Skip to main content
Log in

Reynolds-Averaged, Scale-Adaptive and Large-Eddy Simulations of Premixed Bluff-Body Combustion Using the Eddy Dissipation Concept

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

Abstract

A lean premixed propane/air bluff-body stabilized flame (Volvo test rig) is calculated using the Scale-Adaptive Simulation turbulence model (SAS) and Large-Eddy simulations (LES) as well as the conventional Reynolds-averaged approach (RAS). RAS and SAS are closed by the standard k-𝜖 and the k-ω Shear Stress Transport (SST) turbulence models, respectively. The conventional Smagorinsky and the k-equation sub-grid scales models are used for the LES closure. Effects of the sub-grid scalar flux modeling using the classical gradient hypothesis and Clark’s tensor diffusivity closures both for the inert and reactive LES flows are discussed. The Eddy Dissipation Concept (EDC) is used for the turbulence-chemistry interaction. It assumes that molecular mixing and the subsequent combustion occur in the ’fine structures’ (smaller dissipative eddies, which are close to the Kolmogorov scales). Assuming the full turbulence energy cascade, the characteristic length and velocity scales of the ’fine structures’ are evaluated using different turbulence models (RAS, SAS and LES). The finite-rate chemical kinetics is taken into account by treating the ’fine structures’ as constant pressure and adiabatic homogeneous reactors, calculated as a system of ordinary-differential equations (ODEs) described by a Perfectly Stirred Reactor (PSR) concept. Several further enhancements to model the PSRs are proposed, including a new Livermore Solver (LSODA) for integrating stiff ODEs and a new correction to calculate the PSR time scales. All models have been implemented as a stand-alone application \(\text {edcPisoFoam}\) based on the OpenFOAM technology. Additionally, several RAS calculations were performed using the Turbulence Flame Speed Closure model in Ansys Fluent to assess effects of the heat losses by modeling the conjugate heat transfer between the bluff-body and the reactive flow. Effects of the turbulence Schmidt number on RAS results are discussed as well. Numerical results are compared with available experimental data. Reasonable consistency between experimental data and numerical results provided by RAS, SAS and LES is observed. In general, there is satisfactory agreement between present LES-EDC simulations, numerical results by other authors and measurements without any major modification to the EDC closure constants, which gives a quite reasonable indication on the adequacy and accuracy of the method and its further application for turbulent premixed combustion simulations.

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

Similar content being viewed by others

References

  1. Weller, H.G., Tabor, G., Jasak, H., Fureby, C.: Tensorial approach to computational continuum mechanics using object-oriented techniques. A. Comp. Phys. 12(6), 620–631 (1998)

    Article  Google Scholar 

  2. Lysenko, D.A., Ertesvåg, I.S., Rian, K.E.: Modeling of turbulent separated flows using OpenFOAM. Comput. Fluids 80, 408–422 (2013)

    Article  MATH  Google Scholar 

  3. Lilleberg, B., Christ, D., Ertesvåg, I.S., Rian, K.E., Kneer, R.: Numerical simulation with an extinction database for use with the Eddy Dissipation Concept for turbulent combustion. Flow Turbul. Combust. 91, 319–346 (2013)

    Article  Google Scholar 

  4. Lysenko, D.A., Ertesvåg, I.S., Rian, K.E.: Numerical simulation of non-premixed turbulent combustion using the eddy dissipation concept and comparing with the steady laminar flamelet model. Flow Turbul. Combust. 93, 577–605 (2014)

    Article  Google Scholar 

  5. Lysenko, D.A., Ertesvåg, I.S., Rian, K.E.: Numerical simulations of the sandia flame d using the eddy dissipation concept. Flow Turbul. Combust. 93, 665–687 (2014)

    Article  Google Scholar 

  6. Barlow, R.S., Frank, J.H.: Effects of turbulence on species mass fractions in methane/air jet flames. Proc Combust. Inst. 27, 1087–1095 (1998)

    Article  Google Scholar 

  7. Dunn, M.J., Masri, A.R., Bilger, R.W.: A new piloted premixed jet burner to study strong finite-rate chemistry effects. Combust. Flame 151(1-2), 46–60 (2007)

    Article  Google Scholar 

  8. Dunn, M.J., Masri, A.R., Bilger, R.W., Barlow, R.S., Wang, G.H.: The compositional structure of highly turbulent piloted premixed flames issuing into a hot coflow. Proc. Combust. Inst. 32(2), 1779–1786 (2009)

    Article  Google Scholar 

  9. Barlow, R.S., Fiechtner, G.J., Carter, C.D., Chen, J.-Y.: Experiments on the scalar structure of turbulent CO/H2/N2 jet flames. Combust. Flame 120, 549–569 (2000)

    Article  Google Scholar 

  10. Dally, B.B., Masri, A.R., Barlow, R.S., Fiechtner, G.J.: Instantaneous and mean compositional structure of bluff-body stabilised nonpremixed flames. Combust. Flame 114, 119–148 (1998)

    Article  Google Scholar 

  11. Launder, B., Spalding, D.: The numerical computation of turbulent flows. Comput. Methods Appl. Mech. Eng. 3(2), 269–289 (1974)

    Article  MATH  Google Scholar 

  12. Gran, I.R., Magnussen, B.F.: A numerical study of a bluff-body stabilized diffusion flame. Part 2. Influence of combustion modeling and finite-rate chemistry. Combust. Sci. Technol. 119, 191–217 (1996)

    Article  Google Scholar 

  13. Bowman, C.T., Hanson, R.K., Davidson, D.F., Gardiner, W.C., Lissianski, V., Smith, G.P., Golden, D.M., Frenklach, M., Goldenberg, M.: GRI-Mech. http://www.me.berkeley.edu/gri-mech/. Accessed February 2013 (2008)

  14. Menter, F.R., Egorov, Y.: The Scale-Adaptive Simulation method for unsteady turbulent flow predictions. Part 1, Theory and model description. Flow Turbul. Combust. 85, 113–138 (2010)

    Article  MATH  Google Scholar 

  15. Ertesvåg, I.S., Magnussen, B.F.: The eddy dissipation turbulence energy cascade model. Combust. Sci. Technol. 159, 213–235 (2000)

    Article  Google Scholar 

  16. Hairer, E., Wanner, G.: Solving Ordinary Differential Equations II: Stiff and Differential-Algebraic Problems, Springer Series in Computational Mathematics, 2nd rev. ed. Springer-Verlag, Berlin (1996)

    Book  MATH  Google Scholar 

  17. Radhakrishnan, K., Hindmarsh, A.C.: Description and use of LSODE, the Livermore solver for ordinary differential equations, Lawrence Livermore national laboratory report, UCRL-ID-113855 (1993)

  18. Sjunnesson, A., Olovsson, S., Sjöblom, B.: Validation rig – a tool for flame studies, VOLVO Aero AB S-461 81. Trollhättan, Sweden (1991)

    Google Scholar 

  19. Sjunnesson, A., Nelson, C., Max, E.: LDA measurements of velocities and turbulence in a bluff body stabilized flame, Laser Anemometry 3, ASME (1991)

  20. Sjunnesson, A., Henriksson, P., Löfström, C.: CARS measurements and visualization of reacting flows in a bluff body stabilized flame. AIAA 92–3650 (1992)

  21. Jones, W.P., Marquis, A.J., Wang, F.: Large eddy simulation of a premixed propane turbulent bluff body flame using the Eulerian stochastic field method. Fuel 140, 514–525 (2015)

    Article  Google Scholar 

  22. Ma, T., Gao, Y., Kempf, A.M., Chakraborty, N.: Validation and implementation of algebraic LES modelling of scalar dissipation rate for reaction rate closure in turbulent premixed combustion. Combust. Flame 161, 3134–3153 (2014)

    Article  Google Scholar 

  23. Manickam, B., Franke, J., Muppala, S.P.R., Dinkelacker, F.: Large-eddy simulation of triangular-stabilized lean premixed turbulent flames: quality and error assessment. Flow Turbul. Combust. 88, 563–596 (2012)

    Article  MATH  Google Scholar 

  24. Sabelnikov, V., Fureby, C.: LES combustion modeling for high Re flames using a multi-phase analogy. Combust. Flame 160, 83–96 (2013)

    Article  Google Scholar 

  25. Peters, N.: Turbulent combustion. Cambridge University Press, Cambridge (2000)

    Book  MATH  Google Scholar 

  26. Magnussen, B.F., Hjertager, B.H: On mathematical modeling of turbulent combustion with special emphasis on soot formation and combustion. Proc. Combust. Inst. 16, 719–729 (1976)

    Article  Google Scholar 

  27. Magnussen, B.F.: Modeling of NOx and soot formation by the Eddy Dissipation Concept. Int.Flame Research Foundation, 1st topic Oriented Technical Meeting., 17-19. Holland, Amsterdam (1989)

    Google Scholar 

  28. Menter, F.R., Egorov, Y.: Formulation of the Scale-Adaptive Simulation (SAS) model during the DESIDER Project. In: Haase, W., Braza, M., Revell, A. (eds.) Notes on Num. Fluid Mech Multidisc Design, 103, Springer (2009)

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

    Article  MATH  Google Scholar 

  30. Menter, F., Esch, T.: Elements of industrial heat transfer prediction. In: 16th Brazilian Congress of Mechanical Engineering (COBEM) (2001)

  31. Menter, F.R., Kuntz, M., Langtry, R.: Ten years of industrial experience with the SST turbulence model. Turbulence Heat and Mass Transfer 4, 625–632 (2003)

    Google Scholar 

  32. Yoshizawa, A.: Statistical theory for compressible shear flows, with the application to subgrid modelling. Phys. Fluids 29(2152), 1416–1429 (1986)

    Google Scholar 

  33. Smagorinsky, J.S.: General circulation experiments with primitive equations. Mon. Weather Rev. 91(3), 99–164 (1963)

    Article  Google Scholar 

  34. Sagaut, P.: Large Eddy Simulation for Incompressible Flows, 3rd ed. Springer, Berlin (2006)

    MATH  Google Scholar 

  35. Magnussen, B.F.: The Eddy Dissipation Concept a bridge between science and technology. In: ECCOMAS Thermal Conference on Computational Combustion, Lisbon (2005)

  36. Butz, D., Gao, Y., Kempf, A.M., Chakraborty, N.: Large eddy simulation of a turbulent premixed swirl flame using an algebraic scalar dissipation rate closure. Combust. Flame 162, 3180–3196 (2015)

    Article  Google Scholar 

  37. Westbrook, C.K., Dryer, F.L.: Simplified reaction mechanisms for the oxidation of hydrocarbon fuels in flames. Combust. Sci. Technol. 27, 31–43 (1981)

    Article  Google Scholar 

  38. Westbrook, C.K., Dryer, F.L.: Chemical kinetic modeling of hydrocarbon combustion. Prog. Energy Combust. Sci. 10, 1–57 (1984)

    Article  Google Scholar 

  39. http://web.eng.ucsd.edu/mae/groups/combustion/mechanism.html, Update on 2014-10-04

  40. Zimont, V.L., Lipatnikov, A.N.: A numerical model of premixed turbulent combustion of gases. Chem. Phys. Rep. 14, 993–1025 (1995)

    Google Scholar 

  41. Karpov, V.P., Lipatnikov, A.N., Zimont, V.L.: A test of an engineering model of premixed turbulent combustion. Proc. Combust. Inst. 26, 249–257 (1996)

    Article  Google Scholar 

  42. Warnatz, J., Maas, U., Dibble, R.W.: Combustion, 4th ed. Springer, Berlin (2006)

    MATH  Google Scholar 

  43. ANSYS FLUENT R13. Theory guide. Tech. rep., Ansys Inc (2013)

  44. Cheng, P.: Dynamics of a radiating gas with application to flow over a wavy wall. AIAA J. 4(2), 238–245 (1966)

    Article  Google Scholar 

  45. Smith, T.F., Shen, Z.F., Friedman, J.N.: Evaluation of coefficients for the weighted sum of gray gases model. ASME J. Heat Transfer 104(4), 602–608 (1982)

    Article  Google Scholar 

  46. Dunkle, R.V.: Geometric mean beam lengths for radiant heat transfer calculations. ASME J. Heat Transfer 86(1), 75–80 (1964)

    Article  Google Scholar 

  47. Vandoormaal, J.P., Raithby, G.D.: Enhancements of the SIMPLE method for predicting incompressible fluid flows. Numer. Heat Transfer 7, 147–163 (1984)

    MATH  Google Scholar 

  48. Issa, R.: Solution of the implicitly discretized fluid flow equations by operator splitting. J. Comput. Phys. 62, 40–65 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  49. Waterson, N.P., Deconinck, H.: Design principles for bounded higher-order convection schemes – a unified approach. J. Comput. Phys. 224, 182–207 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  50. Harten, A.: High resolution schemes for hyperbolic conservation laws. J. Comput. Phys. 49, 357–393 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  51. Jasak, H., Weller, H.G., Gosman, A.D.: High resolution NVD differencing scheme for arbitrarily unstructured meshes. Int. J. Numer. Meth. Fluids 31, 431–449 (1999)

    Article  MATH  Google Scholar 

  52. Geurts, B.: Elements of direct and large-eddy simulation. R.T.Edwards, Philadelphia (2004)

    Google Scholar 

  53. Rhie, C., Chow, W.: Numerical study of the turbulent flow past an airfoil with trailing edge separation. AIAA J. 21, 1525–32 (1983)

    Article  MATH  Google Scholar 

  54. Lysenko, D.A., Ertesvåg, I.S., Rian, K.E.: Large-eddy simulation of the flow over a circular cylinder at Reynolds number 3900 using the OpenFOAM toolbox. Flow Turbul. Combust. 89, 491–518 (2012)

    Article  Google Scholar 

  55. Lysenko, D.A., Ertesvåg, I.S., Rian, K.E.: Large-eddy simulation of the flow over a circular cylinder at Reynolds number 2 × 104. Flow Turbul. Combust. 92, 673–698 (2014)

    Article  Google Scholar 

  56. Sanquer, S., Bruel, P., Deshaies, B.: Some specific characteristics of turbulence in the reactive wakes of bluff bodies. AIAA J. 36(6), 994–1001 (1998)

    Article  Google Scholar 

  57. Hasse, C., Sohm, V., Wetzel, M., Durst, B.: Hybrid URANS/LES turbulence simulation of vortex shedding behind a triangular flameholder. Flow Turbul. Combust. 83, 1–20 (2009)

    Article  MATH  Google Scholar 

  58. Shanbhogue, S.J., Husain, S., Lieuwen, T.: Lean blowoff of bluff body stabilized flames: scaling and dynamics, Prog. Energy Combust. Sci. 35, 98–120 (2009)

    Google Scholar 

  59. Welch, P.: The use of fast Fourier transform for the estimation of power spectra: a method based on time averaging over short, modified periodograms. IEEE Trans. Audio Electroacoust. 15(6), 70–73 (1967)

    Article  Google Scholar 

  60. Cao, Y., Tamura, T.: Large-eddy simulations of flow past a square cylinder using structured and unstructured grids. Comput. Fluids 137, 36–54 (2016)

    Article  MathSciNet  Google Scholar 

  61. Yasari, E., Verma, S., Lipatnikov, A.N.: RANS simulations of statistically stationary premixed turbulent combustion using flame speed closure model. Flow Turbul. Combust. 94, 381–414 (2015)

    Article  Google Scholar 

  62. Colin, O., Ducros, F., Veynante, D., Poinsot, T.: A thickened flame model for large eddy simulations of turbulent premixed combustion. Phys. Fluids 12, 1843–1863 (2000)

    Article  MATH  Google Scholar 

  63. Sathiah, P., Lipatnikov, A.: Effects of flame development on stationary premixed turbulent combustion. Proc. Comb. Inst. 31, 3115–3122 (2007)

    Article  Google Scholar 

  64. Baudoin, E., Bai, R.Yu., Nogenmyr, K.J., Bai, X.S., Fureby, C.: Comparison of LES models applied to a bluff body stabilized flame. AIAA 2009–1178 (2009)

  65. Allauddin, U., Klein, M., Pfitzner, M., Chakraborty, N.: A priori and a posteriori analyses of algebraic flame surface density modeling in the context of Large Eddy Simulation of turbulent premixed combustion. Numer Heat Transfer, Part A: Appl. 71(2), 153–171 (2017)

    Article  Google Scholar 

  66. Klein, M., Chakraborty, N., Pfitzner, M.: Analysis of the combined modelling of sub-grid transport and filtered flame propagation for premixed turbulent combustion. Flow Turbul. Combust. 96, 921–938 (2016)

    Article  Google Scholar 

  67. Clark, R.A., Ferziger, J.H., Reynolds, W.C.: Evaluation of subgrid-scale models using an accurately simulated turbulent flow. J. Fluid Mech. 91, 1–16 (1979)

    Article  MATH  Google Scholar 

  68. O’Malley, R.E.: Singular perturbation methods for ordinary differential equations. Springer-Verlag, New York (1991)

    Book  MATH  Google Scholar 

  69. Shampine, L.F., Reichelt, M.W.: The MATLAB ODE Suite, SIAM. J. Sci. Comput. 18, 1–22 (1997)

    MathSciNet  MATH  Google Scholar 

  70. Robertson, H.H.: The Solution of a set of reaction rate equations. In: Walsh, J. (ed.) Numerical Analysis: an Introduction, pp. 178-182. Academic Press, London (1966)

  71. Gobbert, M.K.: Robertson’s example for stiff differential equations. Arizona State University, Technical report (1996)

    Google Scholar 

Download references

Acknowledgements

We are grateful to the Norwegian Meta center for Computational Science (NOTUR) for providing the uninterrupted HPC computational resources and useful technical support. Comments and recommendations for three anonymous and very skilled reviewers of the Journal have increased considerably the quality of the paper.

Funding

Except for the computer allowance acknowledged above, this study has not received any funding.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitry A. Lysenko.

Ethics declarations

Conflict of interests

The authors declare that they have no conflict of interest.

Appendices

Appendix A: General Validation of the LSODA and RADAU5 ODEs Solvers

Here, the validation and verification of the RADAU5 [16] and LSODA [17] integrators are provided. Three well-known non-stiff and stiff test problems were chosen for this purpose. The first test is a simple model of flame propagation, which has been introduced in the Matlab ODE suite [68, 69]. The second benchmark is the Van der Pol equation. The third test is the Robertson problem, which consists of a stiff system of three non-linear ODEs for chemical kinetics proposed by Robertson [70]. For sake of completeness, the validation has been provided for both implemented ODE integrators as well as against the state-of-the-art ODE suite available in Matlab [69].

All test cases have been computed with the same absolute and relative error tolerances equal to \(10^{-8}\).

1.1 A.1 One dimensional flame propagation

A mathematical model of one-dimensional flame propagation [68] can be described as

$$ \frac{dy}{dt} = y^{2} - y^{3}, \;\; y(0) = \delta, \;\; \left[ 0 < t < \frac{2}{\delta} \right], $$
(49)

where y represents the radius of the sphere, t is time and terms y 2 and y 3 come from the surface area and the volume. After igniting, the sphere grows rapidly until it reaches a critical size. Then, the radius of sphere stays at the same size because the amount of oxygen being consumed while combustion in the interior of the sphere balances the amount available through the surface. In this case the critical parameter is the initial radius δ, which determines the stiffness of the problem. A solution becomes stiff near y(t) = 1, increasing or decreasing rapidly toward that solution for small values of δ.

Figure 24 presents the results of integration of this problem for δ = 0.01 and δ = 0.0001. As can be seen, the deviations between all integrators were negligible. All codes computed the problem for both initial conditions.

Fig. 24
figure 24

Computed solution of the flame propagation problem obtained by RADAU5, LSODA and ode23td (Matlab) solvers

1.2 A.2 Van der Pol equation

The van der Pol equation is a second order ODE:

$$ \frac{d^{2}y_{1}}{dt^{2}} -\mu \left( 1- {y_{1}^{2}} \right) \frac{dy_{1}}{dt} + y_{1} = 0, $$
(50)

where \(\mu >0\) is a scalar parameter.

The system of first-order equations can be obtained by making the substitution \(dy_{1} / dt = y_{2}\):

$$ \left[ \begin{array}{c} dy_{1} /dt \\ dy_{2} /dt \end{array} \right] = \begin{array}{ll} y_{2} \\ \mu \left( 1 - {y_{1}^{2}} \right) y_{2} - y_{1} \end{array}, \;\; \left[ \begin{array}{c} y_{1}(0) = 2 \\ y_{2}(0) = 0 \end{array} \right], $$
(51)

The nonstiff system (μ = 1) was computed on the time interval [0 20], while the stiff system (μ = 1000) was calculated on the time interval [0 3000]. Figure 25 displays the computed solutions obtained by RADAU5, LSODA and Matlab. All three solutions collapsed well to each other without any significant deviations.

Fig. 25
figure 25

Computed solutions of the nonstiff (μ = 1) and stiff (μ = 1000) van der Pol equation by RADAU5, LSODA and ode23tb (Matlab) solvers

1.3 A.3 Robertson problem

This problem deals with a system of ODEs that describes the kinetics of an auto-catalytic reaction [70]. The structure of the reactions is

$$\begin{array}{@{}rcl@{}} &&A \overset{k_{1}}{\rightarrow} B\\ &&B + B \overset{k_{2}}{\rightarrow} B + C\\ &&B + C \overset{k_{3}}{\rightarrow} A + C. \end{array} $$
(52)

Under some idealized assumptions [71], the following mathematical model can be set up as a set of three ODEs

$$ \left[ \begin{array}{c} dy_{1} /dt \\ dy_{2} /dt \\ dy_{3} /dt \end{array} \right] = \begin{array}{lllllll} -k_{1} \cdot y_{1} & & k_{3} \cdot y_{2} \cdot y_{3} \\ k_{1} \cdot y_{1} & -k_{2} \cdot {y^{2}_{2}} & -k_{3} \cdot y_{2} \cdot y_{3} \\ & k_{2} \cdot {y^{2}_{2}} & \end{array}, \;\; \left[ \begin{array}{c} y_{1}(0) = 1 \\ y_{2}(0) = 0 \\ y_{3}(0) = 0 \end{array} \right], $$
(53)

where y 1, y 2, y 3 are the concentrations of species \(A,B,C\) respectively. The numerical values of the rate constants were k 1 = 0.04, \(k_{2} = 3 \times 10^{7}\) and \(k_{3} = 10^{4}\). The large differences among the reaction rate constants provide the reason for stiffness. Originally the problem was proposed on the time interval \(0 < t \leq 40\), but it is convenient to extend the integration of solution on much longer intervals due to that many codes fail if t becomes very large.

Figure 26 presents the solutions obtained by RADAU5, LSODA and Matlab. Overall, the discrepancies between calculated solutions were negligible for y 1 and y 3. Small deviations between RADAU5, LSODA and Matlab integrators can be observed for y 2, which probably could be explained by the fact that Matlab has solved the rewritten system of differential algebraic equations by using the conservation law in order to determine the state of y 3, meanwhile RADAU5 and LSODA calculated the original system of equations.

Fig. 26
figure 26

Computed solution of the Robertson problem by RADAU5, LSODA and ode15s (Matlab) solvers

Appendix B: Effects of Sub-grid Scaling Modeling on the Inert flow for the Volvo test rig

Three non-reacting LES runs have been carried out to investigate the effects of the SGS models. For this purpose the k-equation and the Smagorinsky model with two different constants (C s = 0.1 and \(C_{s}= 0.053\), respectively were applied. For a quantitative validation of the present LES simulations, the averages have been computed by sampling over 50 vortex shedding periods (N v s ).

Figure 27 shows the measured and predicted mean stream-wise velocity and its fluctuation as well as the normalized turbulence kinetic energy along the central-line behind the obstacle. In general, all three LES runs matched the experimental data by Sjunnesson et al. [19] reasonably well. The discrepancies between all SGS models were negligible. This finding was supported by the fact that all three runs revealed comparatively the same flow patterns shown in Fig. 28. The differences between the LES and the SASI3 results were small as well. The recirculation zone length was predicted as \(<L_{r}>/H = 1.28\) for all LES runs (and the same as for SASI3), which was in a fairly good agreement with experimental data of Sjunnesson et al. [19], \(<L_{r}>/H = 1.35\).

Fig. 27
figure 27

Normalized mean stream-wise velocity (a), its fluctuations (b) and and normalized turbulence kinetic energy (c) in the wake centerline for the Volvo test rig

Fig. 28
figure 28

Flow structures for the Volvo test rig. Iso-surface of the Q-criterion, Q = 1 × 105

Figure 29 compares one-dimensional frequency spectra extracted from the present solutions at the downstream location x/H = 1.75 on the centerline of the wake. About \(6\times 10^{5}\) samples of the cross-flow velocities were collected (or \(N_{vp} \approx 50\)). For sake of completeness, the spectrum obtained by SASI3 was added to assess the dissipative properties of the SAS and LES results. It can be seen clearly that the spectra obtained by LES collapsed well and had the similar distribution, meanwhile the SAS spectrum became more dissipative after f/f v s = 2.5. The Strouhal numbers were \(\text {St} = 0.27\), 0.29 and 0.28 for the LESI1, LESI2 and LESI3 runs, respectively. These values were in reasonable agreement with the experimental data by Sjunnesson et al. [19] and Sanquer et al. [56], who measured St = 0.25 and \(\text {St} = 0.26\), and corresponded well with the SAS runs.

Fig. 29
figure 29

One-dimensional spectra of the transverse velocity in the wake for the Volvo test rig: non reactive LES results

Appendix C: Effects of the Sub-grid Scalar flux Modeling for the Volvo test rig

Here, additional inert and reactive LES cases are considered in conjunction with the SGSF closure based on the classical gradient hypothesis closure and Clark’s tensor diffusivity model.

The LESI1 case was chosen as baseline to investigate the influence of the SGSF modeling. As the first step, the inert LES run (LESI1a) was calculated where the diffusion term in Eq. 14 was replaced in the spirit of Clark’s model [67] as

$$ F_{d} = \frac{{\Delta}^{2}}{12} \frac{\partial}{\partial x_{j}} \left( \frac{\partial\tilde{u_{j}}}{\partial x_{j}} \frac{\partial \tilde{k}}{\partial x_{j}} \right). $$
(54)

As the second step, the sub-grid scalar flux was replaced in the energy transport equation as

$$ \text{\textbf{b}}_{h} = \frac{{\Delta}^{2}}{12} \left( \frac{\partial \tilde{u_{j}}}{\partial x_{j}} \frac{\partial \tilde{h}}{\partial x_{j}} \right), $$
(55)

and the inert LES run was performed including both modifications (LESI1b).

Figure 30 compares three cases. Both LESI1a and LESI1b were calculated using identical setup as LESI1. It is clearly seen that differences between all cases were minor, and no clear advantage could be seen when using the particular SGSF model.

Fig. 30
figure 30

Normalized mean stream-wise velocity and its fluctuations obtained by LESI1, LESI1a and LESI1b runs in the wake of the Volvo test rig

As the third step, the sub-grid scalar flux was replaced in the species transport equation as

$$ \text{\textbf{b}}_{s} = \frac{{\Delta}^{2}}{12} \left( \frac{\partial \tilde{u_{j}}}{\partial x_{j}} \frac{\partial \tilde{Y_{s}}}{\partial x_{j}} \right). $$
(56)

The LESR1 case was chosen to replicate simulations using both Eqs. 5556 (LESR1a). Figure 31 displays comparison of the normalized, mean temperature and its fluctuations at three axial stages x/H = 0.95, 3.75 and 9.4. In general, the discrepancies between two cases are small. The most pronounced difference was observed at x/H = 0.95, where the LESR1a case provided the slightly lower peak temperature without impulses in the shear layer regions. The minor differences related to the temperature fluctuations were pronounced as well, however, had the same qualitative trends as the baseline case. The species mole fraction results in Fig. 31 were marginally affected, similar to the line thickness or less.

Fig. 31
figure 31

Normalized mean stream-wise temperature and its fluctuations obtained by LESR1 and LESR1a runs in the wake of the Volvo test rig

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lysenko, D.A., Ertesvåg, I.S. Reynolds-Averaged, Scale-Adaptive and Large-Eddy Simulations of Premixed Bluff-Body Combustion Using the Eddy Dissipation Concept. Flow Turbulence Combust 100, 721–768 (2018). https://doi.org/10.1007/s10494-017-9880-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10494-017-9880-4

Keywords

Navigation