Skip to main content
Log in

A Stochastic Backscatter Model for Grey-Area Mitigation in Detached Eddy Simulations

  • Published:
Flow, Turbulence and Combustion Aims and scope Submit manuscript

An Erratum to this article was published on 30 May 2017

Abstract

A new stochastic backscatter model is proposed for detached eddy simulations that accelerates the development of resolved turbulence in free shear layers. As a result, the model significantly reduces so-called grey areas in which resolved turbulence is lacking after the computation has switched from a Reynolds-averaged Navier–Stokes simulation to a large eddy simulation. The new stochastic model adds stochastic forcing to the momentum equations with a rate of backscatter from the subgrid to the resolved scales that is consistent with theory. The effectiveness of the stochastic model is enhanced by including spatial and temporal correlations of the stochastic forcing for scales smaller than the cut-off scale. The grey-area mitigation is demonstrated for two canonical test cases: the plane free shear layer and the round jet.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  1. Spalart, P. R., Jou, W. H., Strelets, M., Allmaras, S. R.: Comments on the feasibility of LES for wings, and on a hybrid RANS/LES approach. In: Liu, C., Liu, Z. (eds.) Advances in DNS/LES. Greyden Press (1997)

  2. Spalart, P. R., Deck, S., Shur, M. L., Squires, K. D., Strelets, M. K., Travin, A.: A new version of detached-eddy simulation, resistant to ambiguous grid densities. Theor. Comput. Fluid Dyn. 20, 181–195 (2006)

    Article  MATH  Google Scholar 

  3. Shur, M. L., Spalart, P. R., Strelets, M. K., Travin, A. K.: A hybrid RANS–LES approach with delayed-DES and wall-modelled LES capabilities. Int. J. Heat Fluid Flow 29, 1638–1649 (2008)

    Article  Google Scholar 

  4. Travin, A. K., Shur, M. L., Spalart, P. R., Strelets, M. K.: Improvement of delayed detached-eddy simulation for LES with wall modelling. In: Wesseling, P., Oñate, E., Périaux, J. (eds.) ECCOMAS CFD. Egmond aan Zee, The Netherlands (2006)

    Google Scholar 

  5. Deck, S.: Zonal-detached eddy simulation of the flow around a high-lift configuration. AIAA J. 43, 2372–2384 (2005)

    Article  Google Scholar 

  6. Kok, J.C., Dol, H.S., Oskam, B., van der Ven, H.: Extra-large eddy simulation of massively separated flows. In: 42nd AIAA Aerospace Sciences Meeting. Reno. AIAA paper 2004-264 (2004)

  7. Kok, J.C., van der Ven, H.: Destabilizing free shear layers in X-LES using a stochastic subgrid-scale model. In: Peng, S.H., Doerffer, P., Haase, W. (eds.) Progress in Hybrid RANS–LES Modelling, Notes on Numerical Fluid Mechanics and Multidisciplinary Design. NLR-TP-2009-327, vol. 111, pp 179–189. Springer, (2009)

  8. Spalart, P. R.: Detached-eddy simulation. Annu. Rev. Fluid Mech. 41, 181–202 (2009)

  9. Adamian, D., Travin, A.: An efficient generator of synthetic turbulence at RANS–LES interface in embedded LES of wall-bounded and free shear flows. In: ICCFD6. St. Petersburg (2010)

  10. Deck, S.: Recent improvements in the zonal detached eddy simulation (ZDES) formulation. Theor. Comput. Fluid Dyn.. doi:10.1007/s00162-011-0240-z(2011)

  11. Jarrin, N., Prosser, R., Uribe, J. C., Benhamadouche, S., Laurence, D.: Reconstruction of turbulent fluctuations for hybrid RANS/LES simulations using a synthetic-eddy method. Int. J. Heat Fluid Flow 30, 435–442 (2009)

    Article  Google Scholar 

  12. Chauvet, N., Deck, S., Jacquin, L.: Zonal-detached eddy simulation of a controlled propulsive jet. AIAA J. 45(10), 2458–2473 (2007)

    Article  Google Scholar 

  13. Shur, M. L., Spalart, P. R., Strelets, M. K., Travin, A. K.: An enhanced version of DES with rapid transition from RANS to LES in separated flows. Flow Turb. Combust. 95(4), 709–737 (2015)

    Article  Google Scholar 

  14. Kok, J.C., van der Ven, H.: Capturing free shear layers in hybrid RANS–LES simulations of separated flow. In: Third Symposium ‘Simulation of Wing and Nacelle Stall’. Braunschweig. http://hdl.handle.net/10921/914. NLR-TP-2012-333 (2012)

  15. Nicoud, F., Ducros, F.: Subgrid-scale stress modelling based on the square of the velocity gradient tensor. Flow Turb. Combust. 62(3), 183–200 (1999)

    Article  MATH  Google Scholar 

  16. Nicoud, F., Toda, H. B., Cabrit, O., Bose, S., Lee, J.: Using singular values to build a subgrid-scale model for large eddy simulations. Phys. Fluids 23(085), 106 (2011)

    Google Scholar 

  17. Mockett, C., Fuchs, M., Garbaruk, A., Shur, M., Spalart, P., Strelets, M., Thiele, F., Travin, A.: Two non-zonal approaches to accelerate RANS to LES transition of free shear layers in DES. In: Girimaji, S., Haase, W., Peng, S.H., Schwamborn, D. (eds.) Progress in Hybrid RANS–LES Modelling, Notes on Numerical Fluid Mechanics and Multidisciplinary Design, vol. 130, pp 187–201. Springer, (2015)

  18. Leith, C. E.: Stochastic backscatter in a subgrid-scale model: Plane shear mixing layer. Phys. Fluids A 2(3), 297–299 (1990)

    Article  Google Scholar 

  19. Schumann, U.: Stochastic backscatter of turbulence energy and scalar variance by random subgrid-scale fluxes. Proc. R. Soc. Lond. Series A 451, 293–318 (1995)

    Article  MATH  Google Scholar 

  20. Lesieur, M.: Turbulence in Fluids, Third Revised and Enlarged edn. Kluwer Academic Publishers, Dordrecht (1997)

    Book  Google Scholar 

  21. Keating, A., Piomelli, U.: A dynamic stochastic forcing method as a wall-layer model for large-eddy simulation. J. Turbul. 7(12), 1–24 (2006)

    MathSciNet  MATH  Google Scholar 

  22. Piomelli, U., Balaras, E., Pasinato, H., Squires, K. D., Spalart, P. R.: The inner–outer layer interface in large-eddy simulations with wall-layer models. Int. J. Heat Fluid Flow 24, 538–550 (2003)

    Article  Google Scholar 

  23. Kok, J. C.: Application of a stochastic backscatter model for grey-area mitigation in detached eddy simulations. In: 6th Symposium on Hybrid RANS–LES Methods. Strassbourg (2016)

  24. Kok, J. C., Fuchs, M., Mockett, C.: Delta wing at high angle of attack. In: Mockett, C., Haase, W., Schwamborn, D (eds.) Go4Hybrid – Grey Aera Mitigation for Hybrid RANS–LES, Notes on Numerical Fluid Mechanics and Multidisciplinary Design. To appear. Springer, (2017)

  25. Probst, A., Reuß, S., Schwamborn, D: The 3-element airfoil. In: Mockett, C., Haase, W., Schwamborn, D. (eds.) Go4Hybrid – Grey Aera Mitigation for Hybrid RANS–LES, Notes on Numerical Fluid Mechanics and Multidisciplinary Design. To appear. Springer, (2017)

  26. Kok, J. C.: Resolving the dependence on freestream values for the kω turbulence model. AIAA J. 38(7), 1292–1294 (2000)

    Article  Google Scholar 

  27. Stolz, S.: High-pass filtered eddy-viscosity models for large-eddy simulations of compressible wall-bounded flows. J. Fluids Eng. 127, 666–673 (2005)

    Article  Google Scholar 

  28. Lévêque, E., Toschi, F., Shao, L., Bertoglio, J. P.: Shear-improved smagorinsky model for large-eddy simulation of wall-bounded turbulent flows. J. Fluid Mech. 570, 491–502 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  29. Lesieur, M., Métais, O.: New trends in large-eddy simulations of turbulence. Annu. Rev. Fluid Mech. 28, 45–82 (1996)

    Article  MathSciNet  Google Scholar 

  30. Kok, J.C., van der Ven, H.: A high-order finite-volume method with block-structured local grid refinement. In: IDIHOM: Industrialization of High-Order Methods – A Top-Down Approach, Notes on Numerical Fluid Mechanics and Multidisciplinary Design, vol. 128. Springer. NLR-TP-2015-015 (2015)

  31. Kok, J. C., van der Ven, H.: A high-order finite-volume method for hybrid RANS–LES computations. In: 16th International Conference on Finite Elements in Flow Problems, Munich (2011)

  32. Kok, J.C.: A high-order low-dispersion symmetry-preserving finite-volume method for compressible flow on curvilinear grids. J. Comput. Phys. 228, 6811–6832 (2009). NLR-TP-2008-775

    Article  MathSciNet  MATH  Google Scholar 

  33. Rozema, W.: Low-Dissipation Methods and Models for the Simulation of Turbulent Subsonic Flow. Ph.D. thesis, University of Groningen (2015)

  34. Tam, C. K. W., Webb, J. C.: Dispersion-relation-preserving finite difference schemes for computational acoustics. J. Comput. Phys. 107, 262–281 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  35. Rienstra, S. W., Hirschberg, A.: An introduction to acoustics. Tech. rep., Eindhoven University of Technology (2006)

  36. Comte-Bellot, G., Corrsin, S.: The use of a contraction to improve isotropy of grid generated turbulence. J. Fluid Mech. 25, 657–682 (1966)

    Article  Google Scholar 

  37. Delville, J.: La décomposition orthogonal aux valeurs propres et l’analyse de l’organisation tridimensionnelle de écoulements turbulents cisaillés libres. Ph.D. thesis, Université de Poitiers (1995)

  38. Arakeri, V. H., Krothapalli, A., Siddavaram, V., Alkislar, M. B., Lourenco, L. M.: On the use of microjets to suppress turbulence in a Mach 0.9 axisymmetric jet. J. Fluid Mech. 490, 75–98 (2003)

    Article  MATH  Google Scholar 

  39. Bridges, J., Wernet, M.P.: Establishing consensus turbulence statistics for hot subsonic jets. In: 16th AIAA/CEAS Aeroacoustics Conference. Stockholm. AIAA Paper 2010-3751 (2010)

  40. Lau, J. C.: Effects of exit Mach number and temperature on mean-flow and turbulence characteristics in round jets. J. Fluid Mech. 105, 193–218 (1981)

    Article  Google Scholar 

  41. Lau, J. C., Morris, P. J., Fisher, M. J.: Measurements in subsonic and supersonic free jets using a laser velocimeter. J. Fluid Mech. 93, 1–27 (1979)

    Article  Google Scholar 

  42. Simonich, J. C., Narayanan, S., Barber, T. J., Nishimura, M.: Aeroacoustic characterization, noise reduction and dimensional scaling effects of high subsonic jets. AIAA J. 39, 2062–2069 (2001)

    Article  Google Scholar 

  43. Shur, M. L., Spalart, P. R., Strelets, M.K.: LES-based evaluation of a microjet noise reduction concept in static and flight conditions. J. Sound Vib. 330(17), 4083–4097 (2011)

    Article  MATH  Google Scholar 

  44. Chasnov, J. R.: Simulation of the Kolmogorov inertial subrange using an improved subgrid model. Phys. Fluids A 3, 188–200 (1991)

    Article  MATH  Google Scholar 

  45. Pope, S. B.: Turbulent Flows. Cambridge University Press (2000)

Download references

Acknowledgments

The author would like to thank Dr. M. Strelets (NTS) for kindly providing the grids and the RANS nozzle data for the round jet.

The research leading to these results has received funding from the European Union Seventh Framework Programme FP7/2007-2013 within the project Go4Hybrid (‘Grey Area Mitigation for Hybrid RANS-LES Methods’) under grant agreement no. 605361 and from NLR’s programmatic research ‘Kennis als Vermogen’.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Johan C. Kok.

Additional information

The original version of this article was revised: Due to a mix-up at the publisher figure 5 displayed the same two graphs as figure 8 in the initial publication. This has now been corrected.

The research leading to these results has received funding from the European Union Seventh Framework Programme FP7/2007-2013 within the project Go4Hybrid (‘Grey Area Mitigation for Hybrid RANS-LES Methods’) under grant agreement no. 605361 and from NLR’s programmatic research ‘Kennis als Vermogen’.

An erratum to this article is available at http://dx.doi.org/10.1007/s10494-017-9816-z.

Appendices

Appendix A: Rate of Backscatter

The new stochastic backscatter model includes spatial correlation in the stochastic variables, contrary to the models of Leith and Schumann. In this section, it is shown that this spatial correlation does not alter the scaling of the rate of backscatter as κ 4 for wave numbers κ → 0.

Consider a homogeneous stochastic forcing term f i (x,t) added to the right-hand side of the momentum equation:

$$\frac{\partial u_{i}}{\partial t} + {\ldots} = f_{i}(\boldsymbol{x},t) , $$

with its spatial Fourier transform \(\hat {f}_{i}(\boldsymbol {\kappa },t)\) given by

$$\hat{f}_{i}(\boldsymbol{\kappa},t) = {\int}_{\mathbb{R}^{3}} f_{i}(\boldsymbol{x},t) \mathrm{e}^{-\mathrm{i}\boldsymbol{\kappa}\cdot\boldsymbol{x}} \mathrm{d}\boldsymbol{x} . $$

Let C i j (r,t) be the two-point correlation of f i ,

$$C_{ij}(\boldsymbol{r},t) = \left\langle f_{i}(\boldsymbol{x},t) f_{j}(\boldsymbol{x}+\boldsymbol{r},t) \right\rangle , $$

which is independent of x due to homogeneity, and let \(\hat {R}_{ij}(\boldsymbol {\kappa },t)\) be the covariance of \(\hat {f}_{i}\),

$$\hat{R}_{ij}(\boldsymbol{\kappa},\boldsymbol{\kappa}^{\prime},t) = \left\langle \overline{\hat{f}_{i}(\boldsymbol{\kappa},t)} \hat{f}_{j}(\boldsymbol{\kappa}^{\prime},t) \right\rangle . $$

The two-point correlation and the covariance are related by

$$ \begin{array}{llllllll} \hat{R}_{ij}(\boldsymbol{\kappa},\boldsymbol{\kappa}^{\prime},t) &= {\int}_{\mathbb{R}^{3}} {\int}_{\mathbb{R}^{3}} \left\langle f_{i}(\boldsymbol{x},t) f_{j}(\boldsymbol{x}+\boldsymbol{r},t) \right\rangle \mathrm{e}^{\mathrm{i}\boldsymbol{\kappa}\cdot\boldsymbol{x}} \mathrm{e}^{-\mathrm{i}\boldsymbol{\kappa}^{\prime}\cdot(\boldsymbol{x}+\boldsymbol{r})} \mathrm{d}\boldsymbol{x} \, \mathrm{d}(\boldsymbol{x}+\boldsymbol{r}) \\ &= {\int}_{\mathbb{R}^{3}} \mathrm{e}^{\mathrm{i}(\boldsymbol{\kappa}-\boldsymbol{\kappa}^{\prime})\cdot\boldsymbol{x}} \mathrm{d}\boldsymbol{x} {\int}_{\mathbb{R}^{3}} \mathrm{e}^{-\mathrm{i}\boldsymbol{\kappa}^{\prime}\cdot\boldsymbol{r}} C_{ij}(\boldsymbol{r},t) \mathrm{d}\boldsymbol{r} \\ &= 2\pi\delta(\boldsymbol{\kappa}-\boldsymbol{\kappa}^{\prime}) \hat{C}_{ij}(\boldsymbol{\kappa}^{\prime},t) \end{array} $$
(10)

so that

$$ \hat{C}_{ij}(\boldsymbol{\kappa},t) = \frac{1}{2\pi} {\int}_{\mathbb{R}^{3}} \hat{R}_{ij}(\boldsymbol{\kappa},\boldsymbol{\kappa}^{\prime},t) \mathrm{d}\boldsymbol{\kappa}^{\prime} . $$
(11)

The rate of backscatter at a wave number κ is determined by the spectrum function \(\hat {F}(\kappa ,t)\) of f i [44]. Here, \(\hat {F}\) is defined analogous to the energy spectrum function E(κ) of the velocity field as defined in Pope, section 6.5 [45]:

$$\hat{F}(\kappa,t) = \oint_{S_{\kappa}} \hat{C}_{ii}(\boldsymbol{\kappa},t) \mathrm{d} S_{\kappa} $$

with S κ the sphere around the origin with radius κ. The spectrum function \(\hat {F}\) should scale as κ 4 for wave numbers smaller than the cut-off wave number κ c π/Δ.

If f i is directly defined as a spatially uncorrelated stochastic variable, then C i j = C δ i j Δ3 δ(r) (with C constant in case of homogeneity), implying \(\hat {C}_{ij} = C{\Delta }^{3}\delta _{ij}\) and \(\hat {F} = 12\pi C{\Delta }^{3}\kappa ^{2}\) (as δ i i = 3 and the surface of a sphere equals 4π κ 2). Thus, such an approach would give the wrong scaling of the power spectrum.

The correct scaling is obtained if f i is defined as the gradient of a spatially uncorrelated stochastic variable ξ i , formulated in case of the Leith model as

$$f_{i}(\boldsymbol{x},t) = \varepsilon_{ijk}\frac{\partial \xi_{k}}{\partial x_{j}} , $$

with ε i j k the alternating symbol. In this case, the Fourier transform of the two-point correlation D i j of ξ i is given by \(\hat {D}_{ij} = D{\Delta }^{3}\delta _{ij} = \tfrac {1}{3} \hat {D}_{kk} \delta _{ij}\) and its covariance by \(\hat {S}_{ij} = \tfrac {2}{3} \pi \hat {D}_{kk} \delta _{ij}\delta (\boldsymbol {\kappa }-\boldsymbol {\kappa }^{\prime })\), according to Eq. 10. One then finds that

$$\begin{array}{llllllll} \hat{R}_{ii}(\boldsymbol{\kappa},\boldsymbol{\kappa}^{\prime},t) &= \varepsilon_{ijk}\varepsilon_{ilm}(-\mathrm{i}\kappa_{j})(\mathrm{i}\kappa^{\prime}_{l}) \hat{S}_{km}(\boldsymbol{\kappa},\boldsymbol{\kappa}^{\prime},t) \\ &= (\delta_{jl}\delta_{km} - \delta_{jm}\delta_{kl})\kappa_{j}\kappa^{\prime}_{l} \tfrac{2}{3}\pi\hat{D}_{ii} \delta_{km} \delta(\boldsymbol{\kappa}-\boldsymbol{\kappa}^{\prime}) = \tfrac{4}{3}\pi\hat{D}_{ii} \kappa^{2} \delta(\boldsymbol{\kappa}-\boldsymbol{\kappa}^{\prime}) , \end{array} $$

which implies upon substitution in Eq. 11 that

$$\hat{C}_{ii} = \tfrac{2}{3} \hat{D}_{ii} \kappa^{2} = 2D {\Delta}^{3} \kappa^{2} . $$

It follows that the power spectrum \(\hat {F} = 8\pi D{\Delta }^{3}\kappa ^{4}\) has the correct scaling.

Finally, consider the case that ξ i is spatially correlated according to

$$D_{ij} = \left\langle \xi_{i}(\boldsymbol{x},t) \xi_{j}(\boldsymbol{x}+\boldsymbol{r},t) \right\rangle = D \delta_{ij}\mathrm{e}^{-d^{2}/2} , $$

with d = |r|/b and \(b = \sqrt {C_{\Delta }} {\Delta } \approx \sqrt {C_{\Delta }} \pi / \kappa _{c}\). Taking the Fourier transform of D i j gives

$$\hat{D}_{ij} = D \delta_{ij} (2\pi)^{3/2} b^{3} \mathrm{e}^{-(b\kappa)^{2}/2} , $$

so that

$$\hat{C}_{ii} = \tfrac{2}{3} \hat{D}_{ii} \kappa^{2} = 2D \kappa^{2} (2\pi)^{3/2} b^{3} \mathrm{e}^{-(b\kappa)^{2}/2} , $$

following the same derivation as above, and finally

$$\hat{F} = 8\pi D\kappa^{4} (2\pi)^{3/2} b^{3} \mathrm{e}^{-(b\kappa)^{2}/2} , $$

which again scales as κ 4 for κκ c .

Appendix B: Correlations of Solutions of the Stochastic Differential Equations

Consider the Langevin-type equation (4) and the spatial stochastic differential equation (6). In this appendix, it is shown that the solution of these equations has the desired spatial and temporal correlation of Eq. 3, interpreted in Lagrangian sense, in case of a uniform flow.

Let G(x,t) be the Green’s function satisfying the equation

$$G + \tau \left( \frac{\partial G}{\partial t} + \boldsymbol{u}\cdot\nabla G \right) = \delta(\boldsymbol{x}) \delta(t) , $$

for a constant velocity u. Applying Fourier transformations of this equation both in space and time results in the following expression for the Fourier transform \(\hat {G}\) of G:

$$\hat{G}(\boldsymbol{\kappa},\omega) = \frac{1}{(2\pi)^{2}} \frac{1}{1 + \mathrm{i}\tau\omega + \mathrm{i}\tau\boldsymbol{u}\cdot\boldsymbol{\kappa}} \, , $$

with κ the spatial wave number vector and ω the angular frequency. Applying inverse Fourier transforms, first in time and then in space, gives

$$G(\boldsymbol{x},t) = \frac{1}{\tau} H(t) \mathrm{e}^{-t/\tau} \delta(\boldsymbol{x}-\boldsymbol{u}t) . $$

Using the Green’s function, a general solution of Eq. 4 for arbitrary dW i can be written as

$$\begin{array}{llllllll} \xi_{i}(\boldsymbol{x},t) &= {\int}_{\mathbb{R}^{3}} {\int}_{-\infty}^{\infty} G(\boldsymbol{x}-\boldsymbol{x}^{\prime},t-t^{\prime}) \sqrt{2\tau} \, \mathrm{d}\boldsymbol{x}^{\prime} \, \mathrm{d} W_{i}(\boldsymbol{x}^{\prime},t^{\prime}) \\ &= {\int}_{-\infty}^{\infty} \sqrt{\tfrac{2}{\tau}} H(t-t^{\prime}) \mathrm{e}^{(t^{\prime}-t)/\tau} \, \mathrm{d} W_{i}(\boldsymbol{x}-\boldsymbol{u}(t-t^{\prime}),t^{\prime}) . \end{array} $$

Given the spatio-temporal correlation of dW i defined by Eq. 5, the following spatio-temporal correlation is found for ξ i :

$$\begin{array}{llllllll} \left\langle \xi_{i}(\boldsymbol{x},t) \xi_{j}(\boldsymbol{y},s) \right\rangle &= {\int}_{-\infty}^{\infty} {\int}_{-\infty}^{\infty} \tfrac{2}{\tau} H(t-t^{\prime}) H(s-s^{\prime}) \mathrm{e}^{(t^{\prime}-t)/\tau} \mathrm{e}^{(s^{\prime}-s)/\tau} \\ &\qquad \left\langle \mathrm{d} W_{i}(\boldsymbol{x}-\boldsymbol{u}(t-t^{\prime}),t^{\prime}) \, \mathrm{d} W_{j}(\boldsymbol{y}-\boldsymbol{u}(s-s^{\prime}),s^{\prime}) \right\rangle \\ &= {\int}_{-\infty}^{\infty} {\int}_{-\infty}^{\infty} \delta_{ij} \tfrac{2}{\tau} H(t-t^{\prime}) H(s-s^{\prime}) \mathrm{e}^{(t^{\prime}-t)/\tau} \mathrm{e}^{(s^{\prime}-s)/\tau} \delta(t^{\prime}-s^{\prime}) \\ &\qquad \exp\left( -\tfrac{1}{2b^{2}} \left|\boldsymbol{x}-\boldsymbol{y}-\boldsymbol{u}(t-s-t^{\prime}+s^{\prime})\right|^{2} \right) \mathrm{d} t^{\prime} \, \mathrm{d} s^{\prime} \\ &= \delta_{ij} \exp\left( -\tfrac{1}{2b^{2}} \left|\boldsymbol{x}-\boldsymbol{y}-\boldsymbol{u}(t-s)\right|^{2} \right) {\int}_{-\infty}^{\min(t,s)} \tfrac{2}{\tau} \mathrm{e}^{(2t^{\prime}-t-s)/\tau} \mathrm{d} t^{\prime} \\ &= \delta_{ij} \exp\left( -\tfrac{1}{2b^{2}} \left|\boldsymbol{x}-\boldsymbol{y}-\boldsymbol{u}(t-s)\right|^{2} \right) \mathrm{e}^{-|t-s|/\tau} , \end{array} $$

which is indeed equivalent with Eq. 3 with x and y replaced by the Lagrangian coordinates xu t and yu s.

Next, let G(x) be the 1D Green’s function satisfying the equation

$$\left( 1 - b^{2} \frac{\mathrm{d}^{2}}{\mathrm{d} x^{2}} \right) G = \delta(x) , $$

which is given by

$$G(x) = \frac{1}{2b} \mathrm{e}^{-\left|x\right|/b} . $$

A general solution of Eq. 6 for arbitrary dV i can then be written as

$$\mathrm{d} W_{i}(\boldsymbol{x},t) = {\int}_{\mathbb{R}^{3}} G_{3}(\boldsymbol{x}-\boldsymbol{x}^{\prime}) 8 b^{3/2} \, \mathrm{d} V_{i}(\boldsymbol{x}^{\prime},t) , $$

with \(G_{3}(\boldsymbol {x}-\boldsymbol {x}^{\prime }) = G(x_{1}-x^{\prime }_{1}) G(x_{2}-x^{\prime }_{2}) G(x_{3}-x^{\prime }_{3})\). If dV i is completely uncorrelated both in space and time, as defined by Eq. 7, then it follows that

$$\begin{array}{llllllll} \left\langle \mathrm{d} W_{i}(\boldsymbol{x},t) \mathrm{d} W_{j}(\boldsymbol{y},s) \right\rangle &= 64 b^{3} {\int}_{\mathbb{R}^{3}} {\int}_{\mathbb{R}^{3}} G_{3}(\boldsymbol{x}-\boldsymbol{x}^{\prime}) G_{3}(\boldsymbol{y}-\boldsymbol{y}^{\prime}) \left\langle \mathrm{d} V_{i}(\boldsymbol{x}^{\prime},t) \mathrm{d} V_{j}(\boldsymbol{y}^{\prime},s) \right\rangle \\ &= 64 b^{3} \delta_{ij} \delta(t-s) {\int}_{\mathbb{R}^{3}} G_{3}(\boldsymbol{x}-\boldsymbol{x}^{\prime}) G_{3}(\boldsymbol{y}-\boldsymbol{x}^{\prime}) \mathrm{d} \boldsymbol{x}^{\prime} \\ &= b^{-3} \delta_{ij} \delta(t-s) {\prod}_{m=1}^{3} {\int}_{\mathbb{R}} \mathrm{e}^{-\left|x_{m}-x^{\prime}_{m}\right|/b} \mathrm{e}^{-\left|y_{m}-x^{\prime}_{m}\right|/b} \mathrm{d} x^{\prime}_{m} \\ &= \delta_{ij} \delta(t-s) {\prod}_{m=1}^{3} \left( 1 + \tfrac{1}{b}\left|x_{m}-y_{m}\right| \right) \mathrm{e}^{-\left|x_{m}-y_{m}\right|/b} \\ &= \delta_{ij} \delta(t-s) \left( \mathrm{e}^{-\left|\boldsymbol{x}-\boldsymbol{y}\right|^{2}/(2b^{2})} + \mathcal{O}(\left|\boldsymbol{x}-\boldsymbol{y}\right|^{3}) \right) . \end{array} $$

For small distances, this is the spatio-temporal correlation of dW i as defined by Eq. 5 up to third order in the distance. For large distances, this correlation rapidly decays, ensuring the correct scaling of the rate of backscatter.

Appendix C: Preservation of the Variance in the Discretized Spatial Stochastic Differential Equation

Consider the 1D stochastic differential equation

$$(1-{\beta\delta_{i}^{2}}) \eta^{\prime}_{i} = \zeta_{i} , $$

with i the grid-cell index, β the smoothing coefficient, and \({\delta _{i}^{2}}\) the second-order difference operator. The stochastic variable ζ i = N(0, 1) is spatially uncorrelated,

$$\left\langle\zeta_{i} \zeta_{j}\right\rangle = \delta_{ij} . $$

We wish to determine the variance of the smoothed variable η .

Let G i be the discrete Green’s function satisfying the equation

$$(1-{\beta\delta_{i}^{2}}) G_{i} = \delta_{i0} . $$

Then, the general solution of the 1D stochastic differential equation (assuming periodicity) is given by

$$\eta^{\prime}_{i} = {\sum}_{j} G_{i-j} \zeta_{j} , $$

so that the variance of \(\eta ^{\prime }_{i}\) is given by

$$\left\langle(\eta^{\prime}_{i})^{2}\right\rangle = {\sum}_{j} {\sum}_{k} G_{i-j} G_{i-k} \left\langle\zeta_{j} \zeta_{k}\right\rangle = {\sum}_{j} G_{i-j}^{2} = N \left|G\right|^{2} , $$

with N the number of grid cells and |G| the L 2 norm of G.

Consider the discrete Fourier transform \(\hat {G}_{k}\) of the Green’s function

$$\hat{G}_{k} = \frac{1}{N} {\sum}_{i} G_{i} \mathrm{e}^{-\mathrm{i} \theta_{k} i} $$

with 𝜃 k = 2π k/N. Applying the Fourier transform to the equation for the Green’s function, one finds

$$\hat{G}_{k} = \frac{1}{N} (1 + 2\beta(1-\cos\theta_{k}) )^{-1} . $$

Thus, the variance of \(\eta ^{\prime }_{i}\) is given by

$$\left\langle(\eta^{\prime}_{i})^{2}\right\rangle = N {\sum}_{k} \left|\hat{G}_{k}\right|^{2} = \frac{1}{N} {\sum}_{k} (1 + 2\beta(1-\cos\theta_{k}) )^{-2} . $$

In the limit for zero mesh size (i.e., δ 𝜃 = 2π/N → 0), this summation becomes an integral over the wave number 𝜃, so that

$$\begin{array}{llllllll} \left\langle(\eta^{\prime}_{i})^{2}\right\rangle &= \frac{1}{2\pi} {\int}_{0}^{2\pi} (1 + 2\beta(1-\cos\theta) )^{-2} \mathrm{d}\theta = \frac{1}{\pi} {\int}_{0}^{\pi} (1 + 2\beta(1-\cos\theta) )^{-2} \mathrm{d}\theta \\ &= \frac{1}{\pi} \frac{2}{1+4\beta} \left[ \frac{ \beta \sin\theta }{ 1 + 2\beta (1+\cos\theta) } + \frac{ 1 + 2\beta }{ \sqrt{1+4\beta} } \arctan\left( \sqrt{1+4\beta} \tan\left( \tfrac{1}{2}\theta\right) \right) \right]_{0}^{\pi} \\ &= \frac{ 1 + 2\beta }{ (1+4\beta)^{3/2} } \end{array} $$

Finally, it follows that to obtain a stochastic variable with unit variance, \(\eta ^{\prime }_{i}\) should be scaled as

$$\eta_{i} = \frac{ (1+4\beta)^{3/4} }{ (1 + 2\beta)^{1/2} } \eta^{\prime}_{i} . $$

In 3D, this scaling should be applied for each computational direction.

Appendix D: Preservation of the Variance in the Discretized Langevin-Type Equation

Consider the Langevin-type equation in primitive form, discretized as

$$\xi^{n} + \frac{\tau}{\delta t} \left( \xi^{n+1/2} - \xi^{n-1/2} \right) + \tau \boldsymbol{u}^{n} \cdot \nabla\xi^{n} = \sqrt{\frac{2\tau}{\delta t}} \eta^{n} , $$

with \(\xi ^{n} = \tfrac {1}{2} (\xi ^{n+1/2} + \xi ^{n-1/2})\).

First, no flow (u = 0) is considered, in which case

$$\left( 1 + \frac{\delta t}{2\tau} \right) \xi^{n+1/2} = \left( 1 - \frac{\delta t}{2\tau} \right) \xi^{n-1/2} + \sqrt{\frac{2\delta t}{\tau}} \eta^{n} . $$

Given that 〈(η n)2〉 = 1 and 〈η n ξ n−1/2〉 = 0, and assuming that 〈(ξ n−1/2)2〉 = 1, it follows that

$$\left( 1 + \frac{\delta t}{2\tau} \right)^{2} \left\langle(\xi^{n+1/2})^{2}\right\rangle = \left( 1 - \frac{\delta t}{2\tau} \right)^{2} + \frac{2\delta t}{\tau} , $$

so that

$$\left\langle(\xi^{n+1/2})^{2}\right\rangle = 1 . $$

Thus, if the initial variance of ξ equals one, then this variance is locally conserved by the central time discretization.

Including the convective term, local conservation of the variance cannot be easily proven. However, one can show global conservation up to third order in the time step. Let ξ be the vector with the values of ξ at all grid points as its components. The Langevin-type equation, discretized both in space and time, can then be written as

$$\boldsymbol{\xi}^{n} + \frac{\tau}{\delta t} \left( \boldsymbol{\xi}^{n+1/2} - \boldsymbol{\xi}^{n-1/2} \right) + \tau \boldsymbol{C}(\boldsymbol{u}^{n}) \boldsymbol{\xi}^{n} = \sqrt{\frac{2\tau}{\delta t}} \boldsymbol{\eta}^{n} , $$

with C the discretized convection operator. For a symmetry-preserving discretization, this operator is skew-symmetric, that is, ξ T C ξ = 0 for any ξ. Rewriting the equation as

$$\left( \boldsymbol{I} + \frac{\delta t}{2\tau}\boldsymbol{I} + \tfrac{1}{2}\delta t\boldsymbol{C} \right) \boldsymbol{\xi}^{n+1/2} = \left( \boldsymbol{I} - \frac{\delta t}{2\tau}\boldsymbol{I} - \tfrac{1}{2}\delta t\boldsymbol{C} \right) \boldsymbol{\xi}^{n-1/2} + \sqrt{\frac{2\delta t}{\tau}} \boldsymbol{\eta}^{n} , $$

and taking the square of the L 2-norm of the left-hand and right-hand sides, one finds

$$\begin{array}{@{}rcl@{}} &&\left( 1 + \frac{\delta t}{2\tau} \right)^{2} \left|\boldsymbol{\xi}^{n+1/2}\right|^{2} + \tfrac{1}{4}(\delta t)^{2} \left|\boldsymbol{C}\boldsymbol{\xi}^{n+1/2}\right|^{2} = \left( 1 - \frac{\delta t}{2\tau} \right)^{2} \left|\boldsymbol{\xi}^{n-1/2}\right|^{2} \\ &&+ \tfrac{1}{4}(\delta t)^{2} \left|\boldsymbol{C}\boldsymbol{\xi}^{n-1/2}\right|^{2} + \frac{2\delta t}{\tau} \left|\boldsymbol{\eta}^{n}\right|^{2} + (\ldots) \boldsymbol{\xi}^{n-1/2} \cdot \boldsymbol{\eta}^{n} , \end{array} $$

if C is skew-symmetric. This time, given that 〈|(η n)2|〉 = N and assuming that 〈|(ξ n−1/2)2|〉 = N, with N the total number of grid points, it follows that

$$\begin{array}{@{}rcl@{}} \left( 1 + \frac{\delta t}{2\tau} \right)^{2} \left( \left\langle\left|\boldsymbol{\xi}^{n+1/2}\right|^{2}\right\rangle - N \right) = \tfrac{1}{4}(\delta t)^{2} \left( \left\langle\left|\boldsymbol{C}\boldsymbol{\xi}^{n-1/2}\right|^{2}\right\rangle - \left\langle\left|\boldsymbol{C}\boldsymbol{\xi}^{n+1/2}\right|^{2}\right\rangle \right) . \end{array} $$

This equation allows for the global conservation of the variance of ξ, that is, 〈|(ξ n+1/2)2|〉 = N, if either conservation of the variance of C ξ is also assumed or the right-hand side, which is of \(\mathcal {O}((\delta t)^{3})\), is neglected. Note that if the discretization is not symmetry-preserving, then a right-hand side of \(\mathcal {O}(\delta t)\) is found that leads to considerable dissipation of the total variance of ξ.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kok, J.C. A Stochastic Backscatter Model for Grey-Area Mitigation in Detached Eddy Simulations. Flow Turbulence Combust 99, 119–150 (2017). https://doi.org/10.1007/s10494-017-9809-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10494-017-9809-y

Keywords

Navigation