Rayleigh–Taylor instability, RTI, occurs at the interface separating two fluids subjected to acceleration when the density gradient and the acceleration are in opposite directions. Previous scientific research primarily considered RTI under the incompressible assumption, which may not be valid in many high-energy-density engineering applications and astrophysical phenomena. In this study, the compressibility effects of the background isothermal stratification strength on multi-mode two-dimensional RTI are explored using fully compressible multi-species direct numerical simulations. Cases under three different isothermal Mach numbers – – are investigated to explore weakly, moderately, and strongly stratified compressible RTI, respectively, at an Atwood number of 0.04. Unlike incompressible RTI, an increase in the flow compressibility through the strength of the background stratification can suppress the RTI growth and can lead to a termination of the RTI mixing layer growth with a highly molecularly mixed state. Our findings suggest that even at the chosen relatively low Atwood number, the variable-density effects can be significantly enhanced due to an increase in the background stratification for the compressible RTI as different spatial profiles become noticeably asymmetric across the mixing layer for the strongly stratified case. In addition, this study compares the chaotic behavior of the cases by studying the transport of the turbulent kinetic energy as well as the vortex dynamics. The Reynolds number dependence of the results is also examined with three different Reynolds numbers, and the findings for the large-scale mixing and flow quantities of interest are shown to be universal in the range of the Reynolds numbers studied.
I. INTRODUCTION
The Rayleigh–Taylor (RT) instability, or RTI, is an instability at the interface between two fluids subjected to acceleration when the directions of the density gradient and the acceleration have opposite signs.1,2 The RTI is observed in nature from atmospheric, oceanic, and geologic flows to astrophysical phenomena, such as supernova explosions. It also occurs in engineering applications with high-energy-density (HED) processes such as inertial confinement fusion (ICF).3 As a result, the RTI is of great interest in the mixing community, and it has been the subject of many experimental and numerical research studies.4–8 Although the flow instability is observed between the highly compressible flows in many natural phenomena and engineering systems, most of the numerical studies benefited from using the incompressible assumption to reduce computational expenses. In this study, we present the flow compressibility effects due to the background stratification on the multi-species two-dimensional (2D) multi-mode RTI with results from direct numerical simulations (DNS).
There have been several investigations on single-mode and multi-mode compressible RTI in previous literature. Wieland et al.15 carried out 2D single-mode compressible RTI DNS using adaptive wavelet-based mesh refinement. It was shown that an increase in the background stratification strength, which determines the flow compressibility of the mixing fluids, leads to the ceasing of the mixing layer growth for the single-mode RTI. Simulations of 2D single-mode compressible RTI with compact finite difference scheme and hyper-viscosity model were also performed by Luo et al.16 to examine the stabilizing and destabilizing effects of the instability under different Mach and Atwood numbers. The work is further extended to study the effects of the Atwood number and stratification parameter on 2D compressible RTI with multi-mode perturbations.17 Mellado et al.18 conducted three-dimensional (3D) large-eddy simulations of compressible RTI and found that intrinsic compressibility effects were limited as the turbulent Mach number is bounded in the range of 0.25–0.6. A DNS of 3D multi-mode compressible RT turbulence at a moderate Atwood number, A = 0.25, was conducted using the auto-adaptive multidomain Chebyshev–Fourier–Fourier numerical method by Gauthier,19 and the simulation results were compared to cases with either a lower Reynolds number or a lower Atwood number within the Boussinesq approximation. In a recent work by Luo and Wang,20 the mixing and energy transfer of 3D compressible RT turbulence indicates that the direct and reverse subgrid-scale (SGS) fluxes can be strengthened by compression and expansion motions, respectively.
Multi-species mixing is also of great interest in the scientific community as it is observed in broader engineering applications, and it has been shown that the flow dynamics could be significantly different than single fluid flows.21 In addition to RTI, there is also a wide study on multi-species mixing induced by the Richtmyer–Meshkov instability (RMI),22–24 which is an interfacial instability similar to RTI but induced by impulsive accelerations, such as shock waves. While the mixing of fluids induced by RMI is commonly studied with compressible simulations due to the compressibility effects of shock waves, high-resolution RMI simulations25 show that under a normal configuration, the flows within the RMI mixing layers are quasi-incompressible except around times of shock–interface interactions. The incompressible multi-species variable-density mixing dynamics of the turbulence aiming to mimic the core regions of the RTI and RMI mixing layers have been studied using several different homogeneous flow configurations as well.21,26–30 These studies lack information about compressibility effects, such as those from the background stratification, which can potentially play a significant role in the multi-species mixing processes.
For the compressible RTI, instability and multi-species mixing can be largely affected by many factors, such as background stratification, fluid properties, and acoustic effects. The mixing process can also generate strong shock waves through the piston-like motions of the bubbles and spikes.31 The growth of compressible RTI can be affected either through fluid compressibility or flow compressibility.32 Fluid compressibility is governed by the fluids' properties, such as the ratios of specific heat capacities, while flow compressibility is associated with the thermodynamic states independent of the fluid properties, such as the interface pressure. The former can be considered as dynamic or intrinsic, which is an effect of the speed of sound; and the latter can be viewed as static, as this compressibility effect results in a stable stratification.19,33
In many previous papers on multi-mode RTI, the domain width is commonly used as the characteristic length scale in the definition of the isothermal Mach number [see Eq. (3) and/or stratification parameter]. However, the characteristic wavelength of the initial perturbation is a better length scale to characterize the RTI problem, at least during the early times of the flow evolution, which is similar to the incompressible RTI study by Livescu et al.36 Hence, in the definition of Ma, we choose to use the characteristic wavelength of the perturbations instead of the domain width, which is different than many previous multi-mode compressible RTI studies. For example, the largest background stratification 3D RTI multi-mode cases have Sr = 6 reported based on the domain width in the work by Gauthier.19 This, in fact, corresponds to around or based on the characteristic wavelength (assuming the problem domain width is as they used a pseudo-spectral code). Similarly, Sr = 0.1 − Sr = 1.0 based on the shown domain width in the work by Luo and Wang17 is equivalent to − , or − , if the characteristic wavelength of the perturbations is used as the reference length. In addition, Luo and Wang20 extended their 2D work to 3D for their moderate Atwood number A = 0.5 by performing numerical simulations up to Sr = 3.0 based on a domain width that corresponds to around eight times smaller stratification parameter when the characteristic wavelength is used in the definition of Sr. Our study on the 2D multi-mode RTI includes cases with Ma = 0.45 where the background stratification strength is stronger than the problems in many previous multi-mode compressible RTI studies.
Our choice of the characteristic length scale can also provide a more direct and fair comparison between our multi-mode compressible RTI results with the single-mode compressible RTI results in previous works, such as the Ma = 0.3 case in the work by Wieland et al.15 To confirm that there is no re-growth at the late times of the flow evolution for the moderately and strongly stratified cases, the ensemble-averaged statistical results of the multi-mode cases in this study are obtained with DNS performed up to much longer non-dimensional times compared to the single-mode cases in the aforementioned work. The findings of this study suggest that in the case with a strong background stratification, Eq. (1) loses its validity even for the RTI under constant acceleration.
The paper is structured as follows: in Sec. II, we detail the governing equations and numerical methods, followed by the setup of DNS cases in Sec. III. Then, we present the analysis of the dynamics of the compressible mixing in Sec. IV. Section V shows and discusses the flow energetics and small-scale vortical motions. Finally, in Sec. VI, we summarize the results of this study.
II. PROBLEM DESCRIPTION
A. Governing equations
B. Non-dimensional problem parameters
C. Initial and boundary conditions
Subsonic outflow boundary conditions are used at the left and right boundaries with the Navier–Stokes characteristic nonreflecting method.34,38 The characteristic-based boundary method makes use of the local one-dimensional inviscid (LODI) relations to prevent any spurious inward-propagating acoustic waves. The target pressure values for the nonreflecting method are calculated from the hydrostatic equation at the domain boundaries.
III. DETAILS OF DIRECT NUMERICAL SIMULATIONS
A. Numerical methods
All multi-species simulations in this work are performed using Hydrodynamics Adaptive Mesh Refinement Simulator (HAMeRS).39 It is a hydrodynamics code with Cartesian adaptive mesh refinement and local time stepping (or sub-cycling) capabilities. Conservative flux correction at coarse-fine grid interfaces40 is implemented for any kind of conservative fluxes in the telescoping form. The parallelization of the code and the management of Cartesian grid cells are facilitated by the Structured Adaptive Mesh Refinement Application Infrastructure (SAMRAI) library41–45 from Lawrence Livermore National Laboratory (LLNL). Feature-based sensors are implemented to identify regions of interest for adaptive mesh refinement (AMR). These include gradient and wavelet sensors46 on different flow fields. The 13-point fourth-order dispersion-relation-preserving (DRP) scheme by Bogey and Bailly47 re-cast in the conservative flux difference form39,48 is used for the discretization of the convective flux. Explicit tenth-order finite difference schemes are used to discretize the diffusive and viscous terms as source terms of the equations. Navier–Stokes nonreflecting characteristic boundary conditions34,38 with the ghost cell approach49 are implemented for outflow domain boundaries. A third-order total variation diminishing Runge–Kutta (RK-TVD) scheme50 is adopted for time integration with the local time stepping for each grid level.
B. Grid settings and realizations
Various cases with different combinations of background stratification strength characterized by the isothermal Mach number, Ma, and problem Reynolds number, Re0, are designed for the DNS runs. Table I lists all of these cases for the 2 D multi-mode RTI problem. The index following the first letter, Ma: 015, 030, and 045, denotes the Mach numbers, 0.15, 0.30, and 0.45, respectively; and the index following the second letter, Re: 1, 2, and 3, represents the Reynolds number, Re0: 3187.5, 6375.0, or 12 750.0, respectively. We should note that while the range of Re0 chosen in this work is lower than that in the previous 2D single-mode compressible RTI paper by Wieland et al.15 (in the range of 25 500–102 000), these multi-mode DNS are computationally more expensive, as they are conducted in a much larger domain with each mode number being repeated around 12–20 times in the periodic dimension. Also note that if the domain width Ly instead of λ0 is used in Eqs. (3) and (19), similar to some previous studies17,19,20 on multi-mode RTI, both the problem isothermal Mach numbers and Reynolds numbers would be at higher values. For instance, the case Ma045Re3 would have an isothermal Mach number of 1.8 and a Reynolds number of 816 000. The isothermal Mach number based on Ly is also included in Table I for each case for reference.
Parameters for the DNS cases. Nx and Ny are the numbers of grid cells in the x- and y-directions, respectively, for the base grid. For reference, the values in the parentheses in the second column represent values for the alternative isothermal Mach number, , defined by the domain width in some previous studies17,19,20 instead of the perturbation wavelength λ0.
Cases . | Ma . | Re0 . | . | Base grid resolution, . | Number of mesh levels . |
---|---|---|---|---|---|
Ma015Re1 | 0.15 (0.6) | 4 | 3 | ||
Ma030Re1 | 0.30 (1.2) | 4 | 3 | ||
Ma045Re1 | 0.45 (1.8) | 2 | 2 | ||
Ma015Re2 | 0.15 (0.6) | 4 | 3 | ||
Ma030Re2 | 0.30 (1.2) | 4 | 3 | ||
Ma045Re2 | 0.45 (1.8) | 2 | 2 | ||
Ma015Re3 | 0.15 (0.6) | 12 750.0 | 4 | 3 | |
Ma030Re3 | 0.30 (1.2) | 12 750.0 | 4 | 3 | |
Ma045Re3 | 0.45 (1.8) | 12 750.0 | 2 | 2048 × 1024 | 2 |
Cases . | Ma . | Re0 . | . | Base grid resolution, . | Number of mesh levels . |
---|---|---|---|---|---|
Ma015Re1 | 0.15 (0.6) | 4 | 3 | ||
Ma030Re1 | 0.30 (1.2) | 4 | 3 | ||
Ma045Re1 | 0.45 (1.8) | 2 | 2 | ||
Ma015Re2 | 0.15 (0.6) | 4 | 3 | ||
Ma030Re2 | 0.30 (1.2) | 4 | 3 | ||
Ma045Re2 | 0.45 (1.8) | 2 | 2 | ||
Ma015Re3 | 0.15 (0.6) | 12 750.0 | 4 | 3 | |
Ma030Re3 | 0.30 (1.2) | 12 750.0 | 4 | 3 | |
Ma045Re3 | 0.45 (1.8) | 12 750.0 | 2 | 2048 × 1024 | 2 |
In addition, 16 realizations of simulations are performed for each combination case of Re0 and Ma for better statistical convergence. Three levels of meshes, with two levels of adaptive mesh refinement, are used for the Ma = 0.15 and Ma = 0.30 cases, while the Ma = 0.45 case only has two levels of meshes. For the realization simulations with and , base grids of the Ma = 0.15 and Ma = 0.30 cases have 256 grid points in the y direction, while that of the Ma = 0.45 case has 512 grid points. Finer base grids are used for the highest Reynolds number simulations with to resolve smaller finest scales in the simulations. At this highest Re0, the base grids of the Ma = 0.15 and Ma = 0.30 cases have 512 grid points, and that of the Ma = 0.45 case has 1024 grid points in the y direction. A constant refinement ratio of 1:2 is used across different grid levels in each direction. A finer base grid is used in Ma = 0.45 cases as the flow experiences larger pressure and density gradients even far away from the mixing layer width due to the stronger background stratification. A value sensor based on the mass fraction is used to identify mixing regions for AMR. Acoustic Courant–Friedrichs–Lewy (CFL) number of 0.5 is used for the Ma = 0.15 case, and 0.2 is used for the stronger background stratification cases. The diffusive CFL number is half of that of the acoustic CFL number for each case. The grid convergence analysis is also performed on one of the realizations for each combination case of Re0 and Ma, and the results are compared with those obtained from simulations with base grids refined by a factor of 2 in each direction. Through grid sensitivity analysis, it is verified that all quantities of interest studied in the later sections are well grid converged for the corresponding realization of each case. While only one realization is used in the grid convergence analysis, all statistical results presented in this work are ensemble-averaged results over all realizations for each case.
IV. MIXING DYNAMICS
A. Flow visualization of mass fraction
Effects of the isothermal background stratification strength on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (Ma015Re3), Ma = 0.30 (Ma030Re3), and Ma = 0.45 (Ma045Re3) cases at from left to right at different times. Top row: ; middle row: ; and bottom row: .
Effects of the isothermal background stratification strength on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (Ma015Re3), Ma = 0.30 (Ma030Re3), and Ma = 0.45 (Ma045Re3) cases at from left to right at different times. Top row: ; middle row: ; and bottom row: .
We also present the Reynolds number effects on the heavy fluid mass fraction at three different problem Re0 for the three different Ma cases at and in Figs. 2 and 3, respectively. It should be reminded that the problem Re0 is mainly controlled by the initial interface kinematic shear viscosity, ν0, of the flow. From the figures, it can be seen that an increase in Re0 has a limited effect on the heights of the bubbles and spikes for all Ma cases, while a higher Re0 leads to finer structures in the mass fraction field. The small Re0 dependence of the mixing layer thickness and other large scales suggests that the slow-down in the growth of the mixing layer under strong background stratification, i.e., for the higher Ma cases, is at least valid in the range of Re0 investigated in this study. This small Reynolds number dependence behavior of the large scales is also consistent with the incompressible and compressible single-mode RTI DNS results by Wei and Livescu51 and Wieland et al.,15 respectively, where the former and latter show insignificant Reynolds number dependence behavior starting from Re = 7500.0 and Re = 25 500.0, respectively.
Reynolds number effects on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases at , and from left to right at .
Reynolds number effects on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases at , and from left to right at .
Reynolds number effects on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases at , and from left to right at .
Reynolds number effects on the heavy fluid mass fraction, Yh, of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases at , and from left to right at .
B. Time evolution of mixing quantities
Time evolution of (a) the normalized mixing width, , and (b) the normalized bubble and spike heights, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3). The lines for the spikes are marked with crosses, and those for the bubbles are marked with circles.
Time evolution of (a) the normalized mixing width, , and (b) the normalized bubble and spike heights, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3). The lines for the spikes are marked with crosses, and those for the bubbles are marked with circles.
In addition to the integral mixing width, Fig. 4(b) shows the heights of the bubbles, hb, and spikes, hs, that are defined with the 99% and 1% thresholds of , respectively. Similar to W, both quantities are normalized by λ0 in the figure. As expected, the behavior of is similar to W, although is more sensitive to the dynamics of the mixing layer edges, especially at late times ( . Similar to W, also shows cessation in their growth rates for the moderately and strongly stratified cases (Ma030Re3 and Ma045Re3) after and , respectively. Both hb and hs have almost the same growth in time but in opposite directions for all cases, and this quasi-symmetry in growth is consistent with the incompressible single-mode RTI DNS results of Wei and Livescu51 at the same Atwood number before the reacceleration regime. Interestingly, Wieland et al.15 found that there is an asymmetry in the growth of the bubble and spike heights for the single-mode compressible RTI, where the spikes are consistently at larger heights than the bubbles for cases at different Ma.
Time evolution of (a) the mixedness, Θ, and (b) the normalized domain-integrated scalar dissipation rate, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Time evolution of (a) the mixedness, Θ, and (b) the normalized domain-integrated scalar dissipation rate, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
It is also important to explore the generality of the above results by investigating the effects of the Reynolds number on the mixing statistical quantities. Figure 6 presents the Reynolds number effects on W and for the cases with the weakest (Ma015Re1, Ma015Re2, and Ma015Re3) and strongest (Ma045Re1, Ma045Re2, and Ma045Re3) background stratification strengths using different Re0. As seen from the figure, Re0 has a very limited effect on the growth of the mixing layer as the values of W approach similar heights for all three Re0 no matter whether Ma = 0.15 or Ma = 0.45. This is consistent with the time evolution of the bubble and spikes as shown earlier in the visualizations of the mass fraction field. This lack of dependence of the large-scale characteristics, such as the mixing width and the bubble/spike heights (not shown in the figure), on the Reynolds number was also shown by Wieland et al.15 and Wei and Livescu51 for the single-mode compressible RTI and incompressible RTI, respectively. These findings indicate that the cease of the mixing layer growth due to strong background stratification is highly universal for the multi-mode compressible RTI and independent of the Re0 at least for the investigated range of Re0. While the other two mixing metrics, Θ, and , show some sensitivities to Re0, in particular at early times when the flow is not chaotic and still developing. For example, gets a slightly larger value for the case Ma015Re1 compared to the case Ma015Re3 at as the diffusion transport coefficient is larger for the lower Reynolds number cases with the same Ma. At later times of the flow evolution, both Θ and collapse as the flow becomes fully developed. This further assures that the ceasing of the mixing and the reduction of the mass fraction variations across the mixing layer caused by strong background stratification strength are quite universal behavior as they do not show much Reynolds number dependence, at least for the range of Re0 chosen in this study. It should be noted that does not have significant Reynolds number dependence at late times. Although the gradients of the mass fraction are smaller for the lower Reynolds number cases, is defined as a quantity that consists of the product of the squared mass fraction gradient magnitude and the mass diffusivity, D, which scales with ν0 as Sc is kept constant and is larger for the low Reynolds number cases. As a result, at early times, such as , while the flow is yet to be developed, is slightly larger for the low Re0 cases as these cases have larger mass diffusivity. At later times, becomes independent of the Reynolds number as the flow has developed, and the larger mass fraction gradients compensate for the smaller mass diffusivity for the high Reynolds number cases. This is also consistent with the flow visualization of the heavy fluid mass fraction field, where a higher value of Re0 leads to finer features and larger gradients in the mass fraction field.
Time evolution of (a) the normalized mixing width, , (b) the mixedness, Θ, and (c) the normalized domain-integrated scalar dissipation rate, , of the Ma = 0.15 cases (top row) and the Ma = 0.45 cases (bottom row) with different Reynolds numbers. Red solid line: ; green dashed line: ; and blue dash-dotted line: .
Time evolution of (a) the normalized mixing width, , (b) the mixedness, Θ, and (c) the normalized domain-integrated scalar dissipation rate, , of the Ma = 0.15 cases (top row) and the Ma = 0.45 cases (bottom row) with different Reynolds numbers. Red solid line: ; green dashed line: ; and blue dash-dotted line: .
C. Spatial profiles of mixing quantities
To further investigate the mixing behavior of the compressible RTI, we present the mean spatial profiles of the heavy fluid mass fraction, , in Fig. 7 for all Ma cases at the largest Re0. At the very early time, , the profiles of all cases have a monotonic transition from heavy fluid region to light fluid region. However, at , the profiles of show non-monotonic shapes as the pure fluids exchange locations within the mixing layer due to fluid entrainment while the molecular mixing is limited. When , the enhanced molecular mixing smooths out the profiles toward more monotonic shapes again for all cases. It is important to note that at these late times, increased strength of the background stratification leads to larger gradients of the profiles near the edges of the mixing layer. This is also noticeable in the visualizations in Fig. 1 for the higher Ma cases (Ma030Re3 and Ma045Re3), where there is a more distinct separation between the pure fluid regions and the mixing layers compared to the low Ma case (Ma015Re3).
Effects of the isothermal background stratification strength on the mean spatial profiles of the heavy fluid mass fraction, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the mean spatial profiles of the heavy fluid mass fraction, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Next, we present the spatial profiles of the heavy fluid mass fraction variance, , at different time instants in Fig. 8, where represents the Reynolds fluctuation of a given quantity. At , there is a symmetric distribution with a single peak for each case, and at , there are double peaks which are consistent with the results of the non-monotonic profiles. A more interesting behavior is observed at later times, i.e., when , as the profiles become asymmetric even with our chosen very small Atwood number. Results of the incompressible numerical simulations9,21,29,30,36 and quasi-incompressible experiments55–57 of the buoyancy-driven flows suggest highly symmetric flow evolution when A < 0.2. Recently, the DNS results by Prine et al.58 showed that for single-mode compressible RTI at A = 0.25, an increase in the background stratification strength leads to a significant difference between the evolution of the locations of the spikes and bubbles, as the former grows much faster than the latter. In addition, they showed that an increase in compressibility through the background stratification strength (higher Ma) prevents the formation of vortical structures at the bubble side of the flows, while the vortical structures are more resistant to the effects of background stratification at the spike side. In this study, the Atwood number is kept very small (A = 0.04), so one would not expect any asymmetric behavior on the spatial profiles of the mass fraction variance. Our results of the low Ma case, Ma015Re3, are consistent with the earlier observations on incompressible or quasi-incompressible RTI flows, as there is no significant difference between the evolution of the hb and hs. However, the mixing state of the flow surprisingly exhibits differences between the heavier and lighter fluid regions when Ma is increased, where the mass fraction variance is larger in the heavier fluid regions compared to the lighter fluid regions. Figure 9 shows the normalized mean spatial profiles of the scalar dissipation rate, , between the cases Ma015Re3, Ma030Re3, and Ma045Re3. The scalar dissipation rate profiles stay quasi-symmetric until the late times of the flow evolution as the large-scale asymmetry in has not been reflected in the small-scale dissipation quantity. At , the spatial profile becomes asymmetric for the case Ma045Re3, which can be attributed to the asymmetric profile as the scalar dissipation rate would be enhanced in regions where the variance is larger.
Effects of the isothermal background stratification strength on the spatial profiles of the heavy fluid mass fraction variance, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the spatial profiles of the heavy fluid mass fraction variance, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of the scalar dissipation rate, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of the scalar dissipation rate, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
The normalized mean density profiles, , are shown in Fig. 10 for all Ma cases at the highest Re0. At , the mean density profiles near the interface at are non-monotonic with inverse gradients for all cases because of the initial conditions of unstable stratifications. In addition, due to the exponential-like relation between the density field and the height of the domain for these isothermally stratified cases, the initial density stratification is not symmetric, and such asymmetry becomes dominant by an increase in the isothermal background stratification strength. As a result, the initial background stratification would also contribute to the observed asymmetry in (see Fig. 8) for the high Ma cases. For incompressible RTI under constant acceleration, the unstable stratification persists indefinitely during the simulation or experiment as the pure fluids have constant densities and the pure fluids continuously fill the growing mixing layer. However, for the isothermally stratified compressible RTI, the densities of the pure fluids are not uniform and are dependent on location in the domain, as derived in Eq. (21) and shown in the initial conditions of Fig. 10. As a result, as seen from Fig. 10(c), the density profile rapidly becomes monotonic without any inverse gradient for the strongest Ma case (Ma045Re3) after because of the coupled effects of mixing and background stratification. From the time evolution of the mixing width, we can see that the mixing layer continues to grow beyond until for this case. There is a similar observation for the moderately stratified case Ma030Re3, where the mixing width temporarily still increases after the mean density profile becomes monotonic at around . This may be due to the inertia of bubbles/spikes, the existence of locally unstable density gradients that are not represented by the mean density profile because of the 2D chaotic motions, and also possibly the strong expansion and compression motions.20 It is also notable that even in our weakest background stratification case, Ma015Re3, the mean density profile eventually becomes monotonic without inverse gradient at late time . This suggests that the mixing layer growth may also cease for this low Ma case as well if the simulation is run for a longer time, which is significantly different than incompressible RTI evolution.
Normalized mean spatial profiles of the density, , of cases with different Ma at at different times. Cyan solid line: ; red dashed line: ; green dash-dotted line: ; and blue dotted line: ; thin gray dashed line: initial conditions at .
Normalized mean spatial profiles of the density, , of cases with different Ma at at different times. Cyan solid line: ; red dashed line: ; green dash-dotted line: ; and blue dotted line: ; thin gray dashed line: initial conditions at .
V. FLOW ENERGETICS
The RTI at the interface leads to a rapid and efficient conversion of potential energy into kinetic energy.21,36 For the incompressible case, the continuous growth of the mixing layer of RTI is caused by this energy conversion. This section investigates the effects of background stratification strength on the multi-mode compressible RTI evolution further through flow energetics.
A. Time evolution and spatial profiles of turbulent kinetic energy
Time evolution of (a) the normalized domain-integrated turbulent kinetic energy, , and (b) the Reynolds normal stress anisotropy, , at . Red solid line: Ma = 0.15 at the interface (Ma015Re3); and green dashed line: Ma = 0.30 (Ma030Re3); blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Time evolution of (a) the normalized domain-integrated turbulent kinetic energy, , and (b) the Reynolds normal stress anisotropy, , at . Red solid line: Ma = 0.15 at the interface (Ma015Re3); and green dashed line: Ma = 0.30 (Ma030Re3); blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Figure 12 presents the normalized mean spatial profiles of turbulent kinetic energy, , at four distinct times for all three Ma cases at the highest Re0. Although the compressibility effects through background stratification on the mixing layer growth are small at , we can still observe that a stronger background stratification strength leads to a smaller peak of the spatial profile. Such reduction of the magnitude of due to stronger background stratification strength becomes more and more observable at later times. In addition, Fig. 13 presents the spatial profiles of the Reynolds normal stress anisotropy at different time instants for all cases at the highest problem Re0. Initially, the peaks of around are greater than 10 for all cases, which indicates that the Reynolds normal stress components are highly anisotropic around the middle portion of the mixing layer. Nevertheless, the Reynolds normal stress anisotropy of the flow slightly decreases to the value of roughly 1.5 at the core of the mixing layer at a later time, . At a very late time, , the spatial profiles of the Reynolds normal stress anisotropy fluctuates around unity for the higher Ma cases (Ma030Re3 and Ma045Re3), whereas the flow stays larger than one within a large portion of the mixing layer for the low Ma case (Ma015Re3). This shows that the flows are more isotropic for the higher Ma cases compared to that of the case Ma015Re3, which is also consistent with the time evolution of the anisotropy in Fig. 11(b).
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of turbulent kinetic energy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of turbulent kinetic energy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the spatial profiles of Reynolds normal stress anisotropy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the spatial profiles of Reynolds normal stress anisotropy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
B. Transport of turbulent kinetic energy
Researchers have extensively studied the kinetic energy transport for turbulence induced by incompressible RTI and buoyancy-driven turbulence.21,26,36,53,61,62 On the other hand, the analysis of kinetic energy budgets for compressible RTI is still limited except some very recent works.20,63,64 In this subsection, the budgets of the turbulent kinetic energy are analyzed for the multi-mode compressible RTI under different background stratification strengths.
Time evolution of the normalized mean of turbulent mass flux velocity in the middle part of the mixing layer, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Time evolution of the normalized mean of turbulent mass flux velocity in the middle part of the mixing layer, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
As seen in the figure, is the largest (and is always positive) for the lowest Ma case (Ma015Re3) as the mixing layer continuously grows, which means the pure fluid motions always convert the potential energy to the kinetic energy. However, rapidly decays at for the cases Ma030Re3 and Ma045Re3. More interestingly, becomes negative for these higher Ma cases [see Fig. 14(b)]. A negative value of the mean turbulent mass flux velocity indicates that there are fluid motions in the reserve directions compared to the earlier times in the middle portion of the domain where the heavier fluid moves up in the negative x direction, whereas lighter fluid migrates down in the opposite direction. This motion is consistent among different realizations and is evident after the ensemble averaging. This behavior is attributed to the result of the inertia of the flow when the stronger background stratification terminates the growth of the mixing layer, which leads to a rapid decay in . Such rapid inertial motion causes undershoot in the time evolution of to obtain a negative value.
The time evolution of the normalized domain-integrated production term, , and dissipation term, , of k for different Ma cases at the highest Re0 is shown in Fig. 15. The production and dissipation terms are given by terms (III) and (VI), respectively, in Eq. (39). These two transport terms are normalized with . From the figure, it can be seen that the time evolution of is very similar to that of , as the production is essentially only contributed by the component , which is marked with circles in the plots. As a result, is also negative and acts as an anti-production term, except at very early times for the higher Ma cases (Ma030Re3, Ma045Re3). On the other hand, is always negative to dissipate .
Time evolution of the normalized total production term , the normalized production component associated with the turbulent mass flux velocity, and the normalized dissipation term, of the turbulent kinetic energy at . The terms are given by Eq. (39). Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3). The lines for the normalized total production are not marked with any symbols, while the lines for the normalized dissipation are marked with crosses. The normalized production component associated with the turbulent mass flux velocity is represented by the circles only, and no lines.
Time evolution of the normalized total production term , the normalized production component associated with the turbulent mass flux velocity, and the normalized dissipation term, of the turbulent kinetic energy at . The terms are given by Eq. (39). Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3). The lines for the normalized total production are not marked with any symbols, while the lines for the normalized dissipation are marked with crosses. The normalized production component associated with the turbulent mass flux velocity is represented by the circles only, and no lines.
The spatial profiles of the normalized budgets of the turbulent kinetic energy, (or ), are shown in Fig. 16 for the three Ma cases at the highest problem Re0 at and . The budget terms given by Eq. (39) are normalized by . From the figure, it can be seen that the production, turbulent transport, and dissipation terms are the only relevant terms in the budgets for different cases at the chosen times, while other terms are negligible. At , the production and the turbulent transport terms are dominant in the middle part of the mixing layer for the two lower Ma cases (Ma015Re3 and Ma030Re3). In this core region of the mixing layer, the turbulent transport terms are very negative. While the production term is still positive for the case Ma045Re3 at in the middle part of the layer, it is small compared to the dissipation and the turbulent transport terms, and the overall rate of change of is very negative. At the edge of the mixing layer, the turbulent transport term is the only dominant term and is positive for each case. As a result, the spatial profiles of continue to spread at this time for all cases. At a later time, , the budgets of the Ma015Re3 and Ma030Re3 cases are still quite similar compared to those at , except that there are more fluctuations in the spatial profiles of the turbulent transport term. As the flows are more chaotic at this time, more realizations may be needed for smoother, or statistically better converged, profiles for this term. However, the general shape of this term at can still be observed, which is quite similar to that at for each case. Its effect is still to transfer from the core to the edges of the mixing layers for all cases, including case Ma045Re3. The production term of the cases Ma015Re3 and Ma030Re3 remains largely positive to continuously convert potential energy into kinetic energy, similar to the observation that this term is large over a long period of time in other studies of incompressible RTI and buoyancy-driven mixing.21,26,36 Nevertheless, the production term has become negative across a large portion of the mixing layer for case Ma045Re3 due to negative turbulent mass flux velocity at this time. All relevant terms are negative at the core of the mixing layer, and hence the peak of (or ) decreases rapidly for this case around this time instant.
Comparison of the normalized budget terms of turbulent kinetic energy, ρk, of Ma = 0.15 (Ma015Re3) (top row), Ma = 0.30 (Ma030Re3) (middle row), and Ma = 0.45 (Ma045Re3) (bottom row) cases at at two different times. The terms are given by Eq. (39). Cyan solid line: production [term (III)]; red dashed line: pressure-dilatation [term (V)]; green dash-dotted line: turbulent transport [term (IV)]; blue dash-dot-dotted line: dissipation [term (VI)]; orange dash-triple-dotted line: negative of convection due to streamwise velocity component associated with turbulent mass flux; thin black dotted line: summation of all terms (rate of change in the simulation frame).
Comparison of the normalized budget terms of turbulent kinetic energy, ρk, of Ma = 0.15 (Ma015Re3) (top row), Ma = 0.30 (Ma030Re3) (middle row), and Ma = 0.45 (Ma045Re3) (bottom row) cases at at two different times. The terms are given by Eq. (39). Cyan solid line: production [term (III)]; red dashed line: pressure-dilatation [term (V)]; green dash-dotted line: turbulent transport [term (IV)]; blue dash-dot-dotted line: dissipation [term (VI)]; orange dash-triple-dotted line: negative of convection due to streamwise velocity component associated with turbulent mass flux; thin black dotted line: summation of all terms (rate of change in the simulation frame).
C. Vortex dynamics
The study of vortex dynamics, particularly the evolution of the vorticity, can provide an additional dimension to examine the effects of background stratification on the multi-mode compressible RTI. Gauthier19 studied the vortex dynamics through helicity for the multi-mode compressible RTI, but only one background stratification strength was considered. While the effects of background stratification on the vorticity dynamics of compressible RTI were addressed by Wieland et al.,15 it was only limited to single-mode RTI. In this section, the effects of background stratification strength on the vorticity evolution for the multi-mode compressible RTI are studied. Figure 17 shows a qualitative comparison of the contours of the only nonzero component (z-component) of vorticity for the 2D flows, , normalized by tr for all three Ma cases at the largest Re0, i.e., Ma015Re3, Ma030Re3, and Ma045Re3, at three different time instants: , and . The first-row panel of the figure compares the contours between the three cases at . At this time, the flows are still not very chaotic, as the roll-up of the bubbles and spikes is still very coherent for all cases. In addition, the shapes of the vortices are essentially the same for all cases, similar to the mass fraction contour plots in Fig. 1. However, it is also observed that an increase in the background stratification strength through Ma decreases the magnitude of the vorticity, as indicated by the minimum and maximum values of the colorbars. The reduced strength of vorticity with Ma is an early indication of background stratification effects on compressible RTI. At the later time, , not only the strength but also the shapes of the vortices have become different between the three cases. The slow-down in the growth of the mixing layer due to the background stratification effect is reflected in the vortical structures where the structures are confined in a narrower region with higher Ma. At the much later time, (shown by the third row of Fig. 17), the low Ma case (Ma015Re3) continues to show the largest vorticity magnitude within the mixing layer. However, the strength of the normalized vorticity is reduced compared to its strength at the earlier time, . Although the flow is more chaotic at this time instant, the large-scale bubbles and spikes are still distinguishable and continue to penetrate into the unmixed regions. On the other hand, the flows have less distinguishable large-scale structures for the higher Ma cases (Ma030Re3 and Ma045Re3) due to larger levels of molecular mixing, and this is consistent with the observation of Yh in Fig. 1. The effects of the background stratification strength on the evolution of the vorticity for these multi-mode compressible RTI cases are similar to those found by Wieland et al.15 for single-mode compressible RTI where the vorticity magnitude decreases and the propagation of the “fronts” of nonzero vorticity is suppressed with a larger Ma.
Effects of the isothermal background stratification strength on the normalized vorticity, , of the Ma = 0.15 (Ma015Re3), Ma = 0.30 (Ma030Re3), and Ma = 0.45 (Ma045Re3) cases from left to right at at different times. Top row: ; middle row: ; and bottom row: .
Effects of the isothermal background stratification strength on the normalized vorticity, , of the Ma = 0.15 (Ma015Re3), Ma = 0.30 (Ma030Re3), and Ma = 0.45 (Ma045Re3) cases from left to right at at different times. Top row: ; middle row: ; and bottom row: .
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of enstrophy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Effects of the isothermal background stratification strength on the normalized mean spatial profiles of enstrophy, , at at different times. Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Time evolution of (a) the normalized domain-integrated enstrophy, , and (b) the normalized domain-integrated root mean square of baroclinic torque, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Time evolution of (a) the normalized domain-integrated enstrophy, , and (b) the normalized domain-integrated root mean square of baroclinic torque, , at . Red solid line: Ma = 0.15 (Ma015Re3); green dashed line: Ma = 0.30 (Ma030Re3); and blue dash-dotted line: Ma = 0.45 (Ma045Re3).
Furthermore, the small-scale vortical motions of the flows are sensitive to the Reynolds number, and more details are discussed in Appendix B.
VI. CONCLUSION
In this study, we explored the background isothermal stratification effects on the 2D multi-mode compressible RTI mixing and energetics with fully compressible multi-species direct numerical simulations. The RTI is driven by the initial inverse density gradient at the perturbed two-fluid interface under constant acceleration. Although the background stratification strength has a small effect on the magnitude of the initial density jump at the interface compared with the Atwood number effect, it determines the background density gradient coupled with the initial jump. The background density gradient was found to largely affect the late-time mixing layer evolution using three cases under different strengths of background stratification. In this study, the strength of the stratification is controlled by the background isothermal Mach number, Ma, and they are chosen as 0.15, 0.30, and 0.45 for weakly, moderately, and strongly stratified cases, respectively. Similar to the 2D single-mode compressible isothermally stratified RTI,15 an increase in the background isothermal Mach number leads to a suppression of the growth of the RTI mixing layer. Simulations with durations that are substantially longer than those of the previous single-mode compressible RTI study show that the multi-mode RTI growth ceases even under relatively moderate stratification strength (Ma = 0.30). When the mixing layer grows, the non-monotonic behavior of the mean density gradient disappears rapidly for the strongly stratified cases, and the mixing layer growth slows down quickly. Eventually, after reaching a certain mixing layer height, the flow becomes well-mixed within the mixing layer as the pure fluid entrainment is ceased, and the high degree of molecular dissipation dominates the mixing process. In addition, the suppression of the mixing layer growth due to strong background stratification rapidly redistributes the energy within the two components of the Reynolds normal stress and leads to a quasi-isotropic flow.
The flow energetics of the multi-mode compressible RTI were also found to be significantly different than the incompressible RTI scenario. The TKE reaches a self-similar growing behavior for the incompressible RTI, whereas the TKE follows a non-linear evolution with rapid increase and decay stages due to the background stratification effects for the moderately and strongly stratified multi-mode compressible RTI cases. In contrast to the self-similar behavior of TKE in incompressible RTI,9 where the turbulent mass flux velocity is always in the same direction of acceleration, the turbulent mass flux velocity is in the opposite direction of the acceleration (i.e., leads to a negative value of the turbulent mass flux velocity if the acceleration is positive) at late times for these two cases, so the production term behaves as a TKE sink, rather than a source term as in incompressible RTI. The spatial profiles of turbulent mass flux velocity also suggest that the mass flux velocity plays a role in removing TKE at the core region of the mixing layer for the moderately and strongly stratified cases at late times when the mixing layer growth ceases. Finally, the Reynolds number dependence was investigated using three different problem Reynolds numbers for each isothermal Mach number case. Our results suggest that the findings on the effects of isothermal background stratification through Ma on most of the large-scale statistical quantities, such as the mixing width, bubble/spike heights, and TKE, appear to be universal as they do not show strong dependency on the Reynolds number. In addition, even the scalar dissipation rate was also found to be insensitive to the Reynolds number under proper scaling.
The findings of this study suggest a more complex flow evolution for the compressible RTI than the widely studied incompressible RTI. Future studies may investigate the effects of background stratification strength on the more realistic 3D compressible RTI and buoyancy-driven turbulence.
ACKNOWLEDGMENTS
D.A. would like to thank the U.S. National Science Foundation CBET Fluid Dynamics Program (Award No. 2234415) for the financial support that made this work possible. For the computations, this work used the Extreme Science and Engineering Discovery Environment (XSEDE),68 which is supported by National Science Foundation Grant No. ACI-1548562. This work also used Stampede2 at the Texas Advanced Computing Center (TACC) through Allocation No. PHY200090 from the Advanced Cyber infrastructure Coordination Ecosystem: Services & Support (ACCESS) program, which is supported by National Science Foundation Grant Nos. 2138259, 2138286, 2138307, 2137603, and 2138296.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Denis Aslangil: Funding acquisition (lead); Investigation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Man Long Wong: Investigation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.
APPENDIX A: COMPARISON OF THE SPATIAL PROFILES BASED ON MASS FRACTION AND VOLUME FRACTION
For flows with large density variations, the flow metrics based on mass, volume, or mole fractions could lead to different conclusions. In this study, the Atwood number is kept low, and we do not anticipate any sensitivity for a selection of these three options. The comparison of the mean spatial profiles of the heavy fluid mass fraction, , and heavy fluid volume fraction, , at different time instants is presented in Fig. 20. The heavy fluid volume fraction is given by , where ρh is the species density of the heavy fluid. From the figure, it can be seen that the profiles are basically identical to the profiles. Figures 21 and 22 show that the variance and the normalized mean spatial profiles of the scalar dissipation rate based on mass and volume fractions are also essentially identical, respectively. We believe the spatial profiles based on mass fraction and volume fraction are basically the same because the Atwood number of the flow is low. Note that the volume fraction is equivalent to the mole fraction for species given by the ideal gas equation of state.
Comparison of the mean spatial profiles of the heavy fluid mass fraction, , and volume fraction, , at different times at . The mass fraction mean profiles are represented by lines, and the volume fraction mean profiles are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
Comparison of the mean spatial profiles of the heavy fluid mass fraction, , and volume fraction, , at different times at . The mass fraction mean profiles are represented by lines, and the volume fraction mean profiles are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
Comparison of the spatial profiles of the heavy fluid mass fraction variance, , and volume fraction variance, , at different times at . The mass fraction variance profiles are represented by lines, and the volume fraction variance profiles are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
Comparison of the spatial profiles of the heavy fluid mass fraction variance, , and volume fraction variance, , at different times at . The mass fraction variance profiles are represented by lines, and the volume fraction variance profiles are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
Comparison of the normalized mean spatial profiles of the scalar dissipation rate, , based on mass fraction or volume fraction at different times at . The profiles based on mass fraction are represented by lines and the profiles based on volume fraction are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
Comparison of the normalized mean spatial profiles of the scalar dissipation rate, , based on mass fraction or volume fraction at different times at . The profiles based on mass fraction are represented by lines and the profiles based on volume fraction are represented by symbols. Red solid line (circles): mass fraction (volume fraction) of Ma015Re3; and blue dash-dotted line (crosses): mass fraction (volume fraction) of Ma045Re3.
APPENDIX B: MORE ANALYSIS ON THE REYNOLDS NUMBER DEPENDENCE
In contrast to large scales such as bubble/spike heights, the vorticity has large sensitivity to the Reynolds number. In Figs. 23 and 24, the contours of the normalized vorticity component at different Re0 are compared at two different time instants, and . The magnitude of the vorticity clearly increases with Re0. It was shown that baroclinic torque and viscous diffusion are two major terms in the budgets of vorticity.15 As a result, as Re0 increases, the viscous diffusion has a smaller effect due to reduced viscosity to offset the production effect of the baroclinic term, and thus, the vorticity is generated at a higher rate.
Reynolds number effects on the normalized vorticity, , of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases with , and from left to right at .
Reynolds number effects on the normalized vorticity, , of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases with , and from left to right at .
Reynolds number effects on the normalized vorticity, , of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases with , and from left to right at .
Reynolds number effects on the normalized vorticity, , of the Ma = 0.15 (top row), Ma = 0.30 (middle row), and Ma = 0.45 (bottom row) cases with , and from left to right at .
Figure 25 shows the time evolution of the Reynolds number dependence of the normalized turbulent kinetic energy ( ), enstrophy ( ), and baroclinic torque ( ). Similar to the mixing width or bubble/spike heights, which is also a large-scale quantity, shows limited sensitivity to Re0. On the other hand, both and are mainly contributed by the smallest scales of the flow. Consistent with the large Reynolds number dependence observed for the vorticity contours, these vorticity-related small-scale quantities also show a large dependency on Re0.
Time evolution of (a) the normalized domain-integrated turbulent kinetic energy, , (b) the normalized domain-integrated enstrophy, , and (c) the normalized domain-integrated root mean square of baroclinic torque, , of the Ma = 0.15 cases (top row) and the Ma = 0.45 cases (bottom row) with different Reynolds numbers. Red solid line: ; green dashed line: ; and blue dash-dotted line: .
Time evolution of (a) the normalized domain-integrated turbulent kinetic energy, , (b) the normalized domain-integrated enstrophy, , and (c) the normalized domain-integrated root mean square of baroclinic torque, , of the Ma = 0.15 cases (top row) and the Ma = 0.45 cases (bottom row) with different Reynolds numbers. Red solid line: ; green dashed line: ; and blue dash-dotted line: .