Skip to main content
Log in

Topology optimization method based on the Wray–Agarwal turbulence model

  • Research Paper
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

One of the current challenges in fluid topology optimization is to address these turbulent flows such that industrial or more realistic fluid flow devices can be designed. Therefore, there is a need for considering turbulence models in more efficient ways into the topology optimization framework. From the three possible approaches (DNS, LES, and RANS), the RANS approach is less computationally expensive. However, when considering the RANS models that have already been considered in fluid topology optimization (Spalart–Allmaras, kε, and kω models), they all include the additional complexity of having at least two more topology optimization coefficients (normally chosen in a “trial and error” approach). Thus, in this work, the topology optimization method is formulated based on the Wray–Agarwal model (“WA2018”), which combines modeling advantages of the kε model (“freestream” modeling) and the kω model (“near-wall” modeling), and relies on the solution of a single equation, also not requiring the computation of the wall distance. Therefore, this model requires the selection of less topology optimization parameters, while also being less computationally demanding in a topology optimization iterative framework than previously considered turbulence models. A discrete design variable configuration from the TOBS approach is adopted, which enforces a binary variables solution through a linearization, making it possible to achieve clearly defined topologies (solid–fluid) (i.e., with clearly defined boundaries during the topology optimization iterations), while also lessening the dependency of the material model penalization in the optimization process (Souza et al. 2021) and possibly reducing the number of topology optimization iterations until convergence. The traditional pseudo-density material model for topology optimization is adopted with a nodal (instead of element-wise) design variable, which enables the use of a PDE-based (Helmholtz) pseudo-density filter alongside the TOBS approach. The formulation is presented for axisymmetric flows with rotation around an axis (“2D swirl flow model”). Numerical examples are presented for some turbulent 2D swirl flow configurations in order to illustrate the approach.

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

Similar content being viewed by others

References

  • Alonso DH, de Sá LFN, Saenz JSR, Silva ECN (2018) Topology optimization applied to the design of 2d swirl flow devices. Struct Multidisc Optim 58(6):2341–2364. https://doi.org/10.1007/s00158-018-2078-0

    Article  MathSciNet  Google Scholar 

  • Alonso DH, de Sá LFN, Saenz JSR, Silva ECN (2019) Topology optimization based on a two-dimensional swirl flow model of tesla-type pump devices. Comput Math Appl 77(9):2499–2533. https://doi.org/10.1016/j.camwa.2018.12.035

    Article  MathSciNet  MATH  Google Scholar 

  • Alonso DH, Garcia Rodriguez LF, Silva ECN (2021) Flexible framework for fluid topology optimization with openfoam®and finite element-based high-level discrete adjoint method (fenics/dolfin-adjoint). Struct Multidisc Optim 1:32. https://doi.org/10.1007/s00158-021-03061-4

    Article  Google Scholar 

  • Andreasen CS, Gersborg AR, Sigmund O (2009) Topology optimization of microfluidic mixers. Int J Numer Meth Fluids 61:498–513. https://doi.org/10.1002/fld.1964

    Article  MathSciNet  MATH  Google Scholar 

  • Arnold D, Brezzi F, Fortin M (1984) A stable finite element method for the stokes equations. Calcolo 21:337–344

    Article  MathSciNet  Google Scholar 

  • Bardina JE, Huang PG, Coakley TJ (1997) Turbulence modeling validation, testing and development. Tech. rep, NASA Technical Memorandum, p 110446

  • Borrvall T, Petersson J (2003) Topology optimization of fluids in stokes flow. Int J Numer Meth Fluids 41(1):77–107. https://doi.org/10.1002/fld.426

    Article  MathSciNet  MATH  Google Scholar 

  • Brezzi F, Fortin M (1991) Mixed and hybrid finite element methods. Springer, Berlin

    Book  Google Scholar 

  • Celik I (2003) RANS/LES/DES/DNS: the future prospects of turbulence modeling. J Fluids Eng 127(5):829–830. https://doi.org/10.1115/1.2033011

    Article  Google Scholar 

  • CFD group at Washington University in St Louis (2020) Wrayagarwalmodels. https://github.com/xuhanwustl/WrayAgarwalModels

  • Chen G, Xiong Q, Morris PJ, Paterson EG, Sergeev A, Wang Y (2014) Openfoam for computational fluid dynamics. Not AMS 61(4):354–363

    MathSciNet  MATH  Google Scholar 

  • COMSOL (2018) CFD Module User’s Guide, 5.4. COMSOL

  • De Chant LJ (2005) The venerable 1/7th power law turbulent velocity profile: a classical nonlinear boundary value problem solution and its relationship to stochastic processes. J Appl Math Comput Mech 161:463–474

    MathSciNet  MATH  Google Scholar 

  • Deardorff JW (1970) A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers. J Fluid Mech 41(2):453–480

    Article  Google Scholar 

  • Dilgen CB, Dilgen SB, Fuhrman DR, Sigmund O, Lazarov BS (2018) Topology optimization of turbulent flows. Comput Methods Appl Mech Eng 331:363–393. https://doi.org/10.1016/j.cma.2017.11.029

    Article  MathSciNet  MATH  Google Scholar 

  • Duan X, Li F, Qin X (2016) Topology optimization of incompressible Navier–Stokes problem by level set based adaptive mesh method. Comput Math Appl 72(4):1131–1141. https://doi.org/10.1016/j.camwa.2016.06.034

    Article  MathSciNet  MATH  Google Scholar 

  • Engin T, Özdemir M (2009) Design, testing and two-dimensional flow modeling of a multiple-disk fan. Exp Thermal Fluid Sci 33(8):1180–1187. https://doi.org/10.1016/j.expthermflusci.2009.07.007

    Article  Google Scholar 

  • Evgrafov A (2004) Topology optimization of navier-stokes equations. In: Nordic MPS 2004. The Ninth Meeting of the Nordic Section of the Mathematical Programming Society, Linköping University Electronic Press, 014, pp 37–55

  • Farrell PE, Ham DA, Funke SW, Rognes ME (2013) Automated derivation of the adjoint of high-level transient finite element programs. SIAM J Sci Comput 35(4):C369–C393

    Article  MathSciNet  Google Scholar 

  • Haftka RT, Gürdal Z (1991) Elements of structural optimization, vol 11. Springer, New York

    MATH  Google Scholar 

  • Han X, Wray TJ, Fiola C, Agarwal RK (2015) Computation of flow in s ducts with Wray–Agarwal one-equation turbulence model. J Propul Power 31(5):1338–1349

    Article  Google Scholar 

  • Han X, Wray T, Agarwal RK (2017) Application of a new des model based on wray-agarwal turbulence model for simulation of wall-bounded flows with separation. In: 47th AIAA Fluid Dynamics Conference, p 3966

  • Han X, Rahman M, Agarwal RK (2018) Development and application of wall-distance-free Wray–Agarwal turbulence model (wa2018). In: 2018 AIAA Aerospace Sciences Meeting, p 0593

  • Hasund KES (2017) Topology optimization for unsteady flow with applications in biomedical flows. Master’s thesis, NTNU

  • Huang X, Xie Y (2007) Convergent and mesh-independent solutions for the bi-directional evolutionary structural optimization method. Finite Elements Anal Design 43(14):1039–1049. https://doi.org/10.1016/j.finel.2007.06.006

    Article  Google Scholar 

  • Hyun J, Wang S, Yang S (2014) Topology optimization of the shear thinning non-Newtonian fluidic systems for minimizing wall shear stress. Comput Math Appl 67(5):1154–1170. https://doi.org/10.1016/j.camwa.2013.12.013

    Article  MathSciNet  MATH  Google Scholar 

  • Jensen KE, Szabo P, Okkels F (2012) Topology optimizatin of viscoelastic rectifiers. Appl Phys Lett 100(23):234102

    Article  Google Scholar 

  • Kontoleontos EA, Papoutsis-Kiachagias EM, Zymaris AS, Papadimitriou DI, Giannakoglou KC (2013) Adjoint-based constrained topology optimization for viscous flows, including heat transfer. Eng Optim 45(8):941–961. https://doi.org/10.1080/0305215X.2012.717074

    Article  MathSciNet  Google Scholar 

  • Langtangen HP, Logg A (2016) Solving PDEs in Minutes - The FEniCS Tutorial Volume I. https://fenicsproject.org/book/

  • Lazarov BS, Sigmund O (2010) Filters in topology optimization based on Helmholtz-type differential equations. Int J Numer Meth Eng 86(6):765–781

    Article  MathSciNet  Google Scholar 

  • Logg A, Mardal KA, Wells G (2012) Automated solution of differential equations by the finite element method: the FEniCS book, vol 84. Springer, New York

    Book  Google Scholar 

  • Mitusch S, Funke S, Dokken J (2019) Dolfin-adjoint 2018.1: automated adjoints for fenics and firedrake. J Open Source Softw 4(38):1292

    Article  Google Scholar 

  • Mortensen M, Langtangen HP, Wells GN (2011) A fenics-based programming framework for modeling turbulent flow by the Reynolds-averaged Navier–Stokes equations. Adv Water Res 34(9):1082–1101. https://doi.org/10.1016/j.advwatres.2011.02.013

    Article  Google Scholar 

  • Munson BR, Young DF, Okiishi TH (2009) Fundamentals of fluid mechanics, 6th edn. Wiley, Hoboken

    MATH  Google Scholar 

  • Nørgaard S, Sigmund O, Lazarov B (2016) Topology optimization of unsteady flow problems using the lattice Boltzmann method. J Comput Phys 307:291–307. https://doi.org/10.1016/j.jcp.2015.12.023

    Article  MathSciNet  MATH  Google Scholar 

  • Olesen LH, Okkels F, Bruus H (2006) A high-level programming-language implementation of topology optimization applied to steady-state Navier–Stokes flow. Int J Numer Meth Eng 65(7):975–1001

    Article  MathSciNet  Google Scholar 

  • CFD Online (2020) Turbulence length scale. https://www.cfd-online.com/Wiki/Turbulence_length_scale

  • OpenFOAM Wiki (2014) Openfoam guide/the simple algorithm in openfoam. http://openfoamwiki.net/index.php/The_SIMPLE_algorithm_in_OpenFOAM

  • Orszag SA (1970) Analytical theories of turbulence. J Fluid Mech 41(2):363–386

    Article  Google Scholar 

  • Papoutsis-Kiachagias E, Kontoleontos E, Zymaris A, Papadimitriou D, Giannakoglou K (2011) Constrained topology optimization for laminar and turbulent flows, including heat transfer. CIRA, editor, EUROGEN, Evolutionary and Deterministic Methods for Design, Optimization and Control, Capua, Italy

  • Papoutsis-Kiachagias EM, Giannakoglou KC (2016) Continuous adjoint methods for turbulent flows, applied to shape and topology optimization: industrial applications. Arch Comput Methods Eng 23(2):255–299

    Article  MathSciNet  Google Scholar 

  • Patankar SV (1980) Numerical heat transfer and fluid flow, 1st edn. McGraw-Hill, New York

    MATH  Google Scholar 

  • Pingen G, Maute K (2010) Optimal design for non-Newtonian flows using a topology optimization approach. Comput Math Appl 59(7):2340–2350

    Article  MathSciNet  Google Scholar 

  • Ramalingom D, Cocquet PH, Bastide A (2018) A new interpolation technique to deal with fluid–porous media interfaces for topology optimization of heat transfer. Comput Fluids 168:144–158. https://doi.org/10.1016/j.compfluid.2018.04.005

    Article  MathSciNet  MATH  Google Scholar 

  • Reddy JN, Gartling DK (2010) The finite element method in heat transfer and fluid dynamics, 3rd edn. CRC Press, Boca Raton

    Book  Google Scholar 

  • Rey Ladino AF (2004) Numerical simulation of the flow field in a friction-type turbine (tesla turbine). Diploma thesis, Institute of Thermal Powerplants, Vienna University of Technology

  • Reynolds O (1895) IV on the dynamical theory of incompressible viscous fluids and the determination of the criterion. Philos Trans R Soc Lond (A) 86:123–164

    MATH  Google Scholar 

  • Rice W (1991) Tesla turbomachinery

  • Romero J, Silva E (2014) A topology optimization approach applied to laminar flow machine rotor design. Comput Methods Appl Mech Eng 279(Supplement C):268–300. https://doi.org/10.1016/j.cma.2014.06.029

    Article  MathSciNet  MATH  Google Scholar 

  • Romero JS, Silva ECN (2017) Non-Newtonian laminar flow machine rotor design by using topology optimization. Struct Multidisc Optim 55(5):1711–1732

    Article  MathSciNet  Google Scholar 

  • Sá LF, Romero JS, Horikawa O, Silva ECN (2018) Topology optimization applied to the development of small scale pump. Struct Multidisc Optim 57(5):2045–2059. https://doi.org/10.1007/s00158-018-1966-7

    Article  MathSciNet  Google Scholar 

  • Sá LFN, Amigo RCR, Novotny AA, Silva ECN (2016) Topological derivatives applied to fluid flow channel design optimization problems. Struct Multidisc Optim 54(2):249–264. https://doi.org/10.1007/s00158-016-1399-0

    Article  MathSciNet  Google Scholar 

  • Saad T (2007) Turbulence modeling for beginners. https://www.cfd-online.com/W/images/3/31/Turbulence_Modeling_For_Beginners.pdf

  • Sato Y, Yaji K, Izui K, Yamada T, Nishiwaki S (2018) An optimum design method for a thermal-fluid device incorporating multiobjective topology optimization with an adaptive weighting scheme. J Mech Des 140(3):031402

    Article  Google Scholar 

  • Schwedes T, Ham DA, Funke SW, Piggott MD (2017) Mesh dependence in PDE-constrained optimisation—an application in tidal turbine array layouts, 1st edn. Springer, New York

    Book  Google Scholar 

  • Sivapuram R, Picelli R (2018) Topology optimization of binary structures using integer linear programming. Finite Elements Anal Design 139:49–61. https://doi.org/10.1016/j.finel.2017.10.006

    Article  MathSciNet  Google Scholar 

  • Sivapuram R, Picelli R, Xie YM (2018) Topology optimization of binary microstructures involving various non-volume constraints. Comput Mater Sci 154:405–425. https://doi.org/10.1016/j.commatsci.2018.08.008

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Sokolowski J, Zochowski A (1999) On the topological derivative in shape optimization. SIAM J Control Optim 37(4):1251–1272

    Article  MathSciNet  Google Scholar 

  • Song XG, Wang L, Baek SH, Park YC (2009) Multidisciplinary optimization of a butterfly valve. ISA Trans 48(3):370–377

    Article  Google Scholar 

  • Sonntag RE, Borgnakke C (2013) Fundamentals of thermodynamics, 8th edn. Wiley, Hoboken

    Google Scholar 

  • Souza BC, Yamabe PVM, Sá LFN, Ranjbarzadeh S, Picelli R, Silva ECN (2021) Topology optimization of fluid flow by using integer linear programming. Struct Multidisc Optim 64:1221–1240

  • Spalart PRA, Allmaras S (1994) A one-equation turbulence model for aerodynamic flows. Cla Recherche Aérospatiale 1:5–21

    Google Scholar 

  • Svanberg K (1987) The method of moving asymptotes-a new method for structural optimization. Int J Numer Methods Eng 24(2):359–373

    Article  MathSciNet  Google Scholar 

  • Sá LF, Yamabe PV, Souza BC, Silva EC (2021) Topology optimization of turbulent rotating flows using Spalart–Allmaras model. Comput Methods Appl Mech Eng 373:113551. https://doi.org/10.1016/j.cma.2020.113551

    Article  MathSciNet  MATH  Google Scholar 

  • Tesla N (1913a) Fluid propulsion. US Patent 1,061,142

  • Tesla N (1913b) Turbine. US 1(061):206

  • Tesla N (1921) Improved process of and apparatus for production of high vacua. GB Patent 179:043

  • Wächter A, Biegler LT (2006) On the implementation of an interior-point filter line-search algorithm for large-scale nonlinear programming. Math Progr 106(1):25–57

    Article  MathSciNet  Google Scholar 

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

    Article  Google Scholar 

  • White FM (2011) Fluid mechanics, 7th edn. McGraw-Hill, New York

    Google Scholar 

  • Wray T, Agarwal RK (2016) Application of the Wray–Agarwal model to compressible flows. In: 46th AIAA Fluid Dynamics Conference, p 3641

  • Wray TJ, Agarwal RK (2015) Low-Reynolds-number one-equation turbulence model based on k-\(\omega\) closure. AIAA J 53(8):2216–2227

  • Yoon GH (2016) Topology optimization for turbulent flow with Spalart–Allmaras model. Comput Methods Appl Mech Eng 303:288–311. https://doi.org/10.1016/j.cma.2016.01.014

    Article  MathSciNet  MATH  Google Scholar 

  • Yoon GH (2020) Topology optimization method with finite elements based on the k-\(\varvec \varepsilon\) turbulence model. Comput Methods Appl Mech Eng 361:112784

  • Yu H (2015) Flow design optimization of blood pumps considering hemolysis. PhD thesis, Magdeburg, Universität, Diss., 2015

  • Zauderer E (1989) Partial differential equations of applied mathematics, 2nd edn. Wiley, Hoboken

    MATH  Google Scholar 

  • Zhang X, Wray T, Agarwal RK (2016) Application of a new simple rotation and curvature correction to the wray-agarwal turbulence model. In: 46th AIAA Fluid Dynamics Conference, p 3475

  • Zhou S, Li Q (2008) A variationals level set method for the topology optimization of steady-state Navier–Stokes flow. J Comput Phys 227(24):10178–10195

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This research was partly supported by CNPq (Brazilian Research Council), FAPERJ (Research Foundation of the State of Rio de Janeiro), and FAPESP (São Paulo Research Foundation). The authors thank the supporting institutions. The first author thanks the financial support of FAPESP under Grant 2017/27049-0. The third author thanks FAPESP under the Young Investigators Awards program under Grants 2018/05797-8 and 2019/01685-3, and FUSP (University of São Paulo Foundation) under Project Numbers 314139 and 314137. The fourth author thanks the financial support of CNPq (National Council for Research and Development) under Grant 302658/2018-1 and of FAPESP under Grant 2013/24434-0. The authors also acknowledge the support of the RCGI (Research Centre for Gas Innovation), hosted by the University of São Paulo (USP) and sponsored by FAPESP (2014/50279-4) and Shell Brazil.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emílio Carlos Nelli Silva.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest

Replication of results

The descriptions of the formulation, the numerical implementation, and the numerical results contain all necessary information for reproducing the results of this article.

Additional information

Responsible Editor: Ole Sigmund

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Comparison of sensitivities with finite differences

A comparison of the computed sensitivities (using dolfin-adjoint) with finite differences is presented in this appendix. The comparison is performed for the optimized topology for the rotating nozzle for turbulent flow under 10 L/min and 2500 rpm (Sect. 6.1), by considering the simulations for laminar (0.06 L/min and 25 rpm) and turbulent (10 L/min and 2500 rpm) flows. The set of points selected for comparison with finite differences in the computational domain is shown in Fig. 25. The comparison is performed for the same configurations considered for laminar and turbulent flows in Sect. 6.1. For \(\alpha = 1\) (fluid), the finite difference approximation is considered through the backward difference approximation: \(\frac{\text{{d}}J}{\text{{d}}\alpha } = \frac{J(\alpha ) - J(\alpha - {\Delta }\alpha )}{{\Delta }\alpha }\), where \(J = {\Phi }_{\text {rel}}\). For \(\alpha = 0\) (solid), the finite difference approximation is considered through forward difference approximation: \(\frac{\text{{d}}J}{\text{{d}}\alpha } = \frac{J(\alpha + {\Delta }\alpha ) - J(\alpha )}{{\Delta }\alpha }\). The computed sensitivities are shown in Fig. 26, for a step size of \(10^{-3}\). As can be seen, the computed sensitivities for this work (by using dolfin-adjoint) and finite differences are close to each other. For a better insight about the differences between the two sensitivities, Fig. 27 depicts the relative differences as defined below, which resulted small. The computed relative difference values may be viewed in sight of the fact that smaller objective function values may hinder the computation of finite differences due to computational errors, as observed in Yoon (2020); Haftka and Gürdal (1991). Furthermore, since a discrete algorithm (not continuous) is being considered in this work, this amount of difference does not seem to pose a problem.

$$\begin{aligned} \left. r_d\right| _\text {laminar} = \left. \frac{\left. \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {FD} - \left. \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {p}}{\text {max}\left| \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {p,\ all\ points}}\right| _\text {laminar} \end{aligned}$$
(31)
$$\begin{aligned} \left. r_d\right| _\text {turbulent} = \left. \frac{\left. \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {FD} - \left. \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {p}}{\text {max}\left| \frac{\text{{d}}J}{\text{{d}}\alpha }\right| _\text {p,\ all\ points}}\right| _\text {turbulent} \end{aligned},$$
(32)

where the subscript “p” indicates the “present work” approach (by using dolfin-adjoint) and “FD” indicates “Finite Differences.” The relative differences values are higher in the turbulent case due to the higher non-linearity of the fluid flow problem.

Fig. 25
figure 25

Topology considered for the finite differences comparison

Fig. 26
figure 26

Sensitivity values computed with the approach of the present work (from FEniCS/dolfin-adjoint) and from finite differences, for laminar and turbulent flows

Fig. 27
figure 27

Relative differences for the cases shown in Fig. 26

Appendix B: 2D double pipe

The design of a 2D double pipe is performed in this Appendix, in order to show the Wray–Agarwal model (2018) being considered for topology optimization in a 2D case for a different initial guess configuration for the design variable. The topology optimization for this type of problem has been previously evaluated by Borrvall and Petersson (2003), for laminar Stokes flow. In this work, the inlets are set to be larger, which is reflected in setting the specified fluid volume fraction (f) as 50%, in order to make possible the formation of straight channels connecting the inlets to the outlets. Also, the outlet flow boundary condition is set as “stress free,” which is more generic (Hasund 2017) with respect to imposing fixed outlet velocity profiles as Borrvall and Petersson (2003). The design domain is shown in Fig. 28.

Fig. 28
figure 28

Design domain for the 2D double pipe

The mesh consists of 16,181 nodes and 32,000 elements (i.e., 100 horizontal and 80 vertical rectangular partitions of crossed triangular elements, see Fig. 29). The input parameters, geometric dimensions, and material model parameters that are considered for the design are shown in Table 5. The maximum inlet Reynolds number is 0.375 (laminar flow case) and \(2.8{\times }10^4\) (turbulent flow case). In order to consider a different configuration for the initial guess with respect to the other examples, the initial guess is chosen as shown in Fig. 30, where \(d = 37.5\) mm and \(r_c = 3.75\) mm. The TOBS approach is considered for \(\varepsilon _\text {relax} = 0.1\) and \(\beta _\text {flip\ limit} = 0.1\) (laminar flow case), and for \(\varepsilon _\text {relax} = 0.05\) and \(\beta _\text {flip\ limit} = 0.05\) (turbulent flow case).

Fig. 29
figure 29

Mesh used in the design of the 2D double pipe

Fig. 30
figure 30

Initial guess used in the design of the 2D double pipe

Table 5 Parameters used for the topology optimization of the 2D double pipe

The optimized topologies are shown in Fig. 31, where the maximum local Reynolds number is given as 0.46 (laminar flow case) and \(2.9{\times }10^4\) (turbulent flow case). As can be noticed, both optimized topologies are essentially different, where both channels join in the middle of the design domain for the laminar flow (similarly to the optimized results obtained by Borrvall and Petersson (2003)), but are kept separated for the turbulent flow. This difference in the optimized topologies shown in Fig. 31 can be viewed from the fact that the high inlet fluid flow velocity from the turbulent flow case does not allow creating a bend in the channel without dissipating significantly more energy, while the energy expended for that in the laminar flow case is minimal. This is also shown in the energy dissipation values from Table 6, where it can be seen that the optimized topologies perform better for their respective fluid flow regimes (43% better in the laminar flow case, and 46% better in the turbulent flow case).

Fig. 31
figure 31

Optimized topologies for the 2D double pipes (\(f =\) 50%)

Table 6 Energy dissipation computed for the laminar and turbulent flow optimized topologies operating under each others’ flow configurations for the 2D double pipes

Appendix C: 2D U-bend channel

The design of a 2D U-bend channel is performed in this Appendix, in order to show the Wray–Agarwal model (2018) being considered for topology optimization in a 2D case. This topology optimization problem has been previously evaluated by Dilgen et al. (2018), for the Spalart–Allmaras and k-\(\omega\) models considering the MMA (Method of Moving Asymptotes) algorithm. The 2D U-bend channel consists of an inlet and an outlet next to each other, with a “rod”-like structure in the middle of the channel that forces the optimized channel to go around it (see Fig. 32). The inlet and outlet zones, as well as the “rod”-like structure are kept as “non-optimizable.”

Fig. 32
figure 32

Design domain for the 2D U-bend channel

Since there is a “non-optimizable” zone inside the computational domain (see Fig. 32), when considering the Helmholtz pseudo-density filter, the filter is applied over the whole computational domain (\({\Omega }\)); however, the design variable value outside the design domain (\({\Omega }\setminus {\Omega }_\alpha\)) is enforced as the previous value (\(\alpha\)). Therefore, the variable \({\alpha }_{f,\text {new}}\) is used in the equations in function of the position (\({\varvec{s}}\)):

$${\alpha }_{f,\text {new}} ({\varvec{s}}) = \left\{\begin{array}{ll} \alpha_f, & \quad \text {if}\; {\varvec{s}} \in {\Omega }_{\alpha}\\ \alpha, & \quad \text{if} \;{\varvec{s}} \in {\Omega} \setminus {\Omega}_{\alpha}\end{array} \right.$$
(33)

The local Reynolds number from Eq. (25) is redefined for this case, where the characteristic length is set as the width of the inlet (L, from Fig. 32) in the place of the external diameter (“\(2 r_{{ext}}\)”). As in the other numerical examples, the fluid is being considered as water.

The mesh consists of 92,431 nodes and 184,000 elements (i.e., 230 horizontal and 200 vertical rectangular partitions of crossed triangular elements, see Fig. 33). The input parameters, geometric dimensions, and material model parameters that are considered for the design are shown in Table 7. The maximum inlet Reynolds number is \(5.4{\times }10^4\). In this work, in order to facilitate convergence, the initial guess is given from the reference topology from Fig. 34. The specified fluid volume fraction (f) is set as 30%. The TOBS approach is considered for \(\varepsilon _\text {relax} = 0.2\) and \(\beta _\text {flip\ limit} = 0.001\).

Fig. 33
figure 33

Mesh used in the design of the 2D U-bend channel

Fig. 34
figure 34

Reference topology for the 2D U-bend channel

Table 7 Parameters used for the topology optimization of the 2D U-bend channel

The optimized topology is shown in Fig. 35, with the corresponding sensitivities’ distribution shown in Fig. 36, where the maximum local Reynolds number is given as \(2.7{\times }10^6\). As can be noticed, the optimized channel is expanded outwards with respect to the “rod”-like structure. Also, the optimized topology features a wider curve than the reference topology (Fig. 34). As a comparison, the optimized topology resulted in 35% less energy dissipation (\(1.10{\times }10^2\) W/m vs. \(1.70{\times }10^2\) W/m), and 11% (\(2.72{\times }10^{-2}\) m vs. \(3.04{\times }10^{-2}\) m) less head loss than the reference topology.

Fig. 35
figure 35

Optimized topology for the 2D U-bend channel

Fig. 36
figure 36

Sensitivities in the last optimization iteration, for the optimized 2D U-bend channel

The simulation of the optimized design is shown in Fig. 37. From it, it can be noticed that the small “bumps” of the optimized topology near the inlet and outlet of the design domain do not significantly change the velocity field of the fluid flow simulation, meaning that their contribution to the energy dissipation should be small. Also, the curve in the optimized channel reduces the bending necessary for the fluid to head towards the outlet, reducing energy dissipation and head loss.

Fig. 37
figure 37

Optimized topology and variables for the optimized 2D U-bend channel

Figure 38 shows a comparison of streamlines between the simulations for the modeled solid material, for the Helmholtz filter (\(r_H =\) 1.0 mm) and for the post-processed topology. As can be noticed, some differences arise between the simulations. First, the streamlines for the modeled solid simulation show an acceleration of the fluid flow after the curve, which is not present in the simulation of the post-processed topology. This effect can also be observed in Dilgen et al. (2018) (Figs. 10c and 14b of Dilgen et al. (2018)). This seems to be a drawback of the modeled material used in fluid topology optimization (inverse permeability), since this modeled material intrinsically implies that there will be fluid flow inside the solid material, even at extremely small values, which may possibly lead to small changes in the characteristics of the solid boundaries and may affect the turbulent flow modeling. When considering the Helmholtz filter, the boundaries are blurred, which has the effect of leading the topology optimization to focus on the main flow, giving less emphasis to local effects that may possibly destabilize the fluid flow (such as some specific inclusions in the middle of the channel) or lead the topology optimization to a relatively worse local minimum.

Fig. 38
figure 38

Streamlines in the 2D U-bend channel optimized topology, by considering the simulation for the modeled solid material, the simulation when including the Helmholtz filter (\(r_H =\) 1.0 mm) and the simulation when considering the post-processed topology

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alonso, D.H., Romero Saenz, J.S., Picelli, R. et al. Topology optimization method based on the Wray–Agarwal turbulence model. Struct Multidisc Optim 65, 82 (2022). https://doi.org/10.1007/s00158-021-03106-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00158-021-03106-8

Keywords

Navigation