Skip to main content

Advertisement

Log in

An adaptive mesh refinement approach based on optimal sparse sensing

  • Original Article
  • Published:
Theoretical and Computational Fluid Dynamics Aims and scope Submit manuscript

Abstract

We introduce a new approach for adaptive mesh refinement in which adaptivity is driven by low rank decomposition and optimal sensing of the dynamically evolving flow field. This method seeks an ordered set of locations for mesh adaptation from the instantaneous data-driven basis of an online proper orthogonal decomposition of the velocity, which organizes features into sparse optimal orthogonal modes based on an energy norm. The sensing is achieved via a computationally expedient discrete empirical interpolation method using rank-revealing QR factorization (Drmac and Gugercin SIAM J Sci Comput 38(2):A631–A648, 2016). The methodology is applicable to a wide range of numerical discretizations, and is tested on a spatiotemporally evolving incompressible turbulent jet, a complex wind turbine wake, and supersonic flow over a forward-facing step. The proposed approach is demonstrated to predict accurate velocity statistics and yield significantly smaller grids in comparison to gradient-based methods. The algorithm is seen to focus refinement in the vicinity of dynamically significant regions such as those characterized by high turbulence kinetic energy, coherent structures and shock interactions. Moreover, the approach does not require parameters or thresholds, which may be difficult to obtain for complex flows, to be known a priori to facilitate mesh adaptation.

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.

Institutional subscriptions

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

Similar content being viewed by others

References

  1. Alauzet, F., George, P.L., Mohammadi, B., Frey, P., Borouchaki, H.: Transient fixed point-based unstructured mesh adaptation. Int. J. Numer. Methods Fluids 43(6–7), 729–745 (2003)

    MathSciNet  MATH  Google Scholar 

  2. Alla, A., Kutz, J.N.: Randomized model order reduction. arXiv preprint arXiv:1611.02316 (2016)

  3. Bai, Z., Wimalajeewa, T., Berger, Z., Wang, G., Glauser, M., Varshney, P.K.: Physics based compressive sensing approach applied to airfoil data collection and analysis. In: 51st AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition 2013 (2013)

  4. Baraniuk, R.G., Cevher, V., Duarte, M.F., Hegde, C.: Model-based compressive sensing. IEEE Trans. Inf. Theory 56(4), 1982–2001 (2010)

    MathSciNet  MATH  Google Scholar 

  5. Barrault, M., Maday, Y., Nguyen, N.C., Patera, A.T.: An ‘empirical interpolation’ method: application to efficient reduced-basis discretization of partial differential equations. C. R. Math. 339(9), 667–672 (2004)

    MathSciNet  MATH  Google Scholar 

  6. Bechmann, A., Sørensen, N.N., Zahle, F.: CFD simulations of the MEXICO rotor. Wind Energy 14(5), 677–689 (2011)

    Google Scholar 

  7. Berger, M., Rigoutsos, I.: An algorithm for point clustering and grid generation. IEEE Trans. Syst. Man Cybern. 21(5), 1278–1286 (1991)

    Google Scholar 

  8. Berger, M.J., Colella, P.: Local adaptive mesh refinement for shock hydrodynamics. J. Comput. Phys. 82(1), 64–84 (1989)

    MATH  Google Scholar 

  9. Berger, M.J., Oliger, J.: Adaptive mesh refinement for hyperbolic partial differential equations. J. Comput. Phys. 53(3), 484–512 (1984)

    MathSciNet  MATH  Google Scholar 

  10. Bright, I., Lin, G., Kutz, J.N.: Compressive sensing based machine learning strategy for characterizing the flow around a cylinder with limited pressure measurements. Phys. Fluids 25(12), 127102 (2013)

    MATH  Google Scholar 

  11. Castro-Díaz, M., Hecht, F., Mohammadi, B., Pironneau, O.: Anisotropic unstructured mesh adaption for flow simulations. Int. J. Numer. Methods Fluids 25(4), 475–491 (1997)

    MathSciNet  MATH  Google Scholar 

  12. Chamorro, L.P., Arndt, R., Sotiropoulos, F.: Reynolds number dependence of turbulence statistics in the wake of wind turbines. Wind Energy 15(5), 733–742 (2012)

    Google Scholar 

  13. Chaturantabut, S., Sorensen, D.C.: Nonlinear model reduction via discrete empirical interpolation. SIAM J. Sci. Comput. 32(5), 2737–2764 (2010)

    MathSciNet  MATH  Google Scholar 

  14. Colella, P.: Multidimensional upwind methods for hyperbolic conservation laws. J. Comput. Phys. 87(1), 171–200 (1990)

    MathSciNet  MATH  Google Scholar 

  15. Colella, P., Glaz, H.M.: Efficient solution algorithms for the Riemann problem for real gases. J. Comput. Phys. 59(2), 264–289 (1985)

    MathSciNet  MATH  Google Scholar 

  16. Compere, G., Marchandise, E., Remacle, J.F.: Transient adaptivity applied to two-phase incompressible flows. J. Comput. Phys. 227(3), 1923–1942 (2008)

    MathSciNet  MATH  Google Scholar 

  17. da Silva, C.B., Métais, O.: On the influence of coherent structures upon interscale interactions in turbulent plane jets. J. Fluid Mech. 473, 103–145 (2002)

    MathSciNet  MATH  Google Scholar 

  18. Donoho, D.L.: Compressed sensing. IEEE Trans. Inf. Theory 52(4), 1289–1306 (2006)

    MathSciNet  MATH  Google Scholar 

  19. Donoho, D.L., Gavish, M.: The optimal hard threshold for singular values is 4/\(\sqrt{3}\). IEEE Trans. Inf. Theory 60(8), 5040–5053 (2014)

    MathSciNet  MATH  Google Scholar 

  20. Drmac, Z., Gugercin, S.: A new selection operator for the discrete empirical interpolation method—improved a priori error bound and extensions. SIAM J. Sci. Comput. 38(2), A631–A648 (2016)

    MathSciNet  MATH  Google Scholar 

  21. Fidkowski, K.J., Darmofal, D.L.: Review of output-based error estimation and mesh adaptation in computational fluid dynamics. AIAA J. 49(4), 673–694 (2011)

    Google Scholar 

  22. Fidkowski, K.J., Roe, P.L.: An entropy adjoint approach to mesh refinement. SIAM J. Sci. Comput. 32(3), 1261–1287 (2010)

    MathSciNet  MATH  Google Scholar 

  23. Foti, D., Duraisamy, K.: An investigation of an implicit large-eddy simulation framework for the vorticity transport equations. In: 2018 AIAA Fluid Dynamics Conference, p. 3407 (2018)

  24. Foti, D., Duraisamy, K.: Multi-dimensional finite volume scheme for the vorticity transport equations. Comput. Fluids 167, 17–32 (2018)

    MathSciNet  MATH  Google Scholar 

  25. Foti, D., Duraisamy, K.: Implicit large-eddy simulation of wind turbine wakes and turbine-wake interactions using the vorticity transport equations. In: AIAA Aviation 2019 Forum, p. 2841 (2019)

  26. Foti, D., Yang, X., Campagnolo, F., Maniaci, D., Sotiropoulos, F.: Wake meandering of a model wind turbine operating in two different regimes. Phys. Rev. Fluids 3(5), 054607 (2018)

    Google Scholar 

  27. Foti, D., Yang, X., Guala, M., Sotiropoulos, F.: Wake meandering statistics of a model wind turbine: insights gained by large eddy simulations. Phys. Rev. Fluids 1(4), 044407 (2016)

    Google Scholar 

  28. Foti, D., Yang, X., Sotiropoulos, F.: Similarity of wake meandering for different wind turbine designs for different scales. J. Fluid Mech. 842, 5–25 (2018)

    MathSciNet  MATH  Google Scholar 

  29. Fowler, J.E.: Compressive-projection principal component analysis. IEEE Trans. Image Process. 18(10), 2230–2242 (2009)

    MathSciNet  MATH  Google Scholar 

  30. Golub, G., Kahan, W.: Calculating the singular values and pseudo-inverse of a matrix. J. Soc. Ind. Appl. Math. Ser. B Numer. Anal. 2(2), 205–224 (1965)

    MathSciNet  MATH  Google Scholar 

  31. Goreinov, S.A., Tyrtyshnikov, E.E., Zamarashkin, N.L.: A theory of pseudoskeleton approximations. Linear Algebra Appl. 261(1–3), 1–21 (1997)

    MathSciNet  MATH  Google Scholar 

  32. Gutmark, E., Wygnanski, I.: The planar turbulent jet. J. Fluid Mech. 73(3), 465–495 (1976)

    Google Scholar 

  33. Hartmann, R., Held, J., Leicht, T., Prill, F.: Error estimation and adaptive mesh refinement for aerodynamic flows. In: ADIGMA-A European Initiative on the Development of Adaptive Higher-Order Variational Methods for Aerospace Applications. Springer, pp. 339–353 (2010)

  34. Holmes, P., Lumley, J.L., Berkooz, G., Rowley, C.W.: Turbulence, Coherent Structures, Dynamical Systems and Symmetry. Cambridge University Press, Cambridge (2012)

    MATH  Google Scholar 

  35. Hornung, R.D., Kohn, S.R.: The use of object-oriented design patterns in the SAMRAI structured AMR framework. In: Proceedings of the SIAM Workshop on Object-Oriented Methods for Inter-Operable Scientific and Engineering Computing. Citeseer (1998)

  36. Hornung, R.D., Kohn, S.R.: Managing application complexity in the SAMRAI object-oriented framework. Concurr. Comput. Pract. Exp. 14(5), 347–368 (2002)

    MATH  Google Scholar 

  37. Ivanell, S., Mikkelsen, R., Sørensen, J.N., Henningson, D.: Stability analysis of the tip vortices of a wind turbine. Wind Energy 13(8), 705–715 (2010)

    Google Scholar 

  38. Jolliffe, I.T.: Discarding variables in a principal component analysis. I: artificial data. J. Roy. Stat. Soc. Ser. C (Appl. Stat.) 21(2), 160–173 (1972)

    MathSciNet  Google Scholar 

  39. Jolliffe, I.T.: Discarding variables in a principal component analysis. II: real data. J. Roy. Stat. Soc. Ser. C (Appl. Stat.) 22(1), 21–31 (1973)

    MathSciNet  Google Scholar 

  40. Jones, W., Nielsen, E., Park, M.: Validation of 3D adjoint based error estimation and mesh adaptation for sonic boom prediction. In: 44th AIAA Aerospace Sciences Meeting and Exhibit, p. 1150 (2006)

  41. Kouhi, M., Oñate, E., Mavriplis, D.: Adjoint-based adaptive finite element method for the compressible Euler equations using finite calculus. Aerosp. Sci. Technol. 46, 422–435 (2015)

    Google Scholar 

  42. Kutz, J.N.: Data-Driven Modeling & Scientific Computation: Methods for Complex Systems & Big Data. Oxford University Press, Oxford (2013)

    MATH  Google Scholar 

  43. Leicht, T., Hartmann, R.: Error estimation and anisotropic mesh refinement for 3D laminar aerodynamic flow simulations. J. Comput. Phys. 229(19), 7344–7360 (2010)

    MathSciNet  MATH  Google Scholar 

  44. Lumley, J.L.: Stochastic Tools in Turbulence. Applied Mathematics and Mechanics, vol. 12. Academic Press, New York (1970)

    MATH  Google Scholar 

  45. Manohar, K., Brunton, B.W., Kutz, J.N., Brunton, S.L.: Data-driven sparse sensor placement for reconstruction: demonstrating the benefits of exploiting known patterns. IEEE Control Syst. 38(3), 63–86 (2018)

    MathSciNet  Google Scholar 

  46. Nemec, M., Aftosmis, M., Wintzer, M.: Adjoint-based adaptive mesh refinement for complex geometries. In: 46th AIAA Aerospace Sciences Meeting and Exhibit, p. 725 (2008)

  47. Noack, B.R., Papas, P., Monkewitz, P.A.: The need for a pressure-term representation in empirical galerkin models of incompressible shear flows. J. Fluid Mech. 523, 339–365 (2005)

    MathSciNet  MATH  Google Scholar 

  48. Pain, C., Umpleby, A., De Oliveira, C., Goddard, A.: Tetrahedral mesh optimisation and adaptivity for steady-state and transient finite element calculations. Comput. Methods Appl. Mech. Eng. 190(29–30), 3771–3796 (2001)

    MATH  Google Scholar 

  49. Papailiopoulos, D., Kyrillidis, A., Boutsidis, C.: Provable deterministic leverage score sampling. In: Proceedings of the 20th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pp. 997–1006 (2014)

  50. Parish, E., Duraisamy, K., Chandrashekar, P.: Generalized Riemann problem-based upwind scheme for the vorticity transport equations. Comput. Fluids 132, 10–18 (2016)

    MathSciNet  MATH  Google Scholar 

  51. Peherstorfer, B., Drmač, Z., Gugercin, S.: Stabilizing discrete empirical interpolation via randomized and deterministic oversampling. arXiv preprint arXiv:1808.10473 (2018)

  52. Peraire, J., Vahdati, M., Morgan, K., Zienkiewicz, O.C.: Adaptive remeshing for compressible flow computations. J. Comput. Phys. 72(2), 449–466 (1987)

    MATH  Google Scholar 

  53. Ramaprian, B., Chandrasekhara, M.: LDA measurements in plane turbulent jets. J. Fluids Eng. 107(2), 264–271 (1985)

    Google Scholar 

  54. Richter, T.: A posteriori error estimation and anisotropy detection with the dual-weighted residual method. Int. J. Numer. Methods Fluids 62(1), 90–118 (2010)

    MathSciNet  MATH  Google Scholar 

  55. Shi, L., Wang, Z.J.: Adjoint-based error estimation and mesh adaptation for the correction procedure via reconstruction method. J. Comput. Phys. 295, 261–284 (2015)

    MathSciNet  MATH  Google Scholar 

  56. Sirovich, L.: Turbulence and the dynamics of coherent structures. I. Coherent structures. Q. Appl. Math. 45(3), 561–571 (1987)

    MathSciNet  MATH  Google Scholar 

  57. Snel, H., Schepers, J., Montgomerie, B.: The MEXICO project (model experiments in controlled conditions): the database and first results of data processing and interpretation. J. Phys. Conf. Ser. 75, 012014 (2007)

    Google Scholar 

  58. Sorensen, J.N., Shen, W.Z.: Numerical modeling of wind turbine wakes. J. Fluids Eng. 124(2), 393–399 (2002)

    Google Scholar 

  59. Stanley, S., Sarkar, S.: Influence of nozzle conditions and discrete forcing on turbulent planar jets. AIAA J. 38(9), 1615–1623 (2000)

    Google Scholar 

  60. Stanley, S., Sarkar, S., Mellado, J.: A study of the flow-field evolution and mixing in a planar turbulent jet using direct numerical simulation. J. Fluid Mech. 450, 377–407 (2002)

    MATH  Google Scholar 

  61. Troldborg, N., Sørensen, J.N., Mikkelsen, R.: Actuator line simulation of wake of wind turbine operating in turbulent inflow. J. Phys. Conf. Ser. 75, 012063 (2007)

    Google Scholar 

  62. VerHulst, C., Meneveau, C.: Large eddy simulation study of the kinetic energy entrainment by energetic turbulent flow structures in large wind farms. Phys. Fluids 26(2), 025113 (2014)

    Google Scholar 

  63. Wissink, A., Potsdam, M., Sankaran, V., Sitaraman, J., Yang, Z., Mavriplis, D.: A coupled unstructured-adaptive Cartesian CFD approach for hover prediction. In: American Helicopter Society 66th Annual Forum. AHS International Alexandria, VA, pp. 1300–1317 (2010)

  64. Woodward, P., Colella, P.: The numerical simulation of two-dimensional fluid flow with strong shocks. J. Comput. Phys. 54(1), 115–173 (1984)

    MathSciNet  MATH  Google Scholar 

  65. Wu, J., Zhu, J., Szmelter, J., Zienkiewicz, O.: Error estimation and adaptivity in Navier–Stokes incompressible flows. Comput. Mech. 6(4), 259–270 (1990)

    MATH  Google Scholar 

  66. Yang, X., Sotiropoulos, F.: A new class of actuator surface models for wind turbines. Wind Energy 21(5), 285–302 (2018)

    Google Scholar 

Download references

Acknowledgements

This work was supported through a contract from Continuum Dynamics, Inc. under Navy STTR Phase II Contract N68335-17-C-0158 titled “Advanced Wake Turbulence Modeling for Naval CFD Applications” at the University of Michigan. The authors acknowledge support from Dr. Glen Whitehouse at Continuum Dynamics, Inc., and Drs. David Findlay and James Forsythe at the Naval Air Systems Command.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel Foti.

Additional information

Communicated by Tim Colonius.

Publisher's Note

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

This paper was prepared as an account of work sponsored by an agency of the United States (U.S.) Government. Neither the U.S. Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the U.S. Government or any agency thereof. The views and opinion expressed herein do not necessarily state or reflect those of the U.S. Government or any agency thereof.

Appendices

Governing equations and setup for spatially evolving turbulent jet

The mean streamwise velocity at the inflow is based on a hyperbolic tangent profile [17] and given by

$$\begin{aligned} u_1(x_1=0, x_2, x_3) = \frac{U_1 + U_2}{2} + \frac{U_1 - U_2}{2} \tanh \left( \frac{h}{4\theta _0} \left( 1 - \frac{2 |x_2|}{h} \right) \right) , \end{aligned}$$
(23)

where \(U_1\) is the centerline jet velocity, \(U_2\) is the co-flow, \(\theta _0\) is momentum thickness, and h is the slot width of the inlet. Both transverse mean velocities \(u_2\) and \(u_3\) are set to zero at the inlet. The top-hat inflow profile is obtained by mirroring the profile around the \(x_2\) plane. A small fluctuating velocity is added to each velocity component of the mean inflow profile to include statistics characteristic of isotropic turbulence. The fluctuations are prescribed to impose an energy spectrum given by

$$\begin{aligned} E(k) = k^s \exp \left( -\frac{s}{2} \left( \frac{k}{k_0} \right) ^2 \right) \end{aligned}$$
(24)

where the wavenumber \(k = (k_2^2 + k_3^2)^{1/2}\) and s and peak wavenumber \(k_0\) are the selected such that the energy input is dominated at small scales or high \(k_0\) and typical of decaying isotropic turbulence \(s \le 4\). The fluctuations are only imposed near the shear layers of the inflow.

The modified vorticity transport equations in compact index notation are as follows (\(i,j=1,2,3\)):

$$\begin{aligned} \frac{\partial \omega _i}{\partial t} + \frac{\partial }{\partial x_j} \left( u_j \omega _i - u_i \omega _j \right) + u_i \frac{\partial \omega _j}{\partial x_j} = \frac{1}{\mathrm{Re}} \frac{\partial ^2 \omega _i}{\partial x_j \partial _j} + \epsilon _{ijk} \frac{\partial F_k}{\partial x_j}, \end{aligned}$$
(25)

where \(u_i\) is the velocity, \(\omega _i = \epsilon _{ijk} \partial u_k / \partial x_j\) is the vorticity, and \(F_k\) is a body force. \(\mathrm{Re} = U L/\nu \) is the Reynolds number defined by a characteristic velocity scale U, characteristic length L, and kinematic viscosity \(\nu \).

The velocity is obtained from the vorticity through a Poisson equation given by

$$\begin{aligned} \frac{\partial ^2 u_i}{\partial x_j \partial x_j} = - \epsilon _{ijk} \frac{\partial \omega _k}{\partial x_j}. \end{aligned}$$
(26)

The governing equations are discretized with a multi-dimensional upwind finite volume approach as described in Ref. [24]. This approach uses upwind differences corrected by second-order differences around smooth flow regions to produce a solution that is second-order accurate in space and time. The multi-dimensional scheme solved in a series of dimensional sweeps shown as follows (for clarity vectors are shown in bold):

$$\begin{aligned} \varvec{\omega }^{n+1}_{i,j,k} = \varvec{\omega }^n_{i,j,k} - \frac{\Delta t}{\Delta _x} \sum ^3_{m=1} \left[ \left( (\varvec{F}^l_m)^n_{i+\frac{1}{2},j,k} - (\varvec{F}^r_m)^n_{i-\frac{1}{2},j,k} \right) - \left( (\varvec{G}_m)^n_{i+\frac{1}{2},j,k} - (\varvec{G}_m)^n_{i-\frac{1}{2},j,k} \right) \right] , \end{aligned}$$
(27)

where n is the time iteration and ijk are the indices of the grid cells in three directions. The flux functions \(\varvec{F}^l, \varvec{F}^r\) are normal direction flux function are obtained by solving a generalized Riemann problem developed for the VTE [50]. The transverse direction flux functions \(\varvec{G}\) are fluxes from the transverse direction and are computed using the flux-based wave propagation approach [24]. This method has been shown to accurately predict a number of vortex-dominated and turbulent flows [23,24,25].

The domain is carefully considered to accurately simulate a spatially evolving jet in the \(x_1\) direction with the VTE. The size of the computational domain in each direction \((L_1 \times L_2 \times L_3) = (12h \times 12h \times 3h)\), where inflow and outflow conditions are specified in the \(x_1\) direction and periodic in the other two directions. While the inflow conditions for the velocity are prescribed with Eq. (23), the outflow boundary conditions must be considered separately for vorticity and velocity. Due to the hyperbolicity of Eq. (25), the vorticity at the outflow that convects out of the domain due to a positive velocity at the outlet does not have a significant effect on the vorticity inside the domain; however, the velocity is strongly dependent on the vorticity due to the Biot–Savart law as computed through Eq. (26). To ensure that the boundary conditions do not affect the solution, the domain is extended with a buffer region in the \(x_1\) direction to 36h. The outflow boundary condition at the boundary of the buffer region for velocity is computed using a fast multi-pole method. The domain with buffer region is shown in Fig. 19.

The domain is initially discretized with several refinement levels of uniform structured grid cells, where the coarsest level uses \((N_1 \times N_2 \times N_3) = (96 \times 32 \times 8)\) across a \((36h \times 12h \times 3h)\) region including a computation and buffer region. This corresponds to a coarsest grid resolution of \(\Delta x_i = 0.375h\). A second level of refinement (with a refinement factor of 2 in each direction) is applied in the computation domain \((12h \times 12h \times 3h)\) with \((N_1 \times N_2 \times N_3) = (64 \times 64 \times 16)\), which ensures a uniform grid size of \(\Delta x_i = 0.1875h\). Additional levels of refinement are only applied on the domain with a refinement factor of 2.

Fig. 19
figure 19

The domain outlined in blue and entire computational domain with buffer region outlined red. The buffer region is discretized with the coarsest grid cells (color figure online)

The momentum thickness of the jet at the inlet is set to be \(h/\theta _0 = 30\), and a slot width \(h = 10\). The centerline velocity is \(U_1 = 1.091\) and the co-flow velocity is \(U_2 = 0.091\). The Reynolds number \(\text {Re}_h = \Delta U h / \nu = 3000\), where \(\Delta U = U_1 - U_2\). The magnitude of the inlet velocity fluctuations are given at 10%.

Validation of spatially evolving turbulent jet

The validation of the static grid cases, S1, S2, and S3, for the planar jet using statistics of the flow quantities is detailed. The temporal averaging of the velocity and vorticity flow fields begins after each simulation reaches a statistically steady state, and continued until the averaged flow field reaches statistical convergence. This corresponds to a total elapsed time of temporal averaging over 200 Kelvin–Helmholtz (K-H) vortices to leave the domain if we consider the frequency of K-H vortices to be \(St_{sl} = f_{sl} \theta _0/U_v = 0.0333\) where \(U_v = 1/2(U_1 + U_2)\).

The mean velocity flow field, \(\langle u_1 \rangle \) (also ensemble-averaged over the spanwise \(x_3\) direction), where \(\langle \cdot \rangle \) indicates \(x_3\) direction averaging, is normalized by the difference in the streamwise centerline velocity \(U_c = \langle u_1 \rangle (x_2=0)\) and the local streamwise co-flow velocity \(U_\infty = \langle u_1 \rangle (x_2=6h)\). The jet half-width \(\delta _{1/2}\), which is calculated using the mean streamwise velocity \(\langle u_1 \rangle (x_2=\delta _{1/2}) - U_\infty = \frac{1}{2} \left( U_c - U_\infty \right) \) is used to normalize the streamwise direction \(x_1\).

Profiles of the mean streamwise velocity at \(x_1/h = 10\) for all three static grid cases are shown in Fig. 20a, and compared with experimental measurements [32, 53]. The location is downstream the onset of the self-similar regime (\(x_1/h = 7\)) as shown in DNS [17] and experiments [59, 60] of a planar jet with similar inlet fluctuations and Re.

The mean streamwise velocity profile for S1, the coarsest grid resolution, shows slightly different behavior compared to S2 and S3, which have two and four times more grid cells in each direction, respectively. Because of the coarse resolution, the shear layer is not fully resolved, yielding a slightly higher and lower gradient near and far from the centerline, respectively. On the other hand, the S2 and S3 profiles agree well with the experimental measurements. There are slight discrepancies between the present simulation cases S2 and S3, and experimental measurements near \(x_2/\delta _{1/2} = 2\) where the velocity approaches zero. However, the velocity magnitude is quite small compared to the centerline velocity, and the two experimental measurements show divergent behavior.

The mean out-of-plane vorticity profiles \(\langle \omega _3\rangle \) at \(x_1/h = 10\) are shown in Fig. 20b. Similar to the mean velocity profiles, the S1 case does not have a sufficient number of grid cells to resolve the shear layer. The peak vorticity occurs near \(x_2/\delta _{1/2} = 0.5\), which is inward compared to the expected vorticity peak around \(x_2/\delta _{1/2} = 1\). The S2 and S3 cases have peak vorticity near the expected location. There are some differences between the S2 and S3 cases, where the most resolved S3 case (with sufficient grid cells to be considered a DNS) produces a smooth profile and a vorticity of zero at the origin. Overall, the mean statistics indicate that the S1 case is severely under-resolved, while the grid resolution for the S2 case is able to capture a majority of the mean statistics compared to the fully resolved S3 case.

Fig. 20
figure 20

Profiles at \(x_1/h = 10\) of the a mean streamwise velocity \(\langle u_1 \rangle - U_\infty \) normalized by the centerline velocity \(U_c - U_\infty \) compared with experimental results from Ref. [32] (circles) and Ref. [53] (squares) and b mean out-of-plane vorticity \(\langle \omega _3\rangle \) normalized by the jet half-width \(\delta _{1/2}\) and centerline velocity

Figure 21a, b shows further comparisons of the static grid cases with the root-mean-square (RMS) streamwise and transverse velocity profiles, respectively, at \(x_1/h = 10\). Similar to the mean velocity, the RMS streamwise velocity profile for the S1 case is significantly different from the other cases and the experimental measurements due to the inadequate resolution. The RMS velocity also indicates further discrepancies between the S2 and S3 cases, where the S3 case aligns with the experimental measurements consistently better. Furthermore, the RMS transverse velocity shows how the grid resolution significantly affects statistics, especially in the transverse direction.

Fig. 21
figure 21

Profiles at \(x_1/h = 10\) of the a RMS streamwise velocity \(\langle u^\prime _1 u^\prime _1 \rangle ^{1/2}\) and b RMS transverse velocity \(\langle u^\prime _2 u^\prime _2\rangle ^{1/2}\) normalized by the centerline velocity \(U_c - U_\infty \) compared with experimental results from Ref. [32] (circles) and Ref. [53] (squares)

Effect of snapshot matrix size

The effect of the size of the snapshot matrix on the sensor location is analyzed in Fig. 22. Three new cases with a snapshot matrix size of \(N=50, 200\) and 400 are used with a base grid resolution equal to that of the R1 case and compared with the R1 case, where the snapshot matrix size is \(N=100\). Figure 22a shows the eigenvalue spectrum of each case. Similar to the eigenvalue spectrum in Fig. 4a, there is a significant energy drop after the first mode for each case. This is significant because the low-rank representation of the jet can be approximated with a relatively small snapshot matrix. The sensor placement of the grid refinement tagging is shown for the four cases in Fig. 22b. As the number of POD modes increases with the snapshot matrix, the placement of the sensors begins to scatter over more of the domain. However, the placement remains concentrated at the inflow and outflow boundaries and at downstream locations (\(x_1/h > 6\)). Moreover, the number of total grid cells produced for each case is relatively similar. This is due to the efficiency of the load balancing and grid clustering algorithms employed to refine the grid cells near the sensors. On the other hand, if there is a desire to increase the grid resolution in the larger areas around the sampling points, modifications to the grid clustering and load balancing can be utilized to refine the grid further away from the sampling points.

Fig. 22
figure 22

The a singular value spectra and b sensor placement for the grid cell for refinement with the instantaneous out-of-plane vorticity based on the size of the snapshot matrix (\(N = 50, 100, 200,\) and 400). Overall grid resolution is based on the R1 case

Comparison to leverage score sampling

In this section, we introduce leverage score sampling [38, 39, 49], an alternative sampling approach to selecting columns of a matrix. The leverage scores \(l_i^{(K)} \in {\mathcal {R}}^{N}\) of a rank-K matrix \(\varPsi _K \in {\mathcal {R}}^{N \times K}\), which contains the top right singular vectors of \(U \in {\mathcal {R}}^{N \times M}\), is given by

$$\begin{aligned} l_i^{(K)} = \left\| [\varPsi _K]_{i,:} \right\| _2^2, \end{aligned}$$
(28)

where i denotes the ith row of \(\varPsi _K\). The sampling points are determined by sorting the leverage scores \(l_0^{(K)} \le l_1^{(K)} \le \dots \le l_{N-1}^{(K)}\) and finding the index \(c \in \{0,1,\dots ,N-1\}\) such that

$$\begin{aligned} c = {\mathop {{{\,\mathrm{arg\,min}\,}}}\limits _x} \left( \sum _{i=0}^c l_i^{(K)} > \theta \right) , \end{aligned}$$
(29)

where \(\theta = K - \epsilon \), for \(\epsilon \in (0,1)\) is the stopping criteria [49]. The stopping criteria is such that \(c = \min (c,K)\). The sampling points are the locations of the top c columns of \(\varPsi \). The points obtained via leverage score sampling are determined in a similar manner as Algorithm 1, the basis \(\varPsi \) is obtained using the POD of a dense snapshot matrix. The leverage scores are (i) computed with Eq. (28), (ii) sorted via a quicksort algorithm, and (iii) selected as the top c scores found with Eq. (29). The total complexity is as follows: step (i) is \({\mathcal {O}}(NK)\), step (ii) is on average \({\mathcal {O}}(N \log N)\) or worst case \({\mathcal {O}}(N^2)\), and step (iii) is \({\mathcal {O}}(c)\). In general, for the types of problems presented herein, \(K = M\) and \(N \gg M\), such that the complexity of computing the leverage score sampling is similar or more operations than the sparse sampling algorithm.

The leverage score sampling is tested on the supersonic flow over a forward facing step presented in Sect. 4.3. Figure 23 shows a comparison of the location of the sampling points obtained by sparse sampling and leverage score sampling. The locations are relatively similar, but the sparse sampling selects a few locations downstream.

Fig. 23
figure 23

Pressure \(P/P_1\) contours with grid lines for base grid and locations sampling points using a the sparse sampling method and b leveraged score method (\(|\nabla P| > 100\))

The error with respect to grid size is shown in Fig. 24a. At the lowest grid resolutions, the sparse sampling outperforms leverage score sampling. The errors converge at higher grid resolutions. At the highest grid resolution, the sparse sampling method slightly outperforms against the leverage score approach. The total number of cells after clustering and cell refinement produced by each algorithm is shown in Fig. 24b. The leverage score sampling produces significantly more cells at low grid resolution, but as grid resolution increases, the leverage score sampling produces slightly fewer compared to the sparse sampling method. Overall, at lower resolutions, the sparse sampling method performed well, but at higher resolutions, both algorithms perform similarly.

Fig. 24
figure 24

a L2 error with respect to the ratio of the truth case grid size to case grid size \(\Delta x_s/\Delta x_i\) and b total number of grid cells of the steady-state condition of the forward facing step simulations using the sparse sampling method (solid) and leveraged scores method (dashed). The red line shows \((\Delta x_s/\Delta x_i)^2\) (color figure online)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Foti, D., Giorno, S. & Duraisamy, K. An adaptive mesh refinement approach based on optimal sparse sensing. Theor. Comput. Fluid Dyn. 34, 457–482 (2020). https://doi.org/10.1007/s00162-020-00522-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-020-00522-2

Keywords

Navigation