Analytical self-similar solutions to two-, three-, and four-equation Reynolds-averaged mechanical–scalar turbulence models describing incompressible turbulent Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz instability-induced mixing in planar geometry are derived in the small Atwood number (Boussinesq) limit. The models are based on the turbulent kinetic energy K and its dissipation rate ε, together with the scalar (heavy-fluid mass fraction) variance S and its dissipation rate χ modeled either differentially or algebraically. The models allow for a simultaneous description of mechanical and scalar mixing, i.e., mixing layer growth and molecular mixing, respectively. Mixing layer growth parameters and other physical observables relevant to each instability are obtained explicitly as functions of the model coefficients. The turbulent fields are also expressed in terms of the model coefficients, with their temporal power-law scalings obtained by requiring that the self-similar equations are explicitly time-independent. The model calibration methodology is described and discussed. Expressions for a subset of the various physical observables are used to calibrate each of the two-, three-, and four-equation models, such that the self-similar solutions are consistent with experimental and numerical simulation data corresponding to these values of the observables and to specific canonical Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz turbulent flows. A calibrated four-equation model is then used to reconstruct the mean and turbulent fields, and late-time turbulent equation budgets for each instability-induced flow across the mixing layer. The reference solutions derived here can provide systematic calibrations and better understanding of mechanical–scalar turbulence models and their predictions for instability-induced turbulent mixing in the very large Reynolds number limit.

Advances in computing capability over the past decades have enabled increasingly detailed direct numerical simulation (DNS), large-eddy simulation (LES), and implicit large-eddy simulation (ILES) studies of turbulent mixing layers arising from classical hydrodynamic Rayleigh–Taylor (RT), Richtmyer–Meshkov (RM), and Kelvin–Helmholtz (KH) instabilities. These advances have engendered better fundamental understanding of the linear, nonlinear, transitional, and turbulent flow regimes arising from these interfacial instabilities.1,2 However, these flows are complex and depend on a large number of dimensionless flow parameters (e.g., Atwood, Mach, Reynolds, Schmidt, and Prandtl numbers), fluid properties (e.g., molecular transport coefficients of gases and liquids), and initial conditions (e.g., modal amplitudes and wavelengths of interfacial perturbations). The flow complexity of such turbulent mixing layers is further exacerbated in the high-energy-density regime in which plasma physics and radiation hydrodynamics must also be considered.3 Furthermore, many of these turbulent flows are found in astrophysical, geophysical, and engineering applications in which the range of temporal and spatial scales and other physics preclude experiments or simulations that reach the conditions required to achieve the transition to a fully turbulent state. As a result, there remains a need to develop, verify, and validate computationally less expensive, reduced-order models that adequately represent the effects of turbulence on the mean hydrodynamic flows arising in these applications. Reynolds-averaged turbulence models are thus used to render the simulation of large Reynolds number turbulent mixing computationally tractable in such applications.

A large variety of Reynolds-averaged turbulence models based on different choices of mechanical turbulent fields (e.g., turbulent kinetic energy dissipation rate or lengthscale) and of varying degrees of complexity ranging from two-equation to Reynolds stress models have been developed to model RT, RM, and KH mixing (see Ref. 2). These models contain a large number of dimensionless coefficients that are broadly specified using two different approaches: (1) most models are calibrated by iteratively modifying a subset of the coefficients until the models adequately predict some chosen set of data, and (2) a much smaller number of models are calibrated using analytic solutions to the Reynolds-averaged equations in the self-similar approximation to constrain the model coefficients in order to predict key observables, such as the mixing layer growth parameters and exponents. The models so calibrated are then applied to other, more diverse flows to quantitatively evaluate their overall predictive capability. In the second approach, the emphasis is on determining the values of the model coefficients that give specified growth parameters and other key quantities that have been measured experimentally or obtained numerically. Complete expressions for the turbulent fields, terms in the turbulent transport equations, and other derived self-similar quantities are typically not given. Moreover, previous self-similarity analyses of turbulence models for the applications of interest here primarily considered mechanical turbulence, thereby precluding them from predicting scalar mixing properties, such as molecular mixing parameters, scalar production-to-dissipation/destruction ratios, and mechanical-to-scalar timescale ratios. A recent review of simulation and modeling of RT instability and mixing experiments, including the application of Reynolds-averaged models, is available.4 

Self-similar solutions to the two-equation Kε model were derived5–7 for RT, RM, and KH mixing. Most effort has considered calibrating the coefficients in the buoyancy/shock production terms in the K and ε equations for RT and RM mixing, which are the most important, least understood, and least universal model coefficients.8 Self-similarity analysis was used to examine the growth of a small Atwood number RT mixing layer developing in a water channel experiment,6 solving a mean temperature equation using coefficients similar to those used in the standard Kε model:9,10 the value of the buoyancy production coefficient Cε0 was determined from similarity and the experimental value of the RT growth parameter α. It was shown that a Kε model with calibrated buoyancy production terms could reproduce experimentally measured reshocked RM mixing layer growth.11 This model was later extended to a Reynolds stress model to capture local anisotropy effects.12 Transport equations for the mass flux velocity and density variance were included to better capture buoyancy effects and turbulent mixing. The conditions needed for and implications of self-similarity of RT, RM, and KH mixing modeling in the context of single- and two-fluid Kε models were extensively discussed.8 Model calibration strategies were also presented in Ref. 8, in which the self-similar equations were integrated over the mixing layer to reduce the equations to easily solvable ordinary differential equations (“zero-dimensional models”) in time to obtain constraints between model coefficients and physical observables. These studies only considered mechanical turbulence.

A KL formulation, where L is a turbulent lengthscale—the Besnard–Harlow–Rauenzahn (BHR) model—was developed to model variable-density and buoyancy-driven turbulence,13 where gradient-diffusion and similarity closures were used to model the turbulent transport equations. Transport equations for the mass flux velocity ai and negative density fluctuation–specific volume fluctuation correlation b were included in the model. A single-fluid KL model was also developed,14 in which the buoyancy production term in the K equation was modeled via a relation with buoyancy–drag models. Self-similar solutions to the model equations for RT and RM mixing in the Boussinesq approximation were used to calibrate the model coefficients and demonstrate the ability of the model to reproduce experimentally measured RT and RM mixing layer widths and scaling parameters. A variant of this model that included shear production terms was applied to low energy density15 and high energy density16 turbulent shear layers. A self-similar analysis of the KLa version of the BHR model was later performed for small Atwood number RT flow, and the model was applied to predict RT, RM, and KH mixing quantities (with many model coefficient values adjusted for each instability case).17 

Self-similar solutions to an alternative formulation of the KLa model were presented18 and later extended to Kelvin–Helmholtz flows using a second turbulent lengthscale.19 The BHR and BHR-220 (i.e., KLab) models were subsequently extended to include transport equations for the Reynolds stress tensor21 and for a second lengthscale.22 Conditions for the attainment of self-similarity of turbulent fields and statistics were introduced and investigated using artificial fluid large-eddy simulation of small Atwood RT mixing.23 The mean heavy fluid mass fraction and self-normalized K, a, and b profiles from the simulation and from the KLa and BHR-2 models were compared. The predictions of several related lengthscale-based models applied to RT mixing were compared.24,25 Most recently, the Dimonte–Tipton model14 was further improved for RT, reshocked RM, and KH mixing.26–28 In particular, these studies also proposed systematic self-similar methodologies to obtain a unified set of model coefficients for different mixing cases and made several modifications to improve the accuracy and robustness of the models. The KLa models17,18 using an algebraic two-fluid expression for b describe immiscible mixing and are not appropriate for fluids that molecularly mix.

More recently, an advanced compressible, multicomponent Kε model was developed and shown to well-predict several reshocked RM instability experiments.29,30 Three- and four-equation scalar generalizations of the Kε model were subsequently proposed and calibrated a priori for RT mixing.31 Note that a mechanical turbulence-based Kε or KL model can only predict large-scale quantities, such as the bubble and spike front widths and their growth rates: additional equations describing transport and diffusion of a scalar are needed to predict higher-order mixing statistics, such as the degree of molecular mixing or mechanical-to-scalar timescale ratio. The four-equation model framework discussed here allows for a simultaneous description of large-scale (K and S) and small-scale (ε and χ) fields, while a two-equation KL or three-equation KLa model only describes large-scale fields.14,18

The self-similar model analysis, calibration, and model applications presented here extend and augment previous studies as follows. First, the K and ε equations describing RT, RM, and KH mixing are augmented by S (scalar variance) and χ (scalar variance dissipation rate) equations that model passive scalar mixing. Thus, three- and four-equation models that treat mechanical and scalar turbulence and can model scalar statistics are considered. Expressions are derived for constant self-similar parameters not considered previously, such as molecular mixing parameters, production-to-dissipation/destruction ratios, and the mechanical-to-scalar timescale ratios. Inequalities that must be satisfied by the model coefficients that follow from positivity constraints are noted to avoid unphysical coefficient choices. Explicit space–time-dependent expressions for all turbulent fields are presented in terms of the model coefficients.

Second, new calibrations of the two- and three-equation Kε models based on analytic solutions to the model coefficients as functions of the growth parameters and exponents for RT, RM, and KH mixing treated simultaneously are presented. Several calibrations are described and discussed in detail, including the methodologies for calibrating four-equation models. The predicted values of the self-similar quantities are summarized for each model and discussed. Third, the analytic expressions for the key mean and turbulent fields are evolved in space and time numerically as applied to particular small Atwood number cases for each instability to illustrate the rate of spreading in space and growth or decay in time of the fields, which has not been shown and discussed in previous studies. Furthermore, the analytic solutions are used to reconstruct the production, dissipation/destruction, and turbulent diffusion terms for each turbulent transport equation and shown at the latest evolution time to illustrate the relative importance of these contributions to the equation budgets, which has not been shown and discussed in previous studies.

Finally, a clear, concise, and complete discussion of the assumptions used to formulate the self-similar approximation, as well as of the steps required to derive the solutions, is provided. The detailed derivation of the self-similar solutions is formulated compactly and in a unified manner for all three instabilities. The steps in the derivation are clearly described to allow reproducing the results and to apply the same procedure to other fluid dynamics problems. The procedure for determining the power-laws of each mean and turbulent field is presented in detail, rather than assumed using dimensional analysis. Therefore, the emphasis of the present analysis is the physics of the self-similar solutions rather than numerical applications of calibrated Reynolds-averaged models as presented in previous work.

This paper is organized as follows. The formulation of the two-, three-, and four-equation turbulence models analyzed here is presented in Sec. II. The analytical self-similar solutions (including fields, growth parameters, and other self-similar quantities) of the models for RT, RM, and KH mixing are given in Sec. III, IV, and V, respectively. To illustrate the utility of the solutions and discuss their behavior, the two-, three-, and four-equation model coefficients are calibrated in Sec. VI and a recommended four-equation model is applied to each instability case in Sec. VII. A summary of the results and discussion are given in Sec. VIII. The self-similar solutions of the model equations for RT, RM, and KH mixing, for decaying homogeneous isotropic turbulence, and for shockless rapid compression are derived in detail in  Appendices A–C, respectively.

The three- and four-equation Reynolds-averaged Navier–Stokes (RANS) models evaluated a priori31 using direct numerical simulation (DNS) data32–34 modeling a very small Atwood number, Schmidt number Sc =7 hot/cold water channel Rayleigh–Taylor mixing experiment35 are briefly summarized here. These models are appropriate for miscible binary fluid mixing and are closely related to the models that have been used for reacting turbulent flows. The scalar fields are passive in these models as the mechanical turbulence equations do not depend on the scalar variance or its dissipation rate S or χ.

For a general incompressible field with Reynolds decomposition ϕα=ϕ¯α+ϕα and compressible field with Favre decomposition ϕα=ϕ̃α+ϕα, the Reynolds- and Favre-averaged fields are ϕ¯α and ϕ̃α=ρϕ¯α/ρ¯, respectively, and ϕα and ϕα are the Reynolds and Favre fluctuating fields, respectively. The mean scalar, mean momentum, mean internal energy, turbulent kinetic energy, turbulent kinetic energy dissipation rate, scalar variance, and scalar variance dissipation rate transport equations are (four-equation model and assuming the summation convention for repeated vector indices)

(1)
(2)
(3)
(4)
(5)
(6)
(7)

respectively, where ṽi is the mean velocity vector, gi=g0δiz is the acceleration vector, ρ¯ is the mean density, p¯ is the mean pressure, τij=ρvivj¯=τijiso+τijdev=(2/3)ρ¯Kδij+τijdev is the Reynolds stress tensor (decomposed into its isotropic plus deviatoric part), σ¯ij=2μ¯(S̃ijδijS̃kk/3) is the mean viscous stress tensor, S̃ij=(1/2)(ṽi/xj+ṽj/xi) is the mean strain-rate tensor, ν¯=μ¯/ρ¯ is the mean mixture kinematic viscosity, D¯ is the mean mixture scalar diffusivity, Sc =ν¯/D¯ is the mixture Schmidt number, ai=vi¯=ρvi¯/ρ¯ is the mass flux velocity, Ũ is the mean internal energy, h̃=Ũ+p¯/ρ¯ is the mean enthalpy, K=vivĩ/2 is the turbulent kinetic energy, ε=σijvi/xj¯/ρ¯ is the turbulent kinetic energy dissipation rate, S=s2̃ is the scalar variance, χ=D¯ρ(s/xj)2¯/ρ¯ is the scalar variance dissipation rate, Cε0Cε3 and Cχ0Cχ3 are positive dimensionless coefficients, and σs, σU, σK, σε, σS, and σχ are positive dimensionless turbulent Schmidt numbers. Here, it is assumed that there is no distinction between the coefficients {σK,σε} and {σK,σε} discussed in Ref. 31. Also, Eq. (7) analyzed here differs from that in Ref. 31, which includes Ret factors multiplying the Cχ2 and Cχ3 coefficients, where Ret=νt/(Cμν¯) is the turbulent Reynolds number and

(8)

is the turbulent viscosity (μt is the dynamic turbulent viscosity and Cμ is a positive dimensionless coefficient): as mentioned in  Appendix A, such Ret factors would break self-similarity.

Alternatively, instead of solving Eq. (7), an algebraic closure

(9)

where Cχ is a positive dimensionless coefficient, can be used for the scalar variance dissipation rate in Eq. (6) (three-equation model). A concentration variance equation similar to Eq. (6) was first introduced in the modeling of a round, turbulent free jet and closed using Eq. (9).36 Such three-equation turbulence models based on the K, ε, and S equations have been adopted in most studies of turbulent reacting flows (see Refs. 37–42 for further discussion and details). The mean scalar field can be the mean heavy fluid mass fraction s̃=m̃1 (with m̃2=1m̃1 the mean light fluid mass fraction), the mean concentration s̃=c̃, or a mean progress variable s̃=ζ̃. These fields correspond to the heavy fluid mass fraction variance S=m12̃, the density variance S=ρ2¯, the concentration variance S=c2̃, or a progress variable variance S=ζ2̃. Rather than using Eq. (9), a scalar variance dissipation rate equation similar to Eq. (7) has been used to improve the fidelity of the modeling of turbulent premixed and partially premixed combustion and flames.43–45 

The algebraic closure of the mass flux velocity is (for ideal gases with mixture ratio of specific heats γ¯)

(10)

the buoyancy-modified Reynolds stress tensor is34 

(11)

where CAijkl are dimensionless tensor coefficients. The term in the last parentheses in Eq. (11) is often called the baroclinic tensor and vanishes for a constant density flow.

The turbulent lengthscale is

(12)

where Cε is a dimensionless coefficient (often referred to as the dimensionless turbulent kinetic energy dissipation rate), the mechanical turbulent timescale is (ω is the turbulent frequency)

(13)

and the scalar turbulent timescale is

(14)

which is τs=τm/(2Cχ) for a three-equation model using Eq. (9). For the three-equation model, the mechanical-to-scalar timescale ratio is

(15)

(see Ref. 31).

The mean internal energy equation follows that in Ref. 11, and thermal conduction is neglected. Using the identity p¯vj¯/xj=vj¯p¯/xj(/xj)(p¯vj¯) and the closure p¯vj¯=(μt/σU)(/xj)(p¯/ρ¯) arising from Eq. (10) with γ = 1 and the assumption σρ=σU in this term (which will be verified later using the similarity analysis), the exact equation

(16)

reduces to Eq. (3).

The fifteen dimensionless model coefficients are σs, σU, σK, σε, σS, σχ,CAijkl,σρ,Cμ,Cε0,Cε1,Cε2,Cχ0,Cχ2, and Cχ3. The pressure–dilatation correlation is neglected in the K and ε equations. Although the shear production term τijṽi/xj in Eqs. (4) and (5) was shown to be very small in small Atwood number RT flow,34 they are included above for completeness and are required for KH mixing. The gradient of the deviatoric part of the Reynolds stress component τzzdev was shown to be nonnegligible in very small Atwood number RT mixing.34 

Equations (1)–(7) allow the model to be applied to flows with different Schmidt numbers. It is expected that solving a transport equation for χ will give better predictions than using an algebraic model in which the mechanical-to-scalar timescale ratio (i.e., Cχ) is a constant independent of the model coefficients. It should be noted that mixing in RT and RM flows cannot be predicted solely using a mechanical turbulence model (i.e., only based on K, ε or L): transport equations describing the evolution of a scalar mixing progress variable (and perhaps its dissipation rate) must also be included. This is also true for higher-order (i.e., second-order or second-moment) models.

Consider the one-dimensional form of the model equations (1)–(6) and either the algebraic model (9) or the transport equation (7) for which the mixing layer is within z[hs(t),hb(t)] and hs(t)<0. The mixing layer width is h(t)=hb(t)hs(t)2hb(t) for very small Atwood number At =(ρ1ρ2)/(ρ1+ρ2)1 (i.e., the bubble and spike front growth is nearly symmetric).

To analytically derive the self-similar solutions in the Boussinesq approximation, assume

  1. small density difference ρ1ρ20 (or At 0);

  2. no turbulence outside the mixing layer [K=ε=S=χ=0 for |z|>h(t)/ξ, where the scale parameter ξ is discussed in Sec. 1];

  3. no mean vertical velocity field in a Galilean reference frame moving with the mixing layer (w¯=w̃+w¯=0, which implies that the mean vertical velocity is given by w̃=az);

  4. incompressible, so that ṽi/xi=0;

  5. σ¯ij=σ̃ij=0; and

  6. molecular viscosity and diffusivity contributions to the turbulent transport are negligible compared to the turbulent contributions, i.e., νtν¯ and DtD¯, where Dt=νt/Sct is the turbulent diffusivity (Sct is the constant turbulent Schmidt number, taken to be σS here).

Therefore, in the small Atwood number limit of a statistically one-dimensional RT, RM, or KH mixing layer, the transport equations can be simplified by neglecting the gradients in the homogeneous (x and y) directions, so that the resulting model only considers variations across the mixing layer (the vertical z-direction in the convention adopted here). The mean pressure gradient in Eqs. (4) and (5) is taken to be in hydrostatic equilibrium

(17)

(g0= constant > 0 for canonical RT flow). This approximation decouples the mean momentum equation (2) if advection is neglected and was used previously.5–7 Numerical simulations showed that the mean pressure satisfies hydrostatic equilibrium closely at sufficiently late times (for example, see Ref. 34). The mean pressure gradient term in Eq. (10) is neglected as in the incompressible limit the mean mixture sound speed is cs=γ¯p¯/ρ¯. The resulting simplified equations are (A1)(A7) for all three instability cases.

Analytic self-similar solutions were obtained for the Kε model,5–7 the KL model,14 the KLa model,17,18 and the K–2La model.19 The two- and three-equation model studies only considered Rayleigh–Taylor and Richtmyer–Meshkov mixing, and the four-equation model considered mixing induced by all three instabilities considered here. The steps required to obtain these solutions are as follows:

  1. assume mean fields that are linear in the similarity variable η=η(z,t) and inverse parabolic in η for the turbulent fields;

  2. for each turbulent field, assume a separable form consisting of the assumed spatiotemporal profile multiplied by a power-law in time and a prefactor, ϕα(z,t)=Aϕαf(η)tpϕα;

  3. substitute these fields into the simplified one-dimensional mean and turbulence model equations;

  4. require each resulting equation to be explicitly time-independent, which determines the exponents {pϕα};

  5. in these explicitly time-independent equations require the constant terms and terms proportional to powers of η to separately equal zero; and

  6. solve these equations for the turbulent Schmidt numbers {σϕα}, mixing layer parameters and exponents appearing in the mixing layer widths, and prefactors {Aϕα} as functions of the model coefficients.

This procedure can be performed in (z, t)-space8,14,18,19 or in η-space5–7 and yield equivalent results. As will be shown in  Appendix A, strict self-similarity imposes a strong constraint on the turbulent Schmidt numbers: they must be equal, Eq. (A48). For additional discussion of self-similarity assumptions and solution methods, see Refs. 46 and 47.

1. General mixing layer width

For each instability, the self-similar mixing layer width grows as a generalized power-law

(18)

where Ah is a constant depending on characteristic parameters for each instability and model coefficients, and the dimensional considerations given the flow parameters in Ah determine the value of the exponent θh. The width (18) corresponds to the statistically one-dimensional growth of a three-dimensional mixing layer arising from multimode initial perturbations in a fully developed turbulent state. Note that the corresponding mixing layer velocity and acceleration are

(19)
(20)

respectively.

The dimensionless similarity variable is

(21)

with ξ = 1 and 2 corresponding to the interpretation of h as the layer half-width (i.e., hb) and width, respectively. From the definition of η(z,t), it follows that the temporal and spatial partial derivatives transform as

(22)

2. Self-similar turbulent field profiles

The self-similar turbulent fields have the general form

(23)

with inverted parabolic profile

(24)

which is symmetric about the centerline z=η=0. The requirement that f(·) is zero outside of the mixing layer is needed for realizability. The temporal and spatial derivatives

(25)
(26)

follow directly from Eqs. (23) and (24). The gradients of the turbulent fields are linear and are positive for z < 0 and negative for z > 0.

3. Self-similar mean and turbulent fields

The mean and turbulent fields are expressed in the separable form

(27)

where the time-dependent functions {fϕα(t)}={tpϕα} are determined by requiring that the transformed self-similar equations be explicitly independent of time (see  Appendix A). For the profile Eq. (24), the temporal and spatial derivatives (25) and (26) become

(28)
(29)

The coupled nonlinear equations have compactly supported inverted parabolic solutions

(30)

with Dirichlet boundary conditions ϕα(η=±1)=0. The constant prefactors {Aϕα} are functions of the model coefficients and other parameters (e.g., g0, At etc.). The details of the determination of the relationships between model coefficients and the field prefactors in (30) are given in  Appendix A.

The maximum or minimum value of a turbulent field is given by its centerline value

(31)

Using Eq. (24) and the profile integrals,

(32)

it follows that the time-dependent turbulent fields integrated over the mixing layer are

(33)

The simplest, lowest order linear mean field profiles common to all three instabilities are the mean scalar for miscible mixing

(34)

satisfying s̃(η=1)=0 and s̃(η=1)=1 for s=m1, c, and ξ. The mean density for miscible mixing is

(35)

with average density ρ0=(ρ1+ρ2)/2 and Δρ=ρ1ρ2>0 satisfying ρ¯(η=1)=ρ2 and ρ¯(η=1)=ρ1. In the self-similar equations, the mean field gradients are

(36)

Using Eq. (18), dimensional analysis gives the general temporal scalings

(37)

For each instability considered here, the mean and turbulent fields can be used to analytically reconstruct the self-similar time-evolution, production, dissipation/destruction, turbulent diffusion, and advection terms in the one-dimensional transport equations

(38)

In the gradient-diffusion approximation, the turbulent diffusion term is the negative gradient of the associated turbulent field flux,

(39)

which has an approximate negative sinusoidal profile with Fϕα(z,t)<0 for z < 0 and Fϕα(z,t)>0 for z > 0. The advection term is

(40)

which has the same profile shape as Fϕα(z,t).

The solutions of the self-similar equations are summarized here; see  Appendix A for details on their derivation.

As in previous similarity analyses of incompressible Rayleigh–Taylor mixing, the mean velocity w̃ and shear production τzzw̃/z are assumed to be zero, and the mean density in the turbulent diffusion terms and in the denominator of the buoyancy production terms is approximated by ρ¯ρ0. Thus, Eqs. (1)–(7) reduce to

(41)
(42)
(43)
(44)
(45)
(46)

Equations (45) and (46) have not been previously considered in self-similarity analysis, but Eq. (45) has been previously considered in the similarity analysis of the KLaV model with V the mass fraction variance.48 

Explicit expressions for the RT self-similar turbulent fields corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Two- and three-equation models

For the two- and three-equation models, using the solutions derived in  Appendix A, Eqs. (A56)–(A59), the turbulent kinetic energy, turbulent kinetic energy dissipation rate, scalar variance, and scalar variance dissipation rate are functions of f(·) given in Eq. (24),

(47)
(48)
(49)
(50)

Note that if 4Cε03>0, then Cε2>Cε0 in order for K(z,t)>0. Morgan et al.48 obtained AS=(1θm)/4 (θm is the molecular mixing parameter) but not its dependence on the model coefficients.

The mass flux velocity, turbulent viscosity, turbulent lengthscale, mechanical turbulent timescale, and scalar turbulent timescale are

(51)
(52)
(53)
(54)

Note that az(z,t)<0 and the turbulent timescales are independent of z. The scalar fields depend on the coefficients {Cεq} in the ε equation, while the mechanical fields are independent of S and χ (i.e., the scalar fields are passive).

2. Four-equation model

For the four-equation model, not assuming an algebraic model (9), the coefficients AS and Aχ in

(55)
(56)

are given by Eqs. (A73) and (A74), respectively. The scalar turbulent timescale is

(57)

These expressions are the same for RM and KH mixing, but with the mixing layer width h(t) and coefficients AS and Aχ corresponding to those instabilities.

Explicit expressions for the RT self-similar mixing layer growth parameter, ratio of generated turbulent kinetic energy to released potential energy, and molecular mixing parameter corresponding to the three- and four-equation models are presented here as a function of the model coefficients.

1. Mixing layer growth parameter

For small Atwood number Rayleigh–Taylor instability, the self-similar mixing layer width grows as a power-law (18) with θh=2 for a constant acceleration,

(58)

with Ah=α At g0. The mixing layer velocity and acceleration are

(59)

Using Eq. (A40), the dimensionless mixing layer growth parameter is5–8 

(60)

which exhibits the explicit dependence on the model coefficients; αb=α/2. Note that α is degenerate, in that a given value can be obtained using different combinations of the coefficients. The coefficients must satisfy the inequalities Cε2,Cε0>3/4 to ensure that α>0; the choice Cε0=Cε2 is excluded as it results in no growth.

It can be shown that the buoyancy production term in the ε equation cannot be set to zero as follows. If Cε0=0, Eq. (60) simplifies to α=[2Cμ/(3σρ)]Cε22/(34Cε2) and positivity requires that Cε2<3/4, which violates the requirement that Cε2>1 for positivity of the RM turbulent timescales [see Eq. (82)] and for decaying isotropic turbulence (see  Appendix B). Although similarity requires σρ=σs, this will not be assumed in the presentation of the results for generality and to allow for additional model calibration options.

As expected from dimensional analysis, K(dh/dt)2(g0Att)2,ε(dh/dt)2/t(g0At)2t, S is dimensionless (for a mass fraction variance), χ1/t,azdh/dtg0At2t,νthdh/dt(g0At)2t3,Lhg0 Att2, and τm,τst. In particular, {K,ε,az,νt,L}0 as g0At 0, as required in the limit of a stable interface. Using Eqs. (53), (58), and (33), the turbulent lengthscale-to-mixing layer width ratio is

(61)

2. Ratio of generated turbulent kinetic energy to released potential energy

The dimensionless ratio of the generated turbulent kinetic energy to released potential energy, ρ0K/ΔΦ, has been obtained from RT numerical simulations49,50 and experiments,51,52 and has been used to calibrate turbulent lengthscale-based Reynolds-averaged models for RT mixing.14,18,19

The released potential energy is

(62)

which is proportional to α. Integrating Eq. (47) over the mixing layer using Eq. (32) gives

(63)

and it follows that the ratio of generated turbulent kinetic energy to released potential energy is

(64)

Note that ρ0K/ΔΦ>0 is singular for Cε2=3/4, and the inequality (Cε2Cε0)/(4Cε23)>0 should be satisfied; as Cε2>3/4, it follows that Cε2>Cε0. Both Eqs. (60) and (64) depend on the difference Cε2Cε0, the magnitude of which is related to the competition between turbulent kinetic energy dissipation rate buoyancy production (i.e., Cε0) and destruction (i.e., Cε2).

3. Molecular mixing parameter

The molecular mixing parameter

(65)

is a constant (and positivity requires AS<1/4), as s̃(1s̃)=f(ξz/h(t))/4, and is also equal to the spatially integrated expression (in the self-similar state),

These expressions also apply to Richtmyer–Meshkov and Kelvin–Helmholtz mixing.

Using Eq. (A58), it follows that the molecular mixing parameter is

(66)

for the three-equation model, which must be positive. For specific, fixed values of Cε0 and Cε2, θm is determined by the value of Cχ.

For the four-equation model,

(67)

with AS given by Eq. (A73). Expression (66) is considerably simpler than (67) and depends only on three coefficients. This expression is the same for RM and KH mixing, but with the coefficient AS corresponding to these instabilities.

Explicit expressions for the RT self-similar production-to-dissipation/destruction and mechanical-to-scalar timescale ratios corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Production-to-dissipation/destruction ratios

The turbulent kinetic energy production-to-dissipation ratio is

(68)

and the turbulent kinetic energy dissipation rate production-to-destruction ratio is

(69)

As Cε2>Cε0,Pε/Dε<PK/DK>1. For the three-equation model, the scalar variance production-to-dissipation ratio is

(70)

These quantities must be positive and are singular for Cε0=3/4. Increasing the value of Cχ decreases the value of PS/DS>1.

For the four-equation model,

(71)

and the scalar variance dissipation rate production-to-destruction ratio is

(72)

with AS and Aχ given by Eqs. (A73) and (A74). These quantities must be positive and are singular if the denominators vanish.

2. Mechanical-to-scalar timescale ratio

For the three-equation model, the mechanical-to-scalar timescale ratio is (15) and for the four-equation model,

(73)

which is singular for Cε0=3/4, with AS and Aχ given by Eqs. (A73) and (A74).

The solutions of the self-similar equations are summarized here; see  Appendix A for details on their derivation.

The self-similar equations describing the growth of incompressible Richtmyer–Meshkov mixing are formally the same as those for Rayleigh–Taylor mixing (41)–(46), but with an impulsive, rather than constant, acceleration:

(74)

where Δvs>0 is the velocity jump due to the impulse (or “shock” passage) across the initial interface, and δ(·) is the Dirac delta function. Thus, the “shock” production terms exist only at t = 0 (the time of “shock” passage through the interface), and the equations subsequently describe a decaying turbulence state. Note that this decaying turbulence is inhomogeneous due to the turbulent diffusion terms (which are absent in homogeneous, isotropic decaying turbulence as discussed in Appendix B 1). The production terms set the initial conditions, i.e., Δvs.

Explicit expressions for the RM self-similar turbulent fields corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Two- and three-equation models

For the two- and three-equation models, using the solutions derived in  Appendix A, Eqs. (A60)–(A63), the turbulent kinetic energy, turbulent kinetic energy dissipation rate, scalar variance, and scalar variance dissipation rate are functions of f(·) given in Eq. (24),

(75)
(76)
(77)
(78)

The mass flux velocity, turbulent viscosity, turbulent lengthscale, mechanical turbulent timescale, and scalar turbulent timescale are

(79)
(80)
(81)
(82)

Positivity of the turbulent timescales requires Cε2>1.

2. Four-equation model

For the four-equation model, not assuming an algebraic model (9), the coefficients AS and Aχ in Eqs. (55) and (56) are given by Eqs. (A75) and (A76), respectively. The scalar turbulent timescale is given by Eq. (57) with these coefficients.

Explicit expressions for the RM self-similar mixing layer growth exponent, molecular mixing parameter corresponding to the three- and four-equation models, and mechanical and scalar field decay exponents are presented here as a function of the model coefficients.

1. Mixing layer growth exponent

For small Atwood number Richtmyer–Meshkov instability, the self-similar mixing layer width grows as a power-law (18) with θh=θ,

(83)

with Ah=h0/t0θ=h01θ(Δvs At +/θ)θ,t0=θh0/(Δvs At +), At+ is the post-shock Atwood number, and h(t0)=(ξ/2)h0. The mixing layer velocity and acceleration are

(84)

The dimensionless mixing layer growth exponent that follows from inverting Eq. (A65) is5,7,8

(85)

and only depends on Cε2; growth requires Cε2>3/2. The growth exponent is singular for Cε2=1.

Note that h(t) depends on a timescale t0 and initial lengthscale h0, as distinct from RT mixing; h0 is interpreted as an initial width of a mixing layer front, and not an amplitude (such as in single-mode linear and nonlinear theory). The dependence on At+ ensures that there is no mixing layer growth if the “post-shock” densities are equal (i.e., At+=0). The dependence of t0 on θ is motivated by requiring that the initial velocity due to the impulse is dh/dt|t0=(ξ/2)Δvs At+.

A relationship between the growth rate of RM mixing and the decay of turbulent kinetic energy was proposed53 using the power-laws predicted by the Kε model for decaying isotropic turbulence (see Appendix B 4) (where, however, θ was incorrectly interpreted as θb and was derived for homogeneous decaying turbulence—different from the expression for RM mixing). The analytic result θ=1/3 for three-dimensional RM instability was also derived,54 and is consistent with recent numerical simulation55 and analytical56 results. Note that ah(t)<0 with θ<1.

As expected from dimensional analysis, K(dh/dt)2(h0/t)2(t/t0)2θ,ε(dh/dt)2/t(h02/t3)(t/t0)2θ, S is dimensionless (for a mass fraction variance), χ1/t,azdh/dt(h0/t)(t/t0)θ,νthdh/dt(h02/t)(t/t0)2θ,Lhh0(t/t0)θ, and τm,τst. In particular, {K,ε,az,νt,L}0 as Δvs At+0, as required in the limit of a stable interface. Using Eqs. (81), (83), and (33), the turbulent lengthscale-to-mixing layer width ratio is

(86)

2. Molecular mixing parameter

Using Eqs. (65) and (A62), the molecular mixing parameter is

(87)

for the three-equation model [Cε2=Cε2(θ)] and must be positive.

For the four-equation model,

(88)

with AS given by Eq. (A75). Expression (87) is considerably simpler than (88) and depends only on two coefficients. For a specific, fixed value of Cε2 (or θ), θm is determined by the value of Cχ.

3. Decay exponents of mechanical and scalar turbulent fields

Following “shock” passage at t = 0, the production terms vanish and the turbulence is in a decaying state described by dissipation/destruction and diffusion. The power-law decay exponents of the mechanical–scalar turbulent fields follow by integrating Eqs. (75)–(78) over the mixing layer using Eq. (32) so that

(89)
(90)
(91)
(92)

For the four-equation model,

(93)
(94)

with AS and Aχ given by Eqs. (A75) and (A76).

Explicit expressions for the RM self-similar production-to-dissipation/destruction and mechanical-to-scalar timescale ratios corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Production-to-dissipation/destruction ratios

After “shock” passage at t = 0, the turbulent kinetic energy production-to-dissipation ratio and turbulent kinetic energy dissipation rate production-to-destruction ratio are zero. For the three-equation model, the scalar variance production-to-dissipation ratio is

(95)

and must be positive; this quantity is singular for θ=2/3. Increasing the value of Cχ decreases the value of PS/DS, which is greater than unity.

For the four-equation model, the scalar variance production-to-dissipation ratio is

(96)

and the scalar variance dissipation rate production-to-destruction ratio is

(97)

with AS and Aχ given by Eqs. (A75) and (A76). These quantities must be positive and are singular if the denominators vanish.

2. Mechanical-to-scalar timescale ratio

For the three-equation model, the mechanical-to-scalar timescale ratio is (15) and for the four-equation model,

(98)

which is singular for θ=2/3 (and θ<2/3 for R > 0), with AS and Aχ given by Eqs. (A75) and (A76).

The solutions of the self-similar equations are summarized here; see  Appendix A for details on their derivation.

For incompressible Kelvin–Helmholtz instability, the mean momentum equation reduces to an equation for the mean shear velocity v¯, and w¯=0. Equations (1)–(5) reduce to (approximating ρ¯ρ0 in the diffusion terms)

(99)
(100)
(101)
(102)
(103)

together with Eqs. (45) and (46) for S and χ. The shear Reynolds stress is modeled as

(104)

and the mean pressure gradient vanishes by symmetry. The dimensionless coefficient Cdev allows for a different value of Cμ in the closure of the shear production terms for KH mixing and was previously introduced in the similarity analysis of the K–2La model.19 

The mean shear velocity is assumed to have a linear profile,

(105)

with average stream velocity v0=(v1+v2)/2, satisfying v¯(η=1)=v01At2Δv/2 and v¯(η=1)=v0+1At2Δv/2 [and reducing to v¯(η=1)=v2 and v¯(η=1)=v1 for At =0], where the velocity Atwood number is

(106)

so that (105) has a similar form as the mean density (35) in RT flow. The factor 1At2 (unity in the single-fluid, constant density case) reduces the growth of the instability with increasing density contrast Δρ and multiplies Δv in the expression for the linear instability growth rate.57–60 The constant mean shear velocity gradient is

(107)

and has a similar form as the mean density gradient (36). Note that v¯(z,t) and v¯(z,t)/z are independent of At because η1/h(t)1/1At2 [see Eq. (117)]. Note that there is no overall time dependence for the mean shear velocity as fv=tθh1=1.

Explicit expressions for the KH self-similar turbulent fields corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Two- and three-equation models

For the two- and three-equation model, using the solutions derived in  Appendix A, Eqs. (A66)–(A69), the turbulent kinetic energy, turbulent kinetic energy dissipation rate, scalar variance, and scalar variance dissipation rate are functions of f(·) given in Eq. (24),

(108)
(109)
(110)
(111)

The turbulent viscosity, turbulent lengthscale, and mechanical and scalar turbulent timescale are

(112)
(113)
(114)

The turbulent timescales are independent of z.

The shear Reynolds stress (104) is

(115)

and the centerline normalized shear stress is

(116)

which depends only on δ [see Eq. (119)] and At. The Atwood number dependent factors in Eqs. (108), (109), (112), (113), and (115) reduce the values of each of these quantities with increasing At.

2. Four-equation model

For the four-equation model, not assuming an algebraic model (9), the coefficients AS and Aχ in Eqs. (55) and (56) are given by Eqs. (A77) and (A78), respectively. The scalar turbulent timescale is given by Eq. (57) with these coefficients.

Explicit expressions for the KH self-similar mixing layer growth parameter, normalized turbulent kinetic energy and turbulent kinetic energy dissipation rate, and molecular mixing parameter corresponding to the three- and four-equation models are presented here as a function of the model coefficients.

1. Mixing layer growth parameter

For incompressible Kelvin–Helmholtz instability, the self-similar mixing layer width grows as a power-law (18) with θh=1,

(117)

with Ah=δ1At2|Δv| and Δv=v1v2 the difference between the velocities of the coflowing streams. The mixing layer velocity and acceleration are

(118)

Using Eq. (A40), the dimensionless mixing layer growth parameter is7,8

(119)

and growth requires Cε2>Cε1. Note that δ (like α) is degenerate, in that a given value can be obtained using different combinations of the coefficients.

With the density ratio sρ1/ρ2>1, At (s)=(s1)/(s+1) and

(120)

which very well describes the reduction in mixing layer growth obtained from the experiments61 and DNSs of variable-density Kelvin–Helmholtz mixing with density ratios s = 1, 2, 4, and 8.62 

As expected from dimensional analysis, K(dh/dt)2(Δv)2,ε(dh/dt)2/t(Δv)2/t, S is dimensionless (for a mass fraction variance), χ1/t,νthdh/dt(Δv)2t,Lh|Δv|t, and τm,τst. In particular, {K,ε,νt,L}0 as Δv0, as required in the limit of a stable interface. Using Eqs. (113), (117), and (33), the turbulent lengthscale-to-mixing layer width ratio is

(121)

2. Normalized turbulent kinetic energy and normalized turbulent kinetic energy dissipation rate

The dimensionless centerline turbulent kinetic energy normalized by (Δv)2 is

(122)

Integrating Eq. (108) over the mixing layer and using Eqs. (117) and (32), it follows that the integrated turbulent kinetic energy normalized by |Δv|3 is

(123)

which evolves linearly with time in the self-similar state.

Integrating Eq. (109) over the mixing layer and using Eqs. (117) and (32), it follows that the integrated turbulent kinetic energy dissipation rate normalized by |Δv|3 is

(124)

The values of each of these ratios decrease with increasing Atwood number.

Equations (119), (122), (123), and (124) depend on the difference Cε2Cε1, the magnitude of which is related to the competition between turbulent kinetic energy dissipation rate shear production (i.e., Cε1) and destruction (i.e., Cε2). Equations (122)–(124) must be positive, again requiring Cε2>Cε1.

3. Molecular mixing parameter

Using Eqs. (65) and (A68), the molecular mixing parameter is

(125)

for the three-equation model and must be positive. For specific, fixed values of Cε1 and Cε2, θm is determined by the value of Cχ.

For the four-equation model,

(126)

with AS given by Eq. (A77). Expression (125) is considerably simpler than (126) and depends only on three coefficients.

Explicit expressions for the KH self-similar production-to-dissipation/destruction and mechanical-to-scalar timescale ratios corresponding to the two-, three-, and four-equation models are presented here as a function of the model coefficients.

1. Production-to-dissipation/destruction ratios

The turbulent kinetic energy production-to-dissipation ratio is

(127)

and the turbulent kinetic energy dissipation rate production-to-destruction ratio is

(128)

so that production is exactly balanced by destruction of the turbulent kinetic energy dissipation rate. It follows from Cε2>Cε1 that PK/DK>1.

For the three-equation model, the scalar variance production-to-dissipation ratio is

(129)

Increasing the value of Cχ decreases the value of PS/DS, which is greater than unity. For the four-equation model,

(130)

and the scalar variance dissipation rate production-to-destruction ratio is

(131)

so that production of the scalar variance dissipation rate is exactly balanced by destruction, with AS and Aχ given by Eqs. (A77) and (A78). These quantities must be positive and are singular if the denominators vanish.

2. Mechanical-to-scalar timescale ratio

For the three-equation model, the mechanical-to-scalar timescale ratio is (15) and for the four-equation model,

(132)

with AS and Aχ given by Eqs. (A77) and (A78).

The relationships between the Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz mixing layer parameters and the model coefficients derived in Secs. III–V are used here to calibrate the two-, three-, and four-equation models. It may be useful to calibrate turbulence models to be able to predict the specific values of observables (e.g., in applications where the Reynolds number is very large). The models are calibrated by solving for the coefficient values that are consistent with the values of some chosen set of self-similar physical observables, as discussed below. Similar calibrations have been performed in previous similarity analyses.8,14,18,19 The number of constraints used must equal the number of model coefficients being calibrated: the calibration process is not unique, as there is freedom to choose which coefficient values are fixed a priori and which constraints will be applied.

Different considerations related to the intended application of a model often dictate which constraints are most relevant and what type of model should be used. Assuming consistent calibrations for each type of model, two-, three-, and four-equation models describe mechanical turbulence the same way and with identical profiles in space and time. If only the bulk mechanical turbulence is to be described, a two-equation model will suffice for which sufficient data are available for coefficient calibration for all three instabilities. If it is also important to describe molecular mixing, then a three-equation model is minimally required: the calibration of the additional coefficient is straightforward but the value is generally different for each instability case.

A four-equation model can predict all of the quantities predicted by two- and three-equation models as well as some set of additional scalar turbulence statistics that are used to calibrate the scalar dissipation rate equation coefficients {Cχq}. As discussed in Sec. VI E, calibration of a four-equation model for RT, RM, and KH mixing is currently challenging, as little scalar turbulence data have been obtained from experimental and numerical simulation data (that ideally should be self-similar). The scalar self-similar observable expressions are also highly nonlinear in the model coefficients, making the calibration much more difficult than for a three-equation model. The three- and four-equation models are less universal than the two-equation model, as different scalar equation coefficient values must be used for each instability case. In numerical (rather than analytical) applications, solving the least number of turbulence model equations is often desirable for robustness across a wider range of flow conditions.

The sixteen (eleven independent allowing σsσρ and σs1/Cdev) coefficients to be calibrated are

(133)

and the constant self-similar physical observables available to determine these values are

  1. α and ρ0K/ΔΦ for RT mixing;

  2. θ for RM mixing;

  3. δ, K(0,t)/(Δv)2, and ε0/|Δv|3 for KH mixing;

  4. PK/DK and Pε/Dε for RT mixing and PK/DK for KH mixing;

  5. PS/DS for RT, RM, KH mixing and a three- or four-equation model;

  6. Pχ/Dχ for RT, RM, and KH mixing and a four-equation model;

  7. θm for RT, RM, KH mixing and a three- or four-equation model; and

  8. R for RT, RM, KH mixing and a three- or four-equation model.

An assumed value of ρ0K/ΔΦ has been used in previous calibrations of turbulent lengthscale models.14,18,19 The production-to-dissipation/destruction ratios and mechanical-to-scalar timescale ratios have not been previously derived for these mixing cases for potential use in model calibration. A hierarchy of increasingly more complete coefficient calibrations can be constructed by assuming successively fewer assumed coefficient values and incorporating more physical constraints. From the mixing layer growth parameters (60), (85), and (119), it is clear that RT, RM, and KH mixing are associated with the turbulent kinetic energy dissipation rate equation coefficients {Cε0,Cε2},Cε2, and {Cε1,Cε2}, respectively. From Eq. (C4), Cε3=2, which is common to all of the calibrations and is relevant to compressible applications. For the four-equation model, the scalar dissipation rate equation model coefficients Cχ0,Cχ2, and Cχ3 are associated with θm, R, and the scalar production-to-dissipation/destruction ratios, i.e., three self-similar observable values are needed to uniquely determine the values of these three coefficients.

The steps in the calibration can be summarized as follows to solve the constraint:

  1. equation for θ for Cε2;

  2. equations for δ and K(0,t)/(Δv)2 for CμCdev and Cε1;

  3. equations for α and ρ0K/ΔΦ for σρ and Cε0;

  4. equation for θm for Cχ (three-equation); and

  5. equations for PS/DS,Pχ/Dχ, θm or R for Cχ0,Cχ2, and Cχ3.

For the purposes of model calibration and illustration of the self-similar solutions derived here, the following five physical observable values pertaining to each instability case will be used (universality of these values is not implied):

(134)

A wide range of values of α has been reported from experiments and numerical simulations,1 depending on the specific details of the initial conditions and perhaps other factors. The value α2αb0.12 represents a value typical of small Atwood number RT DNSs (At =0.01, Sc = 1)63 and gas channel RT experiments using air and helium (At 0.04, Sc 0.7).64 The value ρ0K/ΔΦ0.5 is typical of values obtained from RT water and gas channel experiments51,52 and numerical simulations23,49,50,65,66 The exponent θ0.3 is consistent with shock tube experimental RM data prior to reshock,67,68 numerical simulations using narrowband initial perturbations,55,56 and modal modeling54,69 The values δ0.08 (often called the “spreading parameter”10) and K(0,t)/(Δv)20.035 are consistent with incompressible shear layer (KH) experiments in air10,70 and DNSs and LESs of these experiments.62,71–73

Evidence from both experiments and numerical simulations indicates that the asymptotic values of α, θ, and δ depend on the spectral characteristics of the multimode initial conditions (e.g., broadband vs narrowband, laminar vs turbulent, and inflow).61,63,73–80 RT mixing experiments using salt/fresh water with Sc 103 showed that Schmidt number effects also influence the observable values.81 DNSs of RM mixing also showed that θ and other statistics, such as θm depend on the initial Reynolds number.82 This implies that the mixing layer parameters are also sensitive to the initial conditions (and this is currently poorly understood). It is implicitly assumed in the similarity analysis that mixing layer parameters are functions only of the constant model coefficients and are independent of initial conditions. Consequently, applications of models calibrated using similarity to cases with differing initial conditions require recalibration. Note that a model with highly constrained coefficient values is not necessarily a better model in practice than a less constrained model. Adjusting the model coefficient values does not change the shapes of the assumed spatial mean and turbulent field profiles.

Several possible hierarchical calibrations of the Kε model equations are described below. The choice of which calibration should be used depends on the intended use of the model, which is briefly discussed for each option. As in previous studies, a value of Cμ will be assumed.

Various self-similar and nonself-similar calibration options for the mechanical turbulence equations are developed and discussed here.

1. Calibration of Cε0,Cε1, andCε2for Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz mixing

Consider a model intended to describe prescribed mixing layer growth for all three instabilities. With the reasonable assumption that the values of α, θ, and δ are independent and can, therefore, be used to calibrate these coefficients, Eq. (85) gives

(135)

Eq. (119) gives

(136)

and Eq. (60) gives

(137)

From Eq. (B22), the relationship between the ε equation destruction coefficient and isotropic turbulence decay exponent is

(138)

and equating Eqs. (135) and (138) gives the linear relationships between n and θ,

(139)

The value θ=0.3 implies n = 1.10, which is essentially the same value used in previous self-similar analyses of L-based models (which used the slightly smaller value θ=0.25).14,18,19

This calibration is similar to that used for a buoyancy–drag–shear model describing all three instabilities, in which the analogous coefficients were calculated using specific values of α, θ, and δ.83 The nonlinearity of these expressions implies that the coefficient values are quite sensitive to changes in α, θ, or δ. Note the Cμ always appears multiplied by Cdev in the expression for Cε1: the value of Cε1 is unchanged if the values of both Cμ and Cdev are adjusted keeping their product fixed.

The simplest calibration that incorporates the three instabilities considered here is as follows. Assume that

(140)

as required by strict self-similarity, taking the standard value Cdev = 1 for the Kε model. Such a choice is motivated by retaining the standard value of Cμ and σϕαO(1), which would also allow the model to give good predictions for boundary layer and other canonical turbulent free-shear flows (i.e., mixing layers, jets, and wakes) for which the Kε model was originally developed and calibrated.10,84–87 The standard Kε model coefficient values are Cμ=0.09,Cε1=1.44,Cε2=1.92,σK=1.00, and σε=1.30σK (these standard values of σK and σε were used in previous Kε models for RT, RM, and KH mixing.6,8) The values σϕα=1 were also assumed in the calibration of the KL model for RT and RM mixing.14 

With α, θ, and δ from (134), it follows that

(141)

These values of Cε1 and Cε2 are 1.2% larger and 0.6% smaller than their standard values. As this calibration uses three physical observables corresponding to the growth parameters and exponent (and assumed values of Cμ and Cdev), it yields a model that predicts the self-similar mixing layer growths for each of the three instabilities, but other quantities are unconstrained, such as ρ0K/ΔΦ and K(0,t)/(Δv)2: using the values in Eq. (141) gives ρ0K/ΔΦ=0.922σs/σρ and K(0,t)/(Δv)2=0.03(1At2). Calibrations that also predict specific values of these two quantities are described below.

2. Calibration of Cε0,Cε1,Cε2, andσρfor Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz mixing

Let Ξ=ρ0K/ΔΦ,ϒ=K(0,t)/(Δv)2, and σϕα=σs. The value of Cε2 is given by Eq. (135). The values of Cε1 and CμCdev are determined by solving Eqs. (119) and (122):

(142)
(143)

which depend on the Atwood number (this dependence propagates to other coefficients, as shown below). Note that the K and ε shear production terms PKs=Cdevνt(v¯/z)2 and Pεs=Cε1(ε/K)PKs in Eqs. (102) and (103) are proportional to CμCdev(18ϒAt2)(1At2) and CμCdevCε1(18ϒAt2)2, respectively, which decrease with the increasing Atwood number, thereby reducing instability and turbulence growth. The observables Ξ and ϒ are also very likely Atwood number dependent. Note that for KH mixing, Eqs. (A48) and (A50) require σϕα=1/Cdev for strict self-similarity. However, it also is possible to relax this constraint such that the resulting model is not strictly self-similar.

The values of Cε0 and σρ are determined by solving Eqs. (60) and (64),

(144)
(145)

Equations (143) and (145) both involve the ratio Cμ/σρ=CμCdev and cannot be satisfied simultaneously, showing that the strictly self-similar assumption σρ=σϕα=1/Cdev is inconsistent for RT and RM mixing. Therefore, assume a value Cμ=0.087 in Eq. (143).

Assume that

(146)

(breaking strict self-similarity). With the values in (134), it follows that

(147)

For At =0,Cε1=1.375, and Cdev = 0.678, the values of Cε1 and Cε2 are 4.5% smaller, and 0.6% smaller than their standard values. The value of Cdev is much smaller than the value required by the K–2La model.19 This calibration indicates that fully self-similar RT and RM mixing with typical values of α and θ leads to turbulent Schmidt numbers more than a factor 20 smaller than those for incompressible shear flows. Note that Cε0 does not depend on At, and positivity of Cdev requires At <0.72=0.849 (this critical Atwood number is set by the value 18ϒ=0.72, which corresponds to At =0 experimental data).

Instead, assume that

(148)

[again breaking strict self-similarity, but consistent with Eq. (A50)]. In this case, Eqs. (60) and (64) give

(149)
(150)

together with Cε1 and Cμ in Eqs. (142) and (143). Simultaneously solving Eqs. (149) and (150) gives

(151)
(152)

With the values in (134),

(153)

For At =0,Cε0=0.784 and σρ=2.86. Note that 1/σρa+b(1At2)(0.72At2) and Cε0/σρc[d+e(1At2)(0.72At2)] (a, b, c, d, and e are constants), so that the K and ε equation buoyancy production terms in Eqs. (43) and (44)PKb=[g0νt/(σρρ0)]ρ¯/z and Pεb=Cε0(ε/K)PKb are proportional to At[a+b(1At2)(0.72At2)] and Atc[d+e(1At2)(0.72At2)], respectively, and grow nearly linearly with Atwood number.

Finally, assume that

(154)

(again breaking strict self-similarity). In this case, Eqs. (60) and (64) together with the values for Cε1 and Cμ in Eqs. (142) and (143) give Eq. (149) and

(155)

so that

(156)

With the values in (134),

(157)

The K and ε equation buoyancy production terms also grow nearly linearly with Atwood number. For At =0,Cε0=0.800 and σρ=1.914: these values are comparable to the values 0.85 and 2.00, respectively, recommended by Llor.8 As also shown by Llor,8 in the broader context of self-similar variable acceleration RT flows (SSVARTs), in which g0(t)tn (n = 0 corresponds to a constant acceleration), model stability requires that Cε0>1: this again favors (147).

The calibration options above lead to considerably different values of Cε0 and σρ. The calibration option in (147) gives small values of the turbulent Schmidt numbers similar to those obtained in the self-similar calibration of the KLa and K–2La models, σϕα=0.06.18,19 If it is desired to apply the models solely to RT, RM, and KH mixing, then all four calibration options can be used in principle. If it is desirable for the models to give reasonable predictions for broader applications, such as free-shear flows, then the calibration option (141) should be used. If broader applicability of the model is not a strong consideration, then the option (147) is recommended. Calibration (153) is the only option that utilizes the same values of σϕα for all three instabilities. The At-dependence of several model coefficients arising in the calibration options can be regarded as small Atwood number corrections to the At=0 coefficient expressions, and are not expected to be valid for intermediate or large Atwood number flows. All of the calibrations satisfy the inequalities identified in Secs. III–V that must be satisfied by the coefficients.

For a three-equation model, the calibrations given for a two-equation model above are augmented by Eq. (66), which gives for RT mixing,

(158)

by Eq. (87) which gives for RM mixing

(159)

and by Eq. (125) which gives for KH mixing

(160)

with the expressions above for Cε0,Cε1, and Cε2 pertaining to the selected calibration choice. Each expression for Cχθm/(1θm) and depends only on the turbulent kinetic dissipation rate equation coefficients {Cεq}.

At late times, measurements from experiments and DNSs and ILESs (that are not necessarily in a self-similar state) indicate that θm0.8 for incompressible RT,23,52,63,88,89 RM,55,90,91 and incompressible At =0 KH73,92 mixing. In comparison with RT and RM mixing, there is considerably less available data on higher-order scalar turbulence statistics in incompressible, variable-density KH mixing. Most of the experimental studies have focused on the measurements of concentration profiles, chemical product formation, and probability density functions, while most of the numerical simulation studies have focused on the mechanical turbulence (e.g., mixing layer widths, components of the Reynolds stress tensor, energy spectra, scaling of structure functions) and Atwood number effects. A DNS study of passive scalar statistics in a self-similar turbulent shear layer reported93 a peak value of passive scalar fluctuation variance ϕ2¯/(ΔΦ)20.03, corresponding to θm,KH=14ϕ2¯/(ΔΦ)20.88. Later, a DNS and LES study of passive scalar mixing in shear layers reported a late-time scalar variance 0.055 (see Figs. 4 and 5 of Ref. 73), corresponding to a value θm,KH=14×0.0550.78. More research is needed to better establish the values of the molecular mixing parameter in KH mixing.

The value of Cχ is not universal as it has a different value for each instability case, as determined below. With the molecular mixing parameter value

(161)

for all three instability cases, using the coefficient values from the two-equation model calibrations (147), (153), and (157) gives for RT mixing (158),

(162)

for RM mixing (159),

(163)

and for KH mixing (160),

(164)

respectively.

The corresponding scalar variance decay exponents (B23) predicted using the Cχ values above are

(165)
(166)
(167)

respectively. As briefly discussed in  Appendix B 4, there is a range of values ns1.20–2.60, indicating that the first option is the only reasonable option, which gives a range ns = 1.20–2.20 including all three instabilities: this option also corresponds to values Cχ=0.545–1.000 for all three instabilities.

The implications of the following two approximations are of interest. First, assuming that θm=θm,RT=θm,RM=θm,KH, equating the three expressions in Eqs. (158)–(160), and taking Cε2=1.909 gives

which then implies that Cε0=Cε1=1.500, which are both larger than the calibrated coefficient values. Second, using the coefficient values pertaining to the first calibration and taking the average value of Cχ for all three instabilities Cχ=(1.000+0.545+0.778)/3=0.774, it follows that Eqs. (158)–(160) give

(168)

i.e., the resulting RT and KH values are close to the observed values while the RM value is somewhat larger than the observed values 0.8. Therefore, assuming the same value

(169)

for all three instabilities may provide a reasonable compromise for universal modeling of all three instabilities.

Finally, several general relationships between the scalar variance production-over-dissipation ratio, scalar turbulent timescale, mechanical turbulent timescale, and molecular mixing parameter can be established as follows. First, the expressions for the molecular mixing parameters (66), (87), and (125) and the expressions for the scalar variance production-to-dissipation ratios (70), (95), and (129) for RT, RM, and KH mixing, respectively, show that

(170)

for the three-equation models, and the expressions for the scalar variance production-to-dissipation ratios (71), (96), and (130) for RT, RM, and KH mixing, respectively, show that

(171)

for the four-equation models. If the algebraic model (9) is used for χ in Eq. (171), then

(172)

Equation (170) shows that larger values of θm corresponds to smaller values of PS/DS and because

(173)

it follows that PS/DS1, i.e., the scalar variance production must exceed the dissipation if there is molecular mixing. Therefore, for the three-equation models Eq. (173) gives

(174)

which reduces to τs/t=1/(4θh) with θm=0.8 for all three instabilities, or τs,RT/t=1/8,τs,RM/t=1/(4θ)=10/12, and τs,KH/t=1/4. For RM mixing, larger values of θ imply smaller values of τs,RM/t. The inequality in Eq. (173) obviously requires that θhτs/t>0 or θh>0. With (161), it follows that PS/DS=1.25 for all three instabilities.

A model with more transport equations introduces more coefficients, requiring more constraints to specify them, so a four-equation model can predict more quantities than a three-equation model. Of course, the three- and four-equation models can be cross-calibrated to give identical predictions of the same set of common physical observables for both models. A four-equation model is expected to provide better predictions of transition to turbulence of the scalar fields than a three-equation model. The scalar variance dissipation rate equation coefficients must be recalibrated for other values of Sc.

For the four-equation model R=2AKAχ/(AεAS) and θm=14AS, so that both quantities are functions of the mechanical and scalar model coefficients. The mechanical-to-scalar timescale ratio can be written as R=8AKAχ/[Aε(1θm)], which depends on all of the scalar model coefficients. It follows that the turbulent scalar field prefactors can be expressed directly in terms of θm and R,

(175)

so that

(176)

However, these expressions do not provide the dependence of AS and Aχ on the model coefficients.

The four-equation model subtracts the coefficient Cχ in the three-equation model and adds the coefficients Cχ0,Cχ2, and Cχ3, requiring three constraints for their calibration. The molecular mixing parameters (67), (88), and (126), scalar variance production-to-dissipation ratios (71), (96), and (130), scalar variance dissipation rate production-to-destruction ratios (72), (97), and (131), and mechanical-to-scalar timescale ratios (73), (98), and (132) can provide constraints for the three instability cases. However, these quantities have a highly nonlinear algebraic dependence on the model coefficients and Schmidt number and cannot be solved analytically for the coefficients {Cχq}. Moreover, measurements from either experiments or simulations of the scalar observables discussed here (except for θm for RT and RM mixing) are largely unavailable, particularly in a self-similar state—future research should address this gap.

Consider the two-equation calibrations (147), (153), and (157), and let Sc =0.7 (mixing of gases). In general, the calibration of Cχ0,Cχ2, and Cχ3 requires solving the constraints given by the expressions for the self-similar scalar observables for the coefficients {Cχq}. Given the values of these observables, together with the values of the previously calibrated mechanical turbulence coefficients, the expressions for these observables can then be solved simultaneously (numerically) for the values of the scalar coefficients. The exponents for decaying scalar turbulence derived in Appendix B 4 are not required for this calibration.

Recognizing the limitations noted above and that there is currently limited data available for these observables corresponding to RT and KH mixing, the numerical calibration procedure will be simplified as follows. The values of some of the scalar coefficients will be assumed and either one or two observable constraints will be used to determine the remaining coefficient(s). For all three instabilities, θm will constitute one of the physical observables, and for RT mixing one additional observable will be used (PS/DS). Even in this simplified case, the solutions to the equations are not unique (i.e., there are generally multiple roots): solutions for which any coefficient value is negative are inadmissible. In the future, when more scalar mixing data becomes available, it will not be necessary to assume values of any of the coefficients.

In order for a candidate model to also predict isotropic decay of scalar variance and its dissipation rate in homogeneous turbulence, the expression for the decay exponent ns(Cε2,Cχ2,Cχ3) in Appendix B 4 requires that R0/(Cε21)>0 and R0/(Cε21)+1>0, respectively [see Eq. (B13)]. However, it is difficult to obtain numerical solutions to the all of the constraint equations for RT, RM, and KH mixing that satisfy these inequalities and give positive values of {Cχq} that are not unreasonably large.

Comparing the expressions (72), (97), and (131) for the scalar variance production-to-destruction ratios shows that in general

(177)

which shows that this quantity is proportional to 1/R.

1. Rayleigh–Taylor mixing

To simplify the numerical calibration procedure assume a value of Cχ3=0.1 or 1.5. In this case, the coefficients Cχ0 and Cχ2 are determined by two quantities chosen from {θm,R,PS/DS,Pχ/Dχ}. The observable values that can be used are

(178)

which are the latest-time values obtained from DNS data31,34 closely corresponding to a water channel RT experiment.35 The value PS/DS above is also similar to those obtained in Ref. 63. This experimental data were consistent with a slightly larger late-time value of α0.14: there is presently no other data available for these values.

Taking θm and PS/DS [(161) and (178)] as the scalar observables for RT mixing gives for the calibrations (147), (153), and (157),

(179)

respectively. Using other values of Cχ3 or other pairs of observables, such as {θm,R} or {θm,Pχ/Dχ}, yields other coefficient values.

2. Richtmyer–Meshkov mixing

Taking θm(161) as the scalar observable for RM mixing and taking Cχ3=0 or 1.5 give for the calibrations (147), (153), and (157),

(180)

respectively.

3. Kelvin–Helmholtz mixing

Taking θm(161) as the scalar observable for KH mixing and taking Cχ3=0 or 1.5 give for the calibrations (147), (153), and (157),

(181)

respectively.

The predicted values of twenty-two self-similar parameters derived in Secs. III–V can be compared, providing additional insight into the relative merits of the two-, three-, and four-equation models using the calibration values in Table I. For each instability L(t)/h(t)[Cεπ/(4ξ)]σs/Cμ. Tables II–IV give the predicted values of self-similar parameters corresponding to each calibrated model for RT, RM, and KH mixing together with the equations corresponding to them.

TABLE I.

Summary of values of constant self-similar quantities used to calibrate the coefficients in the two-, three-, and four-equation models and their corresponding constraining expressions. The first five values are used to calibrate the two-equation mechanical turbulence model, the sixth, seventh, and eighth values are additionally used to calibrate the three- and four-equation models, and the ninth value is used in the calibration of the four-equation model for RT mixing.

αρ0KΔΦθδK(0,t)(Δv)2θm,RTθm,RMθm,KH(PSDS)RT
Value 0.12 0.50 0.30 0.08 0.035 0.80 0.80 0.80 1.24 
Equation (60) (64) (85) (119) (122) (66) or (67) (87) or (88) (125) or (126) (71) 
αρ0KΔΦθδK(0,t)(Δv)2θm,RTθm,RMθm,KH(PSDS)RT
Value 0.12 0.50 0.30 0.08 0.035 0.80 0.80 0.80 1.24 
Equation (60) (64) (85) (119) (122) (66) or (67) (87) or (88) (125) or (126) (71) 
TABLE II.

Summary of values of constant self-similar quantities predicted by the calibrated two-, three-, and four-equation models for RT mixing. The values of α, ρ0K/ΔΦ, and θm are the same for each model and are not shown. For the four-equation model, values are shown for each of the three calibrations. The total mixing layer width is h (i.e., ξ = 2) and Cϵ=0.5 (see Sec. VII A).

(PKDK)RT(PϵDϵ)RT(PSDS)RT(PχDχ)RTRRTτm,RTtτs,RTt(Lh)RT
Two-equation 2.00 1.39 ⋯ ⋯ ⋯ 0.25 ⋯ 0.07 
Equation (68) (69) ⋯ ⋯ ⋯ (54) ⋯ (61) 
Three-equation 2.00 1.39 1.24 ⋯ 2.00 0.25 0.13 0.07 
Equation (68) (69) (70) ⋯ (15) (54) (54) (61) 
Four-equation (I) 2.00 1.39 1.24 1.37 2.09 0.25 0.12 0.07 
Four-equation (II) 34.09 14.00 1.24 1.29 70.1 8.27 0.12 2.33 
Four-equation (III) 23.18 9.71 1.24 1.24 48.8 5.55 0.12 1.57 
Equation (68) (69) (71) (72) (73) (54) (57) (61) 
(PKDK)RT(PϵDϵ)RT(PSDS)RT(PχDχ)RTRRTτm,RTtτs,RTt(Lh)RT
Two-equation 2.00 1.39 ⋯ ⋯ ⋯ 0.25 ⋯ 0.07 
Equation (68) (69) ⋯ ⋯ ⋯ (54) ⋯ (61) 
Three-equation 2.00 1.39 1.24 ⋯ 2.00 0.25 0.13 0.07 
Equation (68) (69) (70) ⋯ (15) (54) (54) (61) 
Four-equation (I) 2.00 1.39 1.24 1.37 2.09 0.25 0.12 0.07 
Four-equation (II) 34.09 14.00 1.24 1.29 70.1 8.27 0.12 2.33 
Four-equation (III) 23.18 9.71 1.24 1.24 48.8 5.55 0.12 1.57 
Equation (68) (69) (71) (72) (73) (54) (57) (61) 
TABLE III.

Summary of values of constant self-similar quantities predicted by the calibrated two-, three-, and four-equation models for RM mixing. The values of θ and θm are the same for each model and are not shown. For the four-equation model, values are shown for each of the three calibrations. The total mixing layer width is h (i.e., ξ = 2) and Cϵ=1.54 (see Sec. VII B).

(PSDS)RM(PχDχ)RMRRMτm,RMtτs,RMt(Lh)RM
Two-equation ⋯ ⋯ ⋯ 0.91 ⋯ 0.16 
Equation ⋯ ⋯ ⋯ (82) ⋯ (86) 
Three-equation 1.25 ⋯ 1.09 0.91 0.83 0.16 
Equation (95) ⋯ (98) (82) (82) (86) 
Four-equation (I) 1.25 0.71 1.09 0.91 0.83 0.16 
Four-equation (II) 1.24 0.09 1.12 0.91 0.81 0.92 
Four-equation (III) 1.25 0.42 1.09 0.91 0.81 0.76 
Equation (96) (97) (98) (82) (57) (86) 
(PSDS)RM(PχDχ)RMRRMτm,RMtτs,RMt(Lh)RM
Two-equation ⋯ ⋯ ⋯ 0.91 ⋯ 0.16 
Equation ⋯ ⋯ ⋯ (82) ⋯ (86) 
Three-equation 1.25 ⋯ 1.09 0.91 0.83 0.16 
Equation (95) ⋯ (98) (82) (82) (86) 
Four-equation (I) 1.25 0.71 1.09 0.91 0.83 0.16 
Four-equation (II) 1.24 0.09 1.12 0.91 0.81 0.92 
Four-equation (III) 1.25 0.42 1.09 0.91 0.81 0.76 
Equation (96) (97) (98) (82) (57) (86) 
TABLE IV.

Summary of values of constant self-similar quantities predicted by the calibrated two-, three-, and four-equation models for KH mixing. The values of δ, K(0,t)/(Δv)2, and θm are the same for each model and are not shown. For the four-equation model, values are shown for each of the three calibrations. The total mixing layer width is h (i.e., ξ = 2) and Cϵ=1 (see Sec. VII C).

K0|Δv|3tϵ0|Δv|3(PKDK)KH(PSDS)KHRKHτm,KHtτs,KHt(Lh)KH
Two-equation 0.002 0.005 1.39 – – 0.39 – 0.71 
Equation (123) (124) (127) (129) – (114) – (121) 
Three-equation 0.002 0.005 1.39 1.25 1.56 0.39 0.25 0.71 
Equation (123) (124) (127) (129) (15) (114) (114) (121) 
Four-equation (I) 0.002 0.005 1.39 1.49 1.55 0.39 0.25 0.71 
Four-equation (II) 0.002 0.005 1.39 1.49 1.56 0.39 0.25 0.71 
Four-equation (III) 0.002 0.005 1.39 1.49 1.55 0.39 0.25 0.71 
Equation (123) (124) (127) (130) (132) (114) (57) (121) 
K0|Δv|3tϵ0|Δv|3(PKDK)KH(PSDS)KHRKHτm,KHtτs,KHt(Lh)KH
Two-equation 0.002 0.005 1.39 – – 0.39 – 0.71 
Equation (123) (124) (127) (129) – (114) – (121) 
Three-equation 0.002 0.005 1.39 1.25 1.56 0.39 0.25 0.71 
Equation (123) (124) (127) (129) (15) (114) (114) (121) 
Four-equation (I) 0.002 0.005 1.39 1.49 1.55 0.39 0.25 0.71 
Four-equation (II) 0.002 0.005 1.39 1.49 1.56 0.39 0.25 0.71 
Four-equation (III) 0.002 0.005 1.39 1.49 1.55 0.39 0.25 0.71 
Equation (123) (124) (127) (130) (132) (114) (57) (121) 

1. Rayleigh–Taylor mixing

All of the models predict the calibration values of α, ρ0K/ΔΦ, θm, and PS/DS. The four-equation model calibrations give significantly different values of PK/DK,Pε/Dε,Pχ/Dχ, R, τm/t, and L/h, but the same value of τs/t: the variability results from the significantly different values of Cε0 and σρ. The first four-equation model calibration predicts a larger value of Pχ/Dχ than the second and third calibrations [which are larger than that in (178)]. The second and third four-equation calibrations gives unphysically large values of PK/DK,Pε/Dε, R, and τm/t. Future measurements of these quantities can guide which of the three calibrations is best for RT mixing and indicate whether these models can simultaneously predict such a large number of self-similar observables.

DNS data corresponding to RT mixing with At = 0.01 and Sc = 1 exploring self-similar scalings of turbulence statistics and fields63 and with At = 7.5×104 and Sc = 7 modeling a water channel experiment and exploring turbulent transport mechanisms34 confirmed that the spatial profiles of the molecular mixing parameter, production-to-dissipation/destruction ratios, mechanical and scalar turbulent timescales, and mechanical-to-scalar timescale ratio are approximately constant across the mixing layer (varying most near the edges) at late evolution times approaching a self-similar state.34,63,94 The profiles of PK/DK(z,t) and PS/DS(z,t) indicated average values 3.5 (Fig. 33 of Ref. 63) and 1.25 (Fig. 32 of Ref. 63), respectively. Data from a DNS of At =0.5 RT mixing indicated PK/DK(z,t)PK/DK(t)2.0 at the latest time (Fig. 4 of Ref. 94). The latest time values reported in Ref. 34 were PK/DK(t)2.00,Pε/Dε(t)1.10,PS/DS(t)1.24,Pχ/Dχ(t)1.01. This study also indicated that at the latest time ρ0K/ΔΦ0.57 and θm0.6 (and had not attained asymptotic values), somewhat larger and smaller than the calibration values used here. Note that the value PS/DS=1.24 is nearly identical to the self-similar value 1.25 for θm=0.8. The mechanical-to-scalar timescale ratio (15) had late-time values R1.5–1.8 (Fig. 13 of Ref. 63) and R0.9–1.3 (Fig. 20 of Ref. 31) in reasonable agreement with the prediction of the first four-equation model. Ristorcelli and Clark computed the late-time value (L/h)RC=[4/(πCε)](L/h)RT0.35 (Fig. 11 of Ref. 63), indicating a value (L/h)RC0.18 predicted by the first four-equation model. These (and other self-similar) values surely depend on the initial conditions and other simulation details (including Schmidt number effects), but clearly favor the first four-equation model calibration.

2. Richtmyer–Meshkov mixing

All of the models predict the calibration values of θ, θm, and PS/DS (nearly the same as the RT calibration value) and nearly the same value of R, τm/t, and τs/t. As in the RT mixing case, the four-equation model calibrations predict significantly different values of Pχ/Dχ, all of which are less than one, indicating that destruction exceeds production. The two- and three-equation and the first calibration of the four-equation model predict the same value of L/h, while the other four-equation calibrations predict a value that is approximately five to six times larger. Additional data would again be useful for determining which model and calibration is best for RM mixing.

To date, there has been considerably less analysis of self-similarity in RM mixing than for RT and KH mixing. In particular, there have been no experimental or simulation studies of impulsively or shock-accelerated RM mixing that have considered the quantities discussed here (for either nonself-similar or self-similar conditions). Specifically, measurements of the quantities in Table III are not available. DNS and ILES data for RM mixing with finite initial Reynolds numbers80 indicate decay exponents 3θ2=1.25 and −1.41, respectively, for the turbulent kinetic energy integrated over the mixing layers (89), which are 14% and 28% smaller than the model value –1.1.

3. Kelvin–Helmholtz mixing

All of the models predict the calibration values of δ, K(0,t)/(Δv)2, and θm. All of the models predict the same values of PK/DK,PS/DS,τm/t,τs/t, and L/h. All of the models predict K0/(|Δv|3t)=0.002 and ε0/|Δv|3=0.005. The three- and four-equation models predict the same value of R. Additional scalar data are required to guide model selection for KH mixing.

Thus far, there has been more consideration of self-similar evolution of KH (shear) mixing layers than for RT flows, with both At =0 and At 0. DNS data closely corresponding to the Bell–Mehta shear layer experiment62,71,73,75 at late time predicted K(0,t)/(Δv)20.035 and τyz/ρ00.01 in excellent agreement with the current model predictions. DNS data also indicated that at late times ε0/|Δv|30.006,62,72,95 also in very good agreement with the model predictions. The evolution of the normalized integrated turbulent kinetic energy (123) was considered using DNS: estimating the slope of the pertinent curve in Fig. 3 of Ref. 72 gives (1.81.2)/(460200)0.002, in agreement with the model predictions. Estimating the late-time ratio of the pertinent production and dissipation profiles in Fig. 21(e) of Ref. 71 and Fig. 4 of Ref. 72 gives PK/DK0.0065/0.0041.39, which is in excellent agreement with the model predictions.

The previously calibrated and recommended four-equation model will be applied here to miscible Rayleigh–Taylor, Richtmyer–Meshkov, and Kelvin–Helmholtz instability-induced mixing cases with Sc =0.7 in order to illustrate the solutions of the model. The space–time evolution of fields is presented in unscaled coordinates to facilitate comparison of their broadening as the layer width grows in time quadratically for RT mixing, as a power-law for RM mixing, and linearly for KH mixing as well as their relative rates of growth or decay in time. In addition to the four primary turbulent fields K, ε, S, and χ, the evolution of az, νt, and L are also shown: az enters the buoyancy production term and is a principal field in KLa-type models, νt drives the production and diffusion terms through gradient-diffusion closures, and L is also a principal field in turbulent lengthscale-based models. The requirements and challenges involved in comparing self-similar profiles to available RT, RM, and KH experimental and simulation data are discussed in Sec. VIII F.

The scalar variance will correspond to the heavy fluid mass fraction variance, and the model coefficients used are given by Eqs. (147), (179), (180), and (181) (taking At =0),

(182)

corresponding to “Four-equation (I)” in Tables II–IV. It is apparent from the expressions for the turbulent fields in Secs. III–V that they are not valid at t = 0 (the fields are either zero or singular). These expressions will be extrapolated to t = 0 here by introducing a time offset t0 (an initial characteristic timescale) related to the initial mixing layer width, h(0)=h0=(ξ/2)Aht0θh or

(183)

and letting tt+t0 in the expressions for the mixing layer widths and in the power-laws tpϕα of the turbulent fields. This allows the fields to be defined at t = 0 and approach their asymptotic self-similar forms for tt0 (with larger values of h0 giving larger values of t0). The budgets of the four turbulent transport equations (38), with the terms on the right sides reconstructed using the analytic expressions for the fields corresponding to each instability case given in Secs. III–V, are also shown at the latest evolution time to illustrate the relative magnitudes of the production, dissipation/destruction, and turbulent diffusion terms.

As will be seen later, the magnitudes of the turbulent fields are large within the mixing layer core

It is assumed that ξ = 2, i.e., h(t) is the total mixing layer width. Common to all three instabilities, the turbulent lengthscale grows proportionally with h(t), with a profile 14z2/h(t)2. For each budget, the production-to-dissipation/destruction ratios Pϕα/Dϕα (evaluated, for example, at the centerline z = 0) are consistent with the values in Tables II–IV corresponding to “Four-equation (I).” The production and dissipation/destruction terms also have parabolic profiles. The turbulent diffusion terms Tϕα=(/z)[(νt/σϕα)ϕα/z] are positive near the layer boundaries and negative in the core, indicating outward transport from the core toward the layer edges (and integrating to zero). The shear production terms in the mechanical turbulence equations PKs=(τzz/ρ0)w̃/z and Pεs=Cε1(ε/K)PKs are close to zero at very small Atwood number. The advection terms in all of the equations Aϕα=w̃ϕα/z are also nearly zero at very small Atwood number. The magnitudes of the shear production and advection terms, which are positive and negative over the layer, increase with increasing Atwood number.

To illustrate the spatiotemporal evolution of self-similar Rayleigh–Taylor turbulence and mixing, the following parameters are used:

(184)

with a final time tf = 5 s. These parameters are similar to those used in the simulations by Morgan et al.23 investigating the approach to self-similarity of RT mixing as a function of mode coupling generations. In the numerical evaluation of the model, it is assumed that

(185)

with h(0)=h0=0.01 cm and t0=h0/(αAtg0)=0.04 s. The resulting mixing layer width has a constant, linear-in-time, and quadratic-in-time contribution. The very small value of t0 allows a very rapid transition to a t2 self-similar growth. The growth of the mixing layer width with α=0.12 is shown in Fig. 1.

FIG. 1.

Time-evolution of the mixing layer width (185) for RT mixing.

FIG. 1.

Time-evolution of the mixing layer width (185) for RT mixing.

Close modal

1. Evolution of fields

The space–time evolution of the self-similar profiles of ρ¯, K, ε, S, χ, az, νt, and L (with Cε=0.5 the approximate average value obtained from RT simulation data96) given by Eqs. (35) and (47)–(53) is shown in Figs. 2–9. As time evolves, the linear profile of the mean density across the mixing layer widens with a decreasing slope (asymptotically approaching the fully homogeneously mixed density ρ0), as ρ¯/z1/h(t) decreases with time. The mean heavy fluid mass fraction evolves the same way, but is bounded by [0,1] instead of by [ρ2,ρ1] as for the mean density. The peak values (at z = 0) of the turbulent field profiles evolve in time consistently with the power-law exponents 2, 1, 0, and –1, respectively. The early-time profiles are narrow, widening in space rapidly as the mixing layer width increases as t2. The peak value of S(z, t) remains the same, while that of χ(z,t) decreases with time rapidly at early times (asymptotically decreasing to zero when complete mixing has occurred). The mass flux velocity is negative so that the turbulent kinetic energy buoyancy production term is g0az>0. The profiles of ε(z,t) and az(z,t) both evolve linearly in time. The turbulent viscosity profile increases rapidly as t3.

FIG. 2.

Spatiotemporal evolution of the self-similar mean density ρ¯(z,t)=ρ0[1+2Atz/h(t)] for RT mixing.

FIG. 2.

Spatiotemporal evolution of the self-similar mean density ρ¯(z,t)=ρ0[1+2Atz/h(t)] for RT mixing.

Close modal
FIG. 3.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (47) for RT mixing.

FIG. 3.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (47) for RT mixing.

Close modal
FIG. 4.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (48) for RT mixing.

FIG. 4.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (48) for RT mixing.

Close modal
FIG. 5.

Spatiotemporal evolution of the self-similar mass fraction variance (49) for RT mixing.

FIG. 5.

Spatiotemporal evolution of the self-similar mass fraction variance (49) for RT mixing.

Close modal
FIG. 6.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (50) for RT mixing.

FIG. 6.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (50) for RT mixing.

Close modal
FIG. 7.

Spatiotemporal evolution of the self-similar mass flux velocity (51) for RT mixing.

FIG. 7.

Spatiotemporal evolution of the self-similar mass flux velocity (51) for RT mixing.

Close modal
FIG. 8.

Spatiotemporal evolution of the self-similar turbulent viscosity (52) for RT mixing.

FIG. 8.

Spatiotemporal evolution of the self-similar turbulent viscosity (52) for RT mixing.

Close modal
FIG. 9.

Spatiotemporal evolution of the self-similar turbulent lengthscale (53) for RT mixing.

FIG. 9.

Spatiotemporal evolution of the self-similar turbulent lengthscale (53) for RT mixing.

Close modal

The linear and inverted parabolic self-similar profiles used for the mean and turbulent fields in the KL,14 BHR,17KLa,18 and K–2La19 model studies are the same as assumed here. The model is calibrated using the same observable values as in Ref. 18 and therefore predicts the same values a priori. Simulation data consistent with this calibration indicate that the heavy fluid mass fraction is nearly linear and the turbulent kinetic energy normalized by its maximum value is well-approximated by an inverted parabolic profile across the layer (see Fig. 1 of Ref. 18, Fig. 15 of Ref. 23, and Fig. 2 of Ref. 19). Figure 26 of Ref. 63 shows that the scalar variance normalized by its maximum value also has an approximately inverse parabolic profile, and Fig. 36 shows that the turbulent lengthscale [see Eq. (12) but with Cε=1] normalized by its maximum value has a profile that is broader than an inverted parabola, more consistent with a 14z2/h(t)2 profile. Figure 5 also shows that the molecular mixing parameter is consistent with θm=14S(0,t)=0.8. As also discussed in Ref. 17, the profiles of K and az obtained from gas channel experimental data52 are approximately inverse parabolic.

The tpϕα time-dependence of the turbulent fields complicates the comparison of the profiles to data, which is the reason why self-normalized (to their maximum values) fields have been compared previously, i.e., comparing ϕα(z,t)/(Aϕαtpϕα) obtained from a model and data [see Eq. (23)]. Also, the coordinate is scaled by h(t) so that the width of the self-similar profile automatically matches data consistent with the mixing layer growth. Such comparisons remove both the dependence on the model coefficients (Aϕα) and on the time (tpϕα), and thus become comparisons of the profile shapes only.

2. Evolution of transport equation budgets

The budget of each turbulent transport equation (43)–(46) is shown at the final time tf = 5 s in Figs. 10–13. The buoyancy production terms PKb=g0az and Pεb=Cε0(ε/K)PKb are positive, significantly exceeding the dissipation DK=ε and destruction Dε=Cε2ε2/K terms in magnitude, consistent with the rapid growth of K and ε. The turbulent diffusion terms in these equations are smaller in magnitude compared to the production and dissipation/destruction terms. The mass fraction variance production term PS=(2νt/σm)(m̃1/z)2 is balanced by the sum of the mass fraction variance dissipation DS=2χ and the turbulent diffusion term such that S(0, t) remains constant. The mass fraction variance dissipation rate production Pχ=(Cχ0νt/Sc)(ε/K)(m̃1/z)2+Cχ3εχ/K is nearly balanced by the sum of the destruction term Dχ=Cχ2χ2/S and the turbulent diffusion term, consistent with the slow decay of χ at late times.

FIG. 10.

Budget of the self-similar turbulent kinetic energy equation (43) at the final time for RT mixing.

FIG. 10.

Budget of the self-similar turbulent kinetic energy equation (43) at the final time for RT mixing.

Close modal
FIG. 11.

Budget of the self-similar turbulent kinetic energy dissipation rate equation (44) at the final time for RT mixing.

FIG. 11.

Budget of the self-similar turbulent kinetic energy dissipation rate equation (44) at the final time for RT mixing.

Close modal
FIG. 12.

Budget of the self-similar mass fraction variance equation (45) at the final time for RT mixing.

FIG. 12.

Budget of the self-similar mass fraction variance equation (45) at the final time for RT mixing.

Close modal
FIG. 13.

Budget of the self-similar mass fraction variance dissipation rate equation (46) at the final time for RT mixing.

FIG. 13.

Budget of the self-similar mass fraction variance dissipation rate equation (46) at the final time for RT mixing.

Close modal

Despite extensive research on RT mixing, there is exceptionally little data available pertaining to turbulent transport equation budgets.34,94 The relative importance of the terms in the transport equation budget for K in Ref. 94 for At =0.5 and for K, ε, S, and χ in Ref. 34 for RT mixing in the Boussinesq approximation is in generally good agreement with the model predictions shown here.

To illustrate the spatiotemporal evolution of self-similar incompressible impulsively driven Richtmyer–Meshkov turbulence and mixing, the following parameters are used:

(186)

with a final time tf = 5 s. The impulsive acceleration

(187)

(with t=0.1 s) in the K and ε equations decreases to zero very rapidly after t = 0 and is similar to the acceleration used in a buoyancy–shear–drag turbulence model.83 The value of Δvs is that used in the multimode simulation of a Ma =1.8439 RM instability,55 but with Atwood number 0.05 instead of 0.5. Soulard et al.56 investigated the dependence of various RM mixing parameters on the initial large-scale energy spectrum in a self-similar state for At =0.05. In the numerical evaluation of the model, it is assumed that

(188)

with h(0)=h0=0.01 cm and t0=θh0/(ΔvsAt+)=2.06×106 s. The very small value of t0 allows a very rapid transition to a t0.3 self-similar growth. The slow power-law growth in time with exponent θ=0.3 of the mixing layer width is shown in Fig. 14 (see Fig. 4 in the compressible study of Ref. 55 for comparison).

FIG. 14.

Time-evolution of the mixing layer width (188) for RM mixing.

FIG. 14.

Time-evolution of the mixing layer width (188) for RM mixing.

Close modal

1. Evolution of fields

The space–time evolution of the self-similar profiles of ρ¯, K, ε, S, χ, az, νt, and L (with Cε=1.54 the approximate value obtained from RM DNS data82) given by Eqs. (35) (but with At+) and Eqs. (75)–(81) is shown in Figs. 15–22. As time evolves, the linear profile of the mean density across the mixing layer widens with a reduced slope (asymptotically approaching the fully homogeneously mixed density ρ0), as ρ¯/z1/h(t) decreases with time as in RT mixing, but at a significantly reduced rate. The heavy fluid mass fraction evolves the same way, but is bounded by [0,1] instead of by [ρ2,ρ1] as for the density. The peak values (at z = 0) of the turbulent field profiles evolve in time consistently with the power-law exponents –1.4, –2.4, 0, and –1, respectively. The early-time profiles are narrow, widening in space slowly as the width increases as t0.3. The peak value of S(z, t) remains the same, while that of χ(z,t) decreases with time (asymptotically decreasing to zero when complete mixing has occurred) as in RT mixing. Figure 18 also shows that the molecular mixing parameter is consistent with θm=14S(0,t)=0.8. Except for the mass fraction variance and turbulent lengthscale, the other turbulent fields decay rapidly at early times and more slowly at late times. The mass flux velocity is very small in magnitude and also negative as in RT mixing. The profile of az(z,t) evolves as t0.7 and the profile of νt(z,t) evolves as t0.4. Both K(z, t) and ε(z,t) decrease after the initial impulse at t = 0 after which the fields decay due to dissipation and destruction.

FIG. 15.

Spatiotemporal evolution of the self-similar mean density ρ¯(z,t)=ρ0[1+2At+z/h(t)] for RM mixing.

FIG. 15.

Spatiotemporal evolution of the self-similar mean density ρ¯(z,t)=ρ0[1+2At+z/h(t)] for RM mixing.

Close modal
FIG. 16.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (75) for RM mixing.

FIG. 16.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (75) for RM mixing.

Close modal
FIG. 17.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (76) for RM mixing.

FIG. 17.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (76) for RM mixing.

Close modal
FIG. 18.

Spatiotemporal evolution of the self-similar mass fraction variance (77) for RM mixing.

FIG. 18.

Spatiotemporal evolution of the self-similar mass fraction variance (77) for RM mixing.

Close modal
FIG. 19.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (78) for RM mixing.

FIG. 19.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (78) for RM mixing.

Close modal
FIG. 20.

Spatiotemporal evolution of the self-similar mass flux velocity (79) for RM mixing.

FIG. 20.

Spatiotemporal evolution of the self-similar mass flux velocity (79) for RM mixing.

Close modal
FIG. 21.

Spatiotemporal evolution of the self-similar turbulent viscosity (80) for RM mixing.

FIG. 21.

Spatiotemporal evolution of the self-similar turbulent viscosity (80) for RM mixing.

Close modal
FIG. 22.

Spatiotemporal evolution of the self-similar turbulent lengthscale (81) for RM mixing.

FIG. 22.

Spatiotemporal evolution of the self-similar turbulent lengthscale (81) for RM mixing.

Close modal

Groom and Thornber82 also evaluated the normalized scalar dissipation rate (proportional to the mechanical-to-scalar timescale ratio),

(189)

from their DNS data (which is limited to small Re and is not self-similar), or R0.70, which is smaller than the value 1.09 in Table III. Figures 11 and 12 in Ref. 97 show that late-time profiles of the heavy fluid mass fraction are approximately linear, except near the mixing layer edges.

2. Evolution of transport equation budgets

The budget of each turbulent transport equation (43)–(46) is shown at the final time tf = 5 s in Figs. 23–26. The buoyancy production terms PKb and Pεb are nearly zero except at t = 0. The dissipation DK and destruction Dε terms are large, consistent with the decay of K and ε across the inhomogeneous mixing layer. The turbulent diffusion terms in these equations are smaller in magnitude than the dissipation and destruction terms. The mass fraction variance production term PS is balanced by the sum of the dissipation term DS and the diffusion term such that S(0, t) remains constant. The mass fraction variance dissipation rate destruction term Dχ exceeds the production term Pχ such that χ decreases with time across the layer as in RT mixing.

FIG. 23.

Budget of the self-similar turbulent kinetic energy equation (43) with impulsive acceleration (187) at the final time for RM mixing.

FIG. 23.

Budget of the self-similar turbulent kinetic energy equation (43) with impulsive acceleration (187) at the final time for RM mixing.

Close modal
FIG. 24.

Budget of the self-similar turbulent kinetic energy dissipation rate equation (44) with impulsive acceleration (187) at the final time for RM mixing.

FIG. 24.

Budget of the self-similar turbulent kinetic energy dissipation rate equation (44) with impulsive acceleration (187) at the final time for RM mixing.

Close modal
FIG. 25.

Budget of the self-similar mass fraction variance equation (45) at the final time for RM mixing.

FIG. 25.

Budget of the self-similar mass fraction variance equation (45) at the final time for RM mixing.

Close modal
FIG. 26.

Budget of the self-similar mass fraction variance dissipation rate equation (46) at the final time for RM mixing.

FIG. 26.

Budget of the self-similar mass fraction variance dissipation rate equation (46) at the final time for RM mixing.

Close modal

Despite extensive experimental and computational research on RM mixing, there is exceptionally little data available pertaining to turbulent transport equation budgets.97 The relative importance of the terms in the K transport equation budget shown in Fig. 23 in Ref. 97 for RM mixing with At =0.5 (and computed using ILES) is in generally good qualitative agreement with the model prediction shown here: at the latest time, the evolution of K is dominated by turbulent transport and the pressure–dilatation correlation (which is not modeled here), and by numerical dissipation. The S transport equation budget at the latest time in Fig. 17 in Ref. 97 shows that production dominates and the numerical dissipation is of similar magnitude, which qualitatively agrees with the model prediction shown here.

To illustrate the spatiotemporal evolution of self-similar Kelvin–Helmholtz turbulence and mixing, the following parameters are used:

(190)

which are the stream velocities in the incompressible Bell–Mehta experiment,70 and with a final time tf = 4.5 s. Thus, Δv=600 cm/s, At v=0.25, and At =0. In the numerical evaluation of the model, it is assumed that

(191)

with h(0)=h0=0.01 cm and t0=h0/(δ|Δv|)=2×104 s. The resulting mixing layer width has a constant and linear-in-time contribution. The very small value of t0 allows a very rapid transition to a t self-similar growth. The growth of the mixing layer width with δ=0.08 is shown in Fig. 27.

FIG. 27.

Time-evolution of the mixing layer width (191) for KH mixing.

FIG. 27.

Time-evolution of the mixing layer width (191) for KH mixing.

Close modal

1. Evolution of fields

The space–time evolution of the self-similar profiles of v¯, K, ε, S, χ, νt, and L (with Cε=1 assumed in the absence of data) given by Eqs. (105) and (108)–(113) is shown in Figs. 28–34. As time evolves, the linear profile of the mean shear velocity across the mixing layer widens with a reduced slope (asymptotically approaching the average stream velocity v0), as v¯/z1/h(t) decreases with time, analogously with the mean density in RT and RM mixing. The peak values (at z = 0) of the turbulent fields evolve in time consistently with the power-law exponents 0, –1, 0, and –1. The early-time profiles are narrow, widening in space as the width increases as t. The value of S(0, t) remains the same, while χ(z,t) decreases with time (asymptotically decreasing to zero). Figure 31 also shows that the molecular mixing parameter is consistent with θm=14S(0,t)=0.8. The profiles of ε(z,t) and χ(z,t) both decrease at the same rate in time. The turbulent viscosity profile increases as t.

FIG. 28.

Spatiotemporal evolution of the self-similar mean shear velocity v¯(z,t)=v0[1+2Atvz/h(t)] for KH mixing.

FIG. 28.

Spatiotemporal evolution of the self-similar mean shear velocity v¯(z,t)=v0[1+2Atvz/h(t)] for KH mixing.

Close modal
FIG. 29.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (108) for KH mixing.

FIG. 29.

Spatiotemporal evolution of the self-similar turbulent kinetic energy (108) for KH mixing.

Close modal
FIG. 30.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (109) for KH mixing.

FIG. 30.

Spatiotemporal evolution of the self-similar turbulent kinetic energy dissipation rate (109) for KH mixing.

Close modal
FIG. 31.

Spatiotemporal evolution of the self-similar mass fraction variance (110) for KH mixing.

FIG. 31.

Spatiotemporal evolution of the self-similar mass fraction variance (110) for KH mixing.

Close modal
FIG. 32.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (111) for KH mixing.

FIG. 32.

Spatiotemporal evolution of the self-similar mass fraction variance dissipation rate (111) for KH mixing.

Close modal
FIG. 33.

Spatiotemporal evolution of the self-similar turbulent viscosity (112) for KH mixing.

FIG. 33.

Spatiotemporal evolution of the self-similar turbulent viscosity (112) for KH mixing.

Close modal
FIG. 34.

Spatiotemporal evolution of the self-similar turbulent lengthscale (113) for KH mixing.

FIG. 34.

Spatiotemporal evolution of the self-similar turbulent lengthscale (113) for KH mixing.

Close modal

The linear and inverted parabolic self-similar profiles used for the mean and turbulent fields in the K–2La model19 studies are the same as assumed here. The model is calibrated using the same observable values as in Ref. 19 and, therefore, predicts the same values a priori. Simulation data consistent with this calibration indicates that the mean shear velocity is nearly linear and the turbulent kinetic energy is well-approximated by an inverted parabolic profile across the layer (see Figs. 5 and 6 of Ref. 19). Experimental data (see Fig. 2 of Ref. 70) and numerical simulation data (see Fig. 13 of Ref. 71, Fig. 5 of Ref. 62, Fig. 7 of Ref. 73, and Fig. 2 of Ref. 95) are consistent with a nearly linear mean shear velocity profile within the mixing layer core. It follows from Fig. 29 that K(0,t)/(Δv)2=0.035, consistent with the model calibration. The shear Reynolds stress (115) has a negative parabolic profile, and the centerline normalized shear stress (116) is τyz(0,t)/[ρ0(Δv)2]=0.01, which is in very good agreement with the peak value in Fig. 4(d) of Ref. 70 (but with the opposite sign convention) and in Fig. 4(d) of Ref. 71.

2. Evolution of transport equation budgets

The budget of each turbulent transport equation (102)–(103) and (45)–(46) is shown at the final time tf = 4.5 s in Figs. 35–38. The turbulent kinetic energy shear production term PKs=Cdevνt(v¯/z)2 is larger than the dissipation term DK. The turbulent kinetic energy shear production term PKs is balanced by the sum of the dissipation DK and the smaller in magnitude turbulent diffusion term. The turbulent kinetic energy dissipation rate shear production term Pεs=Cε1(ε/K)PKs is balanced by the destruction term Dε, consistent with Eq. (128) and with the very slow late-time decay of ε. The scalar variance production term PS is balanced by the sum of the dissipation term DS and the smaller in magnitude turbulent diffusion term, and the scalar variance dissipation rate production term Pχ is exactly balanced by the destruction term Dχ such that χ decreases very slowly with time across the layer as in RT and RM mixing. The advection and shear production terms associated with the vertical velocity are close to zero.

FIG. 35.

Budget of the self-similar turbulent kinetic energy equation (102) at the final time for KH mixing.

FIG. 35.

Budget of the self-similar turbulent kinetic energy equation (102) at the final time for KH mixing.

Close modal
FIG. 36.