Hostname: page-component-55f67697df-bzg56 Total loading time: 0 Render date: 2025-05-08T14:11:03.075Z Has data issue: false hasContentIssue false

Stochastic wavevector model for rapidly distorted compressible turbulence

Published online by Cambridge University Press:  08 May 2025

Noah Zambrano*
Affiliation:
Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48103, USA
Karthik Duraisamy
Affiliation:
Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48103, USA
*
Corresponding author: Noah Zambrano, [email protected]

Abstract

A stochastic wavevector approach is formulated to accurately represent compressible turbulence subject to rapid deformations. This approach is inspired by the incompressible particle representation model of Kassinos & Reynolds (1994), and preserves the exact nature of compressible rapid distortion theory (RDT). The adoption of a stochastic – rather than Fourier – perspective simplifies the transformation of statistics to physical space and serves as a starting point for the development of practical turbulence models. We assume small density fluctuations and isentropic flow to obtain a transport equation for the pressure fluctuation. This results in four fewer transport equations compared with the compressible RDT model of Yu & Girimaji (Phys. Fluids, vol. 19, 2007, 041702). The final formulation is closed in spectral space and only requires numerical approximation for the transformation integrals. The use of Monte Carlo for unit wavevector integration motivates the representation of the moments as stochastic variables. Consistency between the Fourier and stochastic representation is demonstrated by showing equivalency between the evolution equations for the velocity spectrum tensor in both representations. Sample clustering with respect to orientation allows for different techniques to be used for the wavevector magnitude integration. The performance of the stochastic model is evaluated for axially compressed turbulence, serving as a simplified model for shock–turbulence interaction, and is compared with linear interaction approximations and direct numerical simulation (DNS). Pure and compressed sheared turbulence at different distortion Mach numbers are also computed and compared with RDT/DNS data. Finally, two additional deformations are applied and compared with solenoidal and pressure-released limits to demonstrate the modelling capability for generic rapid deformations.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

A majority of practical turbulence models rely on the Boussinesq approach which relates the Reynolds stress anisotropy $\overline {u'_i u'_{\!j}}-q^{2}\delta _{ij}/3$ , where $q^{2}$ is twice the turbulent kinetic energy, directly to the mean rate of strain $S_{ij}$ through the eddy viscosity $\nu _t$ , such that the Reynolds stress is determined by the rate of strain. This is generally a reasonable approximation at low mean-flow deformation rates because turbulence has time to respond to the mean flow. However, the eddy-viscosity assumption fails to accurately represent turbulence at large or rapid deformations rates $S \tau _t\gg 1$ , where $S$ is the (large) shear strain rate and $\tau _t$ is the turbulent time scale. In these flows, the turbulence reacts to rapid deformations at a rate different from the mean-flow time scales and behaves more like an elastic solid (Taylor Reference Taylor1935; Crow Reference Crow1968). Thus, the viscous fluid-like qualitative behaviour that an eddy-viscosity assumption imposes (Crow Reference Crow1968) is no longer accurate. For large deformations, rapid distortion theory (RDT) provides an accurate representation of the turbulence by linearising the governing equations. However, RDT is unclosed in inhomogeneous turbulent flows due to the non-locality associated with the fluctuating pressure. While most relevant engineering applications are inhomogeneous, it is still valuable to develop models for simplified homogeneous problems and extend them to inhomogeneous cases. The advantage of building a model beginning from homogeneous RDT is that it requires no modelling assumptions. The motivation for this work is to develop such a model from homogeneous RDT specifically for compressible turbulence such that a future, more complete model can be developed with a rigorous foundation.

In contrast to developing exact solutions and analysing the behaviour of compressible turbulence, as was done in many past compressible RDT studies (Goldstein Reference Goldstein1978; Durbin & Zeman Reference Durbin and Zeman1992; Jacquin et al. Reference Jacquin, Cambon and Blin1993; Simone et al. Reference Simone, Coleman and Cambon1997), we develop a modelling framework that pursues RDT analysis from a stochastic perspective, rather than a spectral one. The stochastic designation is different from most common stochastic methods which include random perturbation terms in the model for turbulent evolution (Pope Reference Pope2001). The formulation presented in this work has deterministic evolution equations that evolve a stochastic initial state, as a first step towards a more general modelling framework. A stochastic perspective allows the use of stochastic integration methods for transforming spectral/stochastic statistics to physical space. This technique was pioneered by Kassinos & Reynolds (Reference Kassinos and Reynolds1994) for incompressible turbulence, where more details of the equivalency between the spectral and stochastic perspective are given in the book by Sagaut & Cambon (Reference Sagaut and Cambon2018). The most recent work to extend this idea to compressible turbulence was done by Yu & Girimaji (Reference Yu and Girimaji2007), Bertsch et al. (Reference Bertsch, Suman and Girimaji2012) and Lavin et al. (Reference Lavin, Girimaji, Suman and Yu2012), but requires solving 21 coupled transport equations.

The RDT is valid when the time scale of mean distortion rate is much shorter than the time scale of turbulent interactions (Batchelor & Proudman Reference Batchelor and Proudman1954). This limiting case is achieved by considering very large deformation values in the mean deformation tensor $\partial \bar {U}_i/\partial x_{\!j}$ . The advantage of RDT is that it simplifies the governing equations by linearising the fluctuation equations with respect to the mean flow (Durbin & Pettersson-Reif Reference Durbin and Pettersson-Reif2011). Fortunately, for homogeneous turbulence, this method is exact in that no modelling assumptions are made. Linearisation is accomplished by neglecting higher-order fluctuating terms in the fluctuating equations. The linearised first-moment equations are then used to formulate evolution equations for the second moments. For compressible RDT, the usage of a Helmholtz decomposition and transformation in Craya–Herring coordinates is popular because significant simplifications can be made based on the symmetries of the mean flow or properties of the fluctuating field. In this frame of reference, the dilatational part of the velocity field is aligned with the wavevector and two orthogonal solenoidal components lie in the plane normal to the wavevector (Cambon et al. Reference Cambon, Coleman and Mansour1993; Jacquin et al. Reference Jacquin, Cambon and Blin1993; Simone et al. Reference Simone, Coleman and Cambon1997). Rather than using this decomposition and transformation, we only apply the transformation for deforming coordinates (Rogallo Reference Rogallo1981), and then solve for fluctuating quantities in this reference frame. A similar aspect to the Craya–Herring frame where the dilatational part of the velocity field is aligned with the wavevector while the solenoidal part is orthogonal is also realised in deforming coordinates. In addition, the method of randomly initialising isotropic turbulence in Rogallo (Reference Rogallo1981) is interpretable using the Craya–Herring frame. The deforming coordinate transformation is effectively a transformation from an Eulerian to a Lagrangian perspective and thus gives the basis for a particle representation of turbulence. Solenoidal and dilatational components for the second moments are found after the fact by using both the compressible and solenoidal transport equations for turbulence statistics.

An important parameter for characterising rapidly strained compressible turbulence is the distortion Mach number $M_d = S\ell /a$ (Durbin & Zeman Reference Durbin and Zeman1992), where $\ell$ is an integral length scale and $a=\sqrt {\gamma \bar {p}/\bar {\rho }}$ is the mean speed of sound. This is similar to the gradient Mach number $M_g$ introduced by Sarkar (Reference Sarkar1995) to describe the evolution of nonlinear turbulence. The distortion Mach number gives a quantitative sense of which regime the turbulence lies in. Jacquin et al. (Reference Jacquin, Cambon and Blin1993) and Cambon et al. (Reference Cambon, Coleman and Mansour1993) defined the solenoidal regime for low $M_d$ values and the pressure-released regime for higher $M_d$ values. In these regimes, either the influence of dilatational effects are ignored (solenoidal) or the pressure effects are considered negligible (pressure-released). When $M_d$ is near these limits, the governing equations greatly simplify such that analytical solutions are possible or the computation of exact solutions is straightforward. The distortion Mach number is used to study the performance of the new stochastic model across a range bounded by the limits and can be compared to direct numerical simulation (DNS)/RDT solutions of past studies. Observing how the model predictions change across $M_d$ will also display the unique behaviour of compressible turbulence described in past studies.

In essence, the primary focus of this work is to extend the particle representation model (PRM) of Kassinos & Reynolds (Reference Kassinos and Reynolds1994) to compressible turbulence. One of the first wavevector models for turbulence appears in the thesis by Fung (Reference Fung1990), where the Eulerian velocity field is represented by a large number of Fourier coefficients and the wavevector orientation is chosen randomly. This approach is also called a kinematic simulation and has been extended to anisotropic cases by Cambon et al. (Reference Cambon, Godeferd, Nicolleau and Vassilicos2004) and Favier et al. (Reference Favier, Godeferd and Cambon2010). A similar anisotropic stochastic wavevector model is described by Kassinos & Reynolds (Reference Kassinos and Reynolds1994), where the initial conditions are randomly generated (hence the stochastic part), and evolved in time deterministically. By shifting to a Lagrangian perspective, we can treat the instantaneous variables as stochastic realisations or ‘particles’ and numerically solve the transformation integrals. In compressible flows, however, the fluctuating velocity field is not solenoidal which causes the wavevector magnitude to appear in the evolution equation for the velocity spectrum tensor and prevents closure at the directional velocity spectrum tensor level. This issue is addressed by introducing an additional integration step but requires more particles compared with the directional velocity spectrum integration method. Our model is also an improved version of the compressible RDT model developed by Yu & Girimaji (Reference Yu and Girimaji2007). The new model is able to achieve better results using four fewer transport equations. This is accomplished by defining a new normalised pressure fluctuation variable that enables closure using isentropic relations. Additionally, to demonstrate consistency between the stochastic and Fourier formulations, we show that integration with respect to the wavevector state space of the stochastic velocity spectrum tensor is equivalent to the integration of the velocity spectrum with respect to the Fourier modes. We also address the inconsistency of using the directional spectrum integration and the unnecessary use of the full velocity spectrum tensor rather than just the symmetric part. Finally, the new compressible PRM (CPRM) is tested on various cases and compared with RDT/DNS data. Axially compressed turbulence is verified against the DNS data of Cambon et al. (Reference Cambon, Coleman and Mansour1993) and used as a simplified shock–turbulence model for varying initial Mach numbers. This simple shock–turbulence model is similar to the idea proposed by Jacquin et al. (Reference Jacquin, Cambon and Blin1993) and is compared with post-shock predictions with linear interaction approximations (LIAs). Performance in homogeneous shear without any mean compression is also evaluated against the DNS/RDT data of Simone et al. (Reference Simone, Coleman and Cambon1997) and provides a basis for comparison against the RDT model of Yu & Girimaji (Reference Yu and Girimaji2007). The solenoidal version of the model is verified for a sheared compression case against the RDT data of Mahesh (Reference Mahesh1996). Finally, application to generic rapid mean deformations is demonstrated by applying the model to plane strain and axisymmetric contraction and verifying the limiting solenoidal/pressure-released behaviour for each deformation.

2. Governing equations and basic assumptions

This work develops the evolution equations for homogeneous anisotropic compressible turbulence using the Reynolds decomposition, which states that $u=\bar {u}+u'$ such that $\bar {u}$ is the ensemble average (also called Reynolds average) and $u'$ is the fluctuation from the mean. The fluctuation field in compressible turbulence includes solenoidal/dilatational velocity modes and an entropy mode. The dilatational velocity is coupled to pressure fluctuations such that it satisfies an acoustic wave equation (Simone et al. Reference Simone, Coleman and Cambon1997). In this work, we consider compressible flows where the fluctuation field is composed of both modes (typically called compressible turbulence). In compressible flows, the mean velocity field is not divergence free, i.e. $\nabla \cdot \bar {\textbf{u}}\neq 0$ . This condition arises from the continuity equation when density varies. The fluctuation field is not divergence free either, i.e. $\nabla \cdot \textbf{u}^{\prime}\neq 0$ . This means that the fluctuating velocity field has both solenoidal and dilatational components, where $\nabla \cdot \textbf{u}_s'=0$ and $\textbf{u}_s$ denotes the solenoidal component and $\textbf{u}_d$ the dilatational component.

The restriction of homogeneity on compressible flows is more stringent as compared with the incompressible case. Both homogeneous incompressible and compressible turbulence require that the mean velocity gradient has the form $\partial \bar {u}_i/\partial x_{\!j}=A_{ij}(t)$ . However, as shown by Blaisdell (Reference Blaisdell1991), the mean density and pressure are only functions of time because the right-hand side of the evolution equation for the density and pressure statistics must also only be functions of time (Pope Reference Pope2001). This limitation is only true when the density dilatation $\overline {\rho \partial u'_i/\partial x_i}$ and pressure dilatation $\overline {p\partial u'_i/\partial x_i}$ are not zero (Blaisdell Reference Blaisdell1991), which is true for the compressible flows considered. Another assumption used to simplify the governing equation is that of small density fluctuations compared with the mean density. While some variable-density flows may have non-negligible density fluctuations compared with the mean (as described by Sandoval (Reference Sandoval1995)), this work primarily focuses on turbulence where the ratio $\rho '/\bar {\rho }$ is approximately zero.

The governing equations comprise the inviscid conservation of momentum, continuity and conservation of fluctuating entropy. The equations are

(2.1) \begin{align} \frac {\partial \rho }{\partial t}=-\frac {\partial }{\partial x_i}(\rho u_i), \end{align}

(2.2) \begin{align} \frac {\partial \rho u_i}{\partial t}=-\frac {\partial }{\partial x_{\!j}}(\rho u_iu_{\!j}+p\delta _{ij}), \end{align}

(2.3) \begin{align} \frac {\partial s'}{\partial t}=-u_i\frac {\partial s'}{\partial x_i}, \end{align}

where $s'$ is the specific entropy fluctuation. If we assume an adiabatic process for an ideal gas with constant specific heat, then the conservation of entropy and the isentropic relation $p/\rho ^{\gamma }=\text{constant}$ imply that $\gamma {\rm D}(\rho '/\bar {\rho })/{\rm D}t={\rm D}(p'/\bar {p})/{\rm D}t$ , where ${\rm D}/{\rm D}t$ is the substantial derivative. By using the continuity equation and the former relation to relate the pressure fluctuation to the velocity fluctuation divergence $\partial u'_i/\partial x_i$ , the variables required to close the second-moment equations are considerably reduced when compared with the use of the equation of state and conservation of energy as done by Yu & Girimaji (Reference Yu and Girimaji2007). This technique was used by other authors in studies of compressible turbulence as well (Durbin & Zeman Reference Durbin and Zeman1992; Cambon et al. Reference Cambon, Coleman and Mansour1993; Simone et al. Reference Simone, Coleman and Cambon1997).

3. Representing compressible rapid distortions

The RDT analysis is used to develop closed-form evolution equations for relevant moments of interest. The linear solutions are valid for turbulent flows with distortions that are sufficiently rapid such that $S\tau _t = S\ell /q \gg 1$ . Applicability is also limited to short times and small wavenumber modulus $\kappa$ , such that $\kappa q/S\ll 1$ .

The evolution equations for the velocity and pressure fluctuations are the only equations required to close the Reynolds stress evolution equation. The first step is the linearisation of the governing fluctuation equations. Full fluctuation equations are obtained by traditional methods as described by Wilcox (Reference Wilcox1998). Higher-order fluctuating terms are then removed to yield the linearised equations:

(3.1) \begin{align} \frac {\partial \rho '}{\partial t}+\bar {u}_{\!j}\frac {\partial \rho '}{\partial x_{\!j}}=-\bar {\rho }\frac {\partial u'_{\!j}}{\partial x_{\!j}}-\rho '\frac {\partial \bar {u}_{\!j}}{\partial x_{\!j}}, \end{align}

(3.2) \begin{align} \frac {\partial u'_i}{\partial t}+\bar {u}_{\!j}\frac {\partial u'_i}{\partial x_{\!j}}=-u'_{\!j}\frac {\partial \bar {u}_i}{\partial x_{\!j}}-\frac {1}{\bar {\rho }}\frac {\partial p'}{\partial x_i}, \end{align}

(3.3) \begin{align} \frac {\partial }{\partial t}\left (\frac {p'}{\gamma \bar {p}}\right )+\bar {u}_{\!j}\frac {\partial }{\partial x_{\!j}}\left (\frac {p'}{\gamma \bar {p}}\right )=-\frac {\partial u'_i}{\partial x_i}. \end{align}

For purely solenoidal modes, pressure fluctuation influences are decoupled and equations simplify with the use of a fluctuating pressure Poisson equation derived from (3.2) and the divergence free condition $\partial u'_i/\partial x_i =0$ :

(3.4) \begin{align} \frac {1}{\bar {\rho }}\frac {\partial ^{2} p^{\prime}}{\partial x_i^{2}}=-2\frac {\partial u_{\!j}^{\prime}}{\partial x_i}\frac {\partial \bar {u}_i}{\partial x_{\!j}}-\bar {u}_{\!j}\frac {\partial ^{2} u_{\!j}^{\prime}}{\partial x_i^{2}}. \end{align}

The pressure Poisson equation will be used to derive the solenoidal component of the compressible RDT of turbulence. Next, a Fourier transform is applied and the reference frame is changed from an Eulerian to Lagrangian point of view. This is done by introducing the substantial derivative ${\rm D}/{\rm D}t$ and is equivalent to transforming into deforming coordinates as done by Rogallo (Reference Rogallo1981):

(3.5) \begin{align} \frac {{\rm D}\hat {u}_i}{{\rm D}t}=-\frac {\partial \bar {u}_i}{\partial x_k}\hat {u}_k+\textrm{i}\kappa _i\frac {\hat {p}}{\bar {\rho }}, \end{align}

(3.6) \begin{align} \frac {{\rm D}\kappa _i}{{\rm D}t}=-\kappa _k\frac {\partial \bar {u}_k}{\partial x_i}, \end{align}

(3.7) \begin{align} \frac {\rm D}{{\rm D}t}\left (\frac {\hat {p}}{\gamma \bar {p}}\right ) =\textrm{i}\kappa _k\hat {u}_k. \end{align}

The hat notation denotes a Fourier-transformed variable and $\textrm{i}$ is the imaginary number. Here $\kappa _i$ is the component of the wavevector, where the evolution equation in homogeneous turbulence from a Lagrangian point of view is obtained from Kassinos & Reynolds (Reference Kassinos and Reynolds1994). Again, these RDT first-moment equations are similar to those used in other compressible RDT studies (Durbin & Zeman Reference Durbin and Zeman1992; Cambon et al. Reference Cambon, Coleman and Mansour1993; Simone et al. Reference Simone, Coleman and Cambon1997). For solenoidal modes, the pressure Poisson equation in wavespace is

(3.8) \begin{align} \displaystyle\frac {-\kappa _i^{2}\hat {p}}{\bar {\rho }}=2\textrm{i}\kappa _i\hat {u}^{s}_{\!j}\frac {\partial \bar {u}_i}{\partial x_{\!j}}. \end{align}

The evolution of the one-point Reynolds stress $R_{ij}=\overline {u'_iu'_{\!j}}$ in physical space is the target of this analysis. The one-point Reynolds stress is related to the two-point Reynolds stress correlation $R_{ij}(\textbf{r})=\overline {u'_i(\textbf{x})u'_{\!j}(\textbf{x}^{\prime})}$ , where Durbin & Pettersson-Reif (Reference Durbin and Pettersson-Reif2011) define the two-point correlation through the symmetric part of the velocity spectrum tensor:

(3.9) \begin{align} \overline {u'_i(\textbf{x})u'_{\!j}(\textbf{x}^{\prime})} =\iiint ^{\infty }_{-\infty }\Phi _{ij}({\boldsymbol{\kappa }) {\rm e}}^{{\textrm{i}{\boldsymbol{\kappa }}}\cdot (\textbf{x}^{\prime}-\textbf{x})}\,{\rm d}^{3}{\boldsymbol{\kappa }}. \end{align}

When $\textbf{x}^{\prime}=\textbf{x}$ , then the two-point correlation reduces to the one-point correlation $R_{ij}$ which is then equal to the integral of $\Phi _{ij}$ across all ${\boldsymbol{\kappa }}$ .

3.1. Normalised pressure fluctuations and second moments

A new variable, the normalised pressure fluctuation (denoted as $\hat {\wp }=\textrm{i}\hat {p}/(\gamma \bar {p})$ henceforth), is introduced to simplify the upcoming derivations for the second moments and to remove dependence on imaginary terms. The evolution equation for $\hat {\wp }$ follows (3.7):

(3.10) \begin{align} \frac {{\rm D}\hat {\wp }}{{\rm D}t} =-\kappa _k\hat {u}_k. \end{align}

Next, the second moments must be determined. The velocity spectrum tensor $\Phi _{ij}$ is used to find the one-point Reynolds stress $R_{ij}$ , but its evolution equation is unclosed. We now define the single-point pressure–velocity spectrum vector ( $\beta _i$ ) and pressure spectrum scalar ( $\Gamma$ ) to help close the $\Phi _{ij}$ evolution equation. These additional spectra are defined in a manner similar to that of the velocity spectrum tensor. The symmetric spectra are defined as

(3.11) \begin{align} \Phi _{ij} =\frac {1}{2} \left (\overline { \hat {u}_i\hat {u}^{+}_{\!j}}+\overline { \hat {u}^{+}_i\hat {u}_{\!j}}\right ), \end{align}

(3.12) \begin{align} \beta _{i}=\frac {1}{2}\left (\overline {\hat {\wp }^{+}\hat {u}_i} + \overline {\hat {\wp }\hat {u}_i^{+}}\right ), \end{align}

(3.13) \begin{align} \Gamma =\overline {\hat {\wp }^{+}\hat {\wp }}, \end{align}

where the $+$ superscript represents the complex conjugate. The velocity spectrum tensor $\Phi _{ij}$ is defined as the real components of the ensemble average of the Fourier-transformed velocity fluctuations (Durbin & Pettersson-Reif Reference Durbin and Pettersson-Reif2011). The final first-moment equations are then established by substituting in the normalised pressure fluctuation:

(3.14) \begin{align} \frac {{\rm D}\hat {u}_i}{{\rm D}t}=-\frac {\partial \bar {u}_i}{\partial x_k}\hat {u}_k+\kappa _ia^{2}\hat {\wp }, \end{align}

(3.15) \begin{align} \frac {{\rm D}\kappa _i}{{\rm D}t}=-\kappa _k\frac {\partial \bar {u}_k}{\partial x_i}, \end{align}

(3.16) \begin{align} \frac {{\rm D}\hat {\wp }}{{\rm D}t} =-\kappa _k\hat {u}_k. \end{align}

The appearance of the $\kappa _ia^{2}\hat {\wp }$ term is associated with the dispersion frequency of acoustic waves (Sagaut & Cambon Reference Sagaut and Cambon2018). This introduces new difficulties with the integration of spectral higher-order moments to physical space. With the first-moment evolution equations established, the second-moment equations are derived by substituting in the time derivatives of the first moments and differentiating with respect to time using the product rule:

(3.17) \begin{align} \frac {\rm D}{{\rm D}t}(\Phi _{ij})=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ki}+\kappa _ia^{2}\beta _{\!j}+\kappa _{\!j}a^{2}\beta _i, \end{align}

(3.18) \begin{align} \frac {\rm D}{{\rm D}t}(\beta _i)=-\kappa _k\Phi _{ki}-\frac {\partial \bar {u}_i}{\partial x_k}\beta _{k}+\kappa _ia^{2}\Gamma, \end{align}

(3.19) \begin{align} \frac {\rm D}{{\rm D}t}(\Gamma )=-2\kappa _i\beta _i. \end{align}

These final spectral second-moment equations are fully closed and require no new modelling assumption aside from the assumptions made in RDT.

3.2. Transformation to physical space

The final step is to take an inverse Fourier transform to obtain statistics in physical space. This is an integration over all wavenumber components, which is simplified by transforming into spherical coordinates. For example, the evolution equation for the Reynolds stress becomes

(3.20) \begin{align} \frac {\rm D}{{\rm D}t}(R_{ij})= \!\int ^{2\pi}_0 \!\int ^{\pi} _0 \!\int ^{\infty }_{-\infty } \left (-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ki}+\kappa _ia^{2}\beta _{\!j}+\kappa _{\!j}a^{2}\beta _i\right ) \kappa ^{2}\sin {\theta }\,{\rm d}\kappa \,{\rm d}\theta \,{\rm d}\phi, \end{align}

where $\kappa =\sqrt {\kappa _i\kappa _i}$ and $\kappa _i$ can be decomposed into product of wavevector magnitude $\kappa$ and orientation $\theta, \phi$ . Note that the appearance of $\kappa _i$ means we cannot analytically integrate with respect to the wavevector magnitude as done by Kassinos & Reynolds (Reference Kassinos and Reynolds1994). The solenoidal velocity spectrum tensor, however, can be integrated this way. The latter expression can be solved using a fast Fourier transform to get the most accurate results. However, for this work, we chose to approach the integration through stochastic methods and numerical integration.

3.3. Higher-order tensors

The transformation into Fourier space allows for relatively easier computation of several higher-order tensors. These tensors are used to obtain more information about the structure of turbulence. The structure tensors as defined by Kassinos & Reynolds (Reference Kassinos and Reynolds1996) are of particular interest. The dimensionality tensor is based on the kinetic energy spectrum:

(3.21) \begin{align} D_{ij}=\iiint ^{\infty }_{-\infty }\frac {\kappa _i\kappa _{\!j}}{\kappa ^{2}}\Phi _{mm}\,{\rm d}^{3}{\boldsymbol{\kappa }}, \end{align}

and gives information on whether the turbulence is one- or two-dimensional. The dimensionality gives information on the type of eddies present in the flow. The eddy types are the one-dimensional jettal eddy, the two-dimensional vorticial eddy and the helical eddy (Kassinos & Reynolds Reference Kassinos and Reynolds1994). The structure tensor information is also required to accurately represent rotating flows (Reynolds Reference Reynolds1989). Similarly, the circulicity tensor, which describes the vorticial field, is also found easily in Fourier space:

(3.22) \begin{align} F_{ij}=\epsilon _{inm}\epsilon _{jts}\iiint ^{\infty }_{-\infty }\frac {\kappa _n\kappa _t}{\kappa ^{2}}\Phi _{ms}\,{\rm d}^{3}{\boldsymbol{\kappa }}. \end{align}

In the former equation, $\epsilon _{ijk}$ is the alternating tensor. Finally, higher-rank tensors are also found without requiring additional transport equations. An important fourth-rank tensor,

(3.23) \begin{align} M_{ijpq}=\iiint ^{\infty }_{-\infty }\frac {\kappa _p\kappa _q}{\kappa ^{2}}\Phi _{ij}\,{\rm d}^{3}{\boldsymbol{\kappa }}, \end{align}

gives information regarding the fluctuating pressure field and is often used to represent non-local information as it contains information on two-point statistics (Kassinos & Reynolds Reference Kassinos and Reynolds1994).

3.4. Limitations

The three main assumptions of the current formulation imply limitation to a certain subset of flows. The most substantial assumptions are those of linearisation about a base flow, small density fluctuations compared with the mean and the isentropic assumption. Linearisation is valid only for turbulence subject to rapid deformation. This simplification allows the closure of the equations for the first and second moments without any phenomenological modelling. However, it implicitly prohibits the application of the formulations to slower deformations, as experienced in many engineering flows of interest. We take linearisation as a first step towards more general models, where it can be viewed as a limiting behaviour with the formulation matching predictions from RDT when subject to increasingly large deformations. However, the formulation is expected to gradually become imprecise as deformation decreases.

To represent slower deformations, additional modelling approximations must be added to the formulation and will be explored in upcoming work. The assumption of small density fluctuations also causes discrepancies with application to problems with large compression ratios. The rationale is that as the mean density changes, the magnitude of the fluctuations can become significant. While we do assume that the density fluctuation is small compared with the mean, it is weakly enforced. The simplification is used in the derivation of (3.1) and (3.2), where it is assumed that terms scaled by $\rho '/\bar {\rho }$ are small enough to be ignored when compared in magnitude with the other terms. However, $\rho '/\bar {\rho }$ is not assumed to be zero when considering its rate of change. The isentropic relation uses ${\rm D}(\rho '/\bar {\rho })/{\rm D}t$ to derive the pressure equation, and thus the influence of the density fluctuation with respect to the mean is accounted for in the pressure terms. This weak assumption was similarly applied by Jacquin et al. (Reference Jacquin, Cambon and Blin1993) and Cambon et al. (Reference Cambon, Coleman and Mansour1993) in their studies of compressed homogeneous turbulence.

Finally, the isentropic relation limits the current formulation to adiabatic and entropy-conserving flows. While many compressible flows have non-entropy-conserving features such as shocks, we choose to invoke isentropy due to the significant reduction in equations. If isentropy was not invoked, the pressure equation would need to be derived in a similar manner to that of Yu & Girimaji (Reference Yu and Girimaji2007), where the ideal gas law and energy equation are used to relate temperature and density to pressure evolution. This would result in four more transport equations; two first-rank tensors and two scalars. This amounts to eight more transport equations, which adds to the computational cost. This work seeks a simple formulation which is applicable to most regions where compressible turbulence is important. Non-isentropic parts of the flow usually require additional treatment. For example, turbulence models for flows with shocks require special shock-capturing methods or variable-Prandtl-number ad hoc modifications (Roy et al. Reference Roy, Pathak and Sinha2018) to accurately predict the shock, and non-adiabatic flows such as hypersonic boundary layers require special treatment to account for the inhomogeneity from the wall. We believe that the treatment for various issues that appear in non-isentropic flows can be combined with the required modifications to the current formulation, rather than incorporated in the base formulation.

4. Stochastic modelling

In a stochastic representation, flow variables are treated as random fields. Since the Fourier transform is a linear operation, the Fourier-transformed random fields are also random fields in ${\boldsymbol{\kappa }}$ -space. Formally, the random field can be written as $\phi (x,t)=\int {\rm e}^{{-i{\boldsymbol{\kappa }}}\cdot \textbf{x}}{\rm d}Z({\boldsymbol{\kappa }}, t)$ , where $Z({\boldsymbol{\kappa }}, t )$ is a complex-valued random measure with the property that ${\mathbb{E}}[{\rm d}Z({\boldsymbol{\kappa }}, t)] = 0$ and ${\mathbb{E}}[{\rm d}Z({\boldsymbol{\kappa }}_1, t_1){\rm d}Z^{*}({\boldsymbol{\kappa }}_2, t_2)] = \delta ({\boldsymbol{\kappa }}_1 - {\boldsymbol{\kappa }}_2) \delta (t_1 - t_2)$ $S({\boldsymbol{\kappa }}){\rm d}{\boldsymbol{\kappa }}_1{\rm d}{\boldsymbol{\kappa }}_2{\rm d}t_1{\rm d}t_2$ , where $S({\boldsymbol{\kappa }})$ is the power spectral density. From these properties, the spectral representation of the stochastic fields has the following mode properties: ${\mathbb{E}}[\hat {\phi }({\boldsymbol{\kappa }}, t)] = 0$ and ${\mathbb{E}}[\hat {\phi }({\boldsymbol{\kappa }}_1, t_1)\hat {\phi }^{*}({\boldsymbol{\kappa }}_2, t_2)] = \delta ({\boldsymbol{\kappa }}_1 - {\boldsymbol{\kappa }}_2)\delta (t_1 - t_2)S({\boldsymbol{\kappa }})$ , which match the properties of the physical spectral modes. This is the primary mathematical rationale for why the properties of Fourier modes allow for representation as stochastic realisations. Section 4.1 details the specific flow variables associated with each stochastic representation. From a physical perspective, this is supported by the hypothesis that in homogeneous turbulence, velocity fluctuations at points separated by large distances tend to become uncorrelated. The de-correlation in wavevector space for widely different wavevectors also enforces the idea of a turbulence energy cascade which suggests that energy is transferred predominantly between nearby scales.

Another motivation for shifting to a stochastic representation is the ease of transformation from spectral to physical space. As discussed in § 3.2, the Fourier modes must be integrated with respect to all wavevectors. One way of doing this integration is through a Monte Carlo integral. Monte Carlo integration for some general function $f(\textbf{x})$ over a domain $\Omega$ is given by (Caflisch Reference Caflisch1998)

(4.1) \begin{align} \int _\Omega f(\textbf{x}){\rm d} \textbf{x} \approx \frac {V}{N_s}\sum ^{N_s}_{}{i=1}f(\textbf{x}_i),\quad V=\int _\Omega {\rm d}\textbf{x}, \end{align}

where $N_s$ is the number of samples that are uniformly sampled from $\Omega$ . This integration approximation approaches the expected value ${\mathbb{E}}(f(\textbf{x}))$ as $n\rightarrow \infty$ . An important concern regarding the accuracy of the integration is realised by noting the infinite integration bounds of (3.20). This must be approximated where some cut-off wavevector magnitude is specified a priori. As done previously, we use (3.20) to convert to a spherical integral and explicitly express the wavevector in terms of orientation and magnitude. In the work by Kassinos & Reynolds (Reference Kassinos and Reynolds1994), the infinite integration with respect to magnitude was done analytically as the wavevector magnitude did not appear in the evolution equation for the conditional Reynolds stress, thus avoiding the problem of approximating the infinite integral bounds. This meant that the incompressible PRM evolved the directional velocity spectrum tensor

(4.2) \begin{align} \Phi ^{\nearrow \kappa }_{ij}(\theta, \phi ) =\int ^{\infty }_0 \kappa ^{2} \Phi _{ij}(\kappa, \theta, \phi ) {\rm d}\kappa, \end{align}

instead of the velocity spectrum tensor. This is the lowest level at which the incompressible formulation can be closed (Van Slooten & Pope Reference Van Slooten and Pope1997). Kassinos & Reynolds (Reference Kassinos and Reynolds1994) also scaled the directional velocity spectrum tensor by a factor of $4\pi$ at initialisation rather than at each time step. This constant appears in the Monte Carlo integration with respect to orientation. By evolving this quantity and averaging across all samples, Kassinos & Reynolds (Reference Kassinos and Reynolds1994) obtained an equivalent method to the Monte Carlo integration with respect to the wavevector orientation:

(4.3) \begin{align} R_{ij}=\int ^{2\pi }_0 \int ^{\pi} _0 \Phi ^{\nearrow \kappa }_{ij} \sin {\theta }\,{\rm d}\theta \,{\rm d}\phi \approx \frac {1}{N_s}\sum ^{N_s}_{n=1}4\pi \Phi ^{\nearrow \kappa }_{ij} \big(\theta ^{(n)},\phi ^{(n)}\big), \end{align}

where the $(n)$ superscript denotes the sample number. Integrating with respect to all wavevector magnitudes is advantageous because it reduces the number of samples required to effectively and accurately use Monte Carlo integration. This is so because it avoids integration of the energy spectrum. The step-like shape of the energy spectrum is not well suited for Monte Carlo integration, where the requirement of uniform sampling makes it difficult to resolve the sharp feature and dramatically increases sample count. Unfortunately, integrating analytically with respect to all wavevector magnitudes is not possible for the compressible case as the time derivative of the velocity spectrum tensor is a function of the wavevector orientation and magnitude, whereas in the incompressible/solenoidal case it is independent. We must introduce another Monte Carlo approximation in order to obtain a solution. This involves selecting a range of wavevector magnitudes $\Delta k$ , such that a Monte Carlo integral is used to find the directional velocity spectrum tensor at each particle:

(4.4) \begin{align} \Phi ^{\nearrow \kappa }_{ij}\big(\theta ^{(n)},\phi ^{(n)}\big)\approx \frac {\Delta k}{N_{sh}}\sum ^{N_{sh}}_{{sh}=1}\big(\kappa ^{(sh)}\big)^{2}\Phi _{ij}\big(\kappa ^{(sh)},\theta ^{(n)},\phi ^{(n)}\big), \end{align}

where the $sh$ subscript/superscript represents the ‘shell’ sample of the wavevector magnitude. To simplify the methodology, we can combine the two Monte Carlo integrals into one, where we sample from a set of particles that each have a unique wavevector magnitude and orientation. This particle realisation differs from the cluster method developed by Kassinos & Reynolds (Reference Kassinos and Reynolds1994) as we no longer group particles based on orientation and rather perform the integration for each wavevector component in one large Monte Carlo integration:

(4.5) \begin{align} R_{ij}\approx \frac {4\pi \Delta k}{N_s}\sum ^{N_s}_{n=1}\big(\kappa ^{(n)}\big)^{2}\Phi _{ij}\big({\boldsymbol{\kappa }^{(n)}}\big). \end{align}

The insight from this analysis shows that the PRM is identical to RDT for every step up until the integration. The PRM uses an approximate Monte Carlo (and thus stochastic) integration technique rather than a Fourier transform algorithm to go from the closed set of equations in Fourier space to physical space.

4.1. Stochastic particle properties and identities

To complete the Monte Carlo integration, the velocity spectrum tensor must be sampled. This provides the motivation for representing the velocity spectrum as a stochastic variable from which we treat samples as Lagrangian particles. We call these stochastic samples particles to follow the terminology of Kassinos & Reynolds (Reference Kassinos and Reynolds1994).

Figure 1. Particle properties associated with a cluster of samples.

Under the assumptions previously stated, the stochastic representation is valid for the turbulence considered in this study. We can now formally define the stochastic variables. Suppose there exist multivariate distributions in ${\boldsymbol{\kappa }}$ -space for $\Phi _{ij}({\boldsymbol{\kappa }},t),\ \beta _i({\boldsymbol{\kappa }},t)$ and $\Gamma ({\boldsymbol{\kappa }},t)$ such that ${\mathcal{K}} \subset {{\mathbb{R}}^{3}}$ is the state space of the wavevector. We can obtain samples from ${{\boldsymbol{\kappa }}^{(1)}},{{\boldsymbol{\kappa}}^{(2)}},\ldots, {{\boldsymbol{\kappa}}^{(n)}} \in {\mathcal{K}}$ , where the $(n)$ superscript denotes sample number. The second moments defined previously correspond to realisations of the underlying distributions. Now, we sample (for a range of ${\boldsymbol{\kappa }}$ vectors) from the initial random distributions for the three second-moment probability density functions (PDFs) and advance their solutions forward in time using the evolution equations for the second moments. Note that the stochastic behaviour comes from the initialisation rather than from the evolution equations. Each realisation for $\Phi _{ij}, \beta _i$ , $\kappa _i$ and $\Gamma$ is evolved in time separately.

Particles are assigned properties which are evolved through time and are used to compute second-moment statistics. The assigned properties are the particle velocity vector $\textbf{u}^{*}$ (which can be decomposed into solenoidal and dilatational components), the particle wavevector ${\boldsymbol{\kappa }}^{*}$ and the particle normalised pressure $\wp ^{*}$ . These stochastic particles represent a one-dimensional one-component flow, a basic building block for the turbulence structure. The particle velocity vector, wavevector and normalised pressure can be thought of as stochastic samples of the respective Fourier modes. Note that compared with the incompressible PRM, the particle wavevector and velocity vector are not orthogonal since the velocity field is not divergence-free. However, the solenoidal velocity vector $\textbf{u}^{*s}$ is orthogonal to ${\boldsymbol{\kappa }}^{*}$ and the dilatational velocity vector $\textbf{u}^{*d}$ is parallel to ${\boldsymbol{\kappa }}^{*}$ such that $u_i^{*s}\kappa _i=0$ and $u^{*}_i=u^{*s}_i+u^{*d}_i$ . This is illustrated in figure 1. In addition, the particle wavevector is similar but not identical to the gradient vector defined by Kassinos & Reynolds (Reference Kassinos and Reynolds1994). Since the stochastic particles are analogous to the Fourier modes of the flow variables, the stochastic governing equations are also analogous to the RDT equations. The first-moment equations are

(4.6) \begin{align} \frac {{\rm D}u^{*}_i}{{\rm D}t}=-\frac {\partial \bar {u}_i}{\partial x_k}u^{*}_k+\kappa ^{*}_ia^2\wp ^{*}, \end{align}

(4.7) \begin{align} \frac {{\rm D}\kappa ^{*}_i}{{\rm D}t}=-\kappa ^{*}_k\frac {\partial \bar {u}_k}{\partial x_i}, \end{align}

(4.8) \begin{align} \frac {{\rm D}\wp ^{*}}{{\rm D}t} =-\kappa ^{*}_ku^{*}_k. \\[-12pt] \nonumber \end{align}

The closed set of second-moment stochastic equations are

(4.9) \begin{align} \frac {\rm D}{{\rm D}t}(\Phi ^{*}_{ij})=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi ^{*}_{kj}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi ^{*}_{ki}+\kappa ^{*}_ia^2\beta ^{*}_{\!j}+\kappa ^{*}_{\!j}a^2\beta ^{*}_i, \end{align}

(4.10) \begin{align} \frac {\rm D}{{\rm D}t}(\beta ^{*}_i)=-\kappa ^{*}_k\Phi ^{*}_{ki}-\frac {\partial \bar {u}_i}{\partial x_k}\beta ^{*}_{k}+\kappa ^{*}_ia^2\Gamma ^{*}, \end{align}

(4.11) \begin{align} \frac {\rm D}{{\rm D}t}(\Gamma ^{*})= -2\kappa ^{*}_i\beta ^{*}_i. \\[-12pt] \nonumber \end{align}

4.2. Equivalency with spectral representation

This section presents a more rigorous justification to show equivalency with spectral and stochastic representations. We turn to the PDF formulation of compressible RDT. While PDF methods are typically used to model the one-point, one-time Navier–Stokes equations (Pope Reference Pope2001), they also provide a more grounded justification for using a stochastic point of view for turbulence. In this section, the marginal PDF is used to relate the random stochastic variables to the Fourier spectra. To show equivalency and consistency between the PDF and traditional formulations, the stochastic velocity spectrum tensor must have the same evolution equation form as the Fourier velocity spectrum:

(4.12) \begin{align} \overline {u'_iu'_{\!j}}=\iiint ^{\infty }_{-\infty }\Phi _{ij}({\boldsymbol{\kappa }}){\rm d}^{3}{\boldsymbol{\kappa }}\equiv \iiint ^{\infty }_{-\infty }\Phi ^{*}_{ij}({\boldsymbol{\mathcal{K}}}){\rm d}^{3}{\boldsymbol{\mathcal{K}}}=\overline {u^{*}_iu^{*}_{\!j}}. \end{align}

The stochastic velocity spectrum is defined similarly to that in Van Slooten & Pope (Reference Van Slooten and Pope1997), where it is equal to the integral of the product of the velocity state space and joint PDF $f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}})$ . However, since the velocity also depends on the normalised pressure fluctuation $\wp$ , the full state space of the joint PDF contains a total of nine variables (two vectors and a scalar), $f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}},{\mathcal{P}})$ , for which the normalised pressure state space ${\mathcal{P}}$ must first be integrated out to give the joint PDF dependent only on velocity and wavevector states. The stochastic pressure–velocity and pressure spectra are defined in a similar fashion. The stochastic second-order moments are

(4.13) \begin{align} \Phi ^{*}_{ij}({\boldsymbol{\mathcal{K}}})=\iiiint ^{\infty }_{-\infty } V_iV_{\!j}f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}},{\mathcal{P}}){\rm d}{\mathcal{P}}\,{\rm d}^{3}{\textbf{V}}=\langle u_i^{*}u_{\!j}^{*}|{\boldsymbol{\kappa }}^{*}={\boldsymbol{\mathcal{K}}}\rangle f^{*}({\boldsymbol{\mathcal{K}}}), \end{align}

(4.14) \begin{align} \beta ^{*}_{i}({\boldsymbol{\mathcal{K}}})=\iiiint ^{\infty }_{-\infty } V_i{\mathcal{P}}f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}},{\mathcal{P}}){\rm d}{\mathcal{P}}\, {\rm d}^{3}{\textbf{V}}=\langle u_i^{*}\wp ^{*}|{\boldsymbol{\kappa }}^{*}={\boldsymbol{\mathcal{K}}}\rangle f^{*}({\boldsymbol{\mathcal{K}}}), \end{align}

(4.15) \begin{align} \Gamma ^{*}_{i}({\boldsymbol{\mathcal{K}}})=\iiiint ^{\infty }_{-\infty } {\mathcal{P}}^2f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}},{\mathcal{P}})\,{\rm d}\mathcal{P}\,{\rm d}^{3}{\textbf{V}} =\langle \wp ^{*}\wp ^{*}|{\boldsymbol{\kappa }}^{*}={\boldsymbol{\mathcal{K}}}\rangle f^{*}({\boldsymbol{\mathcal{K}}}). \end{align}

The joint PDF $f^{*}({\textbf{V}},{\boldsymbol{\mathcal{K}}},{\mathcal{P}})$ is derived using traditional methods (Heinz Reference Heinz2003) and is based on the stochastic first-moment equations (equations (4.6)–(4.8)):

(4.16) \begin{align} \frac {\partial f^{*}}{\partial t}=\frac {\partial \bar {u}_k}{\partial x_n}\frac {\partial }{\partial \mathcal{K}_n}(\mathcal{K}_kf^{*})+\frac {\partial \bar {u}_n}{\partial x_k}\frac {\partial }{\partial V_n}(V_kf^{*})+\frac {\partial }{\partial V_n}(a^2 \mathcal{K}_n {\mathcal{P}} f^{*})-\frac {\partial }{\partial {\mathcal{P}}}(\mathcal{K}_kV_kf^{*}) .\end{align}

Since an Eulerian PDF is used, the stochastic velocity spectrum evolution equation must match the non-conservative Eulerian velocity spectrum evolution equation (transformed from equation (3.17)):

(4.17) \begin{align} \frac {\partial \Phi _{ij}}{\partial t}=\kappa _m\frac {\partial \bar {u}_m}{\partial x_n}\frac {\partial \Phi _{ij}}{\partial \kappa _n}+\frac {\partial \bar {u}_k}{\partial x_k}\Phi _{ij}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ki}-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}-\kappa _ia^2\beta _{\!j}-\kappa _{\!j}a^2\beta _i. \end{align}

Now, the derivative of (4.13) with respect to time is taken to find the stochastic evolution equation. Note that the state space variables ${\textbf{V}}, {\mathcal{P}},{\boldsymbol{\mathcal{K}}}$ are constant in time, but the random variables sampled from them are not. Substituting in (4.16) into the time derivative of (4.13) and integrating each term yields the stochastic velocity spectrum transport equation. Integration by parts is used along with linearity to simplify the solution for each term. The first two terms only depend on integrating with respect to ${\textbf{V}}$ since no $\mathcal{P}$ terms appear:

(4.18) \begin{align} \iiint ^{\infty }_{-\infty } V_iV_{\!j}\frac {\partial \bar {u}_k}{\partial x_n}\frac {\partial }{\partial \mathcal{K}_n}(\mathcal{K}_kf^{*}){\rm d}^{3}{\textbf{V}}=\frac {\partial \bar {u}_k}{\partial x_n}\frac {\partial \Phi ^{*}_{ij}}{\partial \mathcal{K}_n}\mathcal{K}_k+\frac {\partial \bar {u}_k}{\partial x_k}\Phi ^{*}_{ij}, \end{align}

(4.19) \begin{align} \iiint ^{\infty }_{-\infty } V_iV_{\!j}\frac {\partial \bar {u}_n}{\partial x_k}\frac {\partial }{\partial V_n}(V_kf^{*}){\rm d}^{3}{\textbf{V}}=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi ^{*}_{kj}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi ^{*}_{ki}. \\[-2pt] \nonumber \end{align}

The third term requires four integration steps. Integration is first carried out on $\mathcal{P}$ and then on ${\textbf{V}}$ . Terms are simplified using the stochastic second-moment definitions (equations (4.13)–(4.15)):

(4.20) \begin{eqnarray} &&\def\lumina\hspace*{-1.8pc} \iiiint ^{\infty }_{-\infty } V_iV_{\!j}\frac {\partial }{\partial V_n}(a^2 \mathcal{K}_n {\mathcal{P}} f^{*}){\rm d}\mathcal{P} \,{\rm d}^{3}{\textbf{V}} \nonumber\\[4pt]&& \qquad =V_iV_{\!j}a^2 \mathcal{K}_n {\mathcal{P}} f^{*}-\iiiint ^{\infty }_{-\infty } (\delta _{ni}V_{\!j}+\delta _{nj}V_i)a^2 \mathcal{K}_n {\mathcal{P}} f^{*}{\rm d}\mathcal{P}\, {\rm d}^{3}{\textbf{V}} \nonumber \\[4pt]&& \qquad =-a^2 \mathcal{K}_i \beta ^{*}_{\!j} -a^2 \mathcal{K}_{\!j}\beta ^{*}_i. \end{eqnarray}

Finally, the fourth term also requires four integration steps. Due to linearity and independence of $\mathcal{P}$ and ${\textbf{V}}$ , the last term is zero:

(4.21) \begin{eqnarray} &&\def\lumina\hspace*{-1.8pc} \iiiint ^{\infty }_{-\infty } -V_iV_{\!j}\frac {\partial }{\partial {\mathcal{P}}}(\mathcal{K}_kV_kf^{*}){\rm d}\mathcal{P}\, {\rm d}^{3}{\textbf{V}} =-\iiint ^{\infty }_{-\infty } V_iV_{\!j}\mathcal{K}_kV_kf^{*}\,{\rm d}^{3}{\textbf{V}} \nonumber\\[4pt]&&\qquad + \iiiint ^{\infty }_{-\infty } \frac {\partial }{\partial {\mathcal{P}}}(V_iV_{\!j}) \mathcal{K}_kV_kf^{*}{\rm d}\mathcal{P} \,{\rm d}^{3}{\textbf{V}}=0.\end{eqnarray}

Putting solutions (4.18)–(4.21) together gives the evolution equation for the stochastic velocity spectrum tensor from an Eulerian frame of reference:

(4.22) \begin{align} \frac {\partial \Phi ^{*}_{ij}}{\partial t}=\mathcal{K}_m\frac {\partial \bar {u}_m}{\partial x_n}\frac {\partial \Phi ^{*}_{ij}}{\partial \mathcal{K}_n}+\frac {\partial \bar {u}_k}{\partial x_k}\Phi ^{*}_{ij}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi ^{*}_{ki}-\frac {\partial \bar {u}_i}{\partial x_k}\Phi ^{*}_{kj}-\mathcal{K}_ia^2\beta ^{*}_{\!j}-\mathcal{K}_{\!j}a^2\beta ^{*}_i. \end{align}

Equations (4.22) and (4.17) show that the two equations are identical. Since the two are identical, the stochastic representation is consistent with the Fourier representation and thus gives the same results when computing turbulent statistics in physical space.

4.3. Particle clusters

The integration with respect to wavevector magnitude presents additional difficulties. Consider a type of particle cluster where particles are grouped together such that they all have the same orientation but different wavevector magnitudes and spectra, as shown in figure 1. Our idea of clusters is nearly identical to that of Kassinos & Reynolds (Reference Kassinos and Reynolds1994), but rather than possessing random directional spectra, the clusters have a specified range of wavevector magnitudes associated with them, which directly changes the spectra. The motivation for clustering is to reduce the number of samples required to compute accurate statistics and separate the particle information on wavevector magnitude and orientation. Clustering also enables the use of non-stochastic methods for the inner integral:

(4.23) \begin{align} R_{ij}\approx \frac {1}{N_c}\sum ^{N_c}_{n=1}4\pi \int _0^{\infty } \big(\kappa ^{(sh))}\big)^2\Phi _{ij}\big(\kappa ^{(sh)},\theta ^{(c)},\phi ^{(c)}\big){\rm d}\kappa, \end{align}

where the $(c)$ superscript denotes the cluster index. Generally, the wavevector magnitude evolution depends on the wavevector orientation. However, because of the wide magnitude integration limits, the evolution of the wavevector magnitude sometimes plays a negligible role. Therefore, by splitting up the integration steps, we can explicitly ignore the magnitude evolution and only need to consider the evolution of the wavevector orientations. The evolution equation for the unit wavevector $\kappa _i/k=e_i$ is

(4.24) \begin{align} \frac {{\rm D}e_i}{{\rm D}t}=-e_k\frac {\partial \bar {u}_k}{\partial x_i}+\frac {\partial \bar {u}_k}{\partial x_r}e_ke_re_i. \end{align}

By only considering the orientation evolution, we are in essence resampling the stochastic particles at each time step such that they fall within the predefined wavevector magnitude range. The integral in equation (4.23) can be computed in various ways, such as trapezoidal, quadrature, etc. In this work, we use the trapezoidal and Gauss–Laguerre quadrature integration methods. Note that this resampling technique only works for some cases. For example, it was found that the pure shear case requires the evolution magnitude for stability reasons.

5. Application to compressible homogeneous turbulence

In this section we present results for various types of homogeneous flows. We compare against DNS and RDT data for axial compression, pure shear and sheared compression. The model is then extended to plane strain and axisymmetric contraction and the limiting behaviour is evaluated. Initialisation of the stochastic particles is also discussed.

5.1. Initialisation

The initialisation method will change depending on whether the velocity spectrum tensor or directional velocity spectrum tensor is being evolved. The simpler initialisation is the velocity spectrum tensor. Turbulence is initialised from homogeneous isotropic turbulence (HIT), unless otherwise stated. The definition for the velocity spectrum tensor in HIT is given by Durbin & Pettersson-Reif (Reference Durbin and Pettersson-Reif2011) in terms of the energy spectrum:

(5.1) \begin{align} \Phi _{ij}\left({\boldsymbol{\kappa }^{(n)}}\right)=\frac {E\left(\kappa ^{(n)}\right)}{4\pi \left(\kappa ^{(n)}\right)^2}\left (\delta _{ij}-\frac {\kappa ^{(n)}_i\kappa ^{(n)}_{\!j}}{(\kappa ^{(n)})^2}\right ), \end{align}

where $n$ denotes the particle number and $E(K)$ is the energy spectrum. Note that is initialisation is only valid for solenoidal turbulence. The velocity spectrum tensor is assigned to each $n{\rm th}$ particle given its wavevector orientation and magnitude. The initial energy spectrum is specified based on desired initial conditions. In the most general cases, the Von Kármán energy spectrum is used. It is given as (Durbin & Pettersson-Reif Reference Durbin and Pettersson-Reif2011)

(5.2) \begin{align} E(\kappa )=q^2LC_{vK}\frac {(\kappa L)^4}{[1+(\kappa L)^2]^p}, \end{align}

where $L$ is the length scale required for dimensional consistency, $q^2=R_{ii}$ and $p=17/6$ to match the $-5/3$ law. Here $C_{vK}$ is found by enforcing

(5.3) \begin{align} \int _{-\infty }^{\infty }E(k){\rm d}\kappa =\frac {1}{2}q^2, \end{align}

such that $C_{vK}\approx 0.4843$ . When comparing results with RDT/DNS data from Simone et al. (Reference Simone, Coleman and Cambon1997) or Cambon et al. (Reference Cambon, Coleman and Mansour1993), we used the same initial energy spectrum defined in those papers:

(5.4) \begin{align} E(\kappa )=q^2A\kappa ^4\exp {\frac {-2\kappa ^2}{\kappa _p^2}}, \end{align}

where $\kappa _p=8$ is the wavevector magnitude where the energy spectrum peaks and $A$ is a constant such that equation (5.3) is satisfied. To evolve the velocity spectrum tensor, the unit wavevector is initialised randomly about a unit sphere for a specified range of $k$ magnitudes. The number of unique orientations is equal to the cluster sample count, while the number of $\kappa$ discretisations is based on the integration method (logarithmic spacing with $\Delta \kappa$ total range for trapezoidal, for example). To initialise the directional velocity spectrum tensor, only the unit vectors in a unit sphere must be initialised. This is also done randomly. Each unit vector has a unique directional velocity spectrum tensor value such that

(5.5) \begin{align} \frac {1}{N}\sum ^N_{c=1}4\pi \Phi ^{\nearrow \kappa }_{kk}\big(\textbf{e}^{(c)}\big)=\frac {1}{2}q^2. \end{align}

This means that the diagonal values of the Reynolds stress are equal to $({1}/{3})q^2$ . The HIT initialisation for the directional velocity spectrum tensor is then

(5.6) \begin{align} 4\pi \Phi ^{\nearrow \kappa }_{ij}\big({\boldsymbol{\kappa }}^{(c)}/\kappa \big)=\left (\delta _{ij}-\frac {\kappa ^{(c)}_i\kappa ^{(c)}_{\!j}}{\kappa ^2}\right )\int ^\infty _{-\infty }E\big(\kappa ^{(c)}\big){\rm d}\kappa . \end{align}

The energy spectrum integral is numerically evaluated using midpoint integration approximation.

Initialisation of the pressure–velocity and pressure spectra will require knowing their respective energy spectra. Generally, these are not known. For the examples in this work, the velocity–pressure spectrum is set to zero for HIT. However, following the initialisation of the solenoidal/dilatational modes in Simone et al. (Reference Simone, Coleman and Cambon1997), the strong form of acoustic equilibrium (Sarkar et al. Reference Sarkar, Erlebacher, Hussaini and Kreiss1991) is enforced when the pressure spectrum is defined by the dilatational component of the energy spectrum. The ratio of solenoidal to dilatational turbulent kinetic energy $q^2_s/q^2_d$ must first be specified and used to scale the energy spectrum $E(k)$ , such that $E_s(\kappa )+E_d(k)=E(\kappa )$ (where $E_s(\kappa )/E_d(\kappa )=q^2_s/q^2_d$ ), to obtain the following initialisation:

(5.7) \begin{align} \Gamma \big({\boldsymbol{\kappa }^{(n)}}\big)=\frac {E_d\left(\kappa ^{(n)}\right)}{2\pi \left(\kappa ^{(n)}\right)^2\! a_0^2}. \end{align}

The influence of solenoidal and dilatational effects depends on the distortion Mach number $M_d$ (Durbin & Zeman Reference Durbin and Zeman1992). For each case, an initial distortion Mach number $M_d=M_tSq^2/\epsilon$ is specified along with a turbulent Mach number $M_t=q/a$ and turbulent Reynolds number $Re_{t}=q^4/(\nu \epsilon )$ . These three conditions along with the relationship between the total turbulent dissipation rate and energy spectrum,

(5.8) \begin{align} \epsilon =2\nu \int _0^{\infty }\kappa ^2E(\kappa ){\rm d}\kappa, \end{align}

are used to initialise the stochastic spectra. Finally, it is noted that for the pure shear and shear compression case, $\gamma =1.4$ is used whereas $\gamma =5/3$ is used for the axisymmetric compression. All cases use a constant initial mean temperature of $\bar {T}=300$ . The temperature is not assumed to remain constant throughout the simulation, and its change is captured through the change in pressure and speed of sound. The density and temperature fluctuations do not explicitly appear in the evolution equations and thus are not required to be explicitly initialised. However, they can be initialised through the isentropic relations between density, temperature and pressure.

5.2. Algorithm for stochastic representation

All the background necessary for a functional algorithm to evolve the turbulent statistics of compressible turbulence has been discussed in the prior sections. The complete algorithm is now presented in the following pseudo-code to further solidify methodology. Depending on the case, the following variables are known/chosen: the final non-dimensional time $St_f$ , the number of particle clusters $n_{p}$ , the number of wavevector modulus discretisations $n_k$ , the number of time steps $n_t$ , the initial mean speed of sound $a_0$ , the initial turbulent Mach number $M_{t0}$ , the initial distortion Mach number $M_{d0}$ , the initial turbulent Reynolds number $Re_{t0}$ , the initial solenoidal-to-dilatational turbulent kinetic energy ratio $q_s^2/q_d^2$ , the initial shear-to-compression ratio $S_0/C_0$ and the maximum and minimum wavevector magnitudes $\kappa _{{min}}, \kappa _{{max}}$ . The governing equations are solved numerically using finite time discretisation to advance the solution forward in time. We used an explicit fourth-order Runge–Kutta scheme.

Algorithm 1 Compressible PRM

5.3. Axial compression

The axial compression considered in this case is defined by

(5.9) \begin{align} G_{ij}=C\delta _{i1}\delta _{j1}, \quad C(t)=\frac {C_0}{1+C_0t}=\frac {C_0\rho }{\rho _0}, \end{align}

where the compression axis is in the $x_1$ direction (as shown in figure 2 a) and $C$ represents the compression rate. Due to the change in volume/density (varying from $\rho /\rho _0=1$ to $\rho /\rho _0=5$ ), the speed of sound also changes. The relationship between the speed of sound and density ratio is derived from isentropic relations (Mahesh Reference Mahesh1996, Jacquin et al. Reference Jacquin, Cambon and Blin1993):

Figure 2. Schematic of the mean velocity deformations for (a) axial compression, (b) pure shear and (c) sheared compression.

(5.10) \begin{align} a=a_0\sqrt {\frac {1+C_0 t}{(1+C_0t)^\gamma }}. \end{align}

As mentioned previously, the distortion Mach number $M_d$ is equal to the ratio between the initial acoustic time scale and the mean distortion or compression time scale. So if $M_d\ll 1$ , the acoustic time scale is much larger than the mean distortion time scale and a pure acoustic/solenoidal regime is obtained (Durbin & Zeman Reference Durbin and Zeman1992). In this solenoidal regime, the dilatational mode is balanced by the pressure gradient. The pressure is found with the pressure Poisson equation and consequentially describes the dilatational modes (Simone et al. Reference Simone, Coleman and Cambon1997). The energy is bounded in this regime and limits are well established (Durbin & Zeman Reference Durbin and Zeman1992; Jacquin et al. Reference Jacquin, Cambon and Blin1993; Simone et al. Reference Simone, Coleman and Cambon1997). On the other hand, if $M_d\gg 1$ , the dilatational mode is no longer constrained by the pressure (hence the name pressure-released) and is fed directly by compression. For the $M_d\gg 1$ pressure-released limit, the turbulent kinetic energy amplification can be expressed as (Jacquin et al. Reference Jacquin, Cambon and Blin1993)

(5.11) \begin{align} \frac {q^2(t)}{q^2(0)}=\frac {2+(\rho (t)/\rho (0))^2}{3}. \end{align}

For the solenoidal limit $M_d\ll 1$ , turbulent kinetic energy amplification is also found analytically using (Simone et al. Reference Simone, Coleman and Cambon1997)

(5.12) \begin{align} \frac {q^2(t)}{q^2(0)}=\frac {1}{2}\left (1+(\rho (t)/\rho (0))^2\frac {\tan ^{-1}\left (\sqrt {(\rho (t)/\rho (0))^2-1}\right )}{\sqrt {(\rho (t)/\rho (0))^2-1}}\right ). \end{align}

Results for the axial compression cases are compared with DNS data of Cambon et al. (Reference Cambon, Coleman and Mansour1993). The initial conditions used for the three intermediate cases are listed in table 1. Note that the distortion Mach number is defined with the negative compression rate for this problem, $M_d=-M_tCq^2/\epsilon$ . Unless otherwise stated, results presented use 800 particles and 60 discretisations of the wavevector magnitude for a range of $10^{-3}$ to $10^{3}$ . The trapezoidal integration with respect to magnitude was used alongside Monte Carlo integration with respect to orientation. Figure 3 shows the histories of the turbulent kinetic energy amplification for cases A–C with different cluster amounts, varying from 200 to 800. With increasing cluster count, predictions tend to improve and agree closer with the DNS data. The optimal cluster count varies based on the distortion Mach number. We recommend a starting point of 500 clusters for most cases. Figure 3 also shows that increased compressibility ( $M_d$ ) leads to an increase in the kinetic energy amplification. Both solenoidal and dilatational contributions to the kinetic energy are shown in figure 4 and compared with DNS results. The solenoidal mode is nearly unaffected by the increased compression, which corroborates with the theory in Cambon et al. (Reference Cambon, Coleman and Mansour1993). A strong increase of the dilatational energy at the end of the compression is also captured by the CPRM.

Table 1. Initial conditions for axial compression cases.

Figure 3. Histories of turbulent kinetic energy amplification for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$ , black; $M_{d0}=29$ , blue; $M_{d0}=87$ , red; DNS of Cambon et al. (Reference Cambon, Coleman and Mansour1993), circles. Multiple predictions are shown for varying particle cluster counts varying from 200 to 800 clusters.

Figure 4. Evolution of solenoidal and dilatational components of turbulent kinetic energy amplification for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$ , black; $M_{d0}=29$ , blue; $M_{d0}=87$ , red; DNS of Cambon et al. (Reference Cambon, Coleman and Mansour1993), circles.

Figure 5. Evolution of solenoidal and dilatational components of the $D_{11}/q^2$ structure tensor component for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$ , black; $M_{d0}=29$ , blue; $M_{d0}=87$ , red; DNS of Cambon et al. (Reference Cambon, Coleman and Mansour1993), circles.

The dimensionality structure tensor $D_{11}$ component is also computed and compared with DNS data. Figure 5 shows the histories of normalised solenoidal and dilatational components of $D_{11}$ and matches the trends of the solenoidal/dilatational components of the DNS data. Note that for axial compression, $D^d_{11}=R_{11}^d$ (Cambon et al. Reference Cambon, Coleman and Mansour1993). The dimensionality provides information on the angular distribution of the spectral energy. In the pressure-released case, the alignment of the wavevector with the compression direction increases $D_{11}/q^2$ towards one, while the angular distribution of the spectral energy remains unchanged. Meanwhile, in the solenoidal limit, the angular distribution of spectral energy in the plane normal to the compression direction opposes the wavevector motion so that a slower positive increase of $D_{11}/q^2$ is displayed. The $M_{d0}=5$ case has poor $D_{11}^d/q^2_d$ prediction, showing that the CPRM does not capture this spectral distribution opposing the wavevector motion. Reasons for this disagreement include wavevector discretisation errors and the resampling technique inaccuracy for the wavevector magnitude, which remains constant due to the infinite integration. In contrast, when evolving the wavevector magnitude, the dimensionality results are over-predicted, but the case near the solenoidal limit was much closer.

While the CPRM is expected to compare well with RDT, the results show some discrepancies between the high-fidelity data. One reason for this is that the comparing data in this section are from the DNS of Cambon et al. (Reference Cambon, Coleman and Mansour1993). We do not expect RDT to match exactly with DNS, and thus small disagreement is reasonable. Cambon et al. (Reference Cambon, Coleman and Mansour1993) also stated that the solenoidal ratio of $D_{11}/q^2$ predicted by DNS was found to be slightly lower than the RDT analytical prediction. This finding is also supported by the results for the solenoidal CPRM $D_{11}/q^2$ predictions. Finally, the CPRM requires different information for initialisation from what is given for DNS. The pressure–velocity spectrum $\beta _i$ was initialised to zero for this case, while the pressure spectrum $\Gamma$ was initialised with the dilatational energy. These initialisations are approximate as the energy spectra for the pressure–velocity and pressure spectra are not known. Therefore, since the evolution of the turbulence statistics depends on initial conditions, the inexact initial conditions contribute to the error seen in the results.

It is worth noting the differences and analogies of the current CPRM formulation to the traditional RDT used by Jacquin et al. (Reference Jacquin, Cambon and Blin1993). The CPRM uses the same set of governing equations and assumptions as Jacquin et al. (Reference Jacquin, Cambon and Blin1993), but ends up with a different and reduced set of evolution equations. This is because the CPRM only used deforming coordinates/Lagrangian perspective as the coordinate transformation whereas the RDT of Jacquin et al. (Reference Jacquin, Cambon and Blin1993) used deforming coordinates and a Craya–Herring transformation. The CPRM also solves for the full-field variables and does not require the Helmholtz decomposition as compared with the RDT which was developed using this decomposition. The normalised pressure variable $\hat {\wp }$ removes the explicit appearance of i in the evolution equation and so the CPRM can be solved without any need to handle imaginary components, whereas the simplified RDT equations presented by Jacquin et al. (Reference Jacquin, Cambon and Blin1993) did have an explicit imaginary number appear in the pressure terms for the general (equations (30) and (32)) but not in the reduced equations for the axial compression case. The last difference is the method of numerical integration. Both the CPRM and RDT of Jacquin et al. (Reference Jacquin, Cambon and Blin1993) require numerical integration, but the CPRM is solved using stochastic methods while the numerical integration method is not specified by Jacquin et al. (Reference Jacquin, Cambon and Blin1993), but likely used some form of fast Fourier transforms.

5.3.1. Axial compression: a surrogate for shock–turbulence interaction

Axial compression can be used as a simplified model for shock–turbulence interaction. Jacquin et al. (Reference Jacquin, Cambon and Blin1993) used this model problem to compare solenoidal and pressure-released limits against the LIA theory of Ribner (Reference Ribner1954). Homogeneous axially compressed turbulence is only a model of shock–turbulence interaction because true shock–turbulence interaction does not meet the quasi-homogeneous requirement that $\ell \ll \delta$ , where $\delta$ is the mean scale (the thickness of the shock). For further evaluation of the CPRM, we follow the canonical problem of Jacquin et al. (Reference Jacquin, Cambon and Blin1993), where an initial flow at Mach number $M_0$ impinges a shock of strength $\Delta \bar {u}/\bar {u}_0=(\rho -\rho _0)/\rho$ , such that the density ratio is given by

(5.13) \begin{align} \frac {\rho }{\rho _0}=\frac {(\gamma +1)M_0^2}{2+(\gamma -1)M_0^2}. \end{align}

Figure 6 shows the turbulent kinetic energy amplification for varying $M_0$ . The disagreement with LIA (computed by Jacquin et al. (Reference Jacquin, Cambon and Blin1993)) is significant, and follows the same results obtained by Jacquin et al. (Reference Jacquin, Cambon and Blin1993). The CPRM performance in figure 6 is identical to figure 3 because the problem dynamics is the same, just with increased compression rate $C$ (where increasing $C$ approaches the pressure-released solution). The CPRM does match the far-field LIA at very short times/low $M_0$ , but does not match the near-field LIA predictions.

Figure 6. Evolution of turbulent kinetic energy amplification for three intermediate cases in axial compression compared with LIA and limiting RDT: $M_{d0}=5$ , black; $M_{d0}=29$ , blue; $M_{d0}=87$ , red.

5.4. Pure shear

The deformation for pure shear is shown in figure 2(b). Pure shear is defined as

(5.14) \begin{align} G_{ij}=S\delta _{i1}\delta _{j2}, \quad\; S(t)=S_0, \end{align}

where $S_0$ is the initial shear rate, which remains constant. No volumetric compression is experienced ( $\rho /\rho _0=1$ ) and thus the speed of sound and mean pressure/density are constant in time: $a=a_0$ , $\bar {p}(t)=\bar {p}_0$ , $\bar {\rho }(t)=\bar {\rho }$ . For anisotropies $b_{ij}=R_{ij}/q^2-\delta _{ij}/3$ , the pressure-released limit is found by neglecting the pressure terms in the evolution equation for $\Phi ^{*}_{ij}$ . The evolution of the turbulent kinetic energy amplification in this limit is, however, found analytically when the initial turbulence conditions are isotropic (Simone et al. Reference Simone, Coleman and Cambon1997):

(5.15) \begin{align} \frac {q^2(t)}{q^2(0)}=1+\frac {(St)^2}{3}. \end{align}

Figure 7. Turbulent kinetic energy amplification evolution in pure shear for varying distortion Mach numbers prescribed in table 2. Solid line given by equation (5.15). Results are comparable with those of Bertsch et al. (Reference Bertsch, Suman and Girimaji2012) and Lavin et al. (Reference Lavin, Girimaji, Suman and Yu2012).

Simple analytical solutions only exist for the pressure-released limit. Solenoidal limits are evaluated by decreasing the distortion Mach number $M_d\ll 1$ and using the solenoidal form of the evolution equations for $\Phi ^s_{ij}$ . Cluster count varied from 500 near the solenoidal limit to 700 near the pressure-released limit. A 20-point trapezoidal integration was used for the wavevector magnitude. Due to instabilities at high $St$ values, the evolution of the wavevector magnitude was taken into account in the evolution equations for the spectra (no resampling used). A very fine time step was required at high $St$ values, reaching as low as $\Delta t/t_f=0.001$ .

Evolution of the turbulent kinetic energy amplification is presented in figure 7. As expected, the turbulent kinetic energy amplification is bounded by the pressure-released limit and increases with initial $M_d$ . Amplification increase is much greater than that of the DNS data from Simone et al. (Reference Simone, Coleman and Cambon1997). The early-time monotonic increase is predicted by the CPRM but the later-time linear behaviour does not match the DNS. This disagreement is likely due to the re-meshing technique used in the DNS that resulted in loss of turbulent kinetic energy and thus lower growth rates. However, the CPRM does predict the three-stage evolution of the turbulent kinetic energy amplification as described by Bertsch et al. (Reference Bertsch, Suman and Girimaji2012), Lavin et al. (Reference Lavin, Girimaji, Suman and Yu2012) and Kumar et al. (Reference Kumar, Bertsch and Girimaji2014). The intermediate flattening of the turbulent kinetic energy evolution at $St \approx 4{-}6$ is seen in figure 7. In the first stage, irrespective of the initial modal Mach number $M_m=S/a|{\boldsymbol{\kappa }}|$ , rapid growth of the turbulent kinetic energy is experienced and matches the rate of the pressure-released solution. This stage is defined by $St/\sqrt {M_m(0)}\lt 1$ , where the fluctuating pressure change is too slow to influence the velocity field, which grows unbounded. Transition from the first to the second stage begins when the pressure starts to influence the velocity field. Through this influence, the kinetic energy evolution begins to depart from the pressure-released solution. We see a sudden reduction/stabilisation in the kinetic energy amplification growth in figure 7 around $St\approx 1.5$ . The flattening of kinetic energy is bounded by $1\lt M_m(t)\lt \sqrt {M_m(0)}$ . The third stage of kinetic energy evolution begins where the kinetic energy grows rapidly again. This occurs when the acoustic time scale $a_0t|\boldsymbol{\kappa _0}|$ becomes smaller than the mean shear time scale $St$ . This occurs at approximately $a_0t|\boldsymbol{\kappa _0}| \approx 3$ . At this stage, the dilatational component of the flow is small and thus the turbulence is predominately governed by the solenoidal dynamics.

Another important feature of compressible pure shear is the ‘crossover point’ displayed in the evolution of the shear anisotropy component $b_{12}$ . A crossover point at $St\approx 4$ , where trends with $M_d$ reverse, is predicted by the CPRM in figure 8. Before the crossover, $b_{12}$ increases with $M_d$ but decreases soon afterwards. The rate of decrease depends on the distortion Mach number $M_d$ . The crossover occurs due to coupling between the solenoidal and dilatational modes (Simone et al. Reference Simone, Coleman and Cambon1997).

Figure 8. Comparison of shear anisotropy evolution in pure shear between the CPRM (a) and RDT (b) for varying distortion Mach numbers prescribed in table 2. The top solid line is the pressure-released limit while the lower solid line is the solenoidal limit.

Predictions in figure 8 do not match as well with the RDT data of Simone et al. (Reference Simone, Coleman and Cambon1997) at large $St$ . It is noted in Simone et al. (Reference Simone, Coleman and Cambon1997) that the ‘waviness’ at large times in the RDT histories is the result of the wavenumber discretisation and resolution at large wavevectors. Since the integration with respect to wavevector magnitude is done numerically, this also impacts the long-time solution of the CPRM and is likely the cause for the mismatch between the CPRM and RDT of Simone et al. (Reference Simone, Coleman and Cambon1997). The histories of the normal $b_{ij}$ components are shown in figure 9. The RDT/DNS data for these quantities are not available, so we simply compare with the pressure-released and solenoidal limits. For most of the range, the predictions stay bounded. Recall that analytical solutions for pressure-released and solenoidal anisotropy limits do not exist, so they are approximated by using large/small $M_d$ values with pressure terms ignored or using the solenoidal model. This approximation is a reason why it does not always bound the lower/higher $M_d$ cases. The oscillation due to wavenumber discretisation is also pronounced and causes the long-time solution to go past the limits.

The equipartition between the pressure fluctuations and dilatational component of the turbulent kinetic energy is also recorded at large mean shear time scales in figure 10. The equipartition variable is defined as

(5.16) \begin{align} \phi _p=\frac {\overline {u'_2u'_2}\bar {p}\gamma \bar {\rho }}{\overline {p'p'}}=\iiint ^{\infty }_{-\infty }\frac {\Phi _{22}}{a^2\Gamma }{\rm d}^{3}{\boldsymbol{\kappa }}. \end{align}

Bertsch et al. (Reference Bertsch, Suman and Girimaji2012) used this definition of equipartition between the pressure and dilatational velocity components because the flow-normal kinetic energy component is nearly equal to the dilatational kinetic energy of the flow field at late times due to the alignment of the wavenumber vector with the flow-normal direction in homogeneous shear. The evolution of the equipartition as predicted by the CPRM shows a similar sharp decay and oscillatory behaviour near $\phi _p=1$ as the RDT analysis done by Bertsch et al. (Reference Bertsch, Suman and Girimaji2012). Since the pressure dilatation is a driving mechanism for distribution of energy between the mechanical and internal modes, it is expected that the equipartition follows a similar trend to that of the three stages of the pressure correlation/turbulent kinetic energy evolution as described previously. Indeed, it is seen that the start of the plateau of turbulent kinetic energy coincides with the equipartition of energy between dilatation and pressure modes ( $St\approx 6$ ). The evolution of the pressure statistics $\overline {p'p'}/\bar {p}^2\gamma ^2$ is also verified by comparing with the RDT data of Lavin et al. (Reference Lavin, Girimaji, Suman and Yu2012) in figure 11.

Figure 9. Evolution of normal anisotropy components in pure shear for varying distortion Mach numbers prescribed in table 2: pressure-released/solenoidal limits, top/bottom solid lines; $b_{11}$ , red; $b_{22}$ , blue; $b_{33}$ , black.

Figure 10. Evolution of the equipartition variable $\phi _p$ for varying higher distortion Mach numbers prescribed in table 2. Relaxation to equipartition is obtained when the turbulent kinetic energy curve reaches the second stage.

Figure 11. Evolution of the normalised pressure statistics $\overline {p'p'}/\bar {p}^2\gamma ^2$ for cases A–J in table 2. Data match the trends seen in Lavin et al. (Reference Lavin, Girimaji, Suman and Yu2012).

Table 2. Homogeneous isotropic turbulence initial conditions for pure-shear cases.

We compare our model results with those of Yu & Girimaji (Reference Yu and Girimaji2007). For this comparison, the histories of $b_{12}$ are compared for the same range of $M_d$ given in table 2. From figure 12, it is clear that the model of Yu & Girimaji (Reference Yu and Girimaji2007) does predict the correct crossover point, but the small details like the decay afterwards or the magnitude of the crossover point peaks do not match the RDT data as well as the CPRM does. This is likely due to the fact that Yu & Girimaji (Reference Yu and Girimaji2007) did not account for the appearance of the wavevector magnitude in the evolution equations which requires one to evolve the velocity spectrum tensor rather than the directional velocity spectrum tensor. The appearance of the $\kappa ^{3}$ factor in the pressure terms after conversion to spherical coordinates renders it impossible to analytically factor out the directional velocity spectrum tensor. The new model also addresses a discrepancy in the past compressible RDT model where the definitions for second moments in Fourier space considered the entire complex expression, which does not align with the symmetric velocity spectrum tensor definition by Durbin & Pettersson-Reif (Reference Durbin and Pettersson-Reif2011) used to compute the Reynolds stress. This resulted in complex evolution equations where real and imaginary components needed to be treated.

While the CPRM does not match the RDT data exactly, its predictions are much improved compared with past models. Again, due to the formulation of the CPRM and the requirement of the $\beta _i$ and $\Gamma$ spectra, the initialisation is not exactly the same as that of the RDT of Simone et al. (Reference Simone, Coleman and Cambon1997) since we do not know the pressure–velocity and pressure energy spectra. This is one reason why the predictions do not match the RDT results exactly, along with possible discretisation errors. Note that the CPRM and the RDT of Simone et al. (Reference Simone, Coleman and Cambon1997) are still very similar methods, sharing the same differences and analogies as the comparisons made between the CPRM and the RDT of Jacquin et al. (Reference Jacquin, Cambon and Blin1993) discussed in the previous section.

The distribution of the directional velocity spectrum over the unit- ${\boldsymbol{\kappa }}$ sphere is visualised in figure 13. These distributions evolve from a homogeneous isotropic (uniform) distribution over the unit sphere. The magnitudes of $\Phi ^{\nearrow \kappa *}_{kk}$ and $\Phi ^{\nearrow \kappa *}_{12}$ tend to decrease with increasing $M_d$ . This matches the behaviour seen in figure 8. Another interesting feature is that the distribution patterns of $\Phi ^{\nearrow \kappa *}_{kk}$ and $\Phi ^{\nearrow \kappa *}_{12}$ are identical for the same case and only vary in magnitude range.

Figure 12. Evolution of shear anisotropy components in pure shear for varying distortion Mach numbers with the new CPRM (black) compared with the model of Yu & Girimaji (Reference Yu and Girimaji2007) (blue).

Figure 13. Visualisation of the directional velocity spectrum trace and shear component in pure shear for varying distortion Mach numbers prescribed in cases (a) A, (b) D, (c) G and (d) J. Results are shown at time $St=10$ .

5.5. Sheared compression

Sheared compression is a combination of the last two cases. It is defined by

(5.17) \begin{align} G_{ij}=C\delta _{i1}\delta _{j1}+S\delta _{i1}\delta _{j2}, \quad C(t)=\frac {C_0}{1+C_0t},\quad S(t)=\frac {S_0}{1+C_0t}. \end{align}

The shear and compression rates change with time to satisfy the compressible Euler equations. To match the results of Mahesh (Reference Mahesh1996), anisotropic initial spectra for the three second moments are required. This is done by applying pure shear for $t_{ps}$ time, defined by the non-dimensional time $\mathcal{S}_0=S_0t_{ps}$ . Note that for this problem, only solenoidal turbulence is considered. Thus the solenoidal model is used to evolve the directional velocity spectrum tensor $\Phi ^{\nearrow \kappa *}_{ij}$ . The initial shear is intended to generate the anisotropic field but is negligible compared with the applied compression, i.e. $S_0/C_0\ll 1$ . Thus a low shear-compression ratio of $S_0/C_0=0.1$ is used (Mahesh Reference Mahesh1996). Only $\mathcal{S}_0$ varies across the different cases, so all other initial variables remain the same. For this case, $M_{t0}=0.025$ , $Re_{t0}=358$ and $M_{d0}=0.5$ . Again, 800 clusters are used, and wavevector magnitude integration is not required since the directional spectrum is evolved. The histories of $b_{12}=R_{12}/q^2$ are plotted in figure 14.

Figure 14. Evolution of solenoidal component of $R_{12}/q^2$ for varying anisotropic initial condition in sheared compression: blue, $\mathcal{S}_0=1$ ; black, $\mathcal{S}_0=2$ ; red, $\mathcal{S}_0=3$ ; circle, RDT of Mahesh (Reference Mahesh1996).

Each curve has different initial total shear, which increases from red to blue. Agreement with the RDT of Mahesh (Reference Mahesh1996) shows that the CPRM solenoidal method works well for anisotropic initial conditions. Slight discrepancies are likely due to imperfect initialisation, noting that the worst results also have the worst initialisation (start off with the most error). The initial anisotropic velocity spectrum was not given by Mahesh (Reference Mahesh1996) and thus the CPRM was used to obtain the initial spectrum by running the pure shear case for $\mathcal{S}_0$ total shear. Therefore, we are unable to match the initial spectrum exactly to that of Mahesh (Reference Mahesh1996). Figure 14 displays important trends of decreasing shear anisotropy and change of sign for the case with sufficiently large total volumetric strain (which correlates with larger initial total shear).

Figure 15. Evolution of solenoidal component of turbulent kinetic energy for varying anisotropic initial condition in sheared compression: blue, $\mathcal{S}_0=1$ ; black, $\mathcal{S}_0=2$ ; red, $\mathcal{S}_0=3$ ; circles, RDT of Mahesh (Reference Mahesh1996).

Figure 16. Evolution of normal components of $R_{ij}/q^2$ in sheared compression for anisotropic initial condition $\mathcal{S}_0=3$ : $R_{11}/q^2$ , blue; $R_{22}/q^2$ , black; $R_{33}/q^2$ , red; RDT of Mahesh (Reference Mahesh1996), circles.

Figure 15 shows the turbulent kinetic energy amplification for the same cases. Again, due to inexact initialisation, the turbulent kinetic energy amplification is also not exactly predicted but still has the same increasing trends for the three different total shear values. Figure 16 shows the normal components of $R_{ij}/q^2$ compared against the RDT of Mahesh (Reference Mahesh1996) and shows nearly exact agreement, where the compression increases the contribution of $u_1'$ to the turbulent kinetic energy and decreases that of $u'_2$ and $u_{3}^{\prime}$ . Only results for the maximum initial total shear $\mathcal{S}_0=3$ are given in figure 16.

5.6. Plane strain and axisymmetric contraction

The CPRM is applicable to any generic mean deformations that are rapid enough such that RDT still applies. To demonstrate this, the CPRM is applied to plane strain and axisymmetric deformations, defined by

(5.18) \begin{align} \frac {\partial \bar {u}_i}{\partial x_{\!j}}_{ps} = \left [ \begin{array}{ccc} S & \ \ 0 \ \ & 0 \\ 0 & \ \ -S \ \ & 0 \\ 0 & \ \ 0 \ \ & 0 \\ \end{array} \right ],\quad \frac {\partial \bar {u}_i}{\partial x_{\!j}}_{axc} = \left [ \begin{array}{ccc} S & \ \ 0 \ \ & 0 \\ 0 & \ \ -S/2 \ \ & 0 \\ 0 & \ \ 0 & \ \ -S/2 \\ \end{array} \right ]. \end{align}

For this problem, DNS/RDT datasets are not available and thus we cannot determine the model’s accuracy. Instead, we simply verify that the model is performing adequately by comparing predictions with the approximate solenoidal and pressure-released limits. For all cases, 800 clusters were used along with a 20-weight Gauss–Laguerre quadrature. Figure 17(a) shows the normal $b_{ij}$ components for plane strain, while figure 17(b) shows $b_{ij}$ for axisymmetric contraction. The results fall within the limits, showing that the the qualitative behaviour of the CPRM is correct. It is important to note that while cases A and J are near the limits, they are not so small/large that we expect them to follow the limiting lines exactly. The results show that the model can be applied to generic rapid deformations, although more high-fidelity data are needed to evaluate the accuracy of the CPRM.

Figure 17. Evolution of normal components of $b_{ij}$ subject to (a) plane strain and (b) axisymmetric contraction for initial conditions A, D, G and J listed in table 2. Same legend as in figure 9.

6. Concluding remarks

We developed a new stochastic wavevector model for rapidly distorted compressible turbulence subject to the restrictions of homogeneity and small density fluctuations. By representing the evolution equations in spectral space and evolving the pressure–velocity and pressure spectra $\beta _i$ and $\Gamma$ , the velocity spectrum tensor evolution equation was closed without any modelling assumptions. This preserves the exact nature of RDT and only requires approximation of the spectral-to-physical transformation integrals. A new variable called the normalised pressure fluctuation $\wp$ enabled the formulation of $\beta _i$ and $\Gamma$ , and simplified the second-moment evolution equations by removing explicit imaginary terms.

Motivation for a stochastic representation was given by considering the use of Monte Carlo integration to transform to physical space. The stochastic model uses stochastic samples or virtual particles to represent realisations of the random variables $u^{*}_i$ , $\kappa ^{*}_i$ and $\wp ^{*}$ , which were then used to compute the second-moment statistics $\Phi _{ij}$ , $\beta _i$ and $\Gamma$ . Consistency between the Fourier and stochastic representation was shown by comparing the Eulerian velocity spectrum evolution equations derived from the linearised stochastic PDF equations and traditional RDT. The idea of sample/particle clustering with respect to wavevector orientation was also motivated and described, where it became advantageous when a non-stochastic method of integration for the infinite integral of wavevector magnitude was used. By working in spectral space, the computation of higher-order tensors was also possible due to the availability of the wavevector information. These higher-order tensors provide more information on the turbulent field and can be used to develop improved models for nonlinear or inhomogeneous problems.

Next, the new model was applied to three homogeneous compressible problems: axially compressed stationary and shear turbulence, and pure shear deformations. Results were generally in agreement with the available DNS/RDT data. The axial compression case was discussed as an idealised shock–turbulence interaction model, where dilatational and solenoidal turbulent kinetic energy amplification histories were shown to be in agreement with DNS data. Evolution of the turbulent kinetic energy amplification was also evaluated for different cluster amounts and showed a converging trend towards true DNS solutions. The pure shear case demonstrated the applicability and performance of the CPRM across a range of distortion Mach numbers $M_d$ with improvement in the prediction of the shear anisotropy evolution compared with the compressible RDT model of Yu & Girimaji (Reference Yu and Girimaji2007). Equipartition of pressure fluctuations and dilatational velocity fluctuations was also observed for long times, along with the three phases of turbulent kinetic energy evolution. The sheared compression case showed that the model is able to predict stresses/amplification accurately when an anisotropic initial shear is axially compressed. All three cases, however, were found to be sensitive to the initial conditions, which could not be matched exactly.

Finally, since the formulation of the model made no assumption on the form of the deformation tensor, it can be applied to any generic rapid mean deformation. Plane strain and axisymmetric contraction deformations were applied to verify that the resulting Reynolds stress anisotropy histories for four different distortion Mach numbers were bounded within the solenoidal (small $M_{d0}$ ) and pressure-released (large $M_{d0}$ ) limits. Initialisation methods were discussed and developed for the velocity spectrum tensor.

We emphasise that this work is merely a starting point for the development of practical turbulence models that maintain consistency in the rapid-distortion regime. The exact limiting behaviour overcomes the limitations of eddy-viscosity and other standard assumptions, and enables the computation of higher-order tensors to bring in more information on the physics of turbulence. Extensions to general deformation and inhomogeneities are the next steps towards development of the CPRM for use in more complex problems. There are several ways of accounting for nonlinear effects in the current formulation. A limiting model can be created similar to the nonlinear incompressible PRM of Kassinos & Reynolds (Reference Kassinos and Reynolds1996), or more sophisticated approaches can be used to directly model the multi-point triadic interactions such as eddy-damped quasi-normal Markovian models (Orszag Reference Orszag1970). Application of the stochastic wavevector formulation to inhomogeneous compressible turbulence can be accomplished in two ways and will be explored in future work. Although the exact equations presented as equations (3.14)–(3.19) cannot be used directly in inhomogeneous cases, due to the requirement of periodicity for Fourier bases, a similar set of equations can be derived by using a non-periodic basis such as Chebyshev polynomials (Canuto et al. Reference Canuto, Hussaini, Quarteroni and Zang1988). This will require a new derivation of the basic formulation, but the use of a spectral representation should still allow exact closure of the linearised equations since the differentiation properties of the Chebyshev polynomials are similar to the Fourier basis (spatial derivatives become point-wise wavevector-basis coefficient products). Alternatively, the CPRM can be extended to inhomogeneous flows using a quasi-homogeneous approach similar to the elliptic-relaxation technique developed by Durbin (Reference Durbin1991). This approach is more direct in that it represents inhomogeneities without requiring the original tensor evolution equations to be modified. This approach was used by Kassinos et al. (Reference Kassinos, Langer, Haire and Reynolds2000) to extend the incompressible PRM to wall-bounded flows.

Funding

N.Z. was supported by the National Science Foundation Graduate Research Fellowship Program. K.D. was supported by OUSD(RE) grant no. N00014-21-1-295.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Solenoidal and dilatational decomposition

The solenoidal–dilatational velocity decomposition $\hat {u}_i=\hat {u}_i^s+\hat {u}_i^d$ is used to define the solenoidal, dilatational and solenoidal–dilatational cross-correlations:

(A1) \begin{align} \Phi ^s_{ij} =\frac {1}{2} \left (\overline { \hat {u}^s_i\hat {u}^{s+}_{\!j}}+\overline { \hat {u}^{s+}_i\hat {u}^s_{\!j}}\right ), \!\!\quad\! \Phi ^{d}_{ij} =\frac {1}{2} \left (\overline { \hat {u}^d_i\hat {u}^{d+}_{\!j}}+\overline { \hat {u}^{d+}_i\hat {u}^d_{\!j}}\right ), \!\!\quad\! \Phi ^{sd}_{ij} =\frac {1}{2} \left (\overline { \hat {u}^s_i\hat {u}^{d+}_{\!j}}+\overline { \hat {u}^{s+}_i\hat {u}^d_{\!j}}\right ), \end{align}

(A2) \begin{align} \beta ^s_{i}=\frac {1}{2}\left (\overline {\hat {\wp }^+\hat {u}^s_i} + \overline {\hat {\wp }\hat {u}_i^{s+}}\right ),\quad \beta ^d_{i}=\frac {1}{2}\left (\overline {\hat {\wp }^+\hat {u}^d_i} + \overline {\hat {\wp }\hat {u}_i^{d+}}\right ). \end{align}

Note that $\Phi ^{sd}_{ij}\neq \Phi ^{ds}_{ij}$ and, in general, $\Phi ^{sd}_{ij}$ are not symmetric (unlike $\Phi ^s_{ij}$ and $\Phi ^d_{ij}$ ). The velocity spectrum solenoidal/dilatational components are related by

(A3) \begin{align} \Phi _{ij}=\Phi ^s_{ij}+\Phi ^d_{ij}+\Phi ^{sd}_{ij}+\Phi ^{ds}_{ij}. \end{align}

When turbulence is solenoidal, there is no need to introduce $\hat {\wp }$ because the imaginary pressure term from equation (3.5) is removed by substituting in the pressure Poisson equation (3.8). The dilatational velocity evolution is determined once ${\rm D}\hat {u}_i/{\rm D}t$ and ${\rm D}\hat {u}_i^s/{\rm D}t$ are known:

(A4) \begin{align} \frac {{\rm D}\hat {u}^s_i}{{\rm D}t}=-\frac {\partial \bar {u}_i}{\partial x_k}\hat {u}^s_k+2\hat {u}^s_{\!j}\frac {\kappa _i\kappa _k}{\kappa ^2}\frac {\partial \bar {u}_k}{\partial x_{\!j}}, \end{align}

(A5) \begin{align} \frac {{\rm D}\hat {u}^d_i}{{\rm D}t}=-\frac {\partial \bar {u}_i}{\partial x_k}\hat {u}^d_k+\kappa _ia^2\hat {\wp }-2\hat {u}^s_{\!j}\frac {\kappa _i\kappa _k}{\kappa ^2}\frac {\partial \bar {u}_k}{\partial x_{\!j}}, \end{align}

The second-moment evolution equations for the solenoidal, dilatational and cross solenoidal–dilatational velocity spectra are now derived as

(A6) \begin{align} \frac {\rm D}{{\rm D}t}(\Phi _{ij}^s)=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}^s-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ki}^s+2\frac {\partial \bar {u}_k}{\partial x_m}\left (\Phi _{im}^s\frac {\kappa _k\kappa _{\!j}}{\kappa ^2}+\Phi _{jm}^s\frac {\kappa _k\kappa _i}{\kappa ^2}\right ), \end{align}

(A7) \begin{align} \frac {\rm D}{{\rm D}t}(\Phi _{ij}^d)=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}^d-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ki}^d+a^2\left (\kappa _i\beta ^d_{\!j}+\kappa _{\!j}\beta ^d_i\right )-2\frac {\partial \bar {u}_k}{\partial x_m}\frac {\kappa _k}{\kappa ^2}\left (\Phi _{mi}^{sd}\kappa _{\!j}+\Phi _{mj}^{sd}\kappa _i\right ), \end{align}

(A8) \begin{align} \frac {\rm D}{{\rm D}t}(\Phi _{ij}^{sd})=-\frac {\partial \bar {u}_i}{\partial x_k}\Phi _{kj}^{sd}-\frac {\partial \bar {u}_{\!j}}{\partial x_k}\Phi _{ik}^{sd}+\kappa _{\!j}a^2\beta ^s_i+2\frac {\partial \bar {u}_k}{\partial x_m}\frac {\kappa _k}{\kappa ^2}\left (\Phi _{mj}^{sd}\kappa _i-\Phi _{mi}^{s}\kappa _{\!j}\right ), \end{align}

(A9) \begin{align} \frac {\rm D}{{\rm D}t}(\beta ^s_i)=-\kappa _k\Phi ^s_{ki}-\kappa _k\Phi ^{sd}_{ik}-\frac {\partial \bar {u}_i}{\partial x_k}\beta ^s_{k}+2\beta ^s_{\!j}\frac {\partial \bar {u}_k}{\partial x_{\!j}}\frac {\kappa _i\kappa _k}{\kappa ^2}, \end{align}

(A10) \begin{align} \frac {\rm D}{{\rm D}t}(\beta ^d_i)=-\kappa _k\Phi ^{sd}_{ki}-\kappa _k\Phi ^{d}_{ki}-\frac {\partial \bar {u}_i}{\partial x_k}\beta ^d_{k}-2\beta ^s_{\!j}\frac {\partial \bar {u}_k}{\partial x_{\!j}}\frac {\kappa _i\kappa _k}{\kappa ^2}+\kappa _ia^2\Gamma . \end{align}

Note that in the solenoidal velocity spectrum tensor, the dependence on wavevector magnitude is removed. The solenoidal velocity spectrum tensor evolution equation is identical to the evolution equation in the incompressible formulation of Kassinos & Reynolds (Reference Kassinos and Reynolds1994). This is because the divergence-free property of the solenoidal modes allows the use of the pressure Poisson equation, the mean compression effects on the fluctuating density are neglected and the mean density does not vary in space (to be homogeneous). This reinforces the observation in other studies (Cambon et al. Reference Cambon, Coleman and Mansour1993) where the mean compressibility has little to no effect on the solenoidal velocity components.

Now, the initialisation of a turbulence field that contains both solenoidal and dilatational components is described. Since equation (5.1) initialises the solenoidal field, the dilatational field must be initialised separately. If the solenoidal–dilatational cross-correlation $\Phi ^{sd}_{ij}$ is assumed to be zero, the relation $\Phi _{ij}\approx \Phi ^s_{ij}+\Phi ^d_{ij}, q^2=q^2_s+q^2_d$ is satisfied. Since $q_s^2/q_d^2$ is specified at the start, then the dilatational field is found using the same method as described by Simone et al. (Reference Simone, Coleman and Cambon1997):

(A11) \begin{align} \Phi ^d_{ij}({\boldsymbol{\kappa }}^n)=\frac {E_d(\kappa ^n)}{2\pi (\kappa ^n)^2}. \end{align}

References

Batchelor, G.K. & Proudman, I. 1954 The effect of rapid distortion of a fluid in turbulent motion. Q. J. Mech. Appl. Maths 7 (1), 83103.CrossRefGoogle Scholar
Bertsch, R.L., Suman, S. & Girimaji, S.S. 2012 Rapid distortion analysis of high mach number homogeneous shear flows: characterization of flow-thermodynamics interaction regimes. Phys. Fluids 24 (12), 125106.CrossRefGoogle Scholar
Blaisdell, G.A. 1991 Numerical simulation of compressible homogeneous turbulence. PhD thesis, Stanford University, USA.Google Scholar
Caflisch, R.E. 1998 Monte Carlo and Quasi-Monte Carlo Methods. Cambridge University Press.Google Scholar
Cambon, C., Coleman, G.N. & Mansour, N.N. 1993 Rapid distortion analysis and direct simulation of compressible homogeneous turbulence at finite mach number. J. Fluid Mech. 257, 641665.CrossRefGoogle Scholar
Cambon, C., Godeferd, F.S., Nicolleau, F.C.G.A. & Vassilicos, J.C. 2004 Turbulent diffusion in rapidly rotating flows with and without stable stratification. J. Fluid Mech. 499, 231255.CrossRefGoogle Scholar
Canuto, C., Hussaini, M.Y., Quarteroni, A. & Zang, T.A. 1988 Spectral Methods in Fluid Dynamics. Springer-Verlagg.CrossRefGoogle Scholar
Crow, S.C. 1968 Viscoelastic properties of fine-grained incompressible turbulence. J. Fluid Mech. 33 (1), 120.CrossRefGoogle Scholar
Durbin, P.A. 1991 Near-wall turbulence closure modeling without “damping functions”. Theor. Comput. Fluid Dyn. 3 (1), 113.CrossRefGoogle Scholar
Durbin, P.A. & Pettersson-Reif, B.A. 2011 Statistical Theory and Modeling for Turbulent Flows. Wiley.Google Scholar
Durbin, P.A. & Zeman, O. 1992 Rapid distortion theory for homogeneous compressed turbulence with application to modelling. J. Fluid Mech. 242, 349370.CrossRefGoogle Scholar
Favier, B., Godeferd, F.S. & Cambon, C. 2010 On space and time correlations of isotropic and rotating turbulence. Phys. Fluids 22 (1), 015101.CrossRefGoogle Scholar
Fung, J.C.H. 1990 Kinematic simulation of turbulent flow and particle motions. PhD thesis, University of Cambridge, USA.Google Scholar
Goldstein, M.E. 1978 Unsteady vortical and entropic distortions of potential flows round arbitrary obstacles. J. Fluid Mech. 89 (3), 433468.CrossRefGoogle Scholar
Heinz, S. 2003 Statistical Mechanics of Turbulent Flows. Springer.CrossRefGoogle Scholar
Jacquin, L., Cambon, C. & Blin, E. 1993 Turbulence amplification by a shock wave and rapid distortion theory. Phys. Fluids 5 (10), 25392550.CrossRefGoogle Scholar
Kassinos, S.C., Langer, C.A., Haire, S.L. & Reynolds, W.C. 2000 Structure-based turbulence modeling for wall-bounded flows. Intl J. Heat Fluid Flow 21 (5), 599605.CrossRefGoogle Scholar
Kassinos, S.C. & Reynolds, W.C. 1994 A structure-based model for the rapid distortion of homogeneous turbulence. PhD thesis, Stanford University, USA.Google Scholar
Kassinos, S.C. & Reynolds, W.C. 1996 A particle representation model for the deformation of homogeneous turbulence. Annu. Res. Briefs>, (Center for Turbulence Research, Stanford University, 1996), pp. 3161.,+(Center+for+Turbulence+Research,+Stanford+University,+1996),+pp.+31–61.>Google Scholar
Kumar, G., Bertsch, R.L. & Girimaji, S.S. 2014 Stabilizing action of pressure in homogeneous compressible shear flows: effect of mach number and perturbation obliqueness. J. Fluid Mech. 760, 540566.CrossRefGoogle Scholar
Lavin, T.A., Girimaji, S.S., Suman, S. & Yu, H. 2012 Flow-thermodynamics interactions in rapidly-sheared compressible turbulence. Theor. Comput Fluid Dyn. 26 (6), 501522.CrossRefGoogle Scholar
Mahesh, K. 1996 The interaction of a shock wave with a turbulent shear flow. PhD thesis, Stanford University, USA.Google Scholar
Orszag, S.A. 1970 Analytical theories of turbulence. J. Fluid Mech. 41 (2), 363386.CrossRefGoogle Scholar
Pope, S.B. 2001 Turbulent Flows. Cambridge University Press.Google Scholar
Reynolds, W.C. 1989 Effects of rotation on homogeneous turbulence. In Proceeding of the Tenth Australasian Fluid Mechanics Conference, vol. 1, pp. KS2-1, University of Melbourne.Google Scholar
Ribner, H.S. 1954 Shock-Turbulence Interaction and the Generation of Noise. NACA Tech Note.Google Scholar
Rogallo, R.S. 1981 Numerical experiments in homogeneous turbulence. Tech. Rep. 81315, NASA Tech. Mem.Google Scholar
Roy, S., Pathak, U. & Sinha, K. 2018 Variable turbulent prandtl number model for shock/boundary-layer interaction. AIAA J. 56 (1), 342355.CrossRefGoogle Scholar
Sagaut, P. & Cambon, C. 2018 Homogeneous Turbulence Dynamics. Springer.CrossRefGoogle Scholar
Sandoval, D.L. 1995 The dynamics of variable-density turbulence. PhD thesis, University of Washington, USA.Google Scholar
Sarkar, S. 1995 The stabilizing effect of compressibility in turbulent shear flow. J. Fluid Mech. 282, 163186.CrossRefGoogle Scholar
Sarkar, S., Erlebacher, G., Hussaini, Y.M. & Kreiss, H.O. 1991 The analysis and modelling of dilatational terms in compressible turbulence. J. Fluid Mech. 227, 473493.CrossRefGoogle Scholar
Simone, A., Coleman, G.N. & Cambon, C. 1997 The effect of compressibility on turbulent shear flow: a rapid-distortion-theory and direct-numerical-simulation study. J. Fluid Mech. 330, 307338.CrossRefGoogle Scholar
Taylor, G.I. 1935 Turbulence in a contracting stream. Z. Angew. Math. Mech. 15 (1-2), 9196.CrossRefGoogle Scholar
Van Slooten, P.R. & Pope, S.B. 1997 Pdf modeling for inhomogeneous turbulence with exact representation of rapid distortions. Phys. Fluids 9 (4), 10851105.CrossRefGoogle Scholar
Wilcox, D.C. 1998 Turbulence Modeling for CFD. DCW industries.Google Scholar
Yu, H. & Girimaji, S.S. 2007 Extension of compressible ideal-gas rapid distortion theory to general mean velocity gradients. Phys. Fluids 19 (4), 041702.CrossRefGoogle Scholar
Figure 0

Figure 1. Particle properties associated with a cluster of samples.

Figure 1

Algorithm 1 Compressible PRM

Figure 2

Figure 2. Schematic of the mean velocity deformations for (a) axial compression, (b) pure shear and (c) sheared compression.

Figure 3

Table 1. Initial conditions for axial compression cases.

Figure 4

Figure 3. Histories of turbulent kinetic energy amplification for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$, black; $M_{d0}=29$, blue; $M_{d0}=87$, red; DNS of Cambon et al. (1993), circles. Multiple predictions are shown for varying particle cluster counts varying from 200 to 800 clusters.

Figure 5

Figure 4. Evolution of solenoidal and dilatational components of turbulent kinetic energy amplification for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$, black; $M_{d0}=29$, blue; $M_{d0}=87$, red; DNS of Cambon et al. (1993), circles.

Figure 6

Figure 5. Evolution of solenoidal and dilatational components of the $D_{11}/q^2$ structure tensor component for three cases of varying distortion Mach number in axial compression: $M_{d0}=5$, black; $M_{d0}=29$, blue; $M_{d0}=87$, red; DNS of Cambon et al. (1993), circles.

Figure 7

Figure 6. Evolution of turbulent kinetic energy amplification for three intermediate cases in axial compression compared with LIA and limiting RDT: $M_{d0}=5$, black; $M_{d0}=29$, blue; $M_{d0}=87$, red.

Figure 8

Figure 7. Turbulent kinetic energy amplification evolution in pure shear for varying distortion Mach numbers prescribed in table 2. Solid line given by equation (5.15). Results are comparable with those of Bertsch et al. (2012) and Lavin et al. (2012).

Figure 9

Figure 8. Comparison of shear anisotropy evolution in pure shear between the CPRM (a) and RDT (b) for varying distortion Mach numbers prescribed in table 2. The top solid line is the pressure-released limit while the lower solid line is the solenoidal limit.

Figure 10

Figure 9. Evolution of normal anisotropy components in pure shear for varying distortion Mach numbers prescribed in table 2: pressure-released/solenoidal limits, top/bottom solid lines; $b_{11}$, red; $b_{22}$, blue; $b_{33}$, black.

Figure 11

Figure 10. Evolution of the equipartition variable $\phi _p$ for varying higher distortion Mach numbers prescribed in table 2. Relaxation to equipartition is obtained when the turbulent kinetic energy curve reaches the second stage.

Figure 12

Figure 11. Evolution of the normalised pressure statistics $\overline {p'p'}/\bar {p}^2\gamma ^2$ for cases A–J in table 2. Data match the trends seen in Lavin et al. (2012).

Figure 13

Table 2. Homogeneous isotropic turbulence initial conditions for pure-shear cases.

Figure 14

Figure 12. Evolution of shear anisotropy components in pure shear for varying distortion Mach numbers with the new CPRM (black) compared with the model of Yu & Girimaji (2007) (blue).

Figure 15

Figure 13. Visualisation of the directional velocity spectrum trace and shear component in pure shear for varying distortion Mach numbers prescribed in cases (a) A, (b) D, (c) G and (d) J. Results are shown at time $St=10$.

Figure 16

Figure 14. Evolution of solenoidal component of $R_{12}/q^2$ for varying anisotropic initial condition in sheared compression: blue, $\mathcal{S}_0=1$; black, $\mathcal{S}_0=2$; red, $\mathcal{S}_0=3$; circle, RDT of Mahesh (1996).

Figure 17

Figure 15. Evolution of solenoidal component of turbulent kinetic energy for varying anisotropic initial condition in sheared compression: blue, $\mathcal{S}_0=1$; black, $\mathcal{S}_0=2$; red, $\mathcal{S}_0=3$; circles, RDT of Mahesh (1996).

Figure 18

Figure 16. Evolution of normal components of $R_{ij}/q^2$ in sheared compression for anisotropic initial condition $\mathcal{S}_0=3$: $R_{11}/q^2$, blue; $R_{22}/q^2$, black; $R_{33}/q^2$, red; RDT of Mahesh (1996), circles.

Figure 19

Figure 17. Evolution of normal components of $b_{ij}$ subject to (a) plane strain and (b) axisymmetric contraction for initial conditions A, D, G and J listed in table 2. Same legend as in figure 9.