell-centered value is calculated by solving the volume average of the temporal derivative using the divergence theorem, viz.,

(B5)

where vol is the averaging volume (typically the computational cell volume) whose bounding surface area is A; vfluid is the fluid velocity; and n̂ is the outward-pointing normal on the boundary surface. The net result of Eq. (B5) demonstrates that the frequency chirp or optical phase modulation suffered by a ray within a cell is solely determined by the volume-averaged temporal derivative of the electron density and not the propagation direction or the direction of any moving surfaces.

OMEGA shot 60000 is used to demonstrate the effects of chirp and electron transport models. The schematic of the target and the pulse used in the DRACO simulations are shown in Fig. 27. If the plasma flow is radial, a graphical interpretation of Eq. (B5) is illustrated in Fig. 28. The balance of the electron density and fluid velocity gradients determine the sign of the frequency chirp as a ray traverses the corona (see Fig. 29). A pulse shape with narrow pickets and a sustained main pulse initiate a predictable sequence of Doppler-shift events as outlined in Fig. 30. Figure 31 (online only) shows a movie describing the sequence coordinated with Fig. 30.

FIG. 27.

OMEGA shot 60000 (a) target schematic and (b) pulse shape. This shot is used to illustrate the effects of electron transport model, CBET, and Doppler shifts.

FIG. 27.

OMEGA shot 60000 (a) target schematic and (b) pulse shape. This shot is used to illustrate the effects of electron transport model, CBET, and Doppler shifts.

Close modal
FIG. 28.

If the plasma expansion fluid flow is radial, vfluid=νfluidr̂, this special case demonstrates the important aspects of using the equation of continuity to solving the volume-average temporal derivative of the normalized electron density, viz., Eq. (B5). The numbered subscripts indicate the index of the boundary surface A, where there are only two contributing surfaces because of the dot product of n̂vfluid..

FIG. 28.

If the plasma expansion fluid flow is radial, vfluid=νfluidr̂, this special case demonstrates the important aspects of using the equation of continuity to solving the volume-average temporal derivative of the normalized electron density, viz., Eq. (B5). The numbered subscripts indicate the index of the boundary surface A, where there are only two contributing surfaces because of the dot product of n̂vfluid..

Close modal
FIG. 29.

The sensitivity of the electron density and fluid velocity gradients. Near the critical surface, the positive radial gradient of velocity * area product overwhelms the negative electron density gradient, which results in dne/dt<0,dnref/dt>0, which generates a red shift. Farther out in the corona, the negative electron density gradient overwhelms the decreased positive radial gradient of velocity * area product, which results in dne/dt>0,dnref/dt<0, which generates a blue shift. The ultimate shift suffered by the ray will depend on the strength of each region and the path length through each region.

FIG. 29.

The sensitivity of the electron density and fluid velocity gradients. Near the critical surface, the positive radial gradient of velocity * area product overwhelms the negative electron density gradient, which results in dne/dt<0,dnref/dt>0, which generates a red shift. Farther out in the corona, the negative electron density gradient overwhelms the decreased positive radial gradient of velocity * area product, which results in dne/dt>0,dnref/dt<0, which generates a blue shift. The ultimate shift suffered by the ray will depend on the strength of each region and the path length through each region.

Close modal
FIG. 30.

The rising edge of a pulse initiates a predicable sequence of (a) three Doppler-shift events that correlate to the (b) pulse and absorption fraction plot. (1) A red shiftdne/dt<0 occurs within a small volume near the critical density. (2) Next, a blue shiftdne/dt>0 develops within this volume and generates a diminishing outward-propagating wave. The blue shift is not related to the falling edge but is caused by the rising edge. (3) (optional) A sustained pulse (e.g., a long main pulse) intensifies the red-shift region, dne/dt<0 as the negative density gradient moves inward. The volume expands and overwhelms the diminishing outward-propagating blue-shift region. As time progresses, the red-shift region expands, which eventually overwhelms the outer blue-shift region.

FIG. 30.

The rising edge of a pulse initiates a predicable sequence of (a) three Doppler-shift events that correlate to the (b) pulse and absorption fraction plot. (1) A red shiftdne/dt<0 occurs within a small volume near the critical density. (2) Next, a blue shiftdne/dt>0 develops within this volume and generates a diminishing outward-propagating wave. The blue shift is not related to the falling edge but is caused by the rising edge. (3) (optional) A sustained pulse (e.g., a long main pulse) intensifies the red-shift region, dne/dt<0 as the negative density gradient moves inward. The volume expands and overwhelms the diminishing outward-propagating blue-shift region. As time progresses, the red-shift region expands, which eventually overwhelms the outer blue-shift region.

Close modal
FIG. 31.

Movie, coordinated with Fig. 30, shows the progression of the chirp through the corona showing all three phases of development. Multimedia view: https://doi.org/10.1063/1.5022181.1

FIG. 31.

Movie, coordinated with Fig. 30, shows the progression of the chirp through the corona showing all three phases of development. Multimedia view: https://doi.org/10.1063/1.5022181.1

Close modal

The dispersed frequency modulated bandwidth applied on the Omega laser to smooth the transverse far-field speckle pattern is smaller than the observed chirp but it can be of similar magnitude, e.g., 2Å,UV on OMEGA, which forms a discrete set of spectral modes and referred to as smoothing by spectral dispersion (SSD). The current levels of SSD on the NIF are much lower, 0.4Å,UV. However, larger bandwidth mFM-SSD48 drivers are planned, but the bandwidth will be degraded during the main drive when CBET is most active. Currently, DRACO does not model the wavelength modal spectrum of SSD in connection with CBET; it does model speckle imprint and smoothing. Future modifications will incorporate the simulation of the dynamic modal spectrum of SSD using a similar method as for wavelength detuning, in that, it will model the effects of wavelength separation of the discrete SSD spectrum in conjunction with any wavelength detuning during CBET modeling but not any dampening related to coherence-disruption of LPI growth.

The simulations of shot 60000 are also used to illustrate the combined effect of CBET and the choice of electron transport model (see Fig. 32). The effect of flux-limited (f/0.06) electron transport45 is shown in Fig. 32(a). The effect of the nonlocal model based on the Schurtz–Nicolaï–Busquet method10 is shown in Fig. 32(b). The role of CBET does not change the overall shape of the electron density, but rather it changes the amplitude of its underlying functional form. The electron transport model is responsible for the gross changes in functional form and not the CBET effect, as is sometimes misunderstood.

FIG. 32.

Combined effect of CBET and electron thermal conduction model; flux-limited electron transport versus the iSNB nonlocal model.10 Simulations of OMEGA shot 60000 with DRACO illustrate the combined effect of CBET and the electron transport models on target dynamics. The CBET FWHM gain region is indicated by the red rectangular region. A similar amount of energy is deposited for both models. The difference lies in how the shell is driven and the resulting electron density gradients. The CBET effect reduces absorption at the radii interior to the interaction region ne1/4, where the rays would have deposited their energy without significantly affecting the larger radii. (a) The flux-limited (f/0.06) electron transport limits the drive on the target and results in steeper electron density gradients for the same amount of absorption. (b) The iSNB model drives the target harder because of the nonlocal effect and widens both the deposition and CBET regions as a result.

FIG. 32.

Combined effect of CBET and electron thermal conduction model; flux-limited electron transport versus the iSNB nonlocal model.10 Simulations of OMEGA shot 60000 with DRACO illustrate the combined effect of CBET and the electron transport models on target dynamics. The CBET FWHM gain region is indicated by the red rectangular region. A similar amount of energy is deposited for both models. The difference lies in how the shell is driven and the resulting electron density gradients. The CBET effect reduces absorption at the radii interior to the interaction region ne1/4, where the rays would have deposited their energy without significantly affecting the larger radii. (a) The flux-limited (f/0.06) electron transport limits the drive on the target and results in steeper electron density gradients for the same amount of absorption. (b) The iSNB model drives the target harder because of the nonlocal effect and widens both the deposition and CBET regions as a result.

Close modal
1.
M. C.
Herrmann
,
M.
Tabak
, and
J. D.
Lindl
,
Nucl. Fusion
41
,
99
(
2001
).
2.
C. J.
Randall
,
J. R.
Albritton
, and
J. J.
Thomson
,
Phys. Fluids
24
,
1474
(
1981
).
3.
T. R.
Boehly
,
D. L.
Brown
,
R. S.
Craxton
,
R. L.
Keck
,
J. P.
Knauer
,
J. H.
Kelly
,
T. J.
Kessler
,
S. A.
Kumpan
,
S. J.
Loucks
,
S. A.
Letzring
,
F. J.
Marshall
,
R. L.
McCrory
,
S. F. B.
Morse
,
W.
Seka
,
J. M.
Soures
, and
C. P.
Verdon
,
Opt. Commun.
133
,
495
(
1997
).
4.
E. M.
Campbell
and
W. J.
Hogan
,
Plasma Phys. Controlled Fusion
41
,
B39
(
1999
).
5.
V. N.
Goncharov
,
T. C.
Sangster
,
R.
Betti
,
T. R.
Boehly
,
M. J.
Bonino
,
T. J. B.
Collins
,
R. S.
Craxton
,
J. A.
Delettrez
,
D. H.
Edgell
,
R.
Epstein
,
R. K.
Follet
,
C. J.
Forrest
,
D. H.
Froula
,
V. Yu.
Glebov
,
D. R.
Harding
,
R. J.
Henchen
,
S. X.
Hu
,
I. V.
Igumenshchev
,
R.
Janezic
,
J. H.
Kelly
,
T. J.
Kessler
,
T. Z.
Kosc
,
S. J.
Loucks
,
J. A.
Marozas
,
F. J.
Marshall
,
A. V.
Maximov
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
D. T.
Michel
,
J. F.
Myatt
,
R.
Nora
,
P. B.
Radha
,
S. P.
Regan
,
W.
Seka
,
W. T.
Shmayda
,
R. W.
Short
,
A.
Shvydky
,
S.
Skupsky
,
C.
Stoeckl
,
B.
Yaakobi
,
J. A.
Frenje
,
M.
Gatu-Johnson
,
R. D.
Petrasso
, and
D. T.
Casey
,
Phys. Plasmas
21
,
056315
(
2014
).
6.
C. J.
McKinstrie
,
J. S.
Li
,
R. E.
Giacone
, and
H. X.
Vu
,
Phys. Plasmas
3
,
2686
(
1996
).
7.
C. J.
McKinstrie
,
A. V.
Kanaev
,
V. T.
Tikhonchuk
,
R. E.
Giacone
, and
H. X.
Vu
,
Phys. Plasmas
5
,
1142
(
1998
).
8.
P. B.
Radha
,
M.
Hohenberger
,
D. H.
Edgell
,
J. A.
Marozas
,
F. J.
Marshall
,
D. T.
Michel
,
M. J.
Rosenberg
,
W.
Seka
,
A.
Shvydky
,
T. R.
Boehly
,
T. J. B.
Collins
,
E. M.
Campbell
,
R. S.
Craxton
,
J. A.
Delettrez
,
S. N.
Dixit
,
J. A.
Frenje
,
D. H.
Froula
,
V. N.
Goncharov
,
S. X.
Hu
,
J. P.
Knauer
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
J.
Moody
,
J. F.
Myatt
,
R. D.
Petrasso
,
S. P.
Regan
,
T. C.
Sangster
,
H.
Sio
,
S.
Skupsky
, and
A.
Zylstra
,
Phys. Plasmas
23
,
056305
(
2016
).
9.
M.
Hohenberger
,
P. B.
Radha
,
J. F.
Myatt
,
S.
LePape
,
J. A.
Marozas
,
F. J.
Marshall
,
D. T.
Michel
,
S. P.
Regan
,
W.
Seka
,
A.
Shvydky
,
T. C.
Sangster
,
J. W.
Bates
,
R.
Betti
,
T. R.
Boehly
,
M. J.
Bonino
,
D. T.
Casey
,
T. J. B.
Collins
,
R. S.
Craxton
,
J. A.
Delettrez
,
D. H.
Edgell
,
R.
Epstein
,
G.
Fiksel
,
P.
Fitzsimmons
,
J. A.
Frenje
,
D. H.
Froula
,
V. N.
Goncharov
,
D. R.
Harding
,
D. H.
Kalantar
,
M.
Karasik
,
T. J.
Kessler
,
J. D.
Kilkenny
,
J. P.
Knauer
,
C.
Kurz
,
M.
Lafon
,
K. N.
LaFortune
,
B. J.
MacGowan
,
A. J.
Mackinnon
,
A. G.
MacPhee
,
R. L.
McCrory
,
P. W.
McKenty
,
J. F.
Meeker
,
D. D.
Meyerhofer
,
S. R.
Nagel
,
A.
Nikroo
,
S.
Obenschain
,
R. D.
Petrasso
,
J. E.
Ralph
,
H. G.
Rinderknecht
,
M. J.
Rosenberg
,
A. J.
Schmitt
,
R. J.
Wallace
,
J.
Weaver
,
C.
Widmayer
,
S.
Skupsky
,
A. A.
Solodov
,
C.
Stoeckl
,
B.
Yaakobi
, and
J. D.
Zuegel
,
Phys. Plasmas
22
,
056308
(
2015
).
10.
D.
Cao
,
G.
Moses
, and
J.
Delettrez
,
Phys. Plasmas
22
,
082308
(
2015
).
11.
J.
Delettrez
,
R.
Epstein
,
M. C.
Richardson
,
P. A.
Jaanimagi
, and
B. L.
Henke
,
Phys. Rev. A
36
,
3926
(
1987
).
12.
P. B.
Radha
,
V. N.
Goncharov
,
T. J. B.
Collins
,
J. A.
Delettrez
,
Y.
Elbaz
,
V. Y.
Glebov
,
R. L.
Keck
,
D. E.
Keller
,
J. P.
Knauer
,
J. A.
Marozas
,
F. J.
Marshall
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
S. P.
Regan
,
T. C.
Sangster
,
D.
Shvarts
,
S.
Skupsky
,
Y.
Srebro
,
R. P. J.
Town
, and
C.
Stoeckl
,
Phys. Plasmas
12
,
032702
(
2005
).
13.
S. P.
Regan
,
V. N.
Goncharov
,
I. V.
Igumenshchev
,
T. C.
Sangster
,
R.
Betti
,
A.
Bose
,
T. R.
Boehly
,
M. J.
Bonino
,
E. M.
Campbell
,
D.
Cao
,
T. J. B.
Collins
,
R. S.
Craxton
,
A. K.
Davis
,
J. A.
Delettrez
,
D. H.
Edgell
,
R.
Epstein
,
C. J.
Forrest
,
J. A.
Frenje
,
D. H.
Froula
,
M.
Gatu Johnson
,
V. Y.
Glebov
,
D. R.
Harding
,
M.
Hohenberger
,
S. X.
Hu
,
D.
Jacobs-Perkins
,
R. T.
Janezic
,
M.
Karasik
,
R. L.
Keck
,
J. H.
Kelly
,
T. J.
Kessler
,
J. P.
Knauer
,
T. Z.
Kosc
,
S. J.
Loucks
,
J. A.
Marozas
,
F. J.
Marshall
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
D. T.
Michel
,
J. F.
Myatt
,
S. P.
Obenschain
,
R. D.
Petrasso
,
R. B.
Radha
,
B.
Rice
,
M.
Rosenberg
,
A. J.
Schmitt
,
M. J.
Schmitt
,
W.
Seka
,
W. T.
Shmayda
,
M. J.
Shoup
III
,
A.
Shvydky
,
S.
Skupsky
,
A. A.
Solodov
,
C.
Stoeckl
,
W.
Theobald
,
J.
Ulreich
,
M. D.
Wittman
,
K. M.
Woo
,
B.
Yaakobi
, and
J. D.
Zuegel
,
Phys. Rev. Lett.
117
,
025001
(
2016
);
[PubMed]
S. P.
Regan
,
V. N.
Goncharov
,
I. V.
Igumenshchev
,
T. C.
Sangster
,
R.
Betti
,
A.
Bose
,
T. R.
Boehly
,
M. J.
Bonino
,
E. M.
Campbell
,
D.
Cao
,
T. J. B.
Collins
,
R. S.
Craxton
,
A. K.
Davis
,
J. A.
Delettrez
,
D. H.
Edgell
,
R.
Epstein
,
C. J.
Forrest
,
J. A.
Frenje
,
D. H.
Froula
,
M.
Gatu Johnson
,
V. Y.
Glebov
,
D. R.
Harding
,
M.
Hohenberger
,
S. X.
Hu
,
D.
Jacobs-Perkins
,
R. T.
Janezic
,
M.
Karasik
,
R. L.
Keck
,
J. H.
Kelly
,
T. J.
Kessler
,
J. P.
Knauer
,
T. Z.
Kosc
,
S. J.
Loucks
,
J. A.
Marozas
,
F. J.
Marshall
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
D. T.
Michel
,
J. F.
Myatt
,
S. P.
Obenschain
,
R. D.
Petrasso
,
R. B.
Radha
,
B.
Rice
,
M.
Rosenberg
,
A. J.
Schmitt
,
M. J.
Schmitt
,
W.
Seka
,
W. T.
Shmayda
,
M. J.
Shoup
, III
,
A.
Shvydky
,
S.
Skupsky
,
A. A.
Solodov
,
C.
Stoeckl
,
W.
Theobald
,
J.
Ulreich
,
M. D.
Wittman
,
K. M.
Woo
,
B.
Yaakobi
, and
J. D.
Zuegel
,
Phys. Rev. Lett.
117
,
059903(E)
(
2016
).
14.
J. A.
Marozas
,
T. J. B.
Collins
,
J. D.
Zuegel
,
P. B.
Radha
,
F. J.
Marshall
, and
W.
Seka
, in the
44th Annual Anomalous Absorption Conference
, Estes Park, CO, 8–13 June
2014
.
15.
B.
Afeyan
and
S.
Hüller
,
EPJ Web Conf.
59
,
05009
(
2013
).
16.
W. L.
Kruer
,
Phys. Fluids B
3
,
2356
(
1991
).
17.
D.
Eimerl
,
W. L.
Kruer
, and
E. M.
Campbell
, “
Ultrabroad bandwidth for suppression of laser driven parametric instabilities
,”
Comm. Plasma Phys. Controlled Fusion
15
,
85
104
(
1992
).
18.
D. S.
Montgomery
,
Phys. Plasmas
23
,
055601
(
2016
).
19.
P.
Michel
,
L.
Divol
,
E. A.
Williams
,
C. A.
Thomas
,
D. A.
Callahan
,
S.
Weber
,
S. W.
Haan
,
J. D.
Salmonson
,
N. B.
Meezan
,
O. L.
Landen
,
S.
Dixit
,
D. E.
Hinkel
,
M. J.
Edwards
,
B. J.
MacGowan
,
J. D.
Lindl
,
S. H.
Glenzer
, and
L. J.
Suter
,
Phys. Plasmas
16
,
042702
(
2009
).
20.
V.
Igumenshchev
,
W.
Seka
,
D. H.
Edgell
,
D. T.
Michel
,
D. H.
Froula
,
V. N.
Goncharov
,
R. S.
Craxton
,
L.
Divol
,
R.
Epstein
,
R.
Follett
,
J. H.
Kelly
,
T. Z.
Kosc
,
A. V.
Maximov
,
R. L.
McCrory
,
D. D.
Meyerhofer
,
P.
Michel
,
J. F.
Myatt
,
T. C.
Sangster
,
A.
Shvydky
,
S.
Skupsky
, and
C.
Stoeckl
,
Phys. Plasmas
19
,
056314
(
2012
).
21.
J. A.
Marozas
,
T. J. B.
Collins
,
J. D.
Zuegel
,
P. W.
McKenty
,
D.
Cao
,
S.
Fochs
, and
P. B.
Radha
,
J. Phys.: Conf. Ser.
717
,
012107
(
2016
).
22.
Inspired by Anishinaabe word Adaawam, meaning “to borrow something from someone,” to describe the CrossBeam Energy Transfer (CBET), just as energy is borrowed from one beam to another. Resource:
J. D.
Nichols
and
E.
Nyholm
,
A Concise Dictionary of Minnesota Ojibwe
(
University of Minnesota Press
,
Minneapolis
,
1995
).
23.
Inspired by Anishinaabe word Mazinisin, meaning “be imprinted, have a design,” to describe laser energy deposition just as the laser imprints on and changes target morphology. Resource:
J. D.
Nichols
and
E.
Nyholm
,
A Concise Dictionary of Minnesota Ojibwe
(
University of Minnesota Press
,
Minneapolis
,
1995
).
24.
J. A.
Marozas
,
F. J.
Marshall
,
R. S.
Craxton
,
I. V.
Igumenshchev
,
S.
Skupsky
,
M. J.
Bonino
,
T. J. B.
Collins
,
R.
Epstein
,
V. Y.
Glebov
,
D.
Jacobs-Perkins
,
J. P.
Knauer
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
S. G.
Noyes
,
P. B.
Radha
,
T. C.
Sangster
,
W.
Seka
, and
V. A.
Smalyuk
,
Phys. Plasmas
13
,
056311
(
2006
).
25.
D.
Froula
, private communication (
2017
).
26.
Laboratory for Laser Energetics, University of Rochester, Rochester, NY, LLE Document No. DOE/SF/19460-295, LLE Review Quarterly Report No. 78,
1999
, pp.
62
81
.
27.
K. E.
Oughstun
and
G. C.
Sherman
,
Electromagnetic Pulse Propagation in Causal Dielectrics, Springer Series on Wave Phenomena
(
Springer-Verlag
,
Berlin
,
1994
), Vol.
16
.
28.
The Control Handbook, The Electrical Engineering Handbook Series, edited by
W. S.
Levine
(
CRC Press
,
Boca Raton, FL
,
1996
).
29.
S. X.
Hu
,
D. T.
Michel
,
A. K.
Davis
,
R.
Betti
,
P. B.
Radha
,
E. M.
Campbell
,
D. H.
Froula
, and
C.
Stoeckl
,
Phys. Plasmas
23
,
102701
(
2016
).
30.
T. P.
Gill
,
The Doppler Effect: An Introduction to the Theory of the Effect
(
Logos
,
London
,
1965
).
31.
J. A.
Marozas
,
M.
Hohenberger
,
M. J.
Rosenberg
,
D.
Turnbull
,
T. J. B.
Collins
,
P. B.
Radha
,
P. W.
McKenty
,
J. D.
Zuegel
,
F. J.
Marshall
,
S. P.
Regan
,
T. C.
Sangster
,
W.
Seka
,
E. M.
Campbell
,
V. N.
Goncharov
,
M. W.
Bowers
,
J.-M. G.
DiNicola
,
G.
Erbert
,
B. J.
MacGowan
,
L. J.
Pelz
, and
S. T.
Yang
, “
First observation of cross-beam energy transfer mitigation for direct-drive inertial confinement fusion implosions using wavelength detuning at the national ignition facility
,”
Phys. Rev. Lett.
120
,
085001
(
2018
).
32.
M. L.
Spaeth
,
K. R.
Manes
,
D. H.
Kalantar
,
P. E.
Miller
,
J. E.
Heebner
,
E. S.
Bliss
,
D. R.
Speck
,
T. G.
Parham
,
P. K.
Whitman
,
P. J.
Wegner
,
P. A.
Baisden
,
J. A.
Menapace
,
M. W.
Bowers
,
S. J.
Cohen
,
T. I.
Suratwala
,
J. M.
Di Nicola
,
M. A.
Newton
,
J. J.
Adams
,
J. B.
Trenholme
,
R. G.
Finucane
,
R. E.
Bonanno
,
D. C.
Rardin
,
P. A.
Arnold
,
S. N.
Dixit
,
G. V.
Erbert
,
A. C.
Erlandson
,
J. E.
Fair
,
E.
Feigenbaum
,
W. H.
Gourdin
,
R. A.
Hawley
,
J.
Honig
,
R. K.
House
,
K. S.
Jancaitis
,
K. N.
LaFortune
,
D. W.
Larson
,
B. J.
Le Galloudec
,
J. D.
Lindl
,
B. J.
MacGowan
,
C. D.
Marshall
,
K. P.
McCandless
,
R. W.
McCracken
,
R. C.
Montesanti
,
E. I.
Moses
,
M. C.
Nostrand
,
J. A.
Pryatel
,
V. S.
Roberts
,
S. B.
Rodriguez
,
A. W.
Rowe
,
R. A.
Sacks
,
J. T.
Salmon
,
M. J.
Shaw
,
S.
Sommer
,
C. J.
Stolz
,
G. L.
Tietbohl
,
C. C.
Widmayer
, and
R.
Zacharias
,
Fusion Sci. Technol.
69
,
25
(
2016
).
33.
S.
Skupsky
,
J. A.
Marozas
,
R. S.
Craxton
,
R.
Betti
,
T. J. B.
Collins
,
J. A.
Delettrez
,
V. N.
Goncharov
,
P. W.
McKenty
,
P. B.
Radha
,
T. R.
Boehly
,
J. P.
Knauer
,
F. J.
Marshall
,
D. R.
Harding
,
J. D.
Kilkenny
,
D. D.
Meyerhofer
,
T. C.
Sangster
, and
R. L.
McCrory
,
Phys. Plasmas
11
,
2763
(
2004
).
34.
J. A.
Marozas
,
T. J. B.
Collins
,
D. H.
Edgell
,
I. V.
Igumenshchev
, and
J. F.
Myatt
,
Bull. Am. Phys. Soc.
56
,
241
(
2011
).
35.
J. A.
Marozas
,
T. J. B.
Collins
,
P. W.
McKenty
,
J. D.
Zuegel
,
P. B.
Radha
,
F. J.
Marshall
,
W.
Seka
,
D. T.
Michel
, and
M.
Hohenberger
,
Bull. Am. Phys. Soc.
59
,
201
(
2014
).
36.
J. A.
Marozas
,
T. J. B.
Collins
,
P. W.
McKenty
, and
J. D.
Zuegel
,
Bull. Am. Phys. Soc.
60
,
167
(
2015
).
37.
J. J.
MacFarlane
,
I. E.
Golovkin
,
P.
Wang
,
P. R.
Woodruff
, and
N. A.
Pereyra
,
High Energy Density Phys.
3
,
181
(
2007
).
38.
T. J. B.
Collins
and
J. A.
Marozas
, “
Mitigation of cross-beam energy transfer in ignition-scale polar-direct-drive target designs for the National Ignition Facility
,”
Phys. Plasmas
(submitted).
39.
Inspired by Anishinaabe word Daashkaa, meaning “be split, be torn apart,” to describe domain decomposition just as the spatial domain is split and divided amongst processing elements. Resource:
J. D.
Nichols
and
E.
Nyholm
,
A Concise Dictionary of Minnesota Ojibwe
(
University of Minnesota Press
,
Minneapolis
,
1995
).
40.
M.
Born
and
E.
Wolf
,
Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light
, 6th ed. (
Pergamon Press
,
Oxford
,
1980
).
41.
M. S.
Sodha
and
A. K.
Ghatak
,
Inhomogeneous Optical Waveguides, Optical Physics and Engineering
, edited by
W. L.
Wolfe
(
Plenum Press
,
New York
,
1977
).
42.
W. L.
Kruer
, in
The Physics of Laser Plasma Interactions, Frontiers in Physics
, edited by
D.
Pines
(
Westview
,
Boulder, CO
,
2003
), Vol.
73
.
43.
W. H.
Press
,
B. P.
Flannery
,
S. A.
Teukolsky
, and
W. T.
Vetterling
,
Numerical Recipes (FORTRAN Version
) (
Cambridge University Press
,
New York
,
1989
).
44.
S.
Atzeni
and
J.
Meyer-ter-Vehn
,
The Physics of Inertial Fusion: Beam Plasma Interaction, Hydrodynamics, Hot Dense Matter
, 1st ed., International Series (
Oxford University Press
,
Oxford
,
2004
), Vol.
125
.
45.
J.
Delettrez
,
Can. J. Phys.
64
,
932
(
1986
).
46.
D. H.
Froula
,
T. J.
Kessler
,
I. V.
Igumenshchev
,
R.
Betti
,
V. N.
Goncharov
,
H.
Huang
,
S. X.
Hu
,
E.
Hill
,
J. H.
Kelly
,
D. D.
Meyerhofer
,
A.
Shvydky
, and
J. D.
Zuegel
,
Phys. Plasmas
20
,
082704
(
2013
).
47.
R.  K.
Follett
,
J.  G.
Shaw
,
J.  F.
Myatt
,
J.  P.
Palastro
,
R.  W.
Short
, and
D.  H.
Froula
,
Phys. Rev. Lett.
120
,
135005
(
2018
).
48.
Laboratory for Laser Energetics, University of Rochester, Rochester, NY, LLE Document No. DOE/NA/28302-826,
LLE Review Quarterly Report No. 114
,
2008
, pp.
73
80
.