We demonstrate a methodology for diagnosing the multiscale dynamics and energy transfer in complex HED flows with realistic driving and boundary conditions. The approach separates incompressible, compressible, and baropycnal contributions to energy scale-transfer and quantifies the direction of these transfers in (generalized) wavenumber space. We use this to compare the kinetic energy (KE) transfer across scales in simulations of 2D axisymmetric vs fully 3D laser-driven plasma jets. Using the FLASH code, we model a turbulent jet ablated from an aluminum cone target in the configuration outlined by Liao et al. [Phys. Plasmas, 26 032306 (2019)]. We show that, in addition to its well known bias for underestimating hydrodynamic instability growth, 2D modeling suffers from significant spurious energization of the bulk flow by a turbulent upscale cascade. In 2D, this arises as vorticity and strain from instabilities near the jet's leading edge transfer KE upscale, sustaining a coherent circulation that helps propel the axisymmetric jet farther ( 25 % by 3.5 ns) and helps keep it collimated. In 3D, the coherent circulation and upscale KE transfer are absent. The methodology presented here may also help with inter-model comparison and validation, including future modeling efforts to alleviate some of the 2D hydrodynamic artifacts highlighted in this study.

High speed jets arise in many astrophysical and high energy density (HED) applications.1 In astrophysics, supersonic hydrodynamic jets can arise as an initially collimated outflow or an otherwise quasi-spherical supersonic blast wave propagating into an inhomogeneous medium. Jets appear on sources of all scales of astrophysics, including supernovae evolution.2 In inertial confinement fusion (ICF), jets result from the interaction between a shock and density perturbations such as those due to the fill-tube3 or due to voids.4,5 These jets can trigger the mixing of ablator and fuel, which ultimately degrade the ignition yield.6,7 Jets can often be characterized as complex multi-scale flows involving instabilities and turbulence, and in many cases, instabilities trigger turbulent mixing.8,9

Turbulence is common in complex HED flows in astrophysics and ICF. In the evolution of core-collapse supernovae, for example, turbulence is generated due to instabilities between the inner core and the outgoing shock.10–12 Energy transfer between scales is believed to play a key role in such simulations and may account for differences between 2D and 3D.13–15 In particular, it is believed that upscale (inverse) transfer of kinetic energy in 2D may help to increase the efficacy of turbulent transport of heat to the shock to prevent it from stalling, making explosions easier to achieve than in 3D.

Instabilities are also common in complex HED flows in astrophysics and ICF. For example, the Rayleigh-Taylor (RT) instability in ICF has been a significant hindrance to ignition.16 In ICF, 2D simulations are the main “work horse” for experimental design17–19 as routine 3D simulations are prohibitively expensive. A capsule-only 3D simulation requires tens to hundreds of millions of CPU hours on supercomputers even without incorporating the hohlraum and laser physics.20,21 The tradeoffs between 2D and 3D flow physics are still not widely appreciated in ICF simulations. The primary objective of this study is to highlight the differences in energy scale-transfer how they affect the travel of laser-driven jets between 2D and 3D.

Length scales in complex flows are generally characterized by a wavenumber energy spectrum. Yet, the energy transfer pathways and interactions between different scales are often subtle and cannot be diagnosed using the spectrum alone. In this work, we investigate two processes that transfer energy across scales: deformation work, Π, and baropycnal work, Λ.

Deformation work is due to the multi-scale nature of the velocity field and is the sole mechanism for energy transfer across scales (or cascade) in constant-density (incompressible) flows described by the Navier–Stokes equation.22,23 In 3D incompressible homogeneous turbulence, the main mechanism behind Π is believed to be vortex stretching,24–28 which transfers energy downscale. Kraichnan29 predicted and subsequent studies30,31 have demonstrated that an incompressible homogeneous turbulent flow, when restricted to two spatial dimensions, will tend to transfer energy in the opposite direction, i.e., upscale rather than downscale, due to vorticity conservation in 2D. However, HED flows involve significant density variations and do not conserve vorticity, even when restricted to 2D, due to baroclinicity. Therefore, extending such phenomenologies29,31 from incompressible homogeneous flows to inhomogeneous flows with variable density remains dubious.32 A main contribution in this work is to demonstrate, via a direct measurement of Π, that an upscale cascade can indeed arise in 2D hydrodynamic models of HED systems, despite the lack of vorticity conservation.

The second energy transfer pathway we analyze here is baropycnal work Λ, which results from pressure and density variations at different scales.33,34 The direction of energy transfer by Λ depends on the source of pressure gradient and density variations.32,35–38 Typical cases include (i) a baroclinic flow wherein Λ transfers energy downscale into small-scale velocity fluctuations due to opposing directions of pressure and density gradients,32 and (ii) a shock wherein pressure and density gradients are aligned and Λ transfers energy upscale into the large-scale pressure field.34,36,38

Previous works have utilized coarse-graining methods that are common in large eddy simulation modeling39 to decompose length scales and analyze deformation work40,41 and baropycnal work.33,38,42 Lees and Aluie42 conducted simulations of compressible turbulence in 3D with varying levels of compressibility. Their results demonstrated that in 3D homogeneous compressible turbulence, Π transfers energy downscale to the small-scale turbulence, while Λ transfers energy upscale into large-scale potential energy residing in the pressure field. Zhao et al.32 analyzed the energy scale-transfer in baroclinically unstable flows by simulating the RT instability and the ensuing late-stage turbulence in 2D and 3D. They found that in both 2D and 3D, Λ transfers energy downscale by releasing potential energy stored at the largest scales and depositing it over a wide range of smaller scales in the fully turbulent regime. Energy transfer by Π, on the other hand, exhibits opposing behaviors in 2D vs 3D.32 In 2D, upscale transfer by Π leads to higher bulk kinetic energy at large scales, whereas in 3D, the energy transfer is downscale.

Here, we extend and demonstrate the methodology to high energy density (HED) flow regimes by taking advantage of laser-driven jets. We reference the simulation configuration by Liao et al.,43 which proposes a design to set up a turbulent dynamo operating on the OMEGA-EP laser.44,45 This design, sketched in Fig. 1, shines multiple laser beams on the inner wall of a cone target to create a supersonic turbulent jet. In Sec. II, we describe the model configurations, address the sensitivity to initial conditions in 2D and 3D, and report on the jets' evolution and properties. In Sec. III, we introduce the coarse-graining method for scale decomposition and define the energy scale-transfer terms, Π and Λ. In Sec. IV, we analyze KE at large scales and show that the corresponding difference between 2D and 3D can be quantified by Π and Λ. In Sec. V, we characterize the turbulent region at the jet's leading edge using subscale stress and discuss KE scale-transfer in that region. In Sec. VI, we decompose Π and Λ into contributions from the divergence-free and irrotational components of the flow to examine the mechanism causing the 2D jet's leading edge to remain narrow. We summarize the results in the concluding Sec. VII.

FIG. 1.

Configuration of run (left) L5R090 and run (right) L6R720. Label “L” denotes the location of laser deposition along the y-direction (in millimeters), and label “R” denotes the radius of the laser beam (in micrometers). Note that the deposition area is larger than the beam size since the laser propagation direction is not perpendicular to the cone surface.

FIG. 1.

Configuration of run (left) L5R090 and run (right) L6R720. Label “L” denotes the location of laser deposition along the y-direction (in millimeters), and label “R” denotes the radius of the laser beam (in micrometers). Note that the deposition area is larger than the beam size since the laser propagation direction is not perpendicular to the cone surface.

Close modal

FLASH is a multiphysics radiation-hydrodynamics code used frequently for HED applications. It provides a comprehensive environment for full-physics laser driven simulations by solving governing equations in three-temperature unsplit hydrodynamics on an adaptive mesh refinement (AMR) grid.46,47 The laser driven simulation module includes features of laser ray tracing,48 electron thermal conduction with Spitzer conductivities,49 and radiation diffusion with tabulated opacity.50 Liao et al.43 used the FLASH code to develop a test platform for laser-driven turbulent dynamos that matches realizable OMEGA-EP experimental configurations in 3D.43 We conduct our 3D simulations using a similar configuration that generates turbulence by the interaction of laser ablated jets. To simulate the analogous cone target in 2D, we employ cylindrical coordinates, which significantly modifies the configuration of Liao et al. in that the laser deposition becomes azimuthally symmetric. To approximate this modification in 3D, within Cartesian coordinates, we simulate a quasi-azimuthally symmetric configuration in which 256 laser beams deposit energy symmetrically about the center vertical axis of the target. We investigate sensitivity to the laser drive and show the validity of comparing 2D and 3D simulations below. In both 2D and 3D geometries, UV laser beams (351 nm wavelength) deliver a total power of 0.5 TW over a 3.5 ns duration using square pulses. The simulations use multigroup diffusion for radiation transport using six groups.

As shown in Fig. 1, the simulation domain is 0.25 cm wide and 0.75 cm high, and the target occupies 0.3 cm of height from the top boundary of the domain. The main target is 2.7 g/cc aluminum notched into a hollow cone with an interior space of 0.2 cm diameter at the cone base and 0.274 cm of height filled with a pseudo-vacuum of 10 6 g/cc helium. The vacuum region extends from the cone's opening to the bottom of the domain over 0.45 cm. The equations of state (EOS) and opacity of both the aluminum cone and helium pseudo-vacuum are obtained from the atomic database IONMIX.51 

Due to the nature of cylindrical coordinates, the 2D computational domain takes an axisymmetric boundary at x = 0 in Fig. 1 as the center axis, for which both normal and toroidal vector components change sign. The boundaries everywhere else are outflow, which specifies a zero gradient to allow the flow to leave the domain freely. The 3D computational domain takes outflow boundaries in all directions. In this work, all visualizations of the 2D domain are mirrored about the axisymmetric boundary (x = 0) to compare to the slice of the 3D domain.

Applying adaptive mesh refinement,52 we set the initial static grid size of a mesh block to be 16 in all directions, with a maximum refinement level of 5 in both 2D and 3D. Each refinement level doubles the grid size at the finest resolution. Since the computational domain's aspect ratio is 1:6 in 2D and 1:3 in 3D, we use three mesh blocks in 3D aligned in the vertical direction and six mesh blocks in 2D. The effective mesh resolution can be calculated as ( grid size o f single block ) × 2 ( max refinement level 1 ) × ( number o f blocks ), which gives the resolution 256 × 1536 in 2D and 256 × 256 × 768 in 3D. The mesh refinement criterion used is the modified Löhner's estimator,50 for which we set the density and electron temperature to be the refinement variables and the threshold value to trigger refinement is 0.8. This grid mesh configuration resolves turbulence structures that appear around the leading edge of the jet at late times. In the  Appendix, we report on our analysis of numerical convergence (see Fig. 20).

Figure 1 shows initial configurations for runs L5R090 and L6R720, where the label “L” denotes the y-location of laser deposition in millimeters, and the label “R” is the radius of laser deposition in micrometers. Due to the inherent azimuthal symmetry in the 2D simulations, which requires only one laser beam to illuminate the entire target. In 3D, the beams are separated with equal angle azimuthally on the cone. Matching the 2D and 3D configurations of the initial flow is crucial for making comparisons. Here, we list a few laser-drive parameters that play an important role in the initial flow's development.

First, the number of laser rays traced in the simulation significantly affects the ablated flow's smoothness at early times. As shown in Figs. 2(a) and 2(c), the same number of rays reveals distinct early time velocity profile differences. The 2D initial flow using 4096 rays in the single beam manifests large-scale non-uniformity in the reflected-energy region ( y > 0.5325 cm, highlighted by arrows in Fig. 2), whereas this issue is absent in 3D due to using a higher number of laser beams, each with 4096 rays. With a sufficient number of rays shown in Fig. 2(b), the early time velocity profile off the ablated surface agrees in 2D and 3D.

FIG. 2.

The number of laser rays affects the ablated flow's smoothness. Panels show velocity magnitude from run L5R090 at t = 0.3 ns and y > 0.45 cm using different numbers of laser rays. Even though the laser beams are incident on the cone in the region around y = 0.5 cm, a fraction of this laser energy is reflected deeper into the cone around y = 0.6 cm highlighted by the white arrows in the panels. This reflected laser energy induces a flow, which is sensitive to the number laser rays. Comparing panels (a) and (b), we can see that using 4096 rays in the single beam in 2D manifests large-scale non-uniformity in the ablated flow. With 32 768 rays in the single beam in 2D in panel (b), the velocity profile off the ablated surface agrees with 3D in panel (c), which uses 4096 rays/beam × 256 beams. Panel (c) is a slice in the 3D domain at z = 0 cm.

FIG. 2.

The number of laser rays affects the ablated flow's smoothness. Panels show velocity magnitude from run L5R090 at t = 0.3 ns and y > 0.45 cm using different numbers of laser rays. Even though the laser beams are incident on the cone in the region around y = 0.5 cm, a fraction of this laser energy is reflected deeper into the cone around y = 0.6 cm highlighted by the white arrows in the panels. This reflected laser energy induces a flow, which is sensitive to the number laser rays. Comparing panels (a) and (b), we can see that using 4096 rays in the single beam in 2D manifests large-scale non-uniformity in the ablated flow. With 32 768 rays in the single beam in 2D in panel (b), the velocity profile off the ablated surface agrees with 3D in panel (c), which uses 4096 rays/beam × 256 beams. Panel (c) is a slice in the 3D domain at z = 0 cm.

Close modal

Second, the radius of laser deposition has an important role in matching the early time evolution in 2D and 3D. In run L5R090, the reflected laser energy appears in the domain at y > 0.5325 cm in Fig. 2, highlighted by white arrows. However, the fraction of reflected energy differs between 2D and 3D, which leads to differences in the mass ablated at different locations along the cone surface. This can be quantified by comparing the kinetic energy at y > 0.5325 cm, KEup to kinetic energy over the entire domain, KEtot. Their ratio, K E u p / K E tot, is listed in Table I and shows a mismatch between 2D and 3D for run L5R090. Using a larger radius for the laser deposition area minimizes ablation from reflected radiation and, therefore, minimizes differences between 2D and 3D in the early time flow. This is achieved in run L6R720 in Fig. 3, where the fraction K E u p / K E tot is the same in 2D and 3D in Table I. Since the y-momentum, Gy, is central to the jet's traveling distance, we also show a similar ratio G y u p / G y tot in Table II, which agrees between 2D and 3D in run L6R720. Despite increasing the laser radius in run L6R720, it is still necessary to use a sufficient number of rays in 2D to match the early time solutions from the 2D and 3D simulations, as discussed in the previous paragraph.

TABLE I.

Kinetic energy (KE) and internal energy (IE) per volume (in erg/cm3) in different regions of the domain at t = 0.3 ns.

Run KEup(y > 0.5325 cm) KEdown(y < 0.5325 cm) KEtot K E tot + I E tot K E u p / K E tot
2D L5R090  2.27 × 10 7  2.96 × 10 8  3.19 × 10 8  1.26 × 10 9  7% 
3D L5R090  6.12 × 10 6  2.42 × 10 8  2.48 × 10 8  1.26 × 10 9  2% 
2D L6R720  7.58 × 10 7  1.70 × 10 8  2.46 × 10 8  1.29 × 10 9  31% 
3D L6R720  5.82 × 10 7  1.27 × 10 8  1.85 × 10 8  1.29 × 10 9  31% 
Run KEup(y > 0.5325 cm) KEdown(y < 0.5325 cm) KEtot K E tot + I E tot K E u p / K E tot
2D L5R090  2.27 × 10 7  2.96 × 10 8  3.19 × 10 8  1.26 × 10 9  7% 
3D L5R090  6.12 × 10 6  2.42 × 10 8  2.48 × 10 8  1.26 × 10 9  2% 
2D L6R720  7.58 × 10 7  1.70 × 10 8  2.46 × 10 8  1.29 × 10 9  31% 
3D L6R720  5.82 × 10 7  1.27 × 10 8  1.85 × 10 8  1.29 × 10 9  31% 
FIG. 3.

A wider laser beam helps match the early-time evolution in 2D and 3D. Panels show velocity magnitude from run L6R720 at t = 0.3 ns and y > 0.45 cm using a laser beam that is wider than that in L5R090 shown in Figs. 2(b) and 2(c). Panel (b) is a slice in the 3D domain at z = 0 cm.

FIG. 3.

A wider laser beam helps match the early-time evolution in 2D and 3D. Panels show velocity magnitude from run L6R720 at t = 0.3 ns and y > 0.45 cm using a laser beam that is wider than that in L5R090 shown in Figs. 2(b) and 2(c). Panel (b) is a slice in the 3D domain at z = 0 cm.

Close modal
TABLE II.

Axial y-momentum per volume (in g cm/s/cm3) in different regions of the domain at t = 0.3 ns.

Run G y u p(y > 0.5325 cm) G y down(y < 0.5325 cm) G y tot G y u p / G y tot
2D L5R090  4.90  16.69  21.58  23% 
3D L5R090  2.10  14.84  16.94  12% 
2D L6R720  8.89  16.86  25.75  35% 
3D L6R720  6.29  12.71  19.01  33% 
Run G y u p(y > 0.5325 cm) G y down(y < 0.5325 cm) G y tot G y u p / G y tot
2D L5R090  4.90  16.69  21.58  23% 
3D L5R090  2.10  14.84  16.94  12% 
2D L6R720  8.89  16.86  25.75  35% 
3D L6R720  6.29  12.71  19.01  33% 
Third, to create quasi-azimuthally symmetric laser deposition in 3D, we increase the number of laser beams until the traveling distance of the jet converges. This is characterized by the y-momentum centroid along the y-direction,
(1)
where uy is the y-component of velocity, ρ is mass density, and V is the volume of the physical domain. As shown in Fig. 4, run L5.5R360 in 3D converges using 128 laser beams, while L6R720, which uses a larger beam radius, requires only 64 beams. A larger beam radius makes it easier to approximate azimuthal symmetry with fewer beams as expected. Considering the three factors discussed above, we use run L6R720 in the rest of the paper.
FIG. 4.

To create quasi-azimuthally symmetric laser deposition in 3D requires a sufficiently large number of overlapping laser beams. Convergence of jet's traveled distance as a function laser beam number is quantified by the y-momentum centroid, G y centroid, in Eq. (1). This is plotted here using an increasing number of laser beams labeled with “B”. Runs L5.5R360 converge with 128 laser beams, and runs L6R720 converge with only 64 laser beams due to a wider beam radius.

FIG. 4.

To create quasi-azimuthally symmetric laser deposition in 3D requires a sufficiently large number of overlapping laser beams. Convergence of jet's traveled distance as a function laser beam number is quantified by the y-momentum centroid, G y centroid, in Eq. (1). This is plotted here using an increasing number of laser beams labeled with “B”. Runs L5.5R360 converge with 128 laser beams, and runs L6R720 converge with only 64 laser beams due to a wider beam radius.

Close modal
Figure 5 uses run L6R720 to demonstrate that kinetic energy is almost identical between 2D and 3D, differing by 0.47% at t = 3.5. Figure 5 plots kinetic energy KE, internal energy IE, and the total energy Etot being the sum of the KE and IE. Internal energy is also almost identical between 2D and 3D, differing by 2.1% at t = 3.5 ns. To quantify the similarity between 3D quasi-azimuthal symmetry and 2D azimuthal symmetry, we take the standard deviation of the averaged kinetic energy along the azimuthal direction,
(2)
where θ is the azimuthal angle, θ is the domain average at an azimuthal coordinate, N is the number of θ bins, and the overbar denotes the average over all θ. We find a percentage of 2.9% by comparing K E θ std to the domain averaged kinetic energy, K E , at t = 3.5 ns in 3D. This percentage quantifies the K E variance in the azimuthal direction. Naturally, this variance is identically 0% for the 2D simulation.
FIG. 5.

Comparison of kinetic energy (KE), internal energy (IE), and total energy (Etot) between 2D and 3D from run L6R720. KE in 2D and 3D are almost identical with difference of only 0.47% at t = 3.5 ns.

FIG. 5.

Comparison of kinetic energy (KE), internal energy (IE), and total energy (Etot) between 2D and 3D from run L6R720. KE in 2D and 3D are almost identical with difference of only 0.47% at t = 3.5 ns.

Close modal

All visualizations in the remainder of this paper are shown from run L6R720 at time t = 3.5 ns, which is approximately when the jet reaches the bottom of the domain and are shown at the cross section in the xy-plane at z = 0. In Fig. 6, the jet exhibits turbulent structures in both 2D and 3D, especially at the leading edge, which we characterize using the subscale stress discussed in Sec. V. It is evident from the velocity profiles in Fig. 6 that the jet travels farther in 2D than 3D, despite having almost identical domain-integrated kinetic energy (Fig. 5). This indicates that KE resides at different scales between the two configurations, which is most evident at the leading edge of the jets. A shock that is reflected from the jet's leading edge, traveling toward the cone, can be seen as a discontinuity in the velocity at y 0.15 to 0.2 cm in Fig. 6 (see also Fig. 2 in Ref. 53).

FIG. 6.

Visualization of velocity magnitude, | u |, from run L6R720 at t = 3.5 ns. Despite having almost identical domain-integrated kinetic energy in 2D and 3D, the jet travels a longer distance in 2D than in 3D. This indicates that KE resides at different scales between the two configurations, which is most evident at the leading edge of the jets. Panel (b) is a slice in the 3D domain at z = 0 cm.

FIG. 6.

Visualization of velocity magnitude, | u |, from run L6R720 at t = 3.5 ns. Despite having almost identical domain-integrated kinetic energy in 2D and 3D, the jet travels a longer distance in 2D than in 3D. This indicates that KE resides at different scales between the two configurations, which is most evident at the leading edge of the jets. Panel (b) is a slice in the 3D domain at z = 0 cm.

Close modal
To objectively compare the 2D and 3D jets, we use the y-momentum centroid in the y-direction in Eq. (1). This metric gives a bulk characterization of the jet's movement. A momentum-based metric is more appropriate than a center of mass, which would include a low-velocity material near the cone. Figure 7(a) plots the traveling distance in time and shows that the jet in 2D travels farther than 3D at all times. Another metric we use is the y-velocity centroid in the y-direction,
(3)
to highlight the discrepancy seen in Fig. 6. Plots of u y centroid in Fig. 7(b) show a consistent message that the jet in 2D travels farther than 3D at all times, with a difference of 16.1% at t = 3.5 ns. Differences based on u y centroid are larger than those based on G y centroid since density at the leading edge is a few orders of magnitude smaller compared to the region near the target. This can be gleaned from comparing Figs. 6 and 8. A third metric we use is based on identifying the position of the leading edge, ρ y leading, which we define as the location where density jumps by a factor of 100. We find that the leading edge in 2D has traveled 26.7% farther than in 3D at t = 3.5 ns.
FIG. 7.

Distance traveled by the jet along the y axis, characterized by (a) the y-momentum centroid, G y centroid, and (b) y-velocity centroid, u y centroid. Comparing 2D and 3D at t = 3.5 ns, G y centroid differs by 6.8% and u y centroid differs by 16.1%.

FIG. 7.

Distance traveled by the jet along the y axis, characterized by (a) the y-momentum centroid, G y centroid, and (b) y-velocity centroid, u y centroid. Comparing 2D and 3D at t = 3.5 ns, G y centroid differs by 6.8% and u y centroid differs by 16.1%.

Close modal
FIG. 8.

To complement Fig. 6, this figure visualizes momentum magnitude, | G | = | ρ u |, at t = 3.5 ns (note the log-scale colorbar). Momentum at the leading edge of the jet is a few orders of magnitude smaller relative to that near the cone's surface, which underscores the wide variations in density.

FIG. 8.

To complement Fig. 6, this figure visualizes momentum magnitude, | G | = | ρ u |, at t = 3.5 ns (note the log-scale colorbar). Momentum at the leading edge of the jet is a few orders of magnitude smaller relative to that near the cone's surface, which underscores the wide variations in density.

Close modal

For completeness, we report on the compressibility levels of the flow using two quantities. First, in Fig. 9, we show the pointwise Mach number, M = | u | / c, where | u | is magnitude of the local velocity in the lab frame and c is the local sound speed. The Mach number in 2D is generally greater than in 3D, especially in the pre-shocked region at y 0.15 0.3 cm, which is consistent with the 2D jet being generally faster than in 3D. To quantify the differences in Fig. 9, we show the bulk Mach number, M = | u | / c , in  Appendix Fig. 21(a), where is the domain average. A difference of 12.1 % between 2D and 3D occurs at the end of the simulation ( t = 3.5 ns).

FIG. 9.

Visualization of Mach number, M = | u | / c, at t = 3.5 ns. Here, c is the local sound speed. Higher values of M in 2D indicates higher levels of compressibility relative to 3D.

FIG. 9.

Visualization of Mach number, M = | u | / c, at t = 3.5 ns. Here, c is the local sound speed. Higher values of M in 2D indicates higher levels of compressibility relative to 3D.

Close modal

Second, in Fig. 10, we calculate the time series of | · u | rms and | × u | rms, which are more sensitive to the flow at smaller scales due to shocks and turbulence. Both | · u | rms and | × u | rms are greater in 2D than in 3D. However, Fig. 10(b) shows that | × u | rms in 2D starts decreasing after 1 ns, which we will show is due to an upscale vortical energy transfer discussed in Sec. VI A. In contrast, | × u | rms in 3D increases over the entire duration of the simulation as seen from Fig. 10(b). From the ratio | · u | rms / | × u | rms [see the  Appendix Fig. 21(b)], which is a measure of compressibility levels at small scales, we can infer that the 2D jet also exhibits higher small-scale compressibility levels than its 3D counterpart at t = 3.5 ns.

FIG. 10.

Root mean square (rms) of (a) divergence of velocity and (b) vorticity. The ratio | · u | rms / | × u | rms characterizes compressibility levels at small scales. In panel (b), the vortical component at small scales decreases in 2D due to upscale energy transfer as we discuss in Sec. VI.

FIG. 10.

Root mean square (rms) of (a) divergence of velocity and (b) vorticity. The ratio | · u | rms / | × u | rms characterizes compressibility levels at small scales. In panel (b), the vortical component at small scales decreases in 2D due to upscale energy transfer as we discuss in Sec. VI.

Close modal

The coarse-graining approach allows for analyzing the dynamics at different scales in a complex flow, which has proven to be a natural and versatile framework to understand scale interactions.40,54–57 The approach is standard in partial differential equations and distribution theory.58,59 It became common in large eddy simulation (LES) modeling39 of turbulence, thanks to the foundational works of Leonard60 and Germano.61 Reference 62 provides an overview of coarse-graining and its connection to other methods in physics.

For any field a ( x ), a coarse-grained or (low-pass) filtered version of this field, which contains spatial variations at scales > , is defined in n-dimensions as
(4)
Here, G is a convolution kernel derived from the “father” kernel G ( s ) (borrowing the term from wavelet analysis63) G ( s ) is normalized, d n s G ( s ) = 1. It is an even function such that s G ( s ) d n s = 0, which ensures that local averaging is symmetric. G ( s ) also has its main support (or variance) over a region of order unity in diameter, d n s | s | 2 G ( s ) = O ( 1 ). The dilated version of the kernel, G ( r ) = n G ( r / ), is a function of dimensional position vector r and inherits all those properties, except that its main support is over a region of diameter .
An example is the Boxcar kernel in non-dimensional coordinates,
(5)
and its dilated version in dimensional coordinates,
(6)

The scale decomposition in Eq. (4) is essentially a partitioning of scales for a field into large scales ( ), captured by a ¯ , and small scales ( ), captured by the residual a = a a ¯ . In the following analyses, we use the Boxcar kernel in Eq. (6) for coarse-graining. With non-periodic boundary conditions, such as with our simulation domains here, filtering near the boundary requires a choice for the fields beyond the boundary. It was shown in previous works that a natural choice is to extend the domain beyond the physical boundaries with values compatible with the boundary conditions.64–66 However, over the times we analyze the jets in this work, flow across the domain boundaries is negligible. With consideration to the Helmholtz decomposition we use in Sec. VI and the ease of its implementation, we mirror (or reflect) the variables across the boundary when coarse-graining. In the  Appendix Figs. 22–24, we show that mirroring has negligible effect on the results compared to extending the variables beyond the domain boundaries using the true boundary conditions.

Diagnosing the scale-dependent kinetic energy budget in constant-density turbulence typically focuses on analyzing kinetic energy based on the coarse velocity field, 1 2 ρ 0 | u ¯ | 2 (e.g., Refs. 39 and 62) However, in the presence of significant density variations, kinetic energy 1 2 ρ | u | 2 is non-quadratic and its decomposition as a function of scale is not as straightforward.34 Several different decompositions have been proposed in the literature, such as 1 2 ρ ¯ | u ¯ | 2 (e.g., Refs. 67–70) and 1 2 | ( ρ u ) ¯ | 2 (e.g., Refs. 37 and 71–73). However, Zhao and Aluie65 showed that these scale decompositions violate the “inviscid criterion,” which requires that viscous effect be negligible at sufficiently large scales. They also demonstrated that the Favre decomposition,34,74 1 2 ρ ¯ | u ̃ | 2, satisfies the inviscid criterion, where the Favre filtered velocity is density-weighted according to
(7)
The KE budget at scales larger than an arbitrary is derived from the compressible Navier–Stokes equations,34,75
(8)
Here, J ( x ) is spatial transport of large-scale kinetic energy, P ¯ · u ¯ is large-scale pressure dilatation, D ( x ) is viscous dissipation acting on scales larger than , and ϵ inj ( x ) is the energy injected due to external forcing. These terms are defined by Ref. 34 in Eqs. (16)–(18). Deformation work, Π ( x ), and baropycnal work, Λ ( x ), account for the energy transfer across scales, and are defined as
(9)
(10)
where the 2nd-order generalized central moment61 is
(11)
and its Favre density-weighted analogue34,38 is
(12)
Physically, τ ̃ ( u i , u j ) in Eq. (9) represents the subscale stress tensor (per unit mass) exerted by scales < on the larger scale flow, and τ ¯ ( ρ , u j ) in Eq. (10) represents the subscale mass flux due to scales < . As Π ( x ) and Λ ( x ) appear in Eq. (8), they are defined to be positive energy transfer from large scales to small scales.

Π ( x ) and Λ ( x ) contain all information needed to quantify the exchange of energy between the two sets of scales, > and < . Since we have complete knowledge of the dynamics at all scales resolved in a simulation, Π ( x ) and Λ ( x ) can be calculated exactly at every point x in the domain and at any instant in time t. It is often not possible from simulations or observations to resolve all scales present in the real system. Therefore, computing Π ( x ) and Λ ( x ) is only measuring the dynamical coupling between scales present in the data.

In many instances, standard tools that were developed and used in the turbulence literature to the study of cross-scale transfer are only strictly valid to analyze homogeneous isotropic incompressible flows. Consequently, calculations of the energy transfer rates in HED hydrodynamic applications that use these tools may give ambiguous results for inhomogeneous flows, such as the jet in Fig. 6. The problem arises because there are several possible definitions for the scale-transfer terms, Π ( x ) and Λ ( x ), as we now elaborate.

Definitions (9) and (10) for the scale-transfer of energy in budget (8) are not unique with other definitions possible (see examples in Refs. 41 and 64) The difference between any two of these definitions is a divergence term, · ( ), which amounts to a reinterpretation of which terms in budget (8) represent transfer of energy across scales and which terms redistribute (or transport) energy in space, J ( x ). There is an infinite number of ways to reorganize terms in budget (8) and, thus, an infinite number of possible definitions for the transfer of kinetic energy between scales. This freedom in defining Π ( x ) can be thought of as a gauge freedom.64 

In a homogeneous flow, spatial averages of all these definitions are equal because their difference is a divergence that is zero, · ( ) = 0. On the other hand, if one considers inhomogeneous flows, such as the jet in Fig. 6, or if one wishes to analyze the cascade locally in space without spatial averaging, then such definitions can differ qualitatively as well as quantitatively (see Ref. 64 for examples). Definitions (9) and (10) are proper measures of the energy scale-transfer because they satisfy two important physical criteria: (i) Galilean invariance and (ii) vanish in the absence of subscale motion.76 Using such criteria to choose the scale-transfer definitions may be thought of as gauge fixing. Regarding the second criterion, both scale-transfer terms vanish identically at every location x when is the grid scale or smaller.76 The latter is a physically important constraint since scales smaller than should not influence the larger scale flow if those scales do not exist.

In the remainder of this paper, the subscript is omitted for the variables defined above if there is no risk of ambiguity. In what follows, we shall use k to represent the “filtering wavenumber”77 corresponding to length scale ,
(13)
with L = 0.25 cm being the domain size along the x-direction, which we use here as a reference scale. Note that k is not a Fourier wavenumber, just a proxy for scale from the coarse-graining decomposition.

We have already shown in Fig. 5 that KE in 2D and 3D is almost identical at t = 3.5 ns, differing by only 0.47%. However, the distribution of this KE among length scales differs significantly between 2D and 3D.

Coarse KE in Eq. (8), ρ | u ̃ | 2 / 2, quantifies the cumulative KE at all scales larger than . Taking k = L / = 1, Fig. 11 shows that KE in 2D at scales larger than L = 0.25 cm exceeds that in 3D. This difference of 4.80 × 10 9 erg/cm3 in KE content at large scales is highlighted in Table III, which is greater by more than an order of magnitude compared to the 2.96 × 10 8 erg/cm3 (0.47% difference) in total KE. This is consistent with the bulk jet flow in 2D being faster than in 3D (Fig. 6). Indeed, Fig. 25 in the  Appendix shows that most of this difference between 2D and 3D is from the y-component of KE, ρ | ( u y ) ̃ | 2 / 2, due to the flow in the axial direction. Figure 11 also shows that for larger k = 2 , 4 values, the difference between 2D and 3D diminishes. This is to be expected since increasing k is equivalent to including smaller scales in coarse KE metric, which converges to the total KE, ρ | u ̃ | 2 / 2 ρ | u | 2 / 2 in the limit k or, equivalently, 0. We remind the reader that total KE is almost identical in 2D and 3D (Table III and Fig. 5).

FIG. 11.

Domain-averaged coarse KE, ρ ¯ | u ̃ | 2 / 2 , at all scales > . Here, filtering wavenumber k = L / is not a Fourier wavenumber, just a proxy for scale from the coarse-graining decomposition. The largest scales (blue) in 2D have more KE than in 3D, even though total KE is almost identical in both 2D and 3D (Fig. 5). This can be seen here from having the 2D and 3D plots getting closer (solid and dashed) as we include smaller scales (larger k , yellow and red) in the coarse KE.

FIG. 11.

Domain-averaged coarse KE, ρ ¯ | u ̃ | 2 / 2 , at all scales > . Here, filtering wavenumber k = L / is not a Fourier wavenumber, just a proxy for scale from the coarse-graining decomposition. The largest scales (blue) in 2D have more KE than in 3D, even though total KE is almost identical in both 2D and 3D (Fig. 5). This can be seen here from having the 2D and 3D plots getting closer (solid and dashed) as we include smaller scales (larger k , yellow and red) in the coarse KE.

Close modal
TABLE III.

Domain-averaged cumulative KE at t = 3.5 ns.

Cumulative energy ( erg / cm 3 ) k = 1 KEtot
K E 2 D  4.09 × 10 10  6.89 × 10 10 
K E 3 D  3.61 × 10 10  6.86 × 10 10 
K E 2 D K E 3 D  4.80 × 10 9  2.96 × 10 8 
Cumulative energy ( erg / cm 3 ) k = 1 KEtot
K E 2 D  4.09 × 10 10  6.89 × 10 10 
K E 3 D  3.61 × 10 10  6.86 × 10 10 
K E 2 D K E 3 D  4.80 × 10 9  2.96 × 10 8 

To explain the aforementioned large scale ( > L = 0.25 cm) KE difference between 2D and 3D, we analyze the KE scale-transfer, Π and Λ in Eqs. (9) and (10). Figure 12 shows the time series of domain-averaged Π and Λ evaluated at k = L / = 1. In both 2D and 3D, we find Π < 0 when domain-averaged, indicating a transfer of KE upscale from scales smaller than L to larger scales due to the ablative expansion and spread as the jet propagates. On the other hand, Λ > 0 in both 2D and 3D when domain-averaged, indicating downscale transfer. As discussed in Sec. I and in Refs. 32 and 34, energy by baropycnal work Λ arises from the release (or storage, if Λ < 0) of potential energy due to pressure gradients acting against density variations. Much of the pressure gradient arises near the ablation front as expected.

FIG. 12.

Domain-averaged deformation work, Π , and baropycnal work, Λ , across scale k = 1. In both 2D and 3D, Π transfers energy upscale and Λ transfers energy downscale at all times.

FIG. 12.

Domain-averaged deformation work, Π , and baropycnal work, Λ , across scale k = 1. In both 2D and 3D, Π transfers energy upscale and Λ transfers energy downscale at all times.

Close modal

Table IV shows that the large scale KE difference between 2D and 3D can be attributed to differences in their KE scale-transfer. Π and Λ represent energy transfer rates (in erg/s/cm3). Integrating Π and Λ over the duration of the simulation from t = 0 to 3.5 ns, we see in Table IV that there is 4.60 × 10 9 erg/cm3 excess energy received by the large scales ( > L = 0.25 cm) in 2D compared to 3D. This value matches the large scale KE difference of 4.80 × 10 9 erg/cm3 between 2D and 3D in Table III remarkably well,78 to within 4%. Since Π and Λ are the only terms that transfer KE across scales using our coarse-graining decomposition [Eq. (8)], we can infer that differences between 2D and 3D in the jet's bulk speed and the distance traveled (Fig. 6) are due to differences in KE scale-transfer. In the rest of this paper, we explore a physical origin for these differences in KE scale-transfer between 2D and 3D.

TABLE IV.

Domain-averaged KE scale-transfer Π and Λ at k = 1 integrated over the entire time duration (t = 0–3.5 ns).

Energy transferred ( erg / cm 3 ) Π d t Λ d t Π d t + Λ d t
2D  1.64 × 10 10  1.02 × 10 10  6.23 × 10 9 
3D  1.22 × 10 10  1.06 × 10 10  1.63 × 10 9 
Difference: 2D–3D  4.11 × 10 9  3.87 × 10 8  4.60 × 10 9 
Energy transferred ( erg / cm 3 ) Π d t Λ d t Π d t + Λ d t
2D  1.64 × 10 10  1.02 × 10 10  6.23 × 10 9 
3D  1.22 × 10 10  1.06 × 10 10  1.63 × 10 9 
Difference: 2D–3D  4.11 × 10 9  3.87 × 10 8  4.60 × 10 9 

From Fig. 6, there are notable differences between the 2D and 3D jets near the leading edge, where the flow is suggestive of turbulence. To objectively quantify the turbulence intensity, it is traditional to apply a Reynolds decomposition,23  u i = u i + u i , to separate the velocity field into a mean part u i and a fluctuating component u i , where is an ensemble average. The Reynolds stress tensor, u i u j , is typically used to quantify the turbulent fluctuations.22,23,79

However, there are a few disadvantages of such an approach that make it undesirable for our purpose. For one, it requires an ensemble average, which is computationally costly and in practice is often taken to be a temporal average if the flow is in a statistically steady state, which our jets are not, or a spatial average along a statistically homogeneous direction. While the 3D jet is statistically homogeneous in the azimuthal direction at any time, the 2D jet lacks any fluctuations along that direction by definition, making a comparison between 2D and 3D using a Reynolds decomposition unjustifiable. Another disadvantage of a Reynolds decomposition is its inability to inform us about the scales at which such turbulence exists.80,81 It is possible to use the analog of the Reynolds stress u i u j within the coarse-graining approach, which is the subscale stress (per unit mass) as shown by Germano,61,
(14)
The subscale stress is a rank-2 tensor, where each tensor component is a 2nd-order generalized central moment61 defined in Eq. (11). The subscale stress quantifies the momentum flux contribution from subscales,39,41 i.e., scales smaller than . Taking the length scale = L / 2 (i.e., k = 2), the subscale stress magnitude at 3.5 ns is shown in Fig. 13. Since the dominant velocity component is in the y-direction, the magnitude of τ ¯ ( u i , u j ) in both 2D and 3D is predominantly from by the tensor component τ ¯ ( u y , u y ). Comparing 2D and 3D in Fig. 13, we see that the high-intensity subscale stress occurs at the front of the jets but with a marked difference at the leading edge along the jet axis. The 2D jet has pronounced stress at the leading edge between approximately x = 0.05 and 0.05 cm, which is absent in the 3D jet. The structures giving rise to this stress will be discussed in Sec. VI B. The stress morphology in Fig. 13 justifies a focus on the lower half of the domain (y < 0.375 cm, highlighted boxes in Fig. 13) in an attempt to glean insight into differences in the KE scale-transfer between 2D and 3D.
FIG. 13.

Magnitude of (left) 2D and (right) 3D subscale stress, | τ ¯ ( u i , u j ) | = | ( u i u j ) ¯ ( u i ) ¯ ( u j ) ¯ |, at k = 2. Sub-domain outlined by a box is used to study the energy scale-transfer in the “turbulent” region.

FIG. 13.

Magnitude of (left) 2D and (right) 3D subscale stress, | τ ¯ ( u i , u j ) | = | ( u i u j ) ¯ ( u i ) ¯ ( u j ) ¯ |, at k = 2. Sub-domain outlined by a box is used to study the energy scale-transfer in the “turbulent” region.

Close modal

Applying the Helmholtz decomposition to the velocity, u = u d + u s, partitions it into a dilatational component, u d, and a solenoidal component, u s. Their curl-free ( × u d = 0) and divergence-free ( · u s = 0) properties82 allow us to explore the dominant flow components, such as shocks and vorticity, which may contribute to the jet traveling farther in 2D than in 3D. To perform the Helmholtz decomposition, the domain is mirrored to create periodic boundary conditions. While our choice for dealing with the boundary conditions by mirroring is not unique, we shall see that our results are physically meaningful. Dilatational and solenoidal specific kinetic energy (i.e., per unit mass), K d = u i d u i d and K s = u i s u i s, respectively, are shown in Figs. 14 and 15.

FIG. 14.

Dilatational kinetic energy, K d = u i d u i d. High-intensity Kd represents structures near the shock and have a similar shape between 2D and 3D. The shock front is marked by a horizontal black line.

FIG. 14.

Dilatational kinetic energy, K d = u i d u i d. High-intensity Kd represents structures near the shock and have a similar shape between 2D and 3D. The shock front is marked by a horizontal black line.

Close modal
FIG. 15.

Solenoidal kinetic energy, K s = u i s u i s. High-intensity Ks mainly appear at the leading edge of the jet and displays marked differences between 2D and 3D. Horizontal black lines are the same as in Fig. 14 and indicate the position of shock fronts.

FIG. 15.

Solenoidal kinetic energy, K s = u i s u i s. High-intensity Ks mainly appear at the leading edge of the jet and displays marked differences between 2D and 3D. Horizontal black lines are the same as in Fig. 14 and indicate the position of shock fronts.

Close modal

For both the 2D and 3D jets, we can see from Kd in Fig. 14 that there is a marked discontinuity (underscored by a horizontal black line in Fig. 14) at the same y-location where the velocity is discontinuous in Fig. 6. To leading order, the structure of Kd in Fig. 14 is similar between 2D and 3D aside from a difference in their y-location.

Relative to Kd, the solenoidal flow Ks in Fig. 15 shows obvious differences between 2D and 3D. The highest intensity solenoidal flow in 2D is more collimated along the jet axis compared to that in 3D. There is also significant post-shock ( y 0 to 0.13 cm) solenoidal flow activity near the leading edge of the 2D jet, which suggests the presence of vorticity that is absent near the leading edge of the 3D jet. This suggests that differences in the KE scale-transfer are significantly influenced by the solenoidal component in the turbulent region near the leading edge, which we shall now analyze.

Using the Helmholtz decomposition, we can investigate the contribution of the solenoidal and dilatational flow to Π and Λ, which we denote by Π sol , Π dil , Λ sol, and Λ dil. These are obtained by replacing u with u s or u d in Eqs. (9) and (10), respectively. Note that Π sol + Π dil Π since the decomposition also yields transfer terms involving both u s and u d, which we do not analyze here and focus instead on the pure solenoidal behavior.

The opposing signs of Π sol between 2D and 3D in Fig. 16 show the qualitative contrast of the dynamics. In 3D, deformation work from the vortical flow transfers KE downscale, Π sol > 0, whereas it transfers KE upscale in 2D, which sustains large-scale coherent vortical structures near the leading edge of the 2D jet as we shall see in Figs. 17–19.

FIG. 16.

Deformation work, Π sol , and baropycnal work, Λ sol , due to the solenoidal flow and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. The sign difference in Π sol shows the qualitative contrast of the dynamics between the 2D and 3D jets. In 3D, deformation work from the vortical flow transfers KE downscale, Π sol > 0, whereas it transfers KE upscale in 2D, which sustains large-scale coherent vortical structures near the leading edge of the 2D jet as seen in Figs. 17–19.

FIG. 16.

Deformation work, Π sol , and baropycnal work, Λ sol , due to the solenoidal flow and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. The sign difference in Π sol shows the qualitative contrast of the dynamics between the 2D and 3D jets. In 3D, deformation work from the vortical flow transfers KE downscale, Π sol > 0, whereas it transfers KE upscale in 2D, which sustains large-scale coherent vortical structures near the leading edge of the 2D jet as seen in Figs. 17–19.

Close modal
FIG. 17.

Visualization of solenoidal deformation work, Π sol, in the turbulent domain ( y < 0.375 cm) at k = 2. Marked differences can be seen at the leading edge where Π sol has negative values in 2D panel (a) at y 0.05 cm, which are absent in 3D panel (b).

FIG. 17.

Visualization of solenoidal deformation work, Π sol, in the turbulent domain ( y < 0.375 cm) at k = 2. Marked differences can be seen at the leading edge where Π sol has negative values in 2D panel (a) at y 0.05 cm, which are absent in 3D panel (b).

Close modal
FIG. 18.

Correlating coarse velocity gradients, y ( u s ̃ ) y (red-blue colormap), with subscale stress, τ ̃ ( u y s , u y s ) (gray-scale map), at scale k = 2. Since subscale stress is positive semi-definite, the gray-scale map highlights regions of strong scale-transfer. The main differences are seen at the flanks of the 2D jet's leading edge, | x | = 0.03 to 0.05 cm and y < 0.1 cm, where y ( u s ̃ ) y is mostly positive, implying an upscale transfer in 2D as seen in Fig. 17 that is absent in 3D.

FIG. 18.

Correlating coarse velocity gradients, y ( u s ̃ ) y (red-blue colormap), with subscale stress, τ ̃ ( u y s , u y s ) (gray-scale map), at scale k = 2. Since subscale stress is positive semi-definite, the gray-scale map highlights regions of strong scale-transfer. The main differences are seen at the flanks of the 2D jet's leading edge, | x | = 0.03 to 0.05 cm and y < 0.1 cm, where y ( u s ̃ ) y is mostly positive, implying an upscale transfer in 2D as seen in Fig. 17 that is absent in 3D.

Close modal
FIG. 19.

Correlating coarse velocity gradients, y ( u s ̃ ) y (red-blue colormap), with streamlines (black) superposed. The coherent vortical roll-up structure at the leading edge of the 2D jet is absent from the 3D jet. The roll-up in the 2D jet at y 0.05 cm is co-located with the upscale energy transfer seen at that same location in Fig. 17, indicating that the roll-up in 2D is energized by the smaller scale turbulence.

FIG. 19.

Correlating coarse velocity gradients, y ( u s ̃ ) y (red-blue colormap), with streamlines (black) superposed. The coherent vortical roll-up structure at the leading edge of the 2D jet is absent from the 3D jet. The roll-up in the 2D jet at y 0.05 cm is co-located with the upscale energy transfer seen at that same location in Fig. 17, indicating that the roll-up in 2D is energized by the smaller scale turbulence.

Close modal

The upscale transfer of KE in 2D by Π sol in Fig. 16 is similar to the upscale cascade seen in 2D variable density Rayleigh-Taylor turbulence.32 It is also similar to 2D constant-density turbulence30,31,83 due to the absence of vortex stretching and strain self-amplification.28 Even though the magnitudes of Π sol and Λ sol in Fig. 16 are smaller than total Π and Λ over the entire domain in Fig. 12, the vorticity dynamics at the leading edge can have a disproportionate effect on the jet's traveling distance and bulk flow evolution as we shall discuss below.

To understand the mechanistic cause of differences in Π sol between 2D and 3D, we probe its tensorial components. Figure 17 visualizes Π sol at k = 2. We can see obvious differences between 2D and 3D in the sign of Π sol at the leading edge of the jet, which seem to correlate with differences seen in Fig. 15. As defined in Eq. (9), Π sol is a contraction between the velocity gradient tensor, j ( u s ̃ ) i, and the subscale stress tensor, ρ ¯ τ ̃ ( u i s , u j s ). To understand the sign of Π sol, we focus on the tensor components i , j = 2 (y-components), which make the dominant contribution to the spatially averaged Π sol.

The subscale stress component ρ ¯ τ ̃ ( u y s , u y s ) is mathematically guaranteed to be positive semi-definite at every point in the domain.34,77,84 This is shown in Fig. 18. Therefore, the sign of Π sol from the y-components is the same as that of y ( u s ̃ ) y. Figure 18 visualizes y ( u s ̃ ) y with red/blue indicating positive/negative values, implying upscale/downscale KE transfer ( Π sol from i , j = 2 components is negative/positive). Overlapping the colormap in Fig. 18 is a gray-scale map of τ ̃ ( u y s , u y s ), which, being positive semi-definite, highlights regions of strong sub-scale stress and, therefore, strong scale-transfer.

Both panels of Fig. 18 at y > 0.15 cm are roughly similar to leading order in regions of strong stress. In those highlighted regions at the jet flanks at y > 0.15 cm, we see that y ( u s ̃ ) y is mostly negative, implying downscale transfer in both 2D and 3D. The main differences are seen at the flanks of the 2D jet's leading edge, | x | = 0.03 to 0.05 cm and y < 0.1 cm, where y ( u s ̃ ) y is mostly positive, implying an upscale transfer in 2D that is absent in 3D.

At the flanks of the leading edge ( | x | = 0.03 to 0.05 cm) in 2D, we see in Fig. 18 that y ( u s ̃ ) y switches sign along the y-direction between y 0.05 and 0.1 cm. This is suggestive of vortical motion. Figure 19 visualizes the flow streamlines superposed over y ( u s ̃ ) y. From the streamlines in Fig. 19, we can see clearly two coherent vortical structures at the leading edge of the 2D jet, which are absent from the 3D jet. The streamlines in Fig. 19 are consistent with observations we made above that the flow at y > 0.15 cm is roughly similar between 2D and 3D (see also Fig. 29 in  Appendix). The main difference is at y < 0.15 cm, where the coherent roll-up structure in the 2D jet at y 0.05 cm is co-located with the upscale energy transfer seen at that same location in Fig. 17. This is indicative of the coherent roll-up being energized by the smaller scale turbulence created post-shock. Such an upscale scale-transfer is absent in 3D, which cannot sustain the coherent roll-up. Vorticity within this roll-up creates an effective velocity drift84 that helps propel the 2D jet farther and keep it collimated relative to its 3D counterpart.

In this paper, we demonstrated a methodology for diagnosing the multiscale dynamics and energy transfer in complex HED flows with realistic driving and boundary conditions. While it is well known that 2D modeling underestimates the proneness of the flow to instabilities, what we have shown here is that 2D modeling also suffers from significant spurious energization of the bulk flow by instabilities. The energization of the roll-up structures via an upscale transfer from smaller scale turbulence in Figs. 17–19 brings to the fore some of the hydrodynamics artifacts associated with 2D modeling, reinforcing recent findings.32 

In HED applications such as ICF, 2D simulations remain to be the main “work horse” for experimental design17–19 as routine 3D simulations are prohibitively expensive.3 Our hope from this work is to highlight the tradeoff between 2D and 3D flow physics, which may not be as widely appreciated as tradeoffs from approximating other system components such as the laser drive or the hohlraum. While computational realities may prevent the community from routinely conducting 3D hydrodynamic modeling in the foreseeable future, we hope that this work (see also Ref. 32) highlights the need to alleviate some of the hydrodynamic artifacts associated with 2D models.

By demonstrating the applicability of coarse-graining for comparing the multiscale dynamics and energy transfer between 2D and 3D, we have shown that this methodology can help with inter-model comparison and validation. We believe that future attempts at alleviating some of the 2D hydrodynamic artifacts would have to use this approach, at least in some fashion, for testing and model development. After all, the approach has plenty in common with large eddy simulation modeling, which is a well-established field in fluid dynamics.39 

This work was supported by CMAP, an NSF Physics Frontiers Center, under Grant No. PHY-2020249. Partial support from Grant Nos. DE-SC0020229, DE-SC0019329, and CBET-2143702 is also acknowledged. H.A. was also supported by US DOE Grant Nos. DE-SC0014318, DE-SC0019329, US NSF Grant Nos. OCE-2123496, PHY-2206380, US NASA Grant No. 80NSSC18K0772, and US NNSA Grant Nos. DE-NA0003856, DE-NA0003914, DE-NA0004134. J.S. was supported by Nos. DE-SC0019329, DE-NA0003914, and DE-NA0004134. Computing time was provided by NERSC under Contract No. DE-AC02–05CH11231.

The authors have no conflicts to disclose.

Hao Yin: Conceptualization (equal); Data curation (equal); Formal analysis (lead); Investigation (lead); Methodology (equal); Writing – original draft (lead); Writing – review & editing (equal). Jessica K. Shang: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Supervision (lead); Writing – review & editing (equal). Eric G. Blackman: Conceptualization (equal); Supervision (equal); Writing – review & editing (equal). Gilbert Collins: Funding acquisition (lead); Project administration (lead); Resources (equal); Writing – review & editing (equal). Hussein Aluie: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (lead); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

Figure 20 shows that the average traveling speed converges at refinement level 5, which is chosen for the runtime parameter stated in simulation configurations in Sec. II A.

FIG. 20.

Convergence study taking average traveling speed (the ratio of the traveling distance, characterized by G y centroid, to time) at different refinement levels. Average traveling speed converges at refinement level 5.

FIG. 20.

Convergence study taking average traveling speed (the ratio of the traveling distance, characterized by G y centroid, to time) at different refinement levels. Average traveling speed converges at refinement level 5.

Close modal

Figure 21 provides two compressibility metrics supplemental to Sec. II D, where the 2D jet presents higher compressibility levels at the end of the simulation ( t = 3.5 ns). Figure 21(a) is the bulk Mach number that compares the visualization of the Mach number in Fig. 9. A difference of 12.1 % between 2D and 3D occurs at the end of the simulation ( t = 3.5 ns). Figure 21(b) is the root mean square (rms) ratio of divergence of velocity [Fig. 10(a)] to curl of velocity [Fig. 10(b)]. The ratios for 2D and 3D are both below one implying the significance of the divergence-free part of the flow even with the strong shock in the domain.

FIG. 21.

Compressibility metrics in (a) the bulk Mach number, M, and (b) the root mean square (rms) ratio of divergence of velocity to curl of velocity. The compressibility levels in 2D are larger than in 3D at the end of the simulation (t = 3.5 ns).

FIG. 21.

Compressibility metrics in (a) the bulk Mach number, M, and (b) the root mean square (rms) ratio of divergence of velocity to curl of velocity. The compressibility levels in 2D are larger than in 3D at the end of the simulation (t = 3.5 ns).

Close modal

Figures 22–24 compare the coarse-grained fields with two boundary conditions: zero-gradient boundary and mirrored boundary. Figure 22 visualizes the coarse y-velocity, and Fig. 23 visualizes the y-gradient of coarse y-velocity. Figure 24 compares a lineout for Figs. 22 and 23 and demonstrates the differences between the two boundary conditions are negligible. The mirrored boundary is selected to be the boundary condition for the simulations described in Sec. III A.

FIG. 22.

Comparing coarse y-velocity, u ¯ y, with different boundary conditions at k = 1 at t = 3.5 ns, where the magnitudes are almost identical.

FIG. 22.

Comparing coarse y-velocity, u ¯ y, with different boundary conditions at k = 1 at t = 3.5 ns, where the magnitudes are almost identical.

Close modal
FIG. 23.

Comparing y-gradient of coarse y-velocity, y u ¯ y, with different boundary conditions at k = 1 at t = 3.5 ns. Compared to Fig. 22, differences are amplified close to the boundaries due to the derivatives.

FIG. 23.

Comparing y-gradient of coarse y-velocity, y u ¯ y, with different boundary conditions at k = 1 at t = 3.5 ns. Compared to Fig. 22, differences are amplified close to the boundaries due to the derivatives.

Close modal
FIG. 24.

Comparison of zero-gradient and mirrored boundary conditions at k = 1 and t = 3.5 ns along the axis x = 0 from y = 0 to y = 0.375 cm. Differences occur near the boundaries and are negligible.

FIG. 24.

Comparison of zero-gradient and mirrored boundary conditions at k = 1 and t = 3.5 ns along the axis x = 0 from y = 0 to y = 0.375 cm. Differences occur near the boundaries and are negligible.

Close modal

Figure 25 shows the evolution of dominant component (y-component) of the kinetic energy at various scales. There is more kinetic energy in 2D at large scales compare to 3D, which is consistent with the evolution of total kinetic energy in Fig. 11.

FIG. 25.

The y-component of domain-averaged coarse kinetic energy, K E y ¯ = ρ u y ̃ 2 / 2 , at all scales > . K E y ¯ is the dominant component in total K E ¯ , and large scales in 2D still possess more KE than in 3D like in Fig. 11.

FIG. 25.

The y-component of domain-averaged coarse kinetic energy, K E y ¯ = ρ u y ̃ 2 / 2 , at all scales > . K E y ¯ is the dominant component in total K E ¯ , and large scales in 2D still possess more KE than in 3D like in Fig. 11.

Close modal

Figure 26 provides that, at t = 3.5 ns, Π transfers energy upscale and Λ mainly transfers energy downscale as discussed in Sec. IV B.

FIG. 26.

Visualization of energy scale-transfer by Π and Λ at k = 1 and t = 3.5 ns. Π transfers energy upscale and Λ mainly transfers energy downscale. Π and Λ appear localized near the target due to the dominant density contribution.

FIG. 26.

Visualization of energy scale-transfer by Π and Λ at k = 1 and t = 3.5 ns. Π transfers energy upscale and Λ mainly transfers energy downscale. Π and Λ appear localized near the target due to the dominant density contribution.

Close modal

Figure 27 shows the deformation work and baropycnal work due to the dilatational flow at t = 3.5 ns. Energy transfer directions are the same in 2D and 3D for both Π dil and Λ dil , which follows the dynamics in homogeneous compressible turbulence.42 

FIG. 27.

Deformation work, Π dil , and baropycnal work, Λ dil , due to the dilatational flow and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. Energy transfer directions are the same in 2D and 3D for both Π dil and Λ dil , which follows the dynamics in homogeneous compressible turbulence.

FIG. 27.

Deformation work, Π dil , and baropycnal work, Λ dil , due to the dilatational flow and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. Energy transfer directions are the same in 2D and 3D for both Π dil and Λ dil , which follows the dynamics in homogeneous compressible turbulence.

Close modal

Figure 28 shows the components of deformation work, Π , calculated with cross terms at t = 3.5 ns. All components transfer energy downscale in both 2D and 3D, highlighting the contrast of dynamics of Π sol discussed in Sec. VI A.

FIG. 28.

Components of deformation work, Π , calculated with cross terms and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. All components transfer energy downscale in both 2D and 3D, highlighting the contrast of dynamics of Π sol shown in Fig. 16.

FIG. 28.

Components of deformation work, Π , calculated with cross terms and averaged over the subdomain ( y < 0.375 cm) highlighted in Fig. 13, plotted here as a function of filtering wavenumber k at t = 3.5 ns. All components transfer energy downscale in both 2D and 3D, highlighting the contrast of dynamics of Π sol shown in Fig. 16.

Close modal

Figure 29 overlaps the coarse vorticity and streamlines to show the correlation between the high-intensity vorticity and the generation of the circulations discussed in Sec. VI B.

FIG. 29.

Overlapping coarse vorticity ω ¯ = × u ¯ and streamlines calculated for the lower half of the domain (y < 0.375 cm) at t = 3.5 ns. The ω ¯ is at scale k = 2, and the streamlines are generated from the u ̃ for the band between k = 1 and k = 2. The overlaps imply the correlation between the high-intensity vorticity and the generation of the circulations.

FIG. 29.

Overlapping coarse vorticity ω ¯ = × u ¯ and streamlines calculated for the lower half of the domain (y < 0.375 cm) at t = 3.5 ns. The ω ¯ is at scale k = 2, and the streamlines are generated from the u ̃ for the band between k = 1 and k = 2. The overlaps imply the correlation between the high-intensity vorticity and the generation of the circulations.

Close modal
1.
B. E.
Blue
,
S. V.
Weber
,
S. G.
Glendinning
,
N. E.
Lanier
,
D. T.
Woods
,
M. J.
Bono
,
S. N.
Dixit
,
C. A.
Haynam
,
J. P.
Holder
,
D. H.
Kalantar
,
B. J.
MacGowan
,
A. J.
Nikitin
,
V. V.
Rekow
,
B. M.
Van Wonterghem
,
E. I.
Moses
,
P. E.
Stry
,
B. H.
Wilde
,
W. W.
Hsing
, and
H. F.
Robey
, “
Experimental investigation of high-Mach-number 3D hydrodynamic jets at the National Ignition Facility
,”
Phys. Rev. Lett.
94
,
095005
(
2005
).
2.
A. M.
Khokhlov
,
P. A.
Höflich
,
E. S.
Oran
,
J. C.
Wheeler
,
L.
Wang
, and
A. Y.
Chtchelkanova
, “
Jet-induced explosions of core collapse supernovae
,”
Astrophys. J.
524
,
L107
(
1999
).
3.
D. S.
Clark
,
C. R.
Weber
,
J. L.
Milovich
,
J. D.
Salmonson
,
A. L.
Kritcher
,
S. W.
Haan
,
B. A.
Hammel
,
D. E.
Hinkel
,
O. A.
Hurricane
,
O. S.
Jones
,
M. M.
Marinak
,
P. K.
Patel
,
H. F.
Robey
,
S. M.
Sepke
, and
M. J.
Edwards
, “
Three-dimensional simulations of low foot and high foot implosion experiments on the National Ignition Facility
,”
Phys. Plasmas
23
,
056302
(
2016
).
4.
A.
Zylstra
,
D.
Casey
,
A.
Kritcher
,
L.
Pickworth
,
B.
Bachmann
,
K.
Baker
,
J.
Biener
,
T.
Braun
,
D.
Clark
,
V.
Geppert-Kleinrath
et al, “
Hot-spot mix in large-scale HDC implosions at NIF
,”
Phys. Plasmas
27
,
092709
(
2020
).
5.
B. M.
Haines
,
J. P.
Sauppe
,
B. J.
Albright
,
W. S.
Daughton
,
S. M.
Finnegan
,
J. L.
Kline
, and
J. M.
Smidt
, “
A mechanism for reduced compression in indirectly driven layered capsule implosions
,”
Phys. Plasmas
29
,
042704
(
2022
).
6.
S. R.
Goldman
,
S. E.
Caldwell
,
M. D.
Wilke
,
D. C.
Wilson
,
C. W.
Barnes
,
W. W.
Hsing
,
N. D.
Delamater
,
G. T.
Schappert
,
J. W.
Grove
,
E. L.
Lindman
,
J. M.
Wallace
,
R. P.
Weaver
,
A. M.
Dunne
,
M. J.
Edwards
,
P.
Graham
, and
B. R.
Thomas
, “
Shock structuring due to fabrication joints in targets
,”
Phys. Plasmas
6
,
3327
3336
(
1999
).
7.
S. R.
Goldman
,
C. W.
Barnes
,
S. E.
Caldwell
,
D. C.
Wilson
,
S. H.
Batha
,
J. W.
Grove
,
M. L.
Gittings
,
W. W.
Hsing
,
R. J.
Kares
,
K. A.
Klare
,
G. A.
Kyrala
,
R. W.
Margevicius
,
R. P.
Weaver
,
M. D.
Wilke
,
A. M.
Dunne
,
M. J.
Edwards
,
P.
Graham
, and
B. R.
Thomas
, “
Production of enhanced pressure regions due to inhomogeneities in inertial confinement fusion targets
,”
Phys. Plasmas
7
,
2007
2013
(
2000
).
8.
Y.
Zhou
, “
Rayleigh–Taylor and Richtmyer–Meshkov instability induced flow, turbulence, and mixing—I
,”
Phys. Rep.
720
,
1
136
(
2017
).
9.
Y.
Zhou
, “
Rayleigh–Taylor and Richtmyer–Meshkov instability induced flow, turbulence, and mixing—II
,”
Phys. Rep.
723
,
1
160
(
2017
).
10.
T.
Foglizzo
,
R.
Kazeroni
,
J.
Guilet
,
F.
Masset
,
M.
González
,
B. K.
Krueger
,
J.
Novak
,
M.
Oertel
,
J.
Margueron
,
J.
Faure
et al, “
The explosion mechanism of core-collapse supernovae: Progress in supernova theory and experiments
,”
Publ. Astron. Soc. Aust.
32
,
e009
(
2015
).
11.
H.-T.
Janka
,
T.
Melson
, and
A.
Summa
, “
Physics of core-collapse supernovae in three dimensions: A sneak preview
,”
Annu. Rev. Nucl. Part. Sci.
66
,
341
375
(
2016
).
12.
D.
Radice
,
E.
Abdikamalov
,
C. D.
Ott
,
P.
Mösta
,
S. M.
Couch
, and
L. F.
Roberts
, “
Turbulence in core-collapse supernovae
,”
J. Phys. G
45
,
053003
(
2018
).
13.
F.
Hanke
,
B.
Müller
,
A.
Wongwathanarat
,
A.
Marek
, and
H.-T.
Janka
, “
Sasi activity in three-dimensional neutrino-hydrodynamics simulations of supernova cores
,”
Astrophys. J.
770
,
66
(
2013
).
14.
I.
Tamborra
,
F.
Hanke
,
H.-T.
Janka
,
B.
Müller
,
G. G.
Raffelt
, and
A.
Marek
, “
Self-sustained asymmetry of lepton-number emission: A new phenomenon during the supernova shock-accretion phase in three dimensions
,”
Astrophys. J.
792
,
96
(
2014
).
15.
S. M.
Couch
and
C. D.
Ott
, “
The role of turbulence in neutrino-driven core-collapse supernova explosions
,”
Astrophys. J.
799
,
5
(
2015
).
16.
R.
Betti
and
O. A.
Hurricane
, “
Inertial-confinement fusion with lasers
,”
Nat. Phys.
12
,
435
448
(
2016
).
17.
D. J.
Schlossberg
,
G. P.
Grim
,
D. T.
Casey
,
A. S.
Moore
,
R.
Nora
,
B.
Bachmann
,
L. R.
Benedetti
,
R. M.
Bionta
,
M. J.
Eckart
,
J. E.
Field
et al, “
Observation of hydrodynamic flows in imploding fusion plasmas on the national ignition facility
,”
Phys. Rev. Lett.
127
,
125001
(
2021
).
18.
A. L.
Kritcher
,
C. V.
Young
,
H. F.
Robey
,
C. R.
Weber
,
A. B.
Zylstra
,
O. A.
Hurricane
,
D. A.
Callahan
,
J. E.
Ralph
,
J. S.
Ross
,
K. L.
Baker
et al, “
Design of inertial fusion implosions reaching the burning plasma regime
,”
Nat. Phys.
18
,
251
258
(
2022
).
19.
A. L.
Kritcher
,
A. B.
Zylstra
,
D. A.
Callahan
,
O. A.
Hurricane
,
C. R.
Weber
,
D. S.
Clark
,
C. V.
Young
,
J. E.
Ralph
,
D. T.
Casey
,
A.
Pak
et al, “
Design of an inertial fusion experiment exceeding the Lawson criterion for ignition
,”
Phys. Rev. E
106
,
025201
(
2022
).
20.
B. M.
Haines
,
R. C.
Shah
,
J. M.
Smidt
,
B. J.
Albright
,
T.
Cardenas
,
M. R.
Douglas
,
C.
Forrest
,
V. Y.
Glebov
,
M. A.
Gunderson
,
C. E.
Hamilton
,
K. C.
Henderson
,
Y.
Kim
,
M. N.
Lee
,
T. J.
Murphy
,
J. A.
Oertel
,
R. E.
Olson
,
B. M.
Patterson
,
R. B.
Randolph
, and
D. W.
Schmidt
, “
Observation of persistent species temperature separation in inertial confinement fusion mixtures
,”
Nat. Commun.
11
,
544
(
2020
).
21.
M.
Gatu Johnson
,
B.
Haines
,
P.
Adrian
,
C.
Forrest
,
J.
Frenje
,
V.
Glebov
,
W.
Grimble
,
R.
Janezic
,
J.
Knauer
,
B.
Lahmann
,
F.
Marshall
,
T.
Michel
,
F.
Séguin
,
C.
Stoeckl
, and
R.
Petrasso
, “
3D xRAGE simulation of inertial confinement fusion implosion with imposed mode 2 laser drive asymmetry
,”
High Energy Density Phys.
36
,
100825
(
2020
).
22.
H.
Tennekes
and
J. L.
Lumley
,
A First Course in Turbulence
(
The MIT Press
,
Cambridge, MA
,
1972
).
23.
S. B.
Pope
,
Turbulent Flows
(
Cambridge University Press
,
New York
,
2000
).
24.
V.
Borue
and
S. A.
Orszag
, “
Local energy flux and subgrid-scale statistics in three-dimensional turbulence
,”
J. Fluid Mech.
366
,
1
31
(
1998
).
25.
G. L.
Eyink
, “
Cascade of circulations in fluid turbulence
,”
Phys. Rev. E
74
,
066302
(
2006
).
26.
G. L.
Eyink
, “
Dissipative anomalies in singular Euler flows
,”
Physica D
237
,
1956
1968
(
2008
).
27.
H.
Xu
,
A.
Pumir
, and
E.
Bodenschatz
, “
The pirouette effect in turbulent flows
,”
Nat. Phys.
7
,
709
712
(
2011
).
28.
P. L.
Johnson
, “
Energy transfer from large to small scales in turbulence by multiscale nonlinear strain and vorticity interactions
,”
Phys. Rev. Lett.
124
,
104501
(
2020
).
29.
R. H.
Kraichnan
, “
Inertial ranges in two-dimensional turbulence
,”
Phys. Fluids
10
,
1417
1423
(
1967
).
30.
Z.
Xiao
,
M.
Wan
,
S.
Chen
, and
G. L.
Eyink
, “
Physical mechanism of the inverse energy cascade of two-dimensional turbulence: A numerical investigation
,”
J. Fluid Mech.
619
,
1
44
(
2009
).
31.
G.
Boffetta
and
R. E.
Ecke
, “
Two-dimensional turbulence
,”
Annu. Rev. Fluid Mech.
44
,
427
451
(
2012
).
32.
D.
Zhao
,
R.
Betti
, and
H.
Aluie
, “
Scale interactions and anisotropy in Rayleigh-Taylor turbulence
,”
J. Fluid Mech.
930
,
A29
(
2022
).
33.
H.
Aluie
, “
Compressible turbulence: The cascade and its locality
,”
Phys. Rev. Lett.
106
,
174502
(
2011
).
34.
H.
Aluie
, “
Scale decomposition in compressible turbulence
,”
Physica D
247
,
54
65
(
2013
).
35.
H.
Aluie
,
S.
Li
, and
H.
Li
, “
Conservative cascade of kinetic energy in compressible turbulence
,”
Astrophys. J.
751
,
L29
(
2012
).
36.
A. G.
Kritsuk
,
R.
Wagner
, and
M. L.
Norman
, “
Energy cascade and scaling in supersonic isothermal turbulence
,”
J. Fluid Mech.
729
,
R1
(
2013
).
37.
J.
Wang
,
Y.
Yang
,
Y.
Shi
,
Z.
Xiao
,
X. T.
He
, and
S.
Chen
, “
Cascade of kinetic energy in three-dimensional compressible turbulence
,”
Phys. Rev. Lett.
110
,
214505
(
2013
).
38.
G. L.
Eyink
and
T. D.
Drivas
, “
Cascades and dissipative anomalies in compressible fluid turbulence
,”
Phys. Rev. X
8
,
011022
(
2018
).
39.
C.
Meneveau
and
J.
Katz
, “
Scale-invariance and turbulence models for large-eddy simulation
,”
Ann. Rev. Fluid Mech.
32
,
1
32
(
2000
).
40.
G. L.
Eyink
, “
Locality of turbulent cascades
,”
Physica D
207
,
91
116
(
2005
).
41.
G. L.
Eyink
and
H.
Aluie
, “
Localness of energy cascade in hydrodynamic turbulence—I: Smooth coarse graining
,”
Phys. Fluids
21
,
115107
(
2009
).
42.
A.
Lees
and
H.
Aluie
, “
Baropycnal work: A mechanism for energy transfer across scales
,”
Fluids
4
,
92
(
2019
).
43.
A. S.
Liao
,
S.
Li
,
H.
Li
,
K.
Flippo
,
D.
Barnak
,
K. V.
Kelso
,
C.
Fiedler Kawaguchi
,
A.
Rasmus
,
S.
Klein
,
J.
Levesque
,
C.
Kuranz
, and
C.
Li
, “
Design of a new turbulent dynamo experiment on the OMEGA-EP
,”
Phys. Plasmas
26
,
032306
(
2019
).
44.
L.
Waxer
,
D.
Maywar
,
J.
Kelly
,
T.
Kessler
,
B.
Kruschwitz
,
S.
Loucks
,
R.
McCrory
,
D.
Meyerhofer
,
S.
Morse
,
C.
Stoeckl
, and
J.
Zuegel
, “
High-energy petawatt capability for the omega laser
,”
Opt. Photonics News
16
,
30
36
(
2005
).
45.
D. N.
Maywar
,
J. H.
Kelly
,
L. J.
Waxer
,
S. F. B.
Morse
,
I. A.
Begishev
,
J.
Bromage
,
C.
Dorrer
,
J. L.
Edwards
,
L.
Folnsbee
,
M. J.
Guardalben
,
S. D.
Jacobs
,
R.
Jungquist
,
T. J.
Kessler
,
R. W.
Kidder
,
B. E.
Kruschwitz
,
S. J.
Loucks
,
J. R.
Marciante
,
R. L.
McCrory
,
D. D.
Meyerhofer
,
A. V.
Okishev
,
J. B.
Oliver
,
G.
Pien
,
J.
Qiao
,
J.
Puth
,
A. L.
Rigatti
,
A. W.
Schmid
,
M. J.
Shoup
,
C.
Stoeckl
,
K. A.
Thorp
, and
J. D.
Zuegel
, “
OMEGA EP high-energy petawatt laser: Progress and prospects
,”
J. Phys.: Conf. Ser.
112
,
032007
(
2008
).
46.
B.
Fryxell
,
K.
Olson
,
P.
Ricker
,
F. X.
Timmes
,
M.
Zingale
,
D. Q.
Lamb
,
P.
MacNeice
,
R.
Rosner
,
J. W.
Truran
, and
H.
Tufo
, “
FLASH: An adaptive mesh hydrodynamics code for modeling astrophysical thermonuclear flashes
,”
Astrophys. J. Suppl. Ser.
131
,
273
334
(
2000
).
47.
P.
Tzeferacos
,
A.
Rigby
,
A. F. A.
Bott
,
A. R.
Bell
,
R.
Bingham
,
A.
Casner
,
F.
Cattaneo
,
E. M.
Churazov
,
J.
Emig
,
F.
Fiuza
,
C. B.
Forest
,
J.
Foster
,
C.
Graziani
,
J.
Katz
,
M.
Koenig
,
C. K.
Li
,
J.
Meinecke
,
R.
Petrasso
,
H. S.
Park
,
B. A.
Remington
,
J. S.
Ross
,
D.
Ryu
,
D.
Ryutov
,
T. G.
White
,
B.
Reville
,
F.
Miniati
,
A. A.
Schekochihin
,
D. Q.
Lamb
,
D. H.
Froula
, and
G.
Gregori
, “
Laboratory evidence of dynamo amplification of magnetic fields in a turbulent plasma
,”
Nat. Commun.
9
,
591
(
2018
).
48.
P.
Tzeferacos
,
M.
Fatenejad
,
N.
Flocke
,
C.
Graziani
,
G.
Gregori
,
D. Q.
Lamb
,
D.
Lee
,
J.
Meinecke
,
A.
Scopatz
, and
K.
Weide
, “
Flash MHD simulations of experiments that study shock-generated magnetic fields
,”
High Energy Density Phys.
17
(
Part A
),
24
31
(
2015
).
49.
P.
Tzeferacos
,
A.
Rigby
,
A.
Bott
,
A. R.
Bell
,
R.
Bingham
,
A.
Casner
,
F.
Cattaneo
,
E. M.
Churazov
,
J.
Emig
,
N.
Flocke
,
F.
Fiuza
,
C. B.
Forest
,
J.
Foster
,
C.
Graziani
,
J.
Katz
,
M.
Koenig
,
C. K.
Li
,
J.
Meinecke
,
R.
Petrasso
,
H. S.
Park
,
B. A.
Remington
,
J. S.
Ross
,
D.
Ryu
,
D.
Ryutov
,
K.
Weide
,
T. G.
White
,
B.
Reville
,
F.
Miniati
,
A. A.
Schekochihin
,
D. H.
Froula
,
G.
Gregori
, and
D. Q.
Lamb
, “
Numerical modeling of laser-driven experiments aiming to demonstrate magnetic field amplification via turbulent dynamo
,”
Phys. Plasmas, Phys. Plasmas
24
,
041404
(
2017
).
50.
FLASH Center
,
Flash User's Guide
(
FLASH Center for Computational Science
,
2019
).
51.
J.
Macfarlane
, “
IONMIX—A code for computing the equation of state and radiative properties of LTE and non-LTE plasmas
,”
Comput. Phys. Commun.
56
,
259
278
(
1989
).
52.
P.
MacNeice
,
K. M.
Olson
,
C.
Mobarry
,
R.
de Fainchtein
, and
C.
Packer
, “
PARAMESH: A parallel adaptive mesh refinement community toolkit
,”
Comput. Phys. Commun.
126
,
330
354
(
2000
).
53.
J. M.
Foster
,
B. H.
Wilde
,
P. A.
Rosen
,
T. S.
Perry
,
M.
Fell
,
M. J.
Edwards
,
B. F.
Lasinski
,
R. E.
Turner
, and
M. L.
Gittings
, “
Supersonic jet and shock interactions
,”
Phys. Plasmas
9
,
2251
2263
(
2002
).
54.
C.
Meneveau
, “
Statistics of turbulence subgrid-scale stresses: Necessary conditions and experimental tests
,”
Phys. Fluids
6
,
815
833
(
1994
).
55.
G. L.
Eyink
, “
Local energy flux and the refined similarity hypothesis
,”
J. Stat. Phys.
78
,
335
351
(
1995
).
56.
M.
Buzzicotti
,
M.
Linkmann
,
H.
Aluie
,
L.
Biferale
,
J.
Brasseur
, and
C.
Meneveau
, “
Effect of filter type on the statistics of energy transfer between resolved and subfilter scales from a-priori analysis of direct numerical simulations of isotropic turbulence
,”
J. Turbul.
19
,
167
197
(
2018
).
57.
S.
Du
,
H.
Li
,
X.
Fu
, and
Z.
Gan
, “
Anisotropic energy transfer and conversion in magnetized compressible turbulence
,”
Astrophys. J.
948
,
72
(
2023
).
58.
R. S.
Strichartz
,
A Guide to Distribution Theory and Fourier Transforms
(
World Scientific Publishing Company
,
2003
).
59.
L. C.
Evans
,
Partial Differential Equations
(
American Mathematical Society
,
2010
).
60.
A.
Leonard
, “
Energy cascade in large-eddy simulations of turbulent fluid flows
,”
Adv. Geophys.
18
,
237
(
1974
).
61.
M.
Germano
, “
Turbulence: The filtering approach
,”
J. Fluid Mech.
238
,
325
336
(
1992
).
62.
H.
Aluie
, “
Coarse-grained incompressible magnetohydrodynamics: Analyzing the turbulent cascades
,”
New J. Phys.
19
,
025008
(
2017
).
63.
I.
Daubechies
,
Ten Lectures on Wavelets
(
Society for Industrial and Applied Mathematics
,
1992
)
64.
H.
Aluie
,
M.
Hecht
, and
G. K.
Vallis
, “
Mapping the energy cascade in the North Atlantic Ocean: The coarse-graining approach
,”
J. Phys. Oceanogr.
48
,
225
244
(
2018
).
65.
D.
Zhao
and
H.
Aluie
, “
Inviscid criterion for decomposing scales
,”
Phys. Rev. Fluids
3
,
054603
(
2018
).
66.
B. A.
Storer
,
M.
Buzzicotti
,
H.
Khatri
,
S. M.
Griffies
, and
H.
Aluie
, “
Global energy spectrum of the general oceanic circulation
,”
Nat. Commun.
13
,
5314
(
2022
).
67.
P.
Chassaing
, “
An alternative formulation of the equations of turbulent motion for a fluid of variable density
,”
J. Mec. Theor. Appl.
4
,
375
389
(
1985
).
68.
D. J.
Bodony
and
S. K.
Lele
, “
On using large-eddy simulation for the prediction of noise from cold and heated turbulent jets
,”
Phys. Fluids
17
,
085103
(
2005
).
69.
G. C.
Burton
, “
Study of ultrahigh Atwood-number Rayleigh-Taylor mixing dynamics using the nonlinear large-eddy simulation method
,”
Phys. Fluids
23
,
045106
(
2011
).
70.
M.
Karimi
and
S. S.
Girimaji
, “
Influence of orientation on the evolution of small perturbations in compressible shear layers with inflection points
,”
Phys. Rev. E
95
,
033112
(
2017
).
71.
S.
Kida
and
S. A.
Orszag
, “
Energy and spectral dynamics in forced compressible turbulence
,”
J. Sci. Comput.
5
,
85
125
(
1990
).
72.
A. W.
Cook
and
Y.
Zhou
, “
Energy transfer in Rayleigh-Taylor instability
,”
Phys. Rev. E
66
,
026312
(
2002
).
73.
P.
Grete
,
B. W.
O'Shea
,
K.
Beckwith
,
W.
Schmidt
, and
A.
Christlieb
, “
Energy transfer in compressible magnetohydrodynamic turbulence
,”
Phys. Plasmas
24
,
092311
(
2017
).
74.
A. J.
Favre
,
J. J.
Gaviglio
, and
R. J.
Dumas
, “
Further space-time correlations of velocity in a turbulent boundary layer
,”
J. Fluid Mech.
3
,
344
356
(
1958
).
75.
T. D.
Drivas
and
G. L.
Eyink
, “
An Onsager singularity theorem for turbulent solutions of compressible Euler equations
,”
Commun. Math. Phys.
359
,
733
763
(
2018
).
76.
H.
Aluie
and
S.
Kurien
, “
Joint downscale fluxes of energy and potential enstrophy in rotating stratified Boussinesq flows
,”
Europhys. Lett.
96
,
44006
(
2011
).
77.
M.
Sadek
and
H.
Aluie
, “
Extracting the spectrum of a flow by spatial filtering
,”
Phys. Rev. Fluids
3
,
124610
(
2018
).
78.
We note that the 4.1% difference between the two highlighted values in Table III ( 4.80 × 10 9 erg/cm3) and Table IV ( 4.60 × 10 9 erg/cm3) is, at least in part, due to temporal discretization errors when calculating the integrals Γ .
79.
J. O.
Hinze
,
Turbulence
,
2nd ed.
(
McGraw-Hill
,
1975
).
80.
X.
Bian
,
J. K.
Shang
,
E. G.
Blackman
,
G. W.
Collins
, and
H.
Aluie
, “
Scaling of turbulent viscosity and resistivity: Extracting a scale-dependent turbulent magnetic Prandtl number
,”
Astrophys. J. Lett.
917
,
L3
(
2021
).
81.
M.
Buzzicotti
,
B. A.
Storer
,
H.
Khatri
,
S. M.
Griffies
, and
H.
Aluie
, “
Spatio-temporal coarse-graining decomposition of the global ocean geostrophic kinetic energy
,”
J. Adv. Model. Earth Syst.
15
,
e2023MS003693
(
2023
).
82.
J. V.
Bladel
,
On Helmholtz's Theorem in Finite Regions
(
Midwestern Universities Research Association
,
Wisconsin
,
1958
).
83.
J.
Qian
, “
Inverse energy cascade in two-dimensional turbulence
,”
Phys. Fluids
29
,
3608
3611
(
1986
).
84.
H.
Aluie
,
S.
Rai
,
H.
Yin
,
A.
Lees
,
D.
Zhao
,
S. M.
Griffies
,
A.
Adcroft
, and
J. K.
Shang
, “
Effective drift velocity from turbulent transport by vorticity
,”
Phys. Rev. Fluids
7
,
104601
(
2022
).