Skip to main content
Log in

Development and Validation of a New Formulation of Hybrid Temporal Large Eddy Simulation

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

Abstract

Hybrid RANS–LES approaches have aroused interest for years since they provide unsteady information at a reduced numerical cost compared to LES. In the hybrid context, the use of temporal filtering, to control the energy partition between resolved and modeled scales, ensures a consistent bridging between RANS and LES models. In this regard, a new formulation of Hybrid Temporal Large Eddy Simulation (HTLES) is developed, aiming at improving the theoretical foundation of the model associated with an eddy-viscosity closure. The analytical development is performed, applying the Hybrid-Equivalence criterion, and the model is calibrated in decaying isotropic turbulence. In addition, an upgraded version of the approach is proposed to improve the behavior of the model in near-wall regions, introducing a two-fold shielding function and an internal consistency constraint to provide a suitable control of the RANS-to-LES transition. Applying HTLES to the k\(\omega\) SST model, the validation process is carried out on channel and periodic-hill flows, over a range of grids and Reynolds numbers. The predictive accuracy and the robustness to grid coarsening are assessed in these cases, ensuring that HTLES offers a cost-saving alternative to LES.

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
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29
Fig. 30

Similar content being viewed by others

References

  • Afailal, A.H.: Numerical simulation of non-reactive aerodynamics in internal combustion engines using a hybrid RANS/LES approach. Ph.D. thesis, Université de Pau et des pays de l’Adour (2020)

  • Afailal, H., Galpin, J., Velghe, A., Manceau, R.: Development and validation of a hybrid temporal LES model in the perspective of applications to internal combustion engines. Oil Gas Sci. Technol. 74, (2019)

  • Archambeau, F., Méchitoua, N., Sakiz, M.: Code Saturne: A finite volume code for the computation of turbulent incompressible flows—industrial applications. Int. J. Finite Vol. (2004)

  • Ashton, N., Prosser, R., Revell, A.: A hybrid numerical scheme for a new formulation of delayed detached-eddy simulation (DDES) based on elliptic relaxation. In: Journal of Physics: Conference Series, vol. 318, p. 042043. IOP Publishing (2011)

  • Bentaleb, Y., Manceau, R.: A hybrid temporal LES/RANS formulation based on a two-equation subfilter model. In: TSFP Digital Library Online. Begel House Inc. (2011)

  • Billard, F., Laurence, D.: A robust \(k\)-\(\varepsilon\)-\(\overline{v^2}\)/\(k\) elliptic blending turbulence model applied to near-wall, separated and buoyant flows. Int. J. Heat Fluid Flow 33(1), 45–58 (2012)

    Article  Google Scholar 

  • Breuer, M., Jaffrézic, B.: New reference data for the hill flow test case. ERCOFTAC QNET-CFD knowledge Base Wiki: http://www.ercoftac.org/products_and_services/wiki (2005)

  • Chaouat, B.: Subfilter-scale transport model for hybrid RANS/LES simulations applied to a complex bounded flow. Journal of Turbulence (11), N51 (2010)

  • Chaouat, B.: The state of the art of hybrid RANS/LES modeling for the simulation of turbulent flows. Flow Turbul Combust 99(2), 279–327 (2017)

    Article  Google Scholar 

  • Chaouat, B., Schiestel, R.: A new partially integrated transport model for subgrid-scale stresses and dissipation rate for turbulent developing flows. Phys. Fluids 17(6), (2005)

  • Chaouat, B., Schiestel, R.: Progress in subgrid-scale transport modelling for continuous hybrid non-zonal RANS/LES simulations. Int. J. Heat Fluid Fl. 30(4), 602–616 (2009)

    Article  Google Scholar 

  • Chaouat, B., Schiestel, R.: Hybrid RANS/LES simulations of the turbulent flow over periodic hills at high Reynolds number using the PITM method. Comput. Fluids 84, 279–300 (2013)

    Article  MathSciNet  Google Scholar 

  • Chaouat, B., Schiestel, R.: Partially integrated transport modeling method for turbulence simulation with variable filters. Phys. Fluids 25, (2013)

  • Fadai-Ghotbi, A., Friess, C., Manceau, R., Borée, J.: A seamless hybrid RANS-LES model based on transport equations for the subgrid stresses and elliptic blending. Phys. Fluids 22(5), (2010)

  • Fadai-Ghotbi, A., Friess, C., Manceau, R., Gatski, T.B., Borée, J.: Temporal filtering: a consistent formalism for seamless hybrid RANS-LES modeling in inhomogeneous turbulence. Int. J. HeatFluid Flow 31(3), 378–389 (2010)

  • Friess, C., Manceau, R., Gatski, T.: Toward an equivalence criterion for hybrid RANS/LES methods. Comput. Fluids 122, 233–246 (2015)

    Article  MathSciNet  Google Scholar 

  • Germano, M.: Turbulence: the filtering approach. J. Fluid Mech. 238, 325–336 (1992)

    Article  MathSciNet  Google Scholar 

  • Germano, M.: Properties of the hybrid RANS/LES filter. Theor. Comput. Fluid Dyn. 17, 225–231 (2004)

    Article  Google Scholar 

  • Girimaji, S.S.: Partially-averaged Navier-stokes model for turbulence: a Reynolds-averaged Navier-stokes to direct numerical simulation bridging method. J. Appl. Mech. 73, 413 (2006)

    Article  Google Scholar 

  • Hamba, F.: Log-layer mismatch and commutation error in hybrid RANS/LES simulation of channel flow. Int. J. Heat Fluid Fl. 30(1), 20–31 (2009)

    Article  Google Scholar 

  • Han, Y., He, Y., Le, J.: Modification to improved delayed detached-eddy simulation regarding the log-layer mismatch. AIAA J. 58(2), 712–721 (2020)

    Article  Google Scholar 

  • Hinze, J.: Turbulence, 2nd edn. MacGraw Hill, New-York (1975)

    Google Scholar 

  • Jakirlic, S.: Assessment of the RSM, URANS and hybrid models with respect to the different roadmaps including the industrial application challenges. Deliverable 3, 2–36 (2012)

    Google Scholar 

  • Jamal, T., Walters, D.K.: A dynamic time filtering technique for hybrid RANS-LES simulation of non-stationary turbulent flow. In: Fluids Engineering Division Summer Meeting, vol. 59032, p. V002T02A051. American Society of Mechanical Engineers (2019)

  • Jones, W., Launder, B.E.: The prediction of laminarization with a two-equation model of turbulence. Int. J. Heat Mass Transf 15(2), 301–314 (1972)

    Article  Google Scholar 

  • Kang, H.S., Chester, S., Meneveau, C.: Decaying turbulence in an active-grid-generated flow and comparisons with large-eddy simulation. J. Fluid Mech. 480, 129–160 (2003)

    Article  MathSciNet  Google Scholar 

  • Keating, A., Piomelli, U.: A dynamic stochastic forcing method as a wall-layer model for large-eddy simulation. J. Turbul. (7), N12 (2006)

  • Larsson, J., Kawai, S., Bodart, J., Bermejo-Moreno, I.: Large eddy simulation with modeled wall-stress: recent progress and future directions. Mech. Eng. Rev. 3(1), 15–00418 (2016)

    Article  Google Scholar 

  • Lee, M., Moser, R.D.: Direct numerical simulation of turbulent channel flow up to \({R}e_\tau \approx 5200\). J. Fluid Mech. 774, 395–415 (2015)

    Article  Google Scholar 

  • Manceau, R.: Progress in hybrid temporal LES. In: The 6th Symposium of Hybrid RANS-LES Methods, 26–28 September 2016, Strasbourg, France, Notes on Numerical Fluid Mechanics and Multidisciplinary Design, vol. 137, pp. 9–25. Springer (2018)

  • Manceau, R., Hanjalić, K.: Elliptic blending model: a new near-wall Reynolds-stress turbulence closure. Phys. Fluids 14(2), 744–754 (2002)

    Article  Google Scholar 

  • Menter, F., Egorov, Y.: A scale adaptive simulation model using two-equation models. In: 43rd AIAA Aerospace Sciences Meeting and Exhibit, p. 1095 (2005)

  • Menter, F.R.: Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 32(8), 1598–1605 (1994)

    Article  Google Scholar 

  • Moser, R.D., Kim, J., Mansour, N.N.: Direct numerical simulation of turbulent channel flow up to \({Re}_\tau \) = 590. Phys. Fluids 11(4), 943–945 (1999)

    Article  Google Scholar 

  • Nikitin, N., Nicoud, F., Wasistho, B., Squires, K., Spalart, P.R.: An approach to wall modeling in large-eddy simulations. Phys. Fluids 12(7), 1629–1632 (2000)

    Article  Google Scholar 

  • Pope, S.B.: Turbulent Flows. Cambridge University Press (2001)

  • Pruett, C., Gatski, T., Grosch, C.E., Thacker, W.: The temporally filtered Navier-Stokes equations: properties of the residual stress. Phys. Fluids 15(8), 2127–2140 (2003)

    Article  Google Scholar 

  • Rapp, C., Manhart, M.: Experimental investigations on the turbulent flow over a periodic hill geometry. In: TSFP Digital Library Online. Begel House Inc. (2007)

  • Rapp, C., Manhart, M.: Flow over periodic hills: an experimental study. Exp. Fluids 51(1), 247–269 (2011)

    Article  Google Scholar 

  • Sagaut, P.: Large Eddy Simulation for Incompressible Flows, 3rd edn. Springer (2006)

  • Schiestel, R., Dejoan, A.: Towards a new partially integrated transport model for coarse grid and unsteady turbulent flow simulations. Theor. Comput. Fluid Dyn. 18(6), 443–468 (2005)

    Article  Google Scholar 

  • 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(6), 1638–1649 (2008)

    Article  Google Scholar 

  • Smagorinsky, J.: General circulation experiments with the primitive equations: I. the basic experiment. Mon Weather Rev. 91(3), 99–164 (1963)

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

    Article  Google Scholar 

  • Spalart, P., Jou, W., Strelets, M., Allmaras, S.: Comments on the feasibility of LES for wings, and on a hybrid RANS/LES approach. In: 1st AFOSR International Symposium of Engineering Turbulence. Modelling and Measurements, May, pp. 24–26 (1997)

  • Spalart, P.R.: Strategies for turbulence modelling and simulations. Int. J. Heat Fluid Flow 21(3), 252–263 (2000)

    Article  Google Scholar 

  • 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(3), 181 (2006)

    Article  Google Scholar 

  • Tennekes, H.: Eulerian and Lagrangian time microscales in isotropic turbulence. J. Fluid Mech. 67(3), 561–567 (1975)

    Article  Google Scholar 

  • Tran, T., Manceau, R., Perrin, R., Borée, J., Nguyen, A.: A hybrid temporal LES approach. Application to flows around rectangular cylinders. In: Proceedings of 9th ERCOFTAC International Symposium on Engineering Turbulence Modelling and Measurements, Thessaloniki, Greece (2012)

  • Travin, A., Shur, M., Strelets, M., Spalart, P.: Physical and numerical upgrades in the detached-eddy simulation of complex turbulent flows. In: Advances in LES of Complex Flows, pp. 239–254. Springer (2002)

Download references

Acknowledgements

This work was partially supported by the ANRT (CIFRE contract 2017/0963) and the ANR project MONACO_2025 (ANR-17-CE06-0005-01_ACT). This article is dedicated to our dear colleague and friend Tom Gatski.

Funding

This study was funded by the ANRT (CIFRE contract 2017/0963) and the ANR project MONACO_2025 (reference number ANR-17-CE06-0005-01_ACT).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vladimir Duffal.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Appendix

Appendix

1.1 Perturbation Analysis to Ensure H-Equivalence Between TPITM and HTLES

Postulating that Two hybrid approaches based on the same closure, but using different method of control of the energy partition, yield similar low-order statistics of the resolved velocity fields provided that they yield the same level of subfilter energy (Friess et al. 2015), the H-equivalence between the TPITM system Eq. (9, 10) and the HTLES system Eq. (12, 18) is ensured if the two methods tend to the same closure when the filter width is going to infinity and lead to the same energy partition for a particular situation. Hence, using a k\(\varepsilon\) closure herein, the objective of the analysis is to identify the relation between the hybridization functions \(C_{\varepsilon 2}^P(r)\) and \(\psi (r)\), to provide the same level of modeled energy \(k_\text {m}\).

Departing from the RANS limit where the H-equivalence is ensured (\(k_\text {m}=k\)), infinitesimal perturbations of the hybridization functions \(\delta C_{\varepsilon 2}^P\) and \(\delta \psi\) are introduced into the TPITM and HTLES systems, which remain H-Equivalent as long as the same infinitesimal variation \(\delta k_\text {m}\) of the modeled energy is obtained. The perturbation analysis can be carried out in some particular situations, following (Friess et al. 2015). Calculations are conducted herein in the less restrictive situation: inhomogeneous turbulence in straight duct flows. In this case, both modeled energy and dissipation are in equilibrium along streamlines, which leads to the following system of equations for HTLES,

$$\begin{aligned} \nu _\text {m} = C_\mu \frac{k_\text {m}^2}{\psi (r) \varepsilon _\text {m}^*}, \qquad \left\{ \begin{array}{l} 0 = \displaystyle P_\text {m} + D_{k \text {m}} - \psi (r) \varepsilon _\text {m}^*, \\ 0 = \displaystyle C_{\varepsilon 1} \frac{\varepsilon _\text {m}^*}{k_\text {m}} P_\text {m} + D_{\varepsilon \text {m}}^* - C_{\varepsilon 2} \frac{{\varepsilon _\text {m}^*}^2}{k_\text {m}}. \end{array} \right. \end{aligned}$$
(41)

Introducing an infinitesimal perturbation \(\delta \psi\) of the coefficient \(\psi (r)\) in the equation results in infinitesimal variations of the solution, such that the following relation is obtained,

$$\begin{aligned} \left\{ \begin{array}{l} 0 = \displaystyle \delta P_\text {m} + \delta D_{k \text {m}} - \left( \frac{\delta \psi }{\psi (r)} + \frac{\delta \varepsilon _\text {m}^*}{\varepsilon _\text {m}^*} \right) \psi (r) \varepsilon _\text {m}^*, \\ 0 = \displaystyle C_{\varepsilon 1} \frac{\varepsilon _\text {m}^*}{k_\text {m}} P_\text {m} \left( \frac{\delta P_\text {m}}{P_\text {m}} + \frac{\delta \varepsilon _\text {m}^*}{\varepsilon _\text {m}^*} - \frac{\delta k_\text {m}}{k_\text {m}} \right) + \delta D_{\varepsilon \text {m}}^* - C_{\varepsilon 2} \frac{{\varepsilon _\text {m}^*}^2}{k_\text {m}} \left( 2\frac{\delta \varepsilon _\text {m}^*}{\varepsilon _\text {m}^*} - \frac{\delta k_\text {m}}{k_\text {m}} \right) . \end{array} \right. \end{aligned}$$
(42)

The diffusion terms are given by the gradient-diffusion hypothesis with the usual coefficients \(\sigma _k\) and \(\sigma _\varepsilon\),

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle D_{k \text {m}} = \frac{\partial }{\partial x_k} \left( \frac{C_\mu }{\sigma _k} \frac{k_\text {m}^2}{\psi (r) \varepsilon _\text {m}^*} \frac{\partial k_\text {m}}{\partial x_k} \right) , \\ \displaystyle D_{\varepsilon \text {m}}^* = \frac{\partial }{\partial x_k} \left( \frac{C_\mu }{\sigma _\varepsilon } \frac{k_\text {m}^2}{\psi (r) \varepsilon _\text {m}^*} \frac{\partial \varepsilon _\text {m}^*}{\partial x_k} \right) , \end{array} \right. \quad \text {such that} \quad \left\{ \begin{array}{l} \displaystyle \frac{\delta D_{k \text {m}}}{D_{k \text {m}}} = 3\frac{\delta k_\text {m}}{k_\text {m}} - \frac{\delta \varepsilon _\text {m}^*}{\varepsilon _\text {m}^*} - \frac{\delta \psi }{\psi (r)}, \\ \displaystyle \frac{\delta D_{\varepsilon \text {m}}^*}{D_{\varepsilon \text {m}}^*} = 2\frac{\delta k_\text {m}}{k_\text {m}} - \frac{\delta \psi }{\psi (r)}. \end{array} \right. \end{aligned}$$
(43)

Moreover, in eddy-viscosity models the production term is given by,

$$\begin{aligned} P_\text {m} = 2 \gamma C_\mu \frac{k_\text {m}^2}{\psi (r)\varepsilon _\text {m}^*} S^2, \end{aligned}$$
(44)

where S is the statistical averaged of the rate of strain, and \(\gamma =\overline{P_\text {sfs}}/P_\text {m}\) is a correlation coefficient. Then, the perturbation analysis is applied, leading to

$$\begin{aligned} \frac{\delta P_\text {m}}{P_\text {m}} = \frac{\delta \gamma }{\gamma } + 2 \frac{\delta S}{S} + 2 \frac{\delta k_\text {m}}{k_\text {m}} - \frac{\delta \varepsilon _\text {m}^*}{\varepsilon _\text {m}^*} - \frac{\delta \psi }{\psi (r)}. \end{aligned}$$
(45)

Finally, the previous Eqs. (43, 45) are introduced into the system of Eq. (42), which yields

$$\begin{aligned} \delta \psi = - \frac{\left( C_{\varepsilon 2} - C_{\varepsilon 1} \psi (r) \right) \psi (r)}{C_{\varepsilon 2}} \frac{P_\text {m}}{\psi (r) \varepsilon _\text {m}^*} \left( \frac{\delta k_\text {m}}{k_\text {m}} - \frac{\delta \gamma }{\gamma } - 2 \frac{\delta S}{S} \right) . \end{aligned}$$
(46)

A similar analysis can be conducted for TPITM (see Friess et al. (2015) for details), and the following relation is obtained

$$\begin{aligned} \delta C_{\varepsilon 2}^P = (C_{\varepsilon 2}^P(r) - C_{\varepsilon 1}) \frac{P_\text {m}}{\varepsilon _\text {m}} \left( \frac{\delta k_\text {m}}{k_\text {m}} - \frac{\delta \gamma }{\gamma } - 2 \frac{\delta S}{S} \right) . \end{aligned}$$
(47)

Applying the H-equivalence postulate (Friess et al. 2015), i.e. the same modification \(\delta k_\text {m}\) of the modeled energy leads to the same infinitesimal variation of the low-order statistics (\(\delta S\), \(\delta \gamma\)), and recalling the equality \(\varepsilon _\text {m}=\psi (r) \varepsilon _\text {m}^*\) (Eq. 17), then if HTLES and TPITM are H-Equivalent for some initial values of \(\psi (r)\) and \(C_{\varepsilon 2}^P(r)\), they remain equivalent if the infinitesimal perturbations of their hybridization functions expressed by Eq. (46) and (47) satisfy the relation

$$\begin{aligned} \frac{\delta C_{\varepsilon 2}^P}{\left( C_{\varepsilon 2}^P(r) - C_{\varepsilon 1}\right) } = -\frac{C_{\varepsilon 2} \delta \psi }{\left( C_{\varepsilon 2} - C_{\varepsilon 1}\psi (r)\right) \psi (r) }. \end{aligned}$$
(48)

Lastly, the previous formula is integrated between the RANS state (\(r=1\), \(C_{\varepsilon 2}^*(1)=C_{\varepsilon 2}\), \(\psi (1)=1\)) and some arbitrary LES state, such that

$$\begin{aligned} \int _{C_{\varepsilon 2}}^{C_{\varepsilon 2}^P} \frac{1}{x-C_{\varepsilon 1}} \text {d}x = \displaystyle - \int _{1}^{\psi } \frac{C_{\varepsilon 2}}{(C_{\varepsilon 2}- C_{\varepsilon 1} y) y} \text {d}y, \end{aligned}$$
(49)

and simple algebra leads to

$$\begin{aligned} \psi (r) = \frac{C_{\varepsilon 2}}{C_{\varepsilon 2}^P(r)} = \frac{C_{\varepsilon 2}}{C_{\varepsilon 1} + r (C_{\varepsilon 2} - C_{\varepsilon 1})}. \end{aligned}$$
(50)

In the other situations considered in Friess et al. (2015), it is assumed that the modeled dissipation rate is not affected by the modification of the energy partition, such that \(\delta \varepsilon _\text {m}=\delta (\psi (r) \varepsilon _\text {m}^*)=0\) (Eq. 17). Calculations are much simpler in this way and the same formulation of \(\psi (r)\) is obtained (not shown here) .

1.2 H-Equivalence Between TPITM and HTLES Based on SST Closure

In this section, the k\(\omega\) SST HTLES formulation is derived from k\(\omega\) SST TPITM, while preserving the H-equivalence criterion. In the k\(\omega\) SST TPITM model, developed by Bentaleb and Manceau (2011), the modeled viscosity \(\nu _\text {m}\), energy \(k_\text {m}\) and specific dissipation \(\omega _\text {m}\), satisfy

$$\begin{aligned} \nu _\text {m} = \frac{k_\text {m}}{\omega _\text {m}}, \qquad \left\{ \begin{array}{ccl} \displaystyle \frac{D k_\text {m}}{D t} &{} = &{} \displaystyle P_\text {m} + D_{k \text {m}} - C_\mu k_\text {m} \omega _\text {m}, \\ \displaystyle \frac{D \omega _\text {m}}{D t} &{} = &{} \displaystyle \gamma _\omega \frac{\omega _\text {m}}{k_\text {m}} P_\text {m} + D_{\omega \text {m}} - \beta _\omega ^P {\omega _\text {m}}^2 + C_{\omega \text {m}}, \end{array} \right. \end{aligned}$$
(51)

where \(D_{\omega \text {m}}\) stand for the diffusion of \(\omega _\text {m}\), \(C_{\omega \text {m}}\) is the cross-diffusion term, \(\gamma _\omega =C_{\varepsilon 1}-1\) and \(\beta _\omega =C_\mu (C_{\varepsilon 2}-1)\) are the usual coefficients, and the hybridization term \(\beta _\omega ^P\) is defined by

$$\begin{aligned} \beta _\omega ^P(r) = C_\mu \gamma _\omega + r (\beta _\omega - C_\mu \gamma _\omega ). \end{aligned}$$
(52)

The k\(\omega\) SST HTLES model is derived by transferring the new hybridization function \(\psi '(r)\) in the modeled energy equation and introducing a new variable \(\omega _\text {m}^*\) transported by the second equation, in the same way as for the derivation of the k\(\varepsilon\) HTLES in Sect. 3.1, such that

$$\begin{aligned} \nu _\text {m} = \frac{k_\text {m}}{\psi '(r) \omega _\text {m}^*}, \qquad \left\{ \begin{array}{ccl} \displaystyle \frac{D k_\text {m}}{D t} &{} = &{} \displaystyle P_\text {m} + D_{k \text {m}} - C_\mu k_\text {m} \psi '(r) \omega _\text {m}^*, \\ \displaystyle \frac{D \omega _\text {m}^*}{D t} &{} = &{} \displaystyle \gamma _\omega \frac{\omega _\text {m}^*}{k_\text {m}} P_\text {m} + D_{\omega \text {m}}^* - \beta _\omega {\omega _\text {m}^*}^2 + C_{\omega \text {m}}^*, \end{array} \right. \end{aligned}$$
(53)

where \(D_{\omega \text {m}}^*\) stands for the diffusion of \(\omega _\text {m}^*\).

Similar to the case of the k\(\varepsilon\) model, the perturbation analysis is carried out to ensure H-equivalence. The calculations (not shown here) are very similar to those presented in Appendix 7.1, leading to the new expression of the hybridization function

$$\begin{aligned} \psi '(r) = \frac{\beta _\omega }{\beta _\omega ^P(r)} = \frac{\beta _\omega }{C_\mu \gamma _\omega + r (\beta _\omega - C_\mu \gamma _\omega )}. \end{aligned}$$
(54)

Again, it is noted that the same formulation of \(\psi '(r)\) is obtained, irrespective of the situations proposed by Friess et al. (2015) to conduct the perturbation analysis (not shown here).

1.3 The k\(\omega\) SST HTLES Model

The k\(\omega\) SST HTLES formulation is given by :

$$\begin{aligned}&\nu _\text {sfs} = \frac{a_1 k_\text {sfs}}{\text {max}\left[ a_1 \psi '(r) \omega _\text {sfs}^*, F_2 \widetilde{S} \right] }, \end{aligned}$$
(55)
$$\begin{aligned}&\left\{ \begin{array}{ccl} \displaystyle \frac{\partial k_\text {sfs}}{\partial t} + \widetilde{U}_j \frac{\partial k_\text {sfs}}{\partial x_j} &{} = &{} \displaystyle P_k + \frac{\partial }{\partial x_j} \left[ \left( \nu + \frac{\nu _\text {sfs}}{\sigma _k} \right) \frac{\partial k_\text {sfs}}{\partial x_j} \right] - \frac{k_\text {sfs}}{T_\text {m}}, \\ \displaystyle \frac{\partial \omega _\text {sfs}^*}{\partial t} + \widetilde{U}_j \frac{\partial \omega _\text {sfs}^*}{\partial x_j} &{} = &{} \displaystyle \gamma _\omega \frac{1}{\psi '(r)}\widetilde{S}^2 + \frac{\partial }{\partial x_j} \left[ \left( \nu + \frac{\nu _\text {sfs}}{\sigma _\omega } \right) \frac{\partial \omega _\text {sfs}^*}{\partial x_j} \right] - \beta _\omega {\omega _\text {sfs}^*}^2 \\ &{} &{} \displaystyle \qquad + (1-F_1) 2 \frac{1}{\sigma _{\omega 2}} \frac{1}{\psi '(r) \omega _\text {sfs}^*} \frac{\partial \omega _\text {sfs}^*}{\partial x_j} \frac{\partial k_\text {sfs}}{\partial x_j}, \end{array} \right. \end{aligned}$$
(56)

where the production limiter is defined as :

$$\begin{aligned} P_k = \text {min} \left[ \nu _\text {sfs} \widetilde{S}^2 , a_2 C_\mu k_\text {sfs} \psi '(r) \omega _{\text {sfs}}^* \right] , \end{aligned}$$
(57)

and the function \(F_1\) and \(F_2\) are the blending functions (not modified in the hybrid context), and \(\gamma _\omega\), \(\beta _\omega\), \(a_1\), \(a_2\), \(\sigma _k\), \(\sigma _\omega\) are the usual coefficients dependent of the branch of the k\(\omega\) SST model.

The hybridization of the model is carried out by the energy ratio r, via the function \(\psi '(r)\) and the time scale \(T_m\), applying the shielding functions and the internal consistency constraint, such that

$$\begin{aligned} T_\text {m}= & {} \frac{r}{\psi '(r)} \frac{k_\text {m} + c_r k_\text {r}}{C_\mu k_\text {m} \omega _\text {m}^*} , \qquad \text {where} \ \ c_r = \left\{ \begin{array}{cc} 0 &{} \text {if} \ r = 1, \\ f_s &{} \text {if} \ r < 1, \end{array} \right. \end{aligned}$$
(58)
$$\begin{aligned} \psi '(r)= \,& {} \frac{\beta _\omega }{C_\mu \gamma _\omega + r (\beta _\omega - C_\mu \gamma _\omega )}, \end{aligned}$$
(59)
$$\begin{aligned} r\,=\, & {} (1-f_s) \times 1 + f_s \times \text {min}[1,r_K], \end{aligned}$$
(60)
$$\begin{aligned} r_K\,=\, & {} \frac{1}{\beta _0} \left( \frac{U_s}{\sqrt{k}} \right) ^{2/3} \left( \omega _c \frac{k}{\varepsilon } \right) ^{-2/3} \quad \text {where} \ \ \omega _c = \text {min} \left[ \frac{\pi }{dt}, \frac{U_s \pi }{\varDelta } \right] , \ \ \nonumber \\ U_s\,=\, & {} U + \gamma \sqrt{k} \ \ \text {and} \ \ \varDelta = \varOmega ^{1/3}, \end{aligned}$$
(61)

where \(\varOmega\) is the volume of the cell,

$$\begin{aligned} f_s= & {} 1 - \text {tanh}\left[ \text {max}\left[ \xi _K^{p_1} , \xi _D^{p_2} \right] \right] , \end{aligned}$$
(62)
$$\begin{aligned} \xi _K= & {} C_1 \frac{(\nu ^3/\varepsilon )^{1/4}}{d_w} \quad \text {and} \quad \xi _D = C_2 \frac{\varDelta _\text {max}}{d_w}, \quad \text {where} \ \ \varDelta _\text {max} = \text {max} \left[ \varDelta _x, \varDelta _y, \varDelta _z \right] , \end{aligned}$$
(63)

and the statistically-averaged energy and dissipation terms are estimated as

$$\begin{aligned} k = k_\text {m} + k_\text {r} \quad \text {and} \quad \varepsilon = \varepsilon _\text {m} = C_\mu k_\text {m} \psi '(r) \omega _\text {m}^*. \end{aligned}$$
(64)

The estimates of statistically-averaged quantities \(\overline{f}\) are provided by an exponentially-weighted average (Pruett et al. 2003), with a time filter width \(\varDelta _T\) corresponding to several tens of flow-through times herein, such that

$$\begin{aligned} \frac{\partial \overline{f}(t,\varDelta _T)}{\partial t} = \frac{\widetilde{f}(t) - \overline{f}(t,\varDelta _T)}{\varDelta _T}. \end{aligned}$$
(65)

The main coefficients of the upgraded hybrid method are summarized in Table 3.

Table 3 Coefficients of the k\(\omega\) SST HTLES model

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Duffal, V., de Laage de Meux, B. & Manceau, R. Development and Validation of a New Formulation of Hybrid Temporal Large Eddy Simulation. Flow Turbulence Combust 108, 1–42 (2022). https://doi.org/10.1007/s10494-021-00264-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10494-021-00264-z

Keywords

Navigation