Skip to main content
Log in

Application of the PITM Method Using Inlet Synthetic Turbulence Generation for the Simulation of the Turbulent Flow in a Small Axisymmetric Contraction

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

Abstract

We investigate the turbulence modeling of second moment closure used both in RANS and PITM methodologies from a fundamental point of view and its capacity to predict the flow in a low turbulence wind tunnel of small axisymmetric contraction designed by Uberoi and Wallis. This flow presents a complex phenomenon in physics of fluid turbulence. The anisotropy ratio of the turbulent stresses τ 11/τ 22 initially close to 1.4 returns to unity through the contraction, but surprisingly, this ratio gradually increases to its pre-contraction value in the uniform section downstream the contraction. This point constitutes the interesting paradox of the Uberoi and Wallis experiment. We perform numerical simulations of the turbulent flow in this wind tunnel using both a Reynolds stress model developed in RANS modeling and a subfilter scale stress model derived from the partially integrated transport modeling method. With the aim of reproducing the experimental grid turbulence resulting from the effects of the square-mesh biplane grid on the uniform wind tunnel stream, we develop a new analytical spectral method of generation of pseudo-random velocity fields in a cubic box. These velocity fields are then introduced in the channel using a matching numerical technique. Both RANS and PITM simulations are performed on several meshes to study the effects of the contraction on the mean velocity and turbulence. As a result, it is found that the RANS computation using the Reynolds stress model fails to reproduce the increase of anisotropy in the centerline of the channel after passing the contraction. In the contrary, the PITM simulation predicts fairly well this turbulent flow according to the experimental data, and especially, the “return to anisotropy” in the straight section of the channel downstream the contraction. This work shows that the PITM method used in conjunction with an analytical synthetic turbulence generation as inflow is well suited for simulating this flow, while allowing a drastic reduction of the computational resources.

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

Similar content being viewed by others

References

  1. Uberoi, M.S.: Effect of wind-tunnel contraction on free-stream Turbulence. J. Aeronaut. Sci. 23, 754–764 (1956)

    Article  Google Scholar 

  2. Uberoi, M.S.: Equipartition of energy and local isotropy in turbulent flows. J. Appl. Phys. 28, 1165–1170 (1957)

    Article  MATH  Google Scholar 

  3. Uberoi, M.S., Wallis, S.: Small axisymmetric contraction of grid turbulence. J. Fluid Mech. 24, 539–543 (1966)

    Article  Google Scholar 

  4. Jang, S.J., Sung, H.J., Krogstad, P.: Effects of an axisymmetric contraction on a turbulent pipe flow. J. Fluid Mech. 687, 376–403 (2011)

    Article  MATH  Google Scholar 

  5. Lee, J., Jang, S.J., Sung, H.J.: Direct numerical simulation of turbulent flow in a conical diffuser. J. Turbul. 13, 1–29 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  6. Lumley, J., Newman, G.: The return to isotropy of homogeneous turbulence. J. Fluid Mech. 82, 161–178 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  7. Naot, D.: Rapid distortion solutions for a stress transport turbulence model in a contracting flow. Phys. Fluids 21, 752–756 (1977)

    Article  Google Scholar 

  8. Tsuge, S.: Effects of the flow contraction on evolution of turbulence. Phys. Fluids 27, 1948–1956 (1984)

    Article  MATH  Google Scholar 

  9. Lee, M.J.: Distortion of homogeneous turbulence by axisymmetric strain and dilatation. Phys. Fluids A 9, 1541–1557 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  10. Sjögren, T., Johansson, A.: Measurement and modelling of homogeneous axisymmetric turbulence. J. Fluid Mech. 374, 59–90 (1988)

    Article  Google Scholar 

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

    Article  Google Scholar 

  12. Bennet, J.C., Corrsin, S.: Small Reynolds number nearly isotropic turbulence in a straight duct and a contraction. Phys. Fluids 21, 2129–2140 (1978)

    Article  Google Scholar 

  13. Ertunc, O., Durst, F.: On the high contraction ratio anomaly of axisymmetric contraction of grid-generated turbulence. Phys. Fluids 20(025103), 1–15 (2008)

    MATH  Google Scholar 

  14. Antonia, R.A., Lavoie, P., Djenidi, L., Benaissa, A.: Effect of a small axisymmetric contraction on grid turbulence. Exp. Fluids 49, 3–10 (2010)

    Article  Google Scholar 

  15. Uberoi, M.S., Wallis, S.: Effects of grid geometry on turbulence decay. Phys. Fluids 10, 1216–1224 (1967)

    Article  Google Scholar 

  16. Lesieur, M., Metais, O.: New trends in large-eddy simulations of turbulence. Ann. Rev J. Fluid Mech. 28, 45–82 (1996)

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

  18. Leschziner, M., Li, M.N., Tessicini, F.: Simulating flow separation from continuous surfaces: routes to overcoming the Reynolds number barrier Phil. Trans. R. Soc. A 367, 2885–2903 (2009)

    Article  MATH  Google Scholar 

  19. Choi, H., Moin, P.: Grid-point requirements for large eddy simulation: Chapman’s estimates revisited. Phys. Fluids 24(011702), 1–5 (2012)

    Google Scholar 

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

    Article  Google Scholar 

  21. Chaouat, B.: Simulation of turbulent rotating flows using a subfilter scale stress model derived from the partially integrated transport modeling method. Phys. Fluids 24(045108), 1–35 (2012)

    Google Scholar 

  22. Chaouat, B.: Subfilter-scale transport model for hybrid RANS/LES simulations applied to a complex bounded flow. J. Turbul. 11, 1–30 (2010)

    Article  Google Scholar 

  23. Schiestel, R.: Modeling and Simulation of Turbulent Flows. ISTE Ltd and J. Wiley) (2008)

  24. Gatski, T.B.: Second-moment and scalar flux representations in engineering and geophysical flows. Fluid Dyn. Res. 41(012202), 1–24 (2009)

    MATH  Google Scholar 

  25. Leschziner, M.A., Drikakis, D.: Turbulence modelling and turbulent-flow computation in aeronautics. The Aeronaut. J. 106, 349–383 (2002)

    Google Scholar 

  26. Hanjalic, K., Launder, B.E.: Modelling Turbulence in Engineering and the Environnement. Second-Moment Route to Closure. Cambridge University Press (2011)

  27. Chaouat, B.: Simulations of channel flows with effects of spanwise rotation or wall injection using a Reynolds stress model. J. Fluids Eng. ASME 123, 2–10 (2001)

    Article  Google Scholar 

  28. Chaouat, B.: Numerical predictions of channel flows with fluid injection using Reynolds stress model. J. Propul. Power 18(12), 295–303 (2002)

    Article  Google Scholar 

  29. Chaouat, B.: Reynolds stress transport modeling for high-lift airfoil flows. AIAA J. 44(10), 2390–2403 (2006)

    Article  Google Scholar 

  30. Hanjalic, K., Launder, B.E., Schiestel, R.: Multiple-time scale concepts in turbulent transport modelling. In: Proceedings of the 2th Symposium on Turbulence Shear Flow, edited by Springer Verlag, pp. 36–49 (1890)

  31. Schiestel, R.: Multiple-time scale modeling of turbulent flows in one point closures. Phys. Fluids 30, 722–731 (1987)

    Article  MATH  Google Scholar 

  32. Cadiou, A., Hanjalic, K., Stawiarski, K.A.: two-scale second-moment turbulence closure based on weighted spectrum integration. Theor. Comput. Fluid Dyn. 18, 1–26 (2004)

    Article  MATH  Google Scholar 

  33. Stawiarski, K., Hanjalic, K.: On physical constraints for multi-scale turbulence closure models. Progress Comput. Fluid Dyn. 5, 120–135 (2005)

    Article  MATH  Google Scholar 

  34. Fröhlich, J., Von Terzi, D.: Hybrid LES/RANS methods for the simulation of turbulent flows. Prog. Aerosp. Sci. 44, 349–377 (2008)

    Article  Google Scholar 

  35. Argyropoulos, C.D., Markatos, N.C.: Recent advances on the numerical modelling of turbulent flows. Appl. Math. Modell. 39, 693–732 (2015)

    Article  MathSciNet  Google Scholar 

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

    Article  MATH  Google Scholar 

  37. Chaouat, B., Schiestel, R.: A new partially integrated transport model for subgrid-scale stresses and dissipation rate for turbulent developing flows. Phys. Fluids 17(065106), 1–19 (2005)

    MATH  Google Scholar 

  38. Chaouat, B., Schiestel, R.: From single-scale turbulence models to multiple-scale and subgrid-scale models by Fourier transform. Theor. Comput. Fluid Dyn. 21, 201–229 (2007)

    Article  MATH  Google Scholar 

  39. Chaouat, B., Schiestel, R.: Analytical insights into the partially integrated transport modeling method for hybrid Reynolds averaged Navier-Stokes equations-large eddy simulations of turbulent flows. Phys. Fluids 24(085106), 1–34 (2012)

    Google Scholar 

  40. 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  MATH  Google Scholar 

  41. Chaouat, B.: A new partially integrated transport modelling (PITM) method for continuous hybrid non-zonal RANS-LES simulations. In: Third Hybrid RANS-LES Methods, vol. 111, pp. 213–224. Edited by Springer (2010)

  42. Launder, B.E., Reece, G.J., Rodi, W.: Progress in the development of a Reynolds stress turbulence closure. J. Fluid Mech. 68, 537–566 (1975)

    Article  MATH  Google Scholar 

  43. Hanjalic, K., Launder, B.E.: A Reynolds stress model of turbulence and its application to thin shear flow. J. Fluid Mech. 52, 609–638 (1972)

    Article  MATH  Google Scholar 

  44. Speziale, C.G., Sarkar, S., Gatski, T.B.: Modelling the pressure-strain correlation of turbulence: an invariant dynamical system approach. J. Fluid Mech. 227, 245–272 (1991)

    Article  MATH  Google Scholar 

  45. Schiestel, R.: Sur le concept d’échelles multiples en modélisation des écoulements turbulents, Part I. J. Theor. Appl. Mech. 2(3), 417–449 (1983)

    MATH  Google Scholar 

  46. Schiestel, R.: Sur le concept d’échelles multiples en modélisation des écoulements turbulents, Part II. J. Theor. Appl. Mech. 2(4), 601–628 (1983)

    MATH  Google Scholar 

  47. Chaouat, B., Schiestel, R.: Partially integrated transport modeling method for turbulence simulation with variable filters. Phys. Fluids 25(125102), 1–39 (2013)

    MATH  Google Scholar 

  48. Hamba, F.: Analysis of filtered Navier-Stokes equation for hybrid RANS/LES simulation. Phys. Fluids 23(015108), 1–13 (2011)

    Google Scholar 

  49. Roy, P.h.: Résolution des équations de Navier-Stokes par un schéma de haute précision en espace et en temps. Rech. Aerosp. 6, 373–385 (1980)

    MATH  Google Scholar 

  50. Befeno, I., Schiestel, R.: Non-equilibrium mixing of turbulence scales using a continuous hybrid RANS/LES approach: application to the shearless mixing layer Flow. Turbul. Combust. 78, 129–151 (2007)

    Article  MATH  Google Scholar 

  51. Wengle, H, Schiestel, R., Befeno, I., Meri, A.: Large-eddy simulation of the spatial development of a shearless turbulence mixing layer. In: Hirschel, E. H., Henrich, E. (eds.) Numerical Flow simulation III, vol. 82. 271–286, Berlin Heidelberg (2003)

  52. Shur, M.L., Spalart, P.R., Strelets, M.K., Travin, A.K.: Synthetic turbulence generators for RANS-LES interfaces in zonal simulations of aerodynamic and aeroacoustic problems. Flow, Turbul. Combust. 93, 63–92 (2014)

    Article  Google Scholar 

  53. Chaouat, B.: An efficient numerical method for RANS/LES turbulent simulations using subfilter scale stress transport equations. Int. J. Numer. Meth. Fluids 67, 1207–1233 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  54. Moser, R., Kim, D., Mansour, N.: Direct numerical simulation of turbulent channel flow up to R τ = 590. Phys. Fluids 11, 943–945 (1999)

    Article  MATH  Google Scholar 

  55. Stoellinger, M., Roy, R., Heinz, S.: Unified RANS-LES method based on second-order closure. In: Proceedings of the 9th Symposium on Turbulence Shear Flow Phenomena, edited by The University of Melbourne, 7B5, pp. 1–6 (2015)

  56. Simonsen, A.J., Krogstad, P.A.: Turbulent stress invariant analysis: Classification of existing terminology. Phys. Fluids 17(088103), 1–4 (2005)

    MATH  Google Scholar 

  57. Dubief, Y., Delcayre, F.: On coherent-vortex identification in turbulence. J. Turbul. 1 N11, 1–22 (2000)

    MATH  Google Scholar 

  58. Lee, M.J., Reynolds, W.C.: Numerical experiments on the structure of homogeneous turbulence, NASA Tech. Memo NCC-2-15, Report No. TF-24. Edited by Stanford University, California (1985)

  59. Schiestel, R.: On the modelling of turbulent flows out of spectral equilibrium. C. R. Acad. Sci. Paris 302(11), 727–730 (1986)

    MATH  Google Scholar 

  60. Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P.: Numerical Recipes. Cambridge University Press (1992)

  61. Jeandel, D., Brison, J.F., Mathieu, J.: Modeling methods in physical and spectral space. Phys. Fluids 21, 169–182 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  62. Cambon, C., Jeandel, D., Mathieu, J.: Spectral modelling of homogeneous non-isotropic turbulence. J. Fluid Mech. 104, 247–262 (1981)

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bruno Chaouat.

Appendices

Appendix A: Computational Resources for DNS

It is worth evaluating, as a rough guide, the necessary computer resources for simulating the flow in the experimental wind tunnel of Uberoi and Wallis [3] at the Reynolds number R e ≈ 4.47 × 105 in terms of number of grid-points and computational times. DNS simulations require that the grid-size is at least of order of magnitude of the Kolmogorov scale η computed as η = (ν 3/𝜖)1/4. We consider here a simple model where the turbulent kinetic and the dissipation-rate evolve along the centerline of the channel with the coordinate x 1 according to the power law decay of homogeneous turbulence leading to the equations

$$ k(x_{1}) = k_{0} \left[ 1+ \frac{\epsilon_{0}} {n k_{0} U_{b}} x_{1} \right]^{-n} $$
(30)

and

$$ \epsilon(x_{1}) = \epsilon_{0} \left[ 1+ \frac{\epsilon_{0}} {n k_{0} U_{b}} x_{1} \right]^{-(n+1)} $$
(31)

where \(n= 1/(c_{{\epsilon }_{2}}-1)\) and k(x 0) = k 0, 𝜖(x 0) = 𝜖 0. In the case of an infinitesimal domain of the channel between x 1 and x 1 + d x 1, the number of grid-points d N 1(x 1)d N 2 d N 3 of the mesh is then given by

$$ dN_{1}(x_{1}) dN_{2} dN_{3} =\frac{64\, {D^{2}_{1}} dx_{1}} {\eta^{3}(x_{1})} $$
(32)

in order to describe a “minimal” sine curve on a full period using at least four grid points. The number of grid-points is then computed by integration of Eq. 32. As a result, it is found that

$$ N_{1} N_{2} N_{3}= N_{2} N_{3} {\int}_{0}^{L_{1}} dN_{1} dx_{1} = N_{2} N_{3} \frac{64\, {D^{2}_{1}}} {{\eta^{3}_{0}}} {\int}_{0}^{L_{1}} \left[ 1+ \frac{\epsilon_{0}} {n k_{0} U_{b}} x_{1} \right]^{-\frac{3(n+1)} {4}} dx_{1} $$
(33)

Hence,

$$ N_{1} N_{2} N_{3} = \frac{64\, {D^{2}_{1}}} {{\eta^{3}_{0}}} \frac{ k_{0} U_{b}} {\epsilon_{0}} F \left( \frac{ \epsilon_{0} L_{1}} {n k_{0} U_{b}} \right) $$
(34)

where

$$ F = \frac{4n} {1-3n} \left[\left( 1+ \frac{\epsilon_{0}} {n k_{0} U_{b}} L_{1} \right)^{\frac{1-3n} {4}} -1 \right] $$
(35)

The computational time is proportional to the number of grid points N 1 N 2 N 3 and the number of temporal iterations N it and the time required by the central processing unit t CPU per iteration and per grid-point, leading to the result t = N 1 N 2 N 3 N i t t CPU. The number of iteration is given by N it = T/δ t where T is the convective time allowing the eddies to move towards the exit of the channel, whereas δ t is given by the CFL condition δ t = η 0/U b that is here a more stringent constraint than the Kolmogorov time-scale \(\tau _{0}=\sqrt {\nu /\epsilon _{0}}={\eta ^{2}_{0}}/\nu \) by a factor about 100. The convective time is computed from T = L 1/U b so that the computational time is therefore given by

$$ t = N_{1} N_{2} N_{3} \frac{L_{1}} {\eta_{0}}\, t_{\text{CPU}} $$
(36)

The turbulent Reynolds number is \(R_{t}=k^{2}/\nu \epsilon = L_{e}\sqrt {k}/\nu =(L_{e}/\eta )^{4/3}\). Using the preceding values U b = 11.0 m/s, k 0 = 0.46 m2/s and 𝜖 0 = 10 m 2/s 3 the turbulence length-scale and the Kolmogorov scale are computed by \(L_{e}=\nu R_{t}/\sqrt {k}\) so that in the entrance of the channel, L e 0 ≈ 3.43 cm and η 0 ≈ 0.135 mm, respectively. From Eqs. 34 and 35, it is found that the number of grid points is order of N η ≈ 6 × 1012, and that the dimensionless computational time t/t CPU ≈ 1017. These numerical order of magnitudes clearly show that DNS and also by extension highly resolved LES are not at all affordable in term of computational resources even with the rapid increase of super computer power.

Appendix B: Generation of Anisotropic Turbulence with an Imposed Energy Spectrum

1.1 B.1 Generation of isotropic turbulence in the spectral space

The first step consists in generating an homogeneous isotropic field in a cubic box of dimension L using the method developed by Roy [49]. In this aim, we define a random vector stream function in the spectral space [61, 62]

$$ \hat{\boldsymbol{\psi}}(\boldsymbol{\kappa}) = a(\boldsymbol{\kappa}) \boldsymbol{\hat{\psi}_{1}} (\boldsymbol{\kappa}) +j b(\boldsymbol{\kappa}) \boldsymbol{\hat{\psi}_{2}}(\boldsymbol{\kappa}) $$
(37)

where \(\boldsymbol {\hat {\psi }_{1}} (\boldsymbol {\kappa })\) and \(\boldsymbol {\hat {\psi }_{2}} (\boldsymbol {\kappa })\) are two real vectors uniformly distributed on the sphere (see Fig. 22) of radius unity in the half space κ 3 ≥ 0, a and b are two functions defined as a(κ) = α 1(κ) cos(λ κ) and b(κ) = α 2(κ) sin(λ κ) , α 1 and α 2 are real functions and λ is a random number in the interval [0, 2π]. The two real random fields \(\boldsymbol {\hat {\psi }_{1}} (\boldsymbol {\kappa })\) and \(\boldsymbol {\hat {\psi }_{2}} (\boldsymbol {\kappa })\) obtained by drawing lots in the Fourier space define a complex vector stream function \(\boldsymbol {\hat {\psi }} (\boldsymbol {\kappa })\) in the spectral space. If working in spherical coordinates, the random vector Ψ n, p on the sphere of radius unity is obtained by \(\Psi ^{n,p}_{1}= \sin \theta _{n} \cos \phi _{p}\), \(\Psi ^{n,p}_{2}= \sin \theta _{n} \sin \phi _{p}\) and \(\Psi ^{n,p}_{3}= \cos \theta _{n}\), where θ n and ϕ p are the polar and azimuthal angles, respectively. The random angles are then given by ϕ p = 2π p and θ n = arccos(1 − 2n) where p and n are uniform random numbers in the interval [0,1]. The stream function is then computed in the half space κ 3 < 0 by the relation \(\hat {\boldsymbol {\psi }}(-\boldsymbol \kappa ) = \hat {\boldsymbol {\psi }}^{*}(\boldsymbol \kappa )\), where \(\hat {\boldsymbol {\psi }}^{*}\) denotes the conjugate of \(\hat {\boldsymbol {\psi }}\), to ensure that the velocity is real. The velocity is therefore obtained by the relation

$$ \hat{\boldsymbol{u}}(\boldsymbol{\kappa}) = \boldsymbol{\kappa} \wedge \hat{\boldsymbol{\psi}} (\boldsymbol{\kappa}) = \alpha_{1} (\boldsymbol{\kappa}) \cos(\lambda \boldsymbol{\kappa}) (\boldsymbol{\kappa} \wedge \boldsymbol{\hat{\psi}_{1}} (\boldsymbol{\kappa}) + j \alpha_{2} (\boldsymbol{\kappa}) \sin(\lambda \boldsymbol{\kappa}) (\boldsymbol{\kappa} \wedge \boldsymbol{\hat{\psi}_{2}} (\boldsymbol{\kappa}) $$
(38)

verifying automatically the continuity equation. In order to satisfy the energy equation,

$$ k= \frac{1} {2} \sum\limits_{\boldsymbol{\kappa}} \,\, \left<\hat{u}_{m}(\boldsymbol{\kappa}) \hat{u}^{*}_{m}(\boldsymbol{\kappa})\right> = \sum\limits_{\kappa} E(\kappa) \delta \kappa $$
(39)

one finds that the spectral velocity correlation tensor must satisfy the spectral energy equation

$$ \left< \hat{u}_{m}(\boldsymbol \kappa) \hat{u}^{*}_{m}(\boldsymbol\kappa) \right> = \left( \frac{2\pi} {L}\right)^{3} \frac{E(\kappa)} {2 \pi \kappa^{2}} $$
(40)

It is straightforward to see that the two functions a(κ) and b(κ) which appear in Eq. 37 are given by

$$ a(\boldsymbol{\kappa}) = \frac{\cos (\lambda \boldsymbol{\kappa})} {|| \boldsymbol{\kappa} \wedge \boldsymbol{\hat{\psi}_{1}} (\boldsymbol{\kappa}) ||} \left( \frac{2\pi} {L}\right)^{\frac{3} {2}} \left( \frac{E(\kappa)} {2 \pi \kappa^{2}} \right)^{\frac{1}{2}} $$
(41)
$$ b(\boldsymbol{\kappa}) = \frac{\sin(\lambda \boldsymbol{\kappa})} {|| \boldsymbol{\kappa} \wedge \boldsymbol{\hat{\psi}_{2}} (\boldsymbol{\kappa}) ||} \left( \frac{2\pi} {L}\right)^{\frac{3} {2}} \left( \frac{E(\kappa)} {2 \pi \kappa^{2}} \right)^{\frac{1}{2}} $$
(42)

where ||.|| denotes the norm. In Eq. 37, the energy spectrum density is modeled by equations

$$ E(\kappa) = \chi \kappa^{m} \,\,\,\,for\,\,\,\, \kappa < \kappa_{0} $$
(43)
$$ E(\kappa) = C_{\kappa} \epsilon^{2/3} \kappa^{-5/3} \,\,\,\, for\,\,\,\,\, \kappa \ge \kappa_{0} $$
(44)

where the coefficient χ is given by

$$ \chi = C_{\kappa} \epsilon^{2/3} \kappa^{(5+3m)/3}_{0} $$
(45)

and the matching wave number κ 0 is determined by the energy condition \(k={\int }^{\infty }_{0} E(\kappa ) d\kappa \) leading to the value

$$ \kappa_{0}= C_{\kappa}^{3/2} \frac{\epsilon} {k^{3/2}} \left[ \frac{3m+5} {2(m+1)} \right]^{3/2} $$
(46)

For κ c κ 0 where κ c is the cutoff wave number, the ratio value between the subgrid scale energy and the total energy is

$$ \frac{k_{sfs}} {k} = \frac{3(m+1)} {3m+5} \left( \frac{\kappa_{c}} {\kappa_{0}} \right)^{-2/3} $$
(47)
Fig. 22
figure 22

Random vector fields in the spectral space on the sphere of radius unity

1.2 B.2 Generation of anisotropic turbulence in the spectral space

The second step consists in applying a tensor transformation β on the isotropic field produced by the stream function \(\hat {\psi } (\boldsymbol {\kappa })\) to generate an anisotropy field. As for the isotropic case, this procedure allows to satisfy the continuity equation. The velocity is then computed by \(\hat {\boldsymbol {u}}= \boldsymbol {\kappa } \wedge \boldsymbol {\beta } \hat {\boldsymbol {\psi }} (\boldsymbol {\kappa })\). It is possible to compute the coefficient β i j by means of algebra calculus in the spectral space. The starting point consists in applying another tensor transformation α on the velocity itself leading to \(\hat {\boldsymbol {u}}= \boldsymbol {\alpha } (\boldsymbol {\kappa } \wedge \hat {\boldsymbol {\psi }} (\boldsymbol {\kappa }))\). The issue to address is then to determine the tensor β as a function of α. The resulting equation which must be solved finally reads

$$ \hat{\boldsymbol{u}}(\boldsymbol{\kappa)}= \boldsymbol{\alpha} (\boldsymbol{\kappa} \wedge \hat{\boldsymbol{\psi}} (\boldsymbol{\kappa})) = \boldsymbol{\kappa} \wedge \boldsymbol{\beta} \hat{\boldsymbol{\psi}} (\boldsymbol{\kappa}) $$
(48)

or for the i component,

$$ \hat{u}_{i}=j \kappa_{j} \alpha_{im} \epsilon_{mjk} \hat{\psi}_{k} =j \epsilon_{ijk} \kappa_{j} \beta_{km} \hat{\psi}_{m} $$
(49)

Equation 49 allows to work on the velocity as well as on the stream function field. The spectral turbulent energy associated with the wave number κ is then given by

$$ \left< \hat{u}_{i} \hat{u}^{*}_{i} \right> = \alpha_{im} \alpha_{in} \epsilon_{mjk} \epsilon_{npq} \kappa_{j} \kappa_{p} \left< \hat{\psi}_{k} \hat{\psi}^{*}_{q} \right> = \epsilon_{ijk} \epsilon_{ipn} \beta_{km} \beta_{nq} \kappa_{j} \kappa_{p} \left< \hat{\psi}_{m} \hat{\psi}^{*}_{q} \right> $$
(50)

For simplification purposes, we consider now that α and β are diagonal matrices. Moreover, as \(\hat {\boldsymbol {\psi }} (\boldsymbol {\kappa })\) is an isotropic vector, the right hand-side of Eq. 50 becomes

$$ \left< \hat{u}_{i} \hat{u}^{*}_{i} \right> = \epsilon_{ijk} \epsilon_{ipn} \beta_{kk} \beta_{kk} \kappa_{j} \kappa_{p} \left< \hat{\psi}_{k} \hat{\psi}^{*}_{k} \right> $$
(51)

or in a developed form,

$$ \left< \hat{u}_{1}\hat{u}^{*}_{1} \right> = \beta^{2}_{33} {\kappa^{2}_{2}} \left< \hat\psi_{3} \hat\psi^{*}_{3}\right>+ \beta^{2}_{22} {\kappa^{2}_{3}} \left< \hat\psi_{2} \hat\psi^{*}_{2} \right> $$
(52)
$$ \left< \hat{u}_{2} \hat{u}^{*}_{2} \right> = \beta^{2}_{11} {\kappa^{2}_{3}} \left< \hat\psi_{1} \hat\psi^{*}_{1}\right> + \beta^{2}_{33} {\kappa^{2}_{1}} \left< \hat\psi_{3} \hat\psi^{*}_{3} \right> $$
(53)
$$ \left< \hat{u}_{3} \hat{u}^{*}_{3} \right> = \beta^{2}_{22} {\kappa^{2}_{1}} \left< \hat\psi_{2} \hat\psi^{*}_{2}\right> + \beta^{2}_{11} {\kappa^{2}_{2}} \left< \hat\psi_{1} \hat\psi^{*}_{1} \right> $$
(54)

To reproduce the axisymmetric turbulence of the Uberoi and Wallis experiment [3] such as 〈u 2 u 2〉 = 〈u 3 u 3〉, Eqs. 53 and 54 imply that β 22 = β 33. Moreover, to get 〈u 1 u 1〉 > 〈u 2 u 2〉, these equations lead to β 11/β 22 < 1. Taking into account these conditions, it is found that the matrix β can be written as

$$ \mathbf{\boldsymbol{\beta}}= \left( \begin{array} {cccc} 1-2 \gamma & 0 & 0 \\ 0 & 1+ \gamma & 0 \\ 0 & 0 & 1+ \gamma \end{array} \right) $$
(55)

where γ is a numerical coefficient which must be determined to get the desired ratio anisotropy. Equation 50 allows to compute the coefficient γ and leads to the following equations

$$ \left< \hat{u}_{1} \hat{u}^{*}_{1} \right> = \alpha^{2}_{11} \left[ {\kappa^{2}_{2}} \left< \hat \psi_{3} \hat \psi^{*}_{3}\right> +{\kappa^{2}_{3}} \left< \hat\psi_{2} \hat\psi^{*}_{2}\right> \right] = \beta^{2}_{33} {\kappa^{2}_{2}} \left< \hat\psi_{3} \hat\psi^{*}_{3}\right> + \beta^{2}_{22} {\kappa^{2}_{3}} \left< \hat\psi_{2} \hat\psi^{*}_{2} \right> $$
(56)
$$ \left< \hat{u}_{2} \hat{u}^{*}_{2} \right> = \alpha^{2}_{22} \left[ {\kappa^{2}_{3}} \left< \hat \psi_{1} \hat \psi^{*}_{3}\right> +{\kappa^{2}_{1}} \left< \hat\psi_{3} \hat\psi^{*}_{3}\right> \right] = \beta^{2}_{11} {\kappa^{2}_{3}} \left< \hat\psi_{1} \hat\psi^{*}_{1}\right> + \beta^{2}_{33} {\kappa^{2}_{1}} \left< \hat\psi_{3} \hat\psi^{*}_{3} \right> $$
(57)
$$ \left< \hat{u}_{3} \hat{u}^{*}_{3} \right> = \alpha^{2}_{33} \left[ {\kappa^{2}_{1}} \left< \hat \psi_{2} \hat \psi^{*}_{2}\right> +{\kappa^{2}_{2}} \left< \hat\psi_{1} \hat\psi^{*}_{1}\right> \right] = \beta^{2}_{22} {\kappa^{2}_{1}} \left< \hat\psi_{2} \hat\psi^{*}_{2}\right> + \beta^{2}_{11} {\kappa^{2}_{2}} \left< \hat\psi_{1} \hat\psi^{*}_{1} \right> $$
(58)

Now, with the aim to obtain equations which do not depend anymore on the wave number κ, we define the spherical mean of the Fourier transform \(\mathcal {M}\) by the relation [61]

$$ \psi (\kappa) = \mathcal{M}(\psi (\boldsymbol{\kappa})) = \frac{1} {\mathcal{A}} \int \!\!\!\!\ {\int}_{\partial \mathcal{A}} \!\! \psi (\boldsymbol{\kappa}) \,\, d\mathcal{A} $$
(59)

where \(\mathcal {A}\) denotes the area element on the sphere of radius κ = |κ|. Applying this operator \(\mathcal {M}\) on Eqs. 5657 and 58 leads to the resulting equations that read

$$\begin{array}{@{}rcl@{}} && 2 \alpha^{2}_{11} = \beta^{2}_{33}+\beta^{2}_{22} \\ &&2 \alpha^{2}_{22} = \beta^{2}_{11}+\beta^{2}_{33} \\ && 2 \alpha^{2}_{33} = \beta^{2}_{11}+\beta^{2}_{22} \end{array} $$
(60)

As \(\boldsymbol {\kappa } \wedge \hat {\boldsymbol {\psi }} (\boldsymbol {\kappa })\) is an isotropic vector whereas \(\hat {\boldsymbol {u}}(\boldsymbol {\kappa }) = \boldsymbol {\alpha } \big (\boldsymbol {\kappa } \wedge \hat {\boldsymbol {\psi }}(\boldsymbol {\kappa })\big )\) is an anisotropic vector, the anisotropy of the turbulence in the spectral space can be measured by the ratio

$$ \frac{ \left< \hat{u}_{1}(\kappa) \hat{u}^{*}_{1}(\kappa) \right> } { \left< \hat{u}_{2}(\kappa) \hat{u}^{*}_{2}(\kappa) \right> } = \frac{\alpha^{2}_{11}} {\alpha^{2}_{22}} =\zeta $$
(61)

In this case, for a given value of ζ, the solving of the system (60) gives the results α 11 = 1 + γ 0, \(\alpha _{22}=\alpha _{33}=((5 {\gamma ^{2}_{0}} -2 \gamma _{0} +2)/2)^{1/2}\) and the coefficient γ 0 is solution of the second degree equation

$$ {\gamma_{0}^{2}}\left( 1-\frac{5} {2} \zeta\right) + \gamma_{0} (2+\zeta)+(1-\zeta) = 0 $$
(62)

Starting with this initial value γ 0, it is then possible to compute its exact value γ by successive approximations to get the prescribed value 〈u 1(x)u 1(x)〉 / 〈u 2(x)u 2(x)〉 = ζ.

1.3 B. 3 Return to the physical space

As usually, the velocity in the physical space u(x) is then computed from its inverse Fourier transform given by

$$ u_{i}(x_{1},x_{2},x_{3}) = \sum\limits_{\kappa_{1}=\frac{2 \pi n_{1}} {L}} \sum\limits_{\kappa_{2}=\frac{2 \pi n_{2}} {L}} \sum\limits_{ \kappa_{3}=\frac{2 \pi n_{3}} {L}} \hat{u}_{i}(\kappa_{1},\kappa_{2},\kappa_{3}) \exp{(j \kappa_{m} x_{m}) } $$
(63)

for \((n_{1},n_{2},n_{3}) \in \left [-\frac {N} {2}+1,\frac {N} {2}\right ]\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chaouat, B. Application of the PITM Method Using Inlet Synthetic Turbulence Generation for the Simulation of the Turbulent Flow in a Small Axisymmetric Contraction. Flow Turbulence Combust 98, 987–1024 (2017). https://doi.org/10.1007/s10494-016-9794-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10494-016-9794-6

Keywords

Navigation