Skip to main content
Log in

On the Feasibility of Using Large-Eddy Simulations for Real-Time Turbulent-Flow Forecasting in the Atmospheric Boundary Layer

  • Research Article
  • Published:
Boundary-Layer Meteorology Aims and scope Submit manuscript

Abstract

We investigate the feasibility of using large-eddy simulation (LES) for real-time forecasting of instantaneous turbulent velocity fluctuations in the atmospheric boundary layer. Although LES is generally considered computationally too expensive for real-time use, wall-clock time can be significantly reduced by using very coarse meshes. Here, we focus on forecasting errors arising on such coarse grids, and investigate the trade-off between computational speed and accuracy. We omit any aspects related to state estimation or model bias, but rather look at the size and evolution of restriction errors, subgrid-scale errors, and chaotic divergence, to obtain a first idea of the feasibility of LES as a forecasting tool. To this end, we set-up an idealized test scenario in which the forecasting error in a neutral atmospheric boundary layer is investigated based on a fine reference simulation, and a series of coarser LES grids. We find that errors only slowly increase with grid coarsening, related to restriction errors that increase. Unexpectedly, modelling errors slightly decrease with grid coarsening, as both chaotic divergence and subgrid-scale error sources decrease. A practical example, inspired by wind-energy applications, reveals that there is a range of forecasting horizons for which the variance of the forecasting error is significantly reduced compared to the turbulent background variance, while at the same time, associated LES wall times are up to 300 times smaller than simulated time.

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

Similar content being viewed by others

Notes

  1. A node consists of two 10-core “Ivy Bride” Xeon E5-2680v2 central processing units with 64 GB of random-access memory, which are interconnected with a quad data-rate infiniband network.

References

  • Abe H, Kawamura H, Choi H (2004) Very large-scale structures and their effects on the wall shear–stress fluctuations in a turbulent channel flow up to \(\text{ Re }_\tau = 640\). J Fluids Eng 126(5):835–843

    Article  Google Scholar 

  • Ainslie JF (1988) Calculating the flowfield in the wake of wind turbines. J Wind Eng Ind Aerodyn 27(1–3):213–224

    Article  Google Scholar 

  • Aurell E, Boffetta G, Crisanti A, Paladin G, Vulpiani A (1997) Predictability in the large: an extension of the concept of Lyapunov exponent. J Phys A Math Gen 30(1):1–26

    Article  Google Scholar 

  • Basu S, Foufoula-Georgiou E, Porté-Agel F (2002) Predictability of atmospheric boundary-layer flows as a function of scale. Geophys Res Lett 29(21):2038

    Article  Google Scholar 

  • Beare RJ, Macvean MK, Holtslag AA, Cuxart J, Esau I, Golaz JC, Jimenez MA, Khairoutdinov M, Kosovic B, Lewellen D et al (2006) An intercomparison of large-eddy simulations of the stable boundary layer. Boundary-Layer Meteorol 118(2):247–272

    Article  Google Scholar 

  • Belcher S, Coceal O, Goulart E, Rudd A, Robins A (2015) Processes controlling atmospheric dispersion through city centres. J Fluid Mech 763:51–81

    Article  Google Scholar 

  • Bou-Zeid E, Meneveau C, Parlange M (2005) A scale-dependent Lagrangian dynamic model for large eddy simulation of complex turbulent flows. Phys Fluids 17(2):025105

    Article  Google Scholar 

  • Brijs T (2017) Electricity storage participation and modeling in short-term electricity markets. PhD thesis, KU Leuven

  • Calaf M, Meneveau C, Meyers J (2010) Large eddy simulation study of fully developed wind-turbine array boundary layers. Phys Fluids 22(1):015110

    Article  Google Scholar 

  • Canuto C, Quarteroni A, Hussaini MY, Zang TA (1988) Spectral methods in fluid dynamics. Springer, Berlin

    Book  Google Scholar 

  • de Roode SR, Jonker HJ, van de Wiel BJ, Vertregt V, Perrin V (2017) A diagnosis of excessive mixing in smagorinsky subfilter-scale turbulent kinetic energy models. J Atmos Sci 74(5):1495–1511

    Article  Google Scholar 

  • Fang J, Porté-Agel F (2015) Large-eddy simulation of very-large-scale motions in the neutrally stratified atmospheric boundary layer. Boundary-Layer Meteorol 155(3):397–416

    Article  Google Scholar 

  • Frigo M, Johnson SG (2005) The design and implementation of FFTW3. Proc IEEE 93(2):216–231

    Article  Google Scholar 

  • Fuhrer O, Chadha T, Hoefler T, Kwasniewski G, Lapillonne X, Leutwyler D, Lüthi D, Osuna C, Schär C, Schulthess TC et al (2018) Near-global climate simulation at 1 km resolution: establishing a performance baseline on 4888 GPUs with COSMO 5.0. Geosci Model Dev 11(4):1665–1681

    Article  Google Scholar 

  • Gebraad P, Teeuwisse F, Wingerden J, Fleming PA, Ruben S, Marden J, Pao L (2016) Wind plant power optimization through yaw control using a parametric model for wake effectstest—a CFD simulation study. Wind Energy 19(1):95–114

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Goit JP, Meyers J (2015) Optimal control of energy extraction in wind-farm boundary layers. J Fluid Mech 768:5–50

    Article  Google Scholar 

  • Goit JP, Munters W, Meyers J (2016) Optimal coordinated control of power extraction in les of a wind farm with entrance effects. Energies 9(1):29

    Article  Google Scholar 

  • Hirth BD, Schroeder JL, Irons Z, Walter K (2016) Dual-Doppler measurements of a wind ramp event at an Oklahoma wind plant. Wind Energy 19(5):953–962

    Article  Google Scholar 

  • Holmes NS, Morawska L (2006) A review of dispersion modelling and its application to the dispersion of particles: an overview of different dispersion models available. Atmos Environ 40(30):5902–5928

    Article  Google Scholar 

  • Hutchins N, Marusic I (2007) Evidence of very long meandering features in the logarithmic region of turbulent boundary layers. J Fluid Mech 579:1–28

    Article  Google Scholar 

  • Jiménez J (1998) The largest scales of turbulent wall flows. CTR Annu Res Briefs 137:54

    Google Scholar 

  • Jung J, Broadwater RP (2014) Current status and future advances for wind speed and power forecasting. Renew Sust Energy Rev 31:762–777

    Article  Google Scholar 

  • Kalman RE et al (1960) A new approach to linear filtering and prediction problems. J Basic Eng 82(1):35–45

    Article  Google Scholar 

  • Katata G, Chino M, Kobayashi T, Terada H, Ota M, Nagai H, Kajino M, Draxler R, Hort M, Malo A et al (2015) Detailed source term estimation of the atmospheric release for the Fukushima Daiichi Nuclear Power Station accident by coupling simulations of an atmospheric dispersion model with an improved deposition scheme and oceanic dispersion model. Atmos Chem Phys 15(2):1029–1070

    Article  Google Scholar 

  • Katic I, Højstrup J, Jensen NO (1986) A simple model for cluster efficiency. In: European wind energy association conference and exhibition, pp 407–410

  • Kim K, Adrian R (1999) Very large-scale motion in the outer layer. Phys Fluids 11(2):417–422

    Article  Google Scholar 

  • Knudsen T, Bak T, Svenstrup M (2015) Survey of wind farm control—power and fatigue optimization. Wind Energy 18(8):1333–1351

    Article  Google Scholar 

  • Lapillonne X, Osterried K, Fuhrer O (2017) Using OpenACC to port large legacy climate and weather modeling code to GPUs. In: Farber R (ed) Parallel programming with OpenACC. Elsevier, Amsterdam, pp 267–290

  • Larsen GC, Madsen HA, Thomsen K, Larsen TJ (2008) Wake meandering: a pragmatic approach. Wind Energy 11(4):377–395

    Article  Google Scholar 

  • Le Dimet FX, Talagrand O (1986) Variational algorithms for analysis and assimilation of meteorological observations: theoretical aspects. Tellus A Dyn Meteorol Oceanogr 38(2):97–110

    Article  Google Scholar 

  • Leelőssy Á, Molnár F, Izsák F, Havasi Á, Lagzi I, Mészáros R (2014) Dispersion modeling of air pollutants in the atmosphere: a review. Open Geosci 6(3):257–278

    Google Scholar 

  • Leonard A (1975) Energy cascade in large-eddy simulations of turbulent fluid flows. In: Frenkiel FN, Munn RE (eds) Advances in geophysics, vol 18. Elsevier, Amsterdam, pp 237–248

  • Li N, Laizet S (2010) 2DECOMP & FFT—a highly scalable 2D decomposition library and FFT interface. In: Cray user group 2010 conference, pp 1–13

  • Lorenc A (1981) A global three-dimensional multivariate statistical interpolation scheme. Mon Weather Rev 109(4):701–721

    Article  Google Scholar 

  • Lorenz EN (1969) The predictability of a flow which possesses many scales of motion. Tellus 21(3):289–307

    Article  Google Scholar 

  • Mason PJ, Thomson D (1992) Stochastic backscatter in large-eddy simulations of boundary layers. J Fluid Mech 242:51–78

    Article  Google Scholar 

  • Meyers J (2011) Error-landscape assessment of large-eddy simulations: a review of the methodology. J Sci Comput 49(1):65–77

    Article  Google Scholar 

  • Meyers J, Meneveau C (2013) Flow visualization using momentum and energy transport tubes and applications to turbulent flow in wind farms. J Fluid Mech 715:335–358

    Article  Google Scholar 

  • Mikkelsen T (2014) Lidar-based research and innovation at DTU wind energy—a review. J Phys Conf Ser 524:012007

    Article  Google Scholar 

  • Moeng CH (1984) A large-eddy-simulation model for the study of planetary boundary-layer turbulence. J Atmos Sci 41(13):2052–2062

    Article  Google Scholar 

  • Mukherjee S, Schalkwijk J, Jonker HJ (2016) Predictability of dry convective boundary layers: an les study. J Atmos Sci 73(7):2715–2727

    Article  Google Scholar 

  • Munters W, Meyers J (2017a) An optimal control framework for dynamic induction control of wind farms and their interaction with the atmospheric boundary layer. Philos Trans R Soc A 375(2091):20160100

    Article  Google Scholar 

  • Munters W, Meyers J (2017b) Optimal coordinated control of wind-farm boundary layers in large-eddy simulations: intercomparison between dynamic yaw control and dynamic induction control. PhD thesis, Dept Mech Eng, KU Leuven

  • Munters W, Meyers J (2018) Dynamic strategies for yaw and induction control of wind farms based on large-eddy simulation and optimization. Energies 11:177

    Article  Google Scholar 

  • Munters W, Meneveau C, Meyers J (2016) Shifted periodic boundary conditions for simulations of wall-bounded turbulent flows. Phys Fluids 28(2):025112

    Article  Google Scholar 

  • Niayifar A, Porté-Agel F (2015) A new analytical model for wind farm power prediction. J Phys Conf Ser 625:012039

    Article  Google Scholar 

  • Pope SB (2000) Turbulent flows. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Rebours YG, Kirschen DS, Trotignon M, Rossignol S (2007) A survey of frequency and voltage control ancillary services—part I: technical features. IEEE Trans Power Syst 22(1):350–357

    Article  Google Scholar 

  • Sathe A, Mann J (2013) A review of turbulence measurements using ground-based wind lidars. Atmos Meas Tech 6(11):3147

    Article  Google Scholar 

  • Schlipf D, Trabucchi D, Bischoff O, Hofsäß M, Mann J, Mikkelsen T, Rettenmeier A, Trujillo JJ, Kühn M (2010) Testing of frozen turbulence hypothesis for wind turbine applications with a scanning lidar system. ISARS

  • Schlipf D, Schlipf DJ, Kühn M (2013) Nonlinear model predictive control of wind turbines using lidar. Wind Energy 16(7):1107–1129

    Article  Google Scholar 

  • Shah S, Bou-Zeid E (2014) Very-large-scale motions in the atmospheric boundary layer educed by snapshot proper orthogonal decomposition. Boundary-Layer Meteorol 153(3):355–387

    Article  Google Scholar 

  • Shapiro CR, Bauweraerts P, Meyers J, Meneveau C, Gayme DF (2017) Model-based receding horizon control of wind farms for secondary frequency regulation. Wind Energy 20(7):1261–1275

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Sullivan PP, Patton EG (2011) The effect of mesh resolution on convective boundary layer statistics and structures generated by large-eddy simulation. J Atmos Sci 68(10):2395–2415

    Article  Google Scholar 

  • van Stratum BJ, Stevens B (2015) The influence of misrepresenting the nocturnal boundary layer on idealized daytime convection in large-eddy simulation. J Adv Mod Earth Syst 7(2):423–436

    Article  Google Scholar 

  • Váňa F, Düben P, Lang S, Palmer T, Leutbecher M, Salmond D, Carver G (2017) Single precision in weather forecasting models: an evaluation with the IFS. Mon Weather Rev 145(2):495–502

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Verstappen R, Veldman A (2003) Symmetry-preserving discretization of turbulent flow. J Comput Phys 187(1):343–368

    Article  Google Scholar 

  • Vervecken L, Camps J, Meyers J (2015) Stable reduced-order models for pollutant dispersion in the built environment. Build Environ 92:360–367

    Article  Google Scholar 

  • Wang Q, Zhang C, Ding Y, Xydis G, Wang J, Østergaard J (2015) Review of real-time electricity markets for integrating distributed energy resources and demand response. Appl Energy 138:695–706

    Article  Google Scholar 

  • Wiernga J (1993) Representative roughness parameters for homogeneous terrain. Boundary-Layer Meteorol 63(4):323–363

    Article  Google Scholar 

Download references

Acknowledgements

The authors acknowledge support from the Agency for Innovation and Entrepreneurship through research Grant No. 141689. The computational resources and services used in this work were provided by the VSC (Flemish Supercomputer Center), funded by the Research Foundation—Flanders (FWO) and the Flemish Government department EWI.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Johan Meyers.

Additional information

Publisher's Note

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

Appendices

A Restriction and Interpolation

We provide further details on the interpolation and restriction operators introduced in Sect. 3.1. First of all, formally, we define \(\varvec{u}^{i} = [u_1^i,u_2^i,u_3^i] \in {\mathbb {R}}^{N^i}\), with \(N^i = 3N_x^iN_y^iN_z^i-N_x^iN_y^i\) (cf. staggered arrangement of variables discussed in Sect. 3.1). Similarly, \(\varvec{u}^{j}\in {\mathbb {R}}^{N^j}\), further using the convention that \(i<j\) (so that j is the coarser grid). Consequently, for the interpolation and restriction operators in Eqs. 5 and 6, we have \({\mathcal {I}}_{j}^{i} \in {\mathbb {R}}^{N^i\times N^j}\), and \({\mathcal {R}}_{i}^{j} \in {\mathbb {R}}^{N^j\times N^i}\).

Since we use a Cartesian mesh, we split the interpolation and restriction operators in three consecutive one-dimensional operators, so that \({\mathcal {I}}_{j}^{i} = I_{j,z}^i I_{j,y}^i I_{j,x}^i\), and \({\mathcal {R}}_{i}^{j}=R_{i,z}^j R_{i,y}^j R_{i,x}^j\). The matrix \(I_{j,z}^i\) has dimensions \(N^i \times N^{iij}\) with \(N^{iij}= 3N_x^i N_y^i N_z^j - N_x^i N_y^i\). The dimensions of \(I_{j,y}^i\) are \(N^{iij} \times N^{ijj}\), with \(N^{ijj}= 3N_x^i N_y^j N_z^j - N_x^i N_y^j\), and the dimensions of \(I_{j,x}^i\) correspond to \(N^{ijj} \times N^j\). Similar dimensions follow straightforwardly for \(R_{i,x}^j\), \(R_{i,y}^j\), and \(R_{i,z}^j\).

The rows of \(I_{j,x}^i\), \(I_{j,y}^i\), \(I_{j,z}^i\) contain one-dimensional interpolation stencils (and similar for the restriction matrices). Therefore, below, we provide the stencils that we use based on a simple scalar function \(\phi ^i\) and \(\phi ^j\) along one-dimensional grids \({\varvec{r}}^i\) and \({\varvec{r}}^j\). The allocation of the different coefficients in these stencils to elements in the different rows of \(I_{j,x}^i\), \(I_{j,y}^i\), etc., is straightforward, and not further detailed for sake of brevity.

For the interpolation in the x- and y-directions, spectral interpolation is used, simply leading to

$$\begin{aligned} \phi ^i_k = \frac{1}{N^j}\sum _{m=0}^{N^j-1}\sum _{l=0}^{N^j-1}\mathrm {exp}\left( {-2\uppi {\mathrm{i}} m\frac{r^i_l-r^j_k}{L}}\right) \phi _l^j. \end{aligned}$$
(20)

where \(\phi ^i_k\) and \(\phi _l^j\) correspond to fine- and coarse-grid values on locations \(r^i_k\) and \(r^j_l\) respectively. In practice, we do not implement the interpolation in real space, but instead perform the operation in Fourier space.

For the interpolation in the z-direction, we use a polynomial interpolation of order p, where we take \(p=4\), in analogy with our vertical discretization scheme. First to simplify notation, we define the operator \(min_c(a,{\varvec{b}})\), which returns a set of the \(c\in {\mathbb {N}}\) closest points in set \({\varvec{b}}\in {\mathbb {R}}^N\) to a scalar \(a\in {\mathbb {R}}\). This simply gives for the interpolation operator

$$\begin{aligned} \phi ^i_k = \sum _{l\in {\mathcal {P}}}\left( \prod _{\begin{array}{c} m\in {\mathcal {P}}\\ m\ne l \end{array}}\frac{r_k^i -r_m^j}{r_l^j-r_m^j}\right) \phi ^j_l, \text { with } {\mathcal {P}} = min_p\left( r_k^i,{\varvec{r}}^j \right) , \end{aligned}$$
(21)

In analogy, the rows of \(R_{i,x}^j\), \(R_{i,y}^j\), \(R_{i,z}^j\), contain the one-dimensional restriction stencils. For the restriction in x- and y-directions, a spectral cut-off filter is used in combination with simple injection to the coarse grid, leading to

$$\begin{aligned} \phi _l^j = \frac{1}{N^j}\sum _{m=0}^{N^j-1}\sum _{k=0}^{N^i-1}\mathrm {exp}\left( {-2\uppi {\mathrm{i}} m\frac{r^j_k-r^i_l}{L}}\right) \phi ^i_k. \end{aligned}$$
(22)

For the restriction in the z-direction, we use a combination of a box filter and an injection. For the box filter we use a width of \(\varDelta _z^j\), which comes down to \(s=\varDelta _z^j/\varDelta _z^i\) cells on the fine grid. It is easily shown that the following relation holds to filter a field \(\phi ^i\), which is assumed to have been filtered with a width \(\varDelta _z^i\), to a field \(\phi ^j\) with a width \(\varDelta _z^j\)

$$\begin{aligned} {\widetilde{\phi }}^i_l = \frac{1}{s}\sum _{k=-s/2}^{s/2-1}\phi ^i_{l+k+1/2}, \end{aligned}$$
(23)

where the interpolation of \(\phi ^i_{l+k+1/2}\) happens with the same fourth-order interpolation as is described above.

For the refinement experiment we use a Gaussian filter where the standard deviation is chosen as \(\sigma ^2=(s^2-1)/12\), and where the factor 1 / 12 is determined such that the second moments of the Gaussian and box filter are equal [see Leonard (1975) for a derivation], and the factor \(-1\) appears due to the successive filtering [see e.g. Pope (2000)], such that choosing \(s=1\) leaves the field unaltered. This leads to the following relation

$$\begin{aligned} {\widetilde{\phi }}^i_l = \frac{\sum _{k=-\infty }^{\infty }\mathrm {exp}\left( {-\frac{1}{2}\left( \frac{k}{\sigma }\right) ^2}\right) \phi ^i_{l+k}}{\sum _{k=-\infty }^{\infty }\mathrm {exp}\left( -\frac{1}{2}\left( \frac{k}{\sigma }\right) ^2\right) }. \end{aligned}$$
(24)

In a further step the field is restricted to the coarser grid. Due to mismatching cell locations for the u and v velocity components an additional interpolation is needed. For this we again use the fourth-order polynomial interpolation, which leads to the following expression

$$\begin{aligned} \phi ^j_l = \sum _{k\in {\mathcal {P}}}\left( \prod _{\begin{array}{c} m\in {\mathcal {P}}\\ m\ne k \end{array}}\frac{r_l^j -r_m^i}{r_k^i-r_m^i}\right) {\widetilde{\phi }}^i_k, \text { with } {\mathcal {P}} = \mathrm {min}_p\left( r_l^j,{\varvec{r}}^i \right) . \end{aligned}$$
(25)

B Comparison of Time-Averaged Mean Fields

For the sake of completeness, we provide a comparison of time-averaged velocity and turbulent kinetic energy fields obtained on the different grids, which is the standard basis for comparing LES results (using different grids, models, or codes). In contrast to the error analysis in the main text, we present long time averages that omit the initial transient that occurs when initializing with a turbulent field that is not in statistical equilibrium on the simulation grid. To this end the simulations on the different grids are spun up until a statistical steady state is reached. Afterwards, averaging is performed over a period of \({8000}\,\hbox {s}\), ensuring sufficient statistical convergence.

Results are shown in Fig. 9, and overall, it is appreciated that profiles of the mean flow match closely. Profiles of turbulent kinetic energy show a more pronounced grid dependency close to the wall. This is quite standard, as the integral length scale decreases proportional with the distance to the wall, so that less large-scale motions are resolved in this region on coarser grids.

Fig. 9
figure 9

Time-averaged equilibrium streamwise velocity component, \(U_1^i\) (left) and turbulent kinetic energy, \(E^{i}\) (right) for the different grids. Grid numbers: 0 ( ), 1 ( ), 2 ( ) and 3 ( )

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bauweraerts, P., Meyers, J. On the Feasibility of Using Large-Eddy Simulations for Real-Time Turbulent-Flow Forecasting in the Atmospheric Boundary Layer. Boundary-Layer Meteorol 171, 213–235 (2019). https://doi.org/10.1007/s10546-019-00428-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10546-019-00428-5

Keywords

Navigation