As turbines continue to grow in hub height and rotor diameter and wind farms grow larger, consideration of stratified atmospheric boundary layer (ABL) processes in wind power models becomes increasingly important. Atmospheric stratification can considerably alter the boundary layer structure and flow characteristics through buoyant forcing. Variations in buoyancy, and corresponding ABL stability, in both space and time impact ABL wind speed shear, wind direction shear, boundary layer height, turbulence kinetic energy, and turbulence intensity. In addition, the presence of stratification will result in a direct buoyant forcing within the wake region. These ABL mechanisms affect turbine power production, the momentum and kinetic energy deficit wakes generated by turbines, and the turbulent mixing and kinetic energy entrainment in wind farms. Presently, state-of-practice engineering models of mean wake momentum utilize highly empirical turbulence models that do not explicitly account for ABL stability. Models also often neglect the interaction between the wake momentum deficit and the turbulence kinetic energy added by the wake, which depends on stratification. In this work, we develop a turbulence model that models the wake-added turbulence kinetic energy, and we couple it with a wake model based on the parabolized Reynolds-averaged Navier–Stokes equations. Comparing the model predictions to large eddy simulations across stabilities (Obukhov lengths) and surface roughness lengths, we find lower prediction error in both power production and the wake velocity field across the ABL conditions and error metrics investigated.

Atmospheric boundary layer (ABL) flow environments with wind shear and stratification can pose challenges for fast-running engineering wake models.1–4 Error and uncertainty in wake modeling can impact predictions of wind farm power production used for both wind farm design5 and wind farm flow control.6 Many state-of-practice engineering wake models were derived under assumptions of uniform or neutral inflow.7–9 These assumptions greatly simplify wake physics by neglecting or parameterizing ABL effects such as atmospheric stratification and the impact of wind shear on wake turbulence. Yet wind turbine wakes and wind farm performance strongly depend on atmospheric stability.10–14 

Recent studies have developed extensions of existing engineering models that account for the effects of wind direction shear (veer) on wind turbine wake advection15 or through a shape function that depends on azimuthal angle with superimposed veer.3,16 These extended models show substantially improved performance relative to standard axisymmetric analytical wake models.3,15,16 However, by relying on an analytical model form as the starting point (i.e., a Gaussian wake model), these models often neglect the nonlinear interactions between the wake momentum and the turbulence added by the wake itself. The turbulence generated by the presence of a wake is often called wake-added turbulence.17,18 While turbulence can have a significant impact on farm power and farm efficiency in analytical wake models,19 turbulence is often parameterized in a simple manner that does not allow for feedback between the wake deficit and the wake-added turbulence. In many analytical wake models, the wake-added turbulence is only relevant in the (often linear) wake spreading parameter for waked turbines.20 This is because the wake recovery is generally assumed to only depend on the turbulence intensity incident to the turbine,21 rather than the turbulence in the wake. In other words, the wake of a wind turbine operating in freestream ABL flow does not account for the effect of wake-added turbulence on wake recovery. In contrast, in the balance of mean momentum in the wake, the wake recovery is controlled by the spatially varying turbulence, which is the combination of inflow ABL turbulence and wake-added turbulence.4 Recognizing this, several recent engineering wake models have incorporated the effect of wake-added turbulence on the wake recovery.22–24 Yet, a challenge in fast-running wake models is how to incorporate the effect of stratification on the wake-added turbulence, and therefore, the spatially varying wake recovery.

Wake-added turbulence is strongly dependent on atmospheric stratification due to the effects of atmospheric stability on ambient ABL turbulence intensity, wind speed shear, and wind direction shear.4 In addition, the buoyant destruction (generation) of turbulence kinetic energy in stable (unstable) ABL flows plays a secondary role.4 As discussed earlier, existing analytical wake models often do not directly account for these effects due to their simplified, analytical form. Instead, to improve wake turbulence and momentum modeling in the stratified ABL, we leverage an alternative engineering modeling approach based on the parabolized Reynolds-averaged Navier–Stokes (RANS) equations, such as the curled wake model.25,26 The curled wake model solves a simplified and parabolized form of the RANS equations to calculate the velocity deficit along with a Boussinesq eddy viscosity model for the turbulence closure. Working within this framework allows for improvement to the physical representation of the flow field through the incorporation of ABL processes and forcings, while still maintaining computational efficiency. The primary benefit of this approach is that the mean wake momentum deficit and wake-added turbulence are naturally coupled in the dynamics. However, a primary gap is in accurately modeling wake-added turbulence in stratified ABL flows.4 Previous works that employ the curled wake model have used simple zero-equation mixing length modeling25,26 and data-driven turbulence modeling.27 In the present work, we utilize the parabolized RANS curled wake model framework, and we develop a novel, physics-based wake-added turbulence model that is applicable to wind turbine wake modeling in a wide range of stably stratified ABL flows based on physical insights from prior work.4 

The remainder of this work is organized as follows: the methods are presented in Sec. II with a description of the extended curled wake model developed in this study (Sec. II A), the development of a k model (Sec. II B), details of the baseline wake models that are used for comparison (Sec. II C), numerical details of the engineering wake model developed here (Sec. II D), and (Sec. II E) details of the large eddy simulation (LES) data used for testing and validation. The results are presented in Sec. III, with validation against the LES data and comparison with the baseline, existing wake models. Finally, conclusions are provided in Sec. IV.

We start from the curled wake model,26 which solves a parabolic equation for the mean streamwise wake deficit derived from the Navier–Stokes equations. To obtain this equation, the flow field is decomposed into the base ABL flow field without the presence of wind turbines, denoted by B, and the wake flow field denoted by Δ. This decomposition is shown for the velocity,
(1)
where uiB is the instantaneous base flow velocity, Δui is the instantaneous wake deficit velocity, and ui is the full instantaneous velocity. The flow is also Reynolds-decomposed into a mean quantity and a fluctuating quantity (with zero mean). For the instantaneous wake velocity deficit, the Reynolds decomposition is given by
(2)
where Δui¯ and Δui are the temporal mean and the fluctuation of the wake velocity deficit, respectively. The curled wake model further neglects the pressure gradient, viscous stresses, Coriolis forces, and gradients of the base flow.26 In Klemmer and Howland,4 the authors found that the pressure gradient, viscous stresses, and Coriolis forces are relatively negligible for a single turbine wake at the relatively high Reynolds number and Rossby number cases investigated (Ro=G/ωD=1306, where G is the geostrophic wind speed, ω is the angular velocity of Earth, and D is the rotor diameter). The present cases considered in this study have Ro=1306, as further described in Sec. II E. The importance of Coriolis forces depends on the Rossby number. At a Rossby number of Ro=1000 in large eddy simulations of conventionally neutral boundary layer flow, the magnitude of the wake deflection from Coriolis effects was found by Heck and Howland28 to be approximately 0.1D at a location 15D downwind of the wind turbine. At lower Rossby numbers, this deflection can exceed 0.5D.28 For a different set of parameters or in a different configuration (e.g., a wind farm array), it is possible that the pressure gradient and Coriolis forcing, in particular, can become non-negligible. This is particularly true when considering the effects of atmospheric response and blockage for large wind farms,29–32 though this is outside the scope of the present work.
Klemmer and Howland4 identified that the base flow gradient terms are important for both neutral and stable ABL flows. Here, we extend the curled wake model26 by incorporating base flow gradients. In cases where yaw and tilt misalignment is equal to zero, the lateral and vertical velocities, Δv¯ and Δw¯, respectively, are necessarily zero in the curled wake model.26 Starting from the full high Reynolds number, Boussinesq incompressible Navier–Stokes equations, inserting the two decompositions described earlier, and solving for Δu¯ under the assumptions described above yields4 
(3)
where uiuj¯wake=uiuj¯uiBujB¯ is the wake-added Reynolds stress tensor. The mean wake deficit derivation details are abridged in this study for brevity, and the reader is directed to Martínez-Tossas et al.26 for further details. In this study, we model the wake Reynolds stresses according to the Boussinesq hypothesis,
(4)
where νT is the unknown eddy viscosity, ΔSij=12(Δui¯xj+Δuj¯xi) is the mean deficit strain rate tensor, and kwake is the wake-added turbulence kinetic energy (TKE). In other words, kwake is the TKE that arises due to the presence of a turbine and is defined as
(5)
where k is the total TKE, and kB is the TKE in the base ABL flow. The mean momentum deficit is integrated from an initial condition at the rotor plane (x=0), which is from momentum theory, further described in Sec. II B.

Previous studies that solve similar parabolized RANS equations for wind turbine wakes as Eq. (3) have utilized analytical or zero-equation eddy viscosity models for the Reynolds stresses.25–27,34 While these models have the advantage of simplicity and computational efficiency, they lack turbulence representation that comes with the additional complexity of traditional RANS models.35 The working hypothesis of this study is that this additional detail in turbulence modeling, especially the three-dimensional spatial information and the incorporation of buoyant forcing and wind shear effects, is especially important for modeling wakes in the stratified ABL.

Here, we propose a k model for the eddy viscosity νT, where k represents TKE, and is a mixing length. From dimensional analysis νTk, here, we divide the flow into the near-wake and far-wake regions based on the near-wake length x0. The near-wake region is the portion of the wake that is approximately inviscid, as described by classical momentum theory, before turbulence is dominant in the wake dynamics.9,33 In this inviscid region, νT is modeled using the base flow TKE kB and the turbine diameter D as the length scale. The wake-added TKE becomes relevant in the model after x0 when νT is modeled via the full flow field TKE and the mixing length. We can combine these two eddy viscosity contributions using a Heaviside function, which yields
(6)
where x0 is the near-wake length,9,33 Cν is a model constant taken to be 0.04, and H is the Heaviside function.

Both existing analytical wake models7,8 and the curled wake model26 take initial conditions from momentum theory,36 such that Δu¯=2auh at x=x0 downwind of the turbine, where uh is the hub-height wind speed, a is the axial induction factor a=12(11CT), and CT is the thrust coefficient. Specifically, the wake initial condition for Δu¯ is applied at the rotor plane (x=0) and the eddy viscosity form in Eq. (6) yields relatively less recovery until x=x0, where wake-added turbulence then contributes to the eddy viscosity. The Heaviside function limits turbulence in the near-wake region as discussed earlier. Wake-added TKE becomes active in the far wake (x>x0) and is modeled via a transport equation. We target the modeling of kwake, rather than the full TKE, as it has been shown that the wake-added turbulence is highly dependent upon stability and by extension speed and direction shear.4 

To derive a model transport equation for kwake, we start from the full prognostic equation given by
(7)
and we leverage conventional methods for modeling individual terms in the k equation,35 detailed as follows. Equation (7) is derived by starting from the full TKE (k) equation and then solving for the prognostic equation for the wake-added TKE kwake. Further details on the derivation are provided in Klemmer and Howland.4 The advection, production, buoyancy, transport, dissipation, and interaction with the base ABL flow TKE are given by Awake, Pwake, Bwake, Twake, εwake, and kBAwake, respectively.
The final model equation for kwake can be found in Eq. (8). The production term is closed with the closure of the wake Reynolds stress [Eq. (4)]. In the Boussinesq hypothesis in Eq. (4) and the eddy viscosity in Eq. (6), νT is representative of the combined effects from the base flow and the wake, while ΔSij only represents wake effects. To best capture the nonlinear interaction between the base and wake flow in the kwake shear production, the modeled wake Reynolds stresses in Pwake are multiplied by the gradient of the full flow field ui¯ as shown in the first term in parentheses in Eq. (8). For the transport terms in Twake, we use the gradient diffusion hypothesis and assume that turbulent transport is dominant. This results in the second term shown in parenthesis in Eq. (8) with the model constant Ck1. Finally, dimensional analysis is used to model the dissipation via a velocity scale (kwake) and a length scale () scaled by a constant Ck2, as is commonly done in RANS turbulence modeling.35 The resulting equation is shown as follows:
(8)

In Eqs. (6) and (8), the mixing length is taken to be the wake width yD, which is defined as the distance from the wake centerline at which the wake has recovered to 5% of the ambient. The wake width is a common choice for the mixing length in wind turbine wake models.37–40 In RANS modeling of the stratified ABL, the mixing length is often influenced by a variety of factors that limit mixing, including the wall normal distance and stability.41 Previously, Klemmer and Howland4 showed that stability has a strong impact on wake-added TKE. Therefore, the primary goal of this study is to develop an improved model for wake-added TKE and to incorporate the wake-added TKE model into a wake model for mean momentum. Since aspects of stability are encoded in the wake-added TKE, we select the wake width as a simple model of the mixing length. The mixing length approximation using the wake width is quantitatively evaluated in  Appendix B, where the results show that the approximation yields reasonable accuracy for neutral and moderate stabilities, but has increased error at higher stability. The model constants Ck1, and Ck2 are set to 1, and Cν is set to 0.04. To arrive at these values, we performed a sensitivity analysis across all stability cases with aerodynamic roughness lengths of z0=1 cm and z0=50 cm (further details on the LES simulation data are in Sec. II E). Details of this parameter sensitivity analysis are given in  Appendix A.

To assess the performance of the proposed k model, we compare this model against the prediction from three other models: the analytical Gaussian wake model,8,9 the analytical Jensen wake model,7 and a turbulence model from Scott et al.27 that uses the curled wake model. Importantly, the parameters used for these models are taken from the literature, meaning that unlike the k model presented in this work, the comparative models are not tuned for the sweep of ABL cases for z0=1 cm and z0=50 cm. Despite this, all models are evaluated on unseen data (z0=10 cm). In other words, the new k model has not been calibrated to the LES data for z0=10 cm, which is the data to which we make comparisons in this study.

1. Analytical wake models

We compare the proposed k model to two existing, analytical wake models: the Gaussian wake model8,9 and the Jensen wake model.7 For the evaluation of both analytical wake models, we use implementation within the open-source FLOw Redirection and Induction in Steady State (FLORIS) tool version 4.1,42 developed by the National Renewable Energy Laboratory (NREL).

In the present work, the wind turbines are aligned with the incident wind direction at hub-height (i.e., tilt and yaw are zero in all cases), which results in the following equation for the wake deficit for the Gaussian model:
(9)
where uh is the hub height wind speed, σ is the wake width, and zh is the turbine hub height. The wake width is given by
(10)
where σ0=1/8 in the yaw-aligned case, x0 is the near-wake length (taken from measurements in LES, see discussion in Sec. II D), and k* is the wake spreading rate. Following Niayifar and Porté-Agel,21 the wake spreading rate is modeled as
(11)
where I0=uu¯/uh is the ambient streamwise turbulence intensity for a yaw-aligned wind turbine. No wake-added turbulence model is necessary since we are modeling a single turbine wake in all cases.43 This means that wake-added turbulence does not affect wake recovery for a single turbine case.
The Jensen model wake deficit is given by
(12)
where D is the turbine diameter, and kw is the rate of wake expansion. In the present work, we model kw as44 
(13)
where κ=0.4 is the von Karman constant, zh is the hub height, z0 is the roughness length, L is the Obukhov length, and ψm is the stability correction for the momentum. The Businger–Dyer relationships44,45 are used to model the stability correction as ψm=4.7zhL. This parameter is the only mechanism of wake recovery in these two analytical wake models, meaning that the model does not consider how wake-added turbulence affects the wake.

2. Scott model

In addition to comparisons against analytical wake models, we also compared the developed approach to evaluations of the curled wake model using existing turbulence modeling approaches in the literature for wind turbine wakes. In the work by Scott et al.,27 the authors develop a model for νT in the curled wake model by assuming a linear relationship between the wake turbulent shear stress uw¯wake and the mean deficit strain rate ΔS13. They further assume that νT in the wake is only dependent on streamwise position x, neglecting lateral and vertical variations, and they use simulation and experimental data to fit a curve to the νT profiles. The resulting functional form for the eddy viscosity is ultimately given by a Rayleigh function,
(14)
where A=RUB1CT/2 is a function of the turbine radius R, inflow velocity at hub height UB, and thrust coefficient CT, and σ=5.5 from Scott et al.27 

For all models, the coefficient of thrust CT is set to 0.75 to match the LES data. For the k and Scott models, which used the curled wake model framework, the base flow is taken from LES. At x/D=0, both the k and Scott models are given an initial condition of Δu¯=2auh in the rotor area (following the example in Martínez-Tossas et al.25) Additionally, for the k model, kwake is given an initial condition of 0, uniform in space, at x/D=0. A no slip condition is applied at the bottom boundary, and a stress-free condition at the top wall.

The near-wake length x0 is a complex function of the turbine control and the incident wind conditions.9,33 There is a limited understanding of the near-wake length in sheared and stratified ABL flow conditions that we investigate in this study. To isolate the evaluation of the modeling of the turbulent portion of the wind turbine, we provide both the baseline engineering models and the proposed parabolized RANS k model with the near-wake length directly measured from LES as the position in the wake where Δu¯>0 for the first time.33 This is the point at which wake recovery begins. This is defined by the location at which the streamtube-averaged wake deficit is at its minimum value. We leverage these LES measured near-wake lengths x0 rather than using an engineering model for x0 (e.g., Liew et al.33) because the existing models do not account for stability. Future work is suggested to improve modeling of x0 in stratified ABL flow conditions.

We use large eddy simulation (LES) of modeled wind turbines immersed in stratified ABL flows for validation and performance assessment for the model. All data are generated using the PadéOps LES code (https://github.com/Howland-Lab/PadeOps),32,46 which is an open-source, pseudo-spectral computational fluid dynamics solver. The solver uses Fourier collocation in the horizontal directions and a sixth-order staggered compact finite difference scheme in the vertical direction.47 For the temporal integration, a fourth-order strong stability-preserving variant of a Runge–Kutta scheme is used.48 The sigma subfilter-scale model is used49 for the subgrid stresses with a turbulent Prandtl number of 0.4 for the scalar diffusivity.

All simulations employ a fringe region in the streamwise and lateral directions to force the inflow to the desired profile.50 The inflow is specified via the concurrent precursor method,51 in which two simulations are run concurrently: a primary simulation with the turbine and a precursor simulation of the same domain as the primary but without the turbine. This ensures that the wind turbine encounters physically consistent stratified ABL turbulence as its inflow condition. Each LES simulation has a single turbine that is modeled with the actuator disk model without rotation52 and has a diameter of D=126 m, a hub height of 90 m, and a thrust coefficient of CT=0.75. The actuator disk model is used here to evaluate engineering wake models that are also built on actuator disk modeling assumptions.9,26,33 For first validation of the wake model and for computational tractability for the 15 different atmospheric conditions investigated here, we only consider the modeling of a single turbine wake within the ABL. Future work, outside the scope of this paper, should evaluate the modeling of multiple wind turbine wakes within wind farms operating in the stratified ABL.

The model developed in this study is calibrated and tested against 15 ABL cases that correspond to a sweep over five surface cooling rates θtz=0=[0,0.25,0.5,0.75,1] K h−1 and three surface roughness lengths z0=[1,10,50] cm. In total, there are three conventionally neutral boundary layer (CNBL) cases and 12 stable boundary layer (SBL) cases. More specifically, z0=[1,50] cm is used for calibration and z0=10 cm is used for testing. Thus, z0=10 cm is not seen by any model for calibration. In this study, we focus on neutral and stable ABLs because wake losses are typically less important in unstable ABL conditions,12,13 but future work may consider unstable ABL conditions. Data for all flows with z0=10 cm come from Klemmer and Howland.4 The additional ten cases with z0=[1,50] cm are generated for this study to provide stratified flows with varied freestream turbulence intensity for the same surface cooling rates. The 15 cases examined in this study encompass a range of flow conditions, spanning from neutral through to very stable stratification. The inclusion of turbulence intensity variations (through the surface roughness parameter) within each boundary layer enables a more detailed investigation into the role of inflow turbulence on wake-added turbulence in stratified flows. The SBLs are named according to their surface cooling rates (SBL 0.25, SBL 0.5, SBL 0.75, SBL 1), and—unless otherwise specified—results are shown for z0=10 cm.

All cases are driven by a geostrophic wind speed of 12 m/s. The grid spacing in the streamwise and lateral directions is 12.5 m, with a streamwise domain length Lx of 4800 m and a lateral domain length Ly of 2400 m. The vertical extent is 2400 m in the CNBLs and 1600 m in the SBLs. All cases have vertical grid spacing of Δz=6.25 m. The Coriolis frequency is 1.03×104 s−1, which corresponds to a latitude of 45° N. A constant surface heat flux wθ¯w=0 K m s−1 is prescribed in the CNBL. In the SBLs, the surface boundary condition is specified as a cooling rate.53 In all cases, the initial potential temperature profile is set to 301 K up to 700 m for the CNBL and 50 m for the SBLs. Above this height is an inversion with strength 0.01 K m−1. Further simulation details are provided in Table I.

TABLE I.

ABL simulation details. Each row is a different surface cooling rate θtz=0, and the columns show the different surface roughness lengths z0. The Obukhov length L=u*3θ0κgwθ¯w, friction velocity u*, and ambient turbulence intensity I0 are shown for each case.

z0=1 cm z0=10 cm z0=50 cm
θtz=0 (K/h) L (m) u* (m/s) I0 L (m) u* (m/s) I0 L (m) u* (m/s) I0
  0.43  0.1    0.52  0.14    0.61  0.19 
−0.25  86  0.31  0.06  118  0.36  0.07  130  0.39  0.09 
−0.5  43  0.28  0.04  51  0.32  0.06  62  0.34  0.07 
−0.75  29  0.27  0.04  36  0.30  0.05  41  0.32  0.06 
−1  21  0.25  0.04  26  0.28  0.05  29  0.30  0.05 
z0=1 cm z0=10 cm z0=50 cm
θtz=0 (K/h) L (m) u* (m/s) I0 L (m) u* (m/s) I0 L (m) u* (m/s) I0
  0.43  0.1    0.52  0.14    0.61  0.19 
−0.25  86  0.31  0.06  118  0.36  0.07  130  0.39  0.09 
−0.5  43  0.28  0.04  51  0.32  0.06  62  0.34  0.07 
−0.75  29  0.27  0.04  36  0.30  0.05  41  0.32  0.06 
−1  21  0.25  0.04  26  0.28  0.05  29  0.30  0.05 

In this section, the proposed k model is validated through a comparison with LES, and performance is assessed through comparison with the three models detailed in Sec. II. For both the validation and performance assessment, we show qualitative visual results generated using the CNBL, SBL 0.25, and SBL 1 cases, all with z0=10 cm. These cases cover the full range of stabilities, and trends remain consistent across roughness lengths as shown in the aggregated metrics, which are shown for all stabilities and roughness lengths.

We assess the models via four metrics: pointwise error in Δu¯, the streamtube-averaged wake deficit Δu¯s, the centerline wake deficit Δu¯c, and the estimated power of a hypothetical waked turbine P. Together, these four metrics illustrate the predictive capabilities of the proposed model over a varied range of levels of description pertinent to flow field and power prediction.

Starting with the most descriptive qualitative assessment of the flow field predictions, we compare the models and LES pointwise through slices of Δu¯ in the yz plane. Figures 1–3 show the wake deficit of the CNBL, SBL 0.25, and SBL 1 at three streamwise locations: 5D, 10D, and 15D downstream of the turbine (and initial condition).

FIG. 1.

Contours of Δu¯ for the CNBL with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

FIG. 1.

Contours of Δu¯ for the CNBL with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

Close modal
FIG. 2.

Contours of Δu¯ for SBL 0.25 with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

FIG. 2.

Contours of Δu¯ for SBL 0.25 with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

Close modal
FIG. 3.

Contours of Δu¯ for SBL 1 with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

FIG. 3.

Contours of Δu¯ for SBL 1 with z0=10 cm from the k, Scott, Gaussian, and Jensen models compared with LES. The black circle denotes the rotor area.

Close modal

Starting with the CNBL (Fig. 1) at x/D=5, all the models underpredict the magnitude of the deficit at this location, meaning that the turbulence, which is responsible for wake recovery, is overpredicted. The k model most closely approximates Δu¯ at this location, while the Jensen and Gaussian models most severely underpredict the deficit. This can be seen moving downstream to x/D=10 and x/D=15, where the Gaussian and Jensen models differ most from the LES field, and the k and Scott models more closely capture the wake velocity deficit and shape.

This pattern is additionally observed in the two SBLs in Figs. 2 and 3, where again we see that the Scott and Gaussian models significantly underpredict the wake deficit magnitude (corresponding to an overprediction of turbulent diffusion). In contrast, the k model more closely approximates the correct magnitude. The Scott and k models have a further advantage of being able to capture the skewing of the wake in the stable cases, which is not present in the Gaussian model predictions. The wake skewing is a direct consequence of the advective terms within the prognostic equation for the mean wake momentum deficit in Eq. (3), rather than a heuristic parameterization. While both the Scott and k models fail to exactly reproduce the wake shape, the k model appears to more faithfully capture the wake skew, likely due to the inclusion of a three-dimensional eddy viscosity.

To further analyze the pointwise accuracy of the wake predictions, we define the pointwise error as the 2 norm of the difference between the model wake deficit and the LES wake deficit normalized by the 2 norm of the LES wake deficit at each x location,
(15)
This error metric provides a measure of the fidelity of the models at each point in the wake. This pointwise error is evaluated over the spanwise and vertical extents shown in Figs. 1–3. Figure 4 shows the pointwise error for the four considered models. In all cases, the analytical models consistently exhibit the highest error, while the k largely has the lowest error across the three boundary layer cases. Some streamwise positions yield very similar pointwise errors to the Scott model.27 
FIG. 4.

Comparison of the normalized pointwise error [Eq. (15)] for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases.

FIG. 4.

Comparison of the normalized pointwise error [Eq. (15)] for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases.

Close modal

We reduce the granularity of the comparison and performance metrics by introducing the streamtube-averaged wake deficit Δu¯s. In contrast to the pointwise comparison of Δu¯ and the associated error, the streamtube average focuses the analysis on the turbine wake and its recovery. However, in contrast to the pointwise error metric, the streamtube-averaged wake deficit does not directly penalize errors in predictions of the wake shape that were shown in Sec. III A. Figure 5 shows Δu¯s for the CNBL, SBL 0.25, and SBL 1 (with z0=10 cm). The k model tends to be relatively well-matched to the LES across the three ABLs. The analytical models tend to underpredict the magnitude of the streamtube-averaged wake deficit (as observed in Sec. III A).

FIG. 5.

Comparison of the mean streamtube velocity deficit for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. LES data are shown by the symbols (○○○). In all cases, z0=10 cm.

FIG. 5.

Comparison of the mean streamtube velocity deficit for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. LES data are shown by the symbols (○○○). In all cases, z0=10 cm.

Close modal

It should be noted that all the models fail to capture the precise streamwise development of the flow, a result that is seen in all error metrics. This is likely due to the difficulty in capturing the near-wake behavior. None of the models have a mechanism for increasing the magnitude of the wake deficit downwind of the initial condition. This would necessitate more accurate near-wake models, particularly for cases with complex shear.

To explore the model performance across the whole sweep of ABL cases (see Sec. II E), we analyze the normalized mean absolute deviation (NMAD) of Δu¯s given by
(16)
where the subscript x indicates an average in streamwise direction. This quantity is shown in Fig. 6 for all z0=10 cm ABL cases. The k model performs best, with the lowest values of εs,x for four out of the five cases. The case in which the k model does not perform best is the SBL 0.5.
FIG. 6.

Comparison of the mean streamtube deficit error [Eq. (16)] for the k, Scott, Gaussian, and Jensen models.

FIG. 6.

Comparison of the mean streamtube deficit error [Eq. (16)] for the k, Scott, Gaussian, and Jensen models.

Close modal

The centerline velocity deficit Δu¯c contains less spatial information than Δu¯s. This reduced granularity allows for a comparison that does not directly encode the effect of stability on the shape of the wake deficit. The centerline corresponds to the location of hub height in each yz slice of the wake, assuming the wake centerline is not deflecting in a mean sense given inflow-rotor alignment and limited Coriolis effects from the high Rossby number.28  Figure 7 compares the centerline wake deficit for each model. In the CNBL in Fig. 7(a) and SBL 0.25 in Fig. 7(b), the k model shows the best agreement with the LES. Interestingly, in the most stable case [SBL 1, Fig. 7(c)], the analytical models exhibit the closest agreement with the LES, although this will depend on the selected tuning parameters for the wake spreading rate kw described in Sec. II C.

FIG. 7.

Comparison of the centerline wake velocity deficit for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. LES data are shown by the symbols (○○○). In all cases, z0=10 cm.

FIG. 7.

Comparison of the centerline wake velocity deficit for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. LES data are shown by the symbols (○○○). In all cases, z0=10 cm.

Close modal
To summarize the centerline wake deficit, we can define a metric analogous to Eq. (16),
(17)
The above equation measures the NMAD of Δu¯c predicted by the models from the LES. Figure 8 compares εc,x for each model for the five ABL cases. Overall, the k model has the lowest εc,x across the majority of the ABL cases, predominantly corresponding to the flows with the largest Obukhov lengths (CNBL, SBL 0.25, and SBL 0.5). As shown in Fig. 7, the analytical models—particularly the Jensen model—tend to perform best in the high stability or lowest Obukhov length cases (SBL 0.75 and SBL 1), outperforming the k model in several of these very stable cases. Interestingly, the Gaussian model demonstrates the worst performance in the CNBLs compared with the other models. This is likely due to the high value of ambient turbulence intensity in the CNBLs (see Table I), which dictates the wake spreading rate. This is in contrast to the other models that either use more sophisticated turbulence models (Scott and k) or use a value for the wake spreading rate based on a combination of surface roughness and Obukhov length (Jensen).
FIG. 8.

Comparison of the mean centerline deficit error [Eq. (17)] for the k, Scott, Gaussian, and Jensen models.

FIG. 8.

Comparison of the mean centerline deficit error [Eq. (17)] for the k, Scott, Gaussian, and Jensen models.

Close modal
In this section, we assess the power prediction capabilities of the models for a hypothetical waked turbine. The utility of this analysis is that it directly quantifies the degree to which the models predict both the wake momentum deficit magnitude and its shape. This analysis quantifies how impactful the wake deficit shape is for a downwind turbine in the realistic application setting of wake modeling. We calculate the power according to the rotor-equivalent power model54,55
(18)
where ρ is the air density, Cp is the power coefficient, and the integral is taken over the area of the rotor disk where (yyT)2+z2<(D/2)2. The base flow uB¯ comes from the LES data for all four wake models, and yT=yh. It should be noted that turbine power production can also be computed via the rotor-equivalent wind speed model,54 in which cube in Eq. (18) is moved outside the integral, thereby computing the disk-averaged wind speed as opposed to disk-averaged power. Both the rotor-equivalent power and rotor-equivalent wind speed models exhibit under and over prediction of power relative to field data in different regimes of speed and direction shear.56 Here, we use the rotor-equivalent power model given in Eq. (18) for power calculated from the wake models normalized by the power calculated from the LES, denoted as P/PLES. The results of relative model performance are the same with the rotor-equivalent wind speed model, with the k model generally outperforming the other models.
Figure 9 shows the power ratio P/PLES for the CNBL, SBL 0.25, and SBL 1 for the four models. In the CNBL and SBL 1, the k model generally outperforms the other models, particularly at x/D5. As in Sec. III B, we define a streamwise NMAD metric to characterize the predictions for each case across the four models. This is given by
(19)
FIG. 9.

Comparison of the power ratio [calculated from Eq. (18)] for a turbine located at yT=yh in the lateral direction for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. Power is shown for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models. The dashed gray line indicates the location where the ratio is equal to 1, and the model exactly matches the LES power prediction.

FIG. 9.

Comparison of the power ratio [calculated from Eq. (18)] for a turbine located at yT=yh in the lateral direction for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. Power is shown for the k ( ), Scott ( ), Gaussian ( ), and Jensen ( ) models. The dashed gray line indicates the location where the ratio is equal to 1, and the model exactly matches the LES power prediction.

Close modal

Figure 10 shows the NMAD for power across the 5  z0=10 cm flows. Except for the SBL-0.25 cases, the k model outperforms all other models in predicting power. To further test the robustness of the k model in its ability to predict the spatial distribution of the wake, we move the hypothetical turbine laterally, such that it is only partially waked. This emphasizes the degree to which the wake models capture the skewing from wind direction shear. The results of this for εP,x are shown in Fig. 11, where the location of yT has been moved from yh to yh+0.5D. Overall, the k model still performs the best, while the Gaussian and Jensen model results tend to worsen as the Obukhov length decreases. This is because the axisymmetric Jensen and Gaussian wake models do not account for wake skewing. Additionally, in both Figs. 10 and 11, the k model consistently outperforms the Scott model, likely due to the spatial information available within the turbulence model.

FIG. 10.

Comparison of the mean power error [Eq. (19)] for the k, Scott, Gaussian, and Jensen models. The hypothetical turbine is located at yT=yh in the lateral direction and averaged over all streamwise locations from x/D=2 to x/D=20.

FIG. 10.

Comparison of the mean power error [Eq. (19)] for the k, Scott, Gaussian, and Jensen models. The hypothetical turbine is located at yT=yh in the lateral direction and averaged over all streamwise locations from x/D=2 to x/D=20.

Close modal
FIG. 11.

Comparison of the mean power error [Eq. (19)] for the k, Scott, Gaussian, and Jensen models. The hypothetical turbine is located at yT=yh+ 0.5D in the lateral direction and averaged over all streamwise locations from x/D=2 to x/D=20.

FIG. 11.

Comparison of the mean power error [Eq. (19)] for the k, Scott, Gaussian, and Jensen models. The hypothetical turbine is located at yT=yh+ 0.5D in the lateral direction and averaged over all streamwise locations from x/D=2 to x/D=20.

Close modal

In Secs. III AIII D, we analyzed four distinct metrics to assess model performance. Here, we aggregate these results to more compactly demonstrate the improved performance of the k model over the other three models across the ABL conditions and error metrics considered. In Table II, we show the average NMAD values (εs,x, εc,x, and εP,x) across the 5  z0=10 cm ABL test cases. Across all metrics, the k model has the lowest average error. For the hypothetical turbine power production, the k model provides a 70% decrease in error relative to the Gaussian model and a 60% decrease relative to the Jensen model. This is important because predicting power for downwind turbines is the typical application of engineering wake models.

TABLE II.

Average error across the five testing cases for the NMAD metrics [Eqs. (16), (17), and (19)] for the four models.

Model Average εs,x Average εc,x Average εP,x
k  0.20  0.22  0.071 
Scott  0.30  0.33  0.11 
Gauss  0.40  0.39  0.24 
Jensen  0.28  0.25  0.17 
Model Average εs,x Average εc,x Average εP,x
k  0.20  0.22  0.071 
Scott  0.30  0.33  0.11 
Gauss  0.40  0.39  0.24 
Jensen  0.28  0.25  0.17 

We additionally look across all four error metrics, and we quantify the fraction of cases and metrics in which each model provided the lowest, second lowest, second highest, and highest error. These results are shown in Table III, where we see that the k model provides the lowest error 80% of the time and the second lowest error 5% of the time. Importantly, the k model never has the highest error, which is in contrast to all other models presented here. These results demonstrate the robustness of the proposed fast-running wake model across ABL conditions and error metrics, including both flow field characteristics and power predictions.

TABLE III.

Percentage of time (across all cases and all four metrics) that a given model has the lowest error (best performance) up through the highest error (worst performance).

Model Lowest error (%) Second lowest error (%) Second highest error (%) Highest error (%)
k  80  15 
Scott  70  15  15 
Gauss  20  30  50 
Jensen  20  40  35 
Model Lowest error (%) Second lowest error (%) Second highest error (%) Highest error (%)
k  80  15 
Scott  70  15  15 
Gauss  20  30  50 
Jensen  20  40  35 

The proposed k model is a new model for wake-added turbulence, in addition to a new fast-running model for mean streamwise velocity deficit. For the proposed k model, we assess the predictions of wake-added turbulence kinetic energy kwake. Figure 12 shows slices of kwake in the yz plane for the CNBL, SBL 0.25, and SBL 1. The k model captures the overall shape of kwake, as well as the magnitude. The other models are unable to predict three-dimensional fields for kwake, which is an inherent advantage of the proposed k model.

FIG. 12.

Contours of kwake for CNBL, SBL 0.25, and SBL 1 with z0=10 cm from the k compared with LES.

FIG. 12.

Contours of kwake for CNBL, SBL 0.25, and SBL 1 with z0=10 cm from the k compared with LES.

Close modal
Separately from the three-dimensional fields, we can compare one-dimensional kwake predictions from the k model and the Crespo–Hernández model20 to LES. The Crespo–Hernández model is used to make prediction of the streamwise wake-added turbulence intensity I+=uu¯uBuB¯/uh. The empirical model equation is given by20,
(20)
and is valid in the far wake region. The Crespo–Hernández model provides predictions of I+, which can be related to kwake via the approximation from van der Laan et al.,57 where the total turbulence intensity and the streamwise turbulence intensity are related by Itotal0.8Iu, giving
(21)
In other words, for the comparison presented here, the Crespo–Hernández model for streamwise wake-added turbulence intensity is used to make predictions of kwake, which is the quantity predicted by the k model and also measured in the LES. The approximation from van der Laan et al.57 was derived for neutral surface layers, which is not strictly applicable to the current wake analysis. Using LES, we evaluated Itotal/Iu (not shown), and it exhibits some stability dependence, but the more pronounced effect is the dependence on the streamwise location (i.e., position in the wake). The results also show that Itotal0.8Iu is a reasonable approximation (i.e., lower error than assuming ItotalIu) for x/D approximately less than 10, which is also where wake-added turbulence intensity is most active. We emphasize that this approximation from van der Laan et al.57 is only used to transform the Crespo–Hernandez empirical model of wake-added streamwise turbulence intensity to wake-added TKE to enable a comparison, as a baseline, to the model developed here. This approximation from van der Laan et al.57 in Eq. (21) is not used in the new k modeling framework.

Figure 13 shows the maximum value of kwake at each x location in the wake for the k and Crespo–Hernández models compared with LES for the CNBL, SBL 0.25, and SBL 1 starting from the start of the far wake region as defined by the near-wake length. It should be noted that due to the asymptotic nature of the Crespo–Hernández model and the non-monotonic behavior of the near-wake length, which is furthest downstream for SBL 0.25 and furthest upstream for SBL 1, the maximum value of the model appears to be non-monotonic. However, the model is only changing between the three flows via I0, so the behavior is actually monotonic as would be expected. The maximum value for the empirical model visualized in this figure is a function of the near-wake length, which is non-monotonic, rather than the Crespo–Hernández model.

FIG. 13.

Maximum kwake for the k model and the empirical Crespo–Hernández20 models compared with LES for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. All cases shown have z0=10 cm.

FIG. 13.

Maximum kwake for the k model and the empirical Crespo–Hernández20 models compared with LES for the (a) CNBL, (b) SBL-0.25, and (c) SBL-1 cases. All cases shown have z0=10 cm.

Close modal

In the CNBL, both models compare well with LES. For the stable cases, the Crespo–Hernández model is unable to predict the changing location of the peak in the maximum kwake profile. While the k model does not perfectly capture the magnitude or location of the peak (particularly in SBL 1), this model does provide results that compare more favorably with the LES.

As a final note on this comparison between the k and Crespo–Hernández models, it is important to emphasize that the current approach in most engineering wake models is to predict the wake expansion rate based on the turbulence intensity incident to the turbine. Therefore, the turbine wake is often not affected by the wake-added turbulence generated by the turbine itself. As a specific example of this, the Gaussian and Jensen model predictions shown throughout this study do not even use a wake-added turbulence model. Several recent engineering wake models have recognized this and incorporated the effect of wake-added turbulence on wake recovery,22–24 but the influence of stratification on the wake-added turbulence has not been a focus. The k model provides dynamically consistent mean velocity predictions, which are based on how the spatially evolving wake-added turbulence, explicitly depending on stability, influences wake recovery.

In this work, a novel wake turbulence model is introduced that utilizes a one-equation k model for the turbulent eddy viscosity within the curled wake model framework that solves the parabolized RANS equations for the wake velocity deficit in an efficient manner. It takes 3.5 s on a standard computer to run the k model for a single turbine case on the given domain, which has 106 grid points. A model equation for the wake-added turbulence kinetic energy is developed with a mixing length based on the wake width and near-wake length. The model is validated against LES data for five ABL test cases corresponding to a sweep over the Obukhov length. The model was also compared against existing engineering wake models to assess performance. Four error metrics (pointwise, streamtube, and centerline wake deficits and turbine power production for a downwind, waked turbine) are used to evaluate the models. The proposed k model exhibited the best performance across the five ABL conditions investigated that are out-of-sample from the calibration of parameters, with a 65% mean reduction in prediction error for the power production of a downwind turbine. The proposed k model also yielded the lowest error 78% of the time across all error metrics and ABL conditions. This illustrates the advantage of this model for wakes in diverse stratified, turbulent, atmospheric boundary layer flow. While the k model was tested in wind conditions out-of-sample from the calibration of parameters, the model calibration was able to benefit from data that come from the present LES using different roughness lengths, which could give the model some advantage over wake models from the literature. However, we anticipate that the qualitative conclusions are robust to this, namely, that incorporating a model for wake-added TKE that depends on stability into the wake momentum solution improves predictions. By improving the fidelity of the turbulence model, especially in its ability to capture three-dimensional effects and stratification, we were able to better capture the wake behavior, even in the most extreme of stratified cases (lowest Obukhov length). This improvement comes from the nonlinear interaction between the mean wake momentum deficit and the stability-dependent wake-added turbulence in space. This spatial modeling is in contrast to the common approach in-state-of-practice engineering wake models where the wake recovery only depends on the turbulence intensity incident to the turbine generating the wake without the feedback inherent to the dynamics or where wake recovery depends on a wake-added turbulence model that does not depend on stability.

Future work should investigate several important avenues for improvements to the given model. Most importantly, a near-wake length model should be developed that accounts for stability so that the near-wake length is not only dependent upon ambient turbulence intensity. Furthermore, the proposed steady model could be extended to dynamic wake modeling. The model should be extended to incorporate Coriolis effects28 to capture lower Rossby number regimes associated with larger wind turbines or lower wind speeds. Finally, the model must be extended and evaluated in predictions of wind farm flows, beyond a single wake, where wake superposition, blockage, and farm-level entrainment play important roles in the turbulence and wind farm dynamics. Future investigation should study appropriate mixing length models for wind farm flows. Future work could consider coupling this parabolized RANS based wake model with a top-down wind farm model to jointly capture turbine wakes and wind farm array effects.58 

The authors gratefully acknowledge support from the National Science Foundation (Fluid Dynamics program, Grant No. FD-2226053, Program Manager: Dr. Ronald D. Joslin). All simulations were performed on the Stampede3 supercomputer under the NSF ACCESS project ATM170028.

The authors have no conflicts to disclose.

Kerry S. Klemmer: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal). Michael F. Howland: Conceptualization (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

The k model has three tunable parameters: Cν, Ck1, and Ck2. As stated in Sec. II B, Ck1 and Ck2 are set to 1 and Cν is set to 0.04 in the present modeling framework. To arrive at these values, we performed sensitivity analysis in which samples were independently drawn from uniform distributions for all three parameters, and we used the streamtube-averaged velocity deficit as our performance metric. The sampling distributions for Ck1 and Ck2 are given by U(0,2) and the sampling distribution for Cν is given by U(0,0.1). These distributions assume little prior knowledge, though the support of the Cν distribution is dictated by known values of this parameter in the literature.35 Across all ABL cases, the results were found to be relatively insensitive to changes in Ck1 and Ck2, so these were set to their mean values. An example of this insensitivity is shown in Fig. 14, which illustrates that in the two cases shown, the error is most strongly affected by changes in Cν. For Cν, we used the range of values across the z0=1,50 cm cases and used the average value that minimized the error across these cases.

FIG. 14.

Sensitivity analysis for Cν, Ck1, and Ck2 for z0=1 cm.

FIG. 14.

Sensitivity analysis for Cν, Ck1, and Ck2 for z0=1 cm.

Close modal

To further assess the parameter values, we tested Cν values of 0.03 and 0.05. The results of these perturbed parameter values are given in Table IV for the averaged NMAD values for streamtube-averaged wake deficit, centerline wake deficit, and hypothetical turbine power production and are compared to the baseline models for reference. Overall, we find that the k model in all cases has the lowest errors for streamtube-averaged wake deficit and power production, and the second lowest error for centerline wake deficit, with the Jensen model yielding the lowest error in this metric when Cν=0.05, but with the k model yielding lower error than the Jensen model when Cν=0.03. These sensitivity experiments show that the k model maintains performance over a range of parameter values.

TABLE IV.

Average error across all 15 cases for the NMAD metrics including the k model with Cν=0.03 and Cν=0.05.

Model Average εs,x Average εc,x Average εP,x
k  0.19  0.24  0.041 
k,Cν=0.03  0.22  0.22  0.039 
k,Cν=0.05  0.20  0.27  0.048 
Scott  0.31  0.35  0.067 
Gauss  0.43  0.39  0.12 
Jensen  0.31  0.24  0.13 
Model Average εs,x Average εc,x Average εP,x
k  0.19  0.24  0.041 
k,Cν=0.03  0.22  0.22  0.039 
k,Cν=0.05  0.20  0.27  0.048 
Scott  0.31  0.35  0.067 
Gauss  0.43  0.39  0.12 
Jensen  0.31  0.24  0.13 

The wake width is a common choice for the mixing length in wind turbine wake models.37–40 This mixing length is an approximation of wakes in stratified ABLs, where the mixing length is often influenced by a variety of factors that limit mixing, including the wall normal distance and stability.41 To evaluate this mixing length model quantitatively, we compute the mixing length from LES for three different stabilities as 2=(Δu¯/z· Δu¯/z)/uw¯wake. The mixing length is computed using least squares regression in a domain that is y[2.5D,2.5D] and z[zh0.6D,zh+D], where zh is the hub-height. The results are shown in Fig. 15. The simple and computationally efficient choice of the wake width as the mixing length provides reasonably accurate predictions in neutral and moderate stabilities. The errors are larger in stronger stability, likely because of more severe effects on the mixing length from the stratification. A comparison of prediction error from the k model using various mixing length models, including the wake width used in this study or constant mixing lengths based on the turbine radius, is shown in Fig. 16. The results show that using the wake width as the mixing length has lower error than using a constant value for the mixing length. More complex mixing length models, especially in the context of wind farm flows with multiple turbine wakes, should be investigated in future work.

FIG. 15.

The mixing lengths calculated from LES data for three stabilities investigated in this study compared with the wake width yD in each case.

FIG. 15.

The mixing lengths calculated from LES data for three stabilities investigated in this study compared with the wake width yD in each case.

Close modal
FIG. 16.

Comparison of the normalized pointwise error [Eq. (15)] for the k model using different mixing length models. We test the mixing length primarily used in this paper, where is equal to the wake width, shown as k. We test two additional mixing length models, =0.08R and =0.16R. Both constant mixing length approaches yield higher error than the mixing length that is equal to the wake width.

FIG. 16.

Comparison of the normalized pointwise error [Eq. (15)] for the k model using different mixing length models. We test the mixing length primarily used in this paper, where is equal to the wake width, shown as k. We test two additional mixing length models, =0.08R and =0.16R. Both constant mixing length approaches yield higher error than the mixing length that is equal to the wake width.

Close modal
1.
T.
Göçmen
,
P.
Van der Laan
,
P.-E.
Réthoré
,
A. P.
Diaz
,
G. C.
Larsen
, and
S.
Ott
, “
Wind turbine wake models developed at the technical university of Denmark: A review
,”
Renewable Sustainable Energy Rev.
60
,
752
769
(
2016
).
2.
R. J.
Stevens
and
C.
Meneveau
, “
Flow structure and turbulence in wind farms
,”
Annu. Rev. Fluid Mech.
49
,
311
339
(
2017
).
3.
G.
Narasimhan
,
D. F.
Gayme
, and
C.
Meneveau
, “
Effects of wind veer on a yawed wind turbine wake in atmospheric boundary layer flow
,”
Phys. Rev. Fluids
7
,
114609
(
2022
).
4.
K. S.
Klemmer
and
M. F.
Howland
, “
Momentum deficit and wake-added turbulence kinetic energy budgets in the stratified atmospheric boundary layer
,”
Phys. Rev. Fluids
9
,
114607
(
2024
).
5.
C. R.
Shapiro
,
G. M.
Starke
,
C.
Meneveau
, and
D. F.
Gayme
, “
A wake modeling paradigm for wind farm design and control
,”
Energies
12
,
2956
(
2019
).
6.
J.
Meyers
,
C.
Bottasso
,
K.
Dykes
,
P.
Fleming
,
P.
Gebraad
,
G.
Giebel
,
T.
Göçmen
, and
J.-W.
van Wingerden
, “
Wind farm flow control: Prospects and challenges
,”
Wind Energy Sci.
7
,
2271
2306
(
2022
).
7.
N. O.
Jensen
, “
A note on wind generator interaction
,” Risø-M Report No. 2411 (
Risø National Laboratory
,
1983
).
8.
M.
Bastankhah
and
F.
Porté-Agel
, “
A new analytical model for wind-turbine wakes
,”
Renewable Energy
70
,
116
123
(
2014
).
9.
M.
Bastankhah
and
F.
Porté-Agel
, “
Experimental and theoretical study of wind turbine wakes in yawed conditions
,”
J. Fluid Mech.
806
,
506
541
(
2016
).
10.
L. P.
Chamorro
and
F.
Porté-Agel
, “
Effects of thermal stability and incoming boundary-layer flow characteristics on wind-turbine wakes: A wind-tunnel study
,”
Boundary Layer Meteorol.
136
,
515
533
(
2010
).
11.
A.
Peña
and
O.
Rathmann
, “
Atmospheric stability-dependent infinite wind-farm models and the wake-decay coefficient
,”
Wind Energy
17
,
1269
1285
(
2014
).
12.
M.
Abkar
and
F.
Porté-Agel
, “
Influence of atmospheric stability on wind-turbine wakes: A large-eddy simulation study
,”
Phys. Fluids
27
,
035104
(
2015
).
13.
M. F.
Howland
,
A. S.
Ghate
,
J. B.
Quesada
,
J. J.
Pena Martínez
,
W.
Zhong
,
F. P.
Larrañaga
,
S. K.
Lele
, and
J. O.
Dabiri
, “
Optimal closed-loop wake steering – Part 2: Diurnal cycle atmospheric boundary layer conditions
,”
Wind Energy Sci.
7
,
345
365
(
2022
).
14.
A.
Kirby
,
T. D.
Dunstan
, and
T.
Nishino
, “
An analytical model of momentum availability for predicting large wind farm power
,”
J. Fluid Mech.
976
,
A24
(
2023
).
15.
M.
Abkar
,
J. N.
Sørensen
, and
F.
Porté-Agel
, “
An analytical model for the effect of vertical wind veer on wind turbine wakes
,”
Energies
11
,
1838
(
2018
).
16.
G.
Narasimhan
,
D. F.
Gayme
, and
C.
Meneveau
, “
Analytical wake modeling in atmospheric boundary layers: Accounting for wind veer and thermal stratification
,”
J. Phys. Conf. Ser.
2767
,
092018
(
2024
).
17.
L. P.
Chamorro
and
F.
Porté-Agel
, “
A wind-tunnel investigation of wind-turbine wakes: Boundary-layer turbulence effects
,”
Boundary Layer Meteorol.
132
,
129
149
(
2009
).
18.
G.-W.
Qian
and
T.
Ishihara
, “
A new analytical wake model for yawed wind turbines
,”
Energies
11
,
665
(
2018
).
19.
K. S.
Klemmer
,
E. P.
Condon
, and
M. F.
Howland
, “
Evaluation of wind resource uncertainty on energy production estimates for offshore wind farms
,”
J. Renewable Sustainable Energy
16
,
013302
(
2024
).
20.
A.
Crespo
and
J.
Hernández
, “
Turbulence characteristics in wind-turbine wakes
,”
J. Wind Eng. Ind. Aerodyn.
61
,
71
85
(
1996
).
21.
A.
Niayifar
and
F.
Porté-Agel
, “
Analytical modeling of wind farms: A new approach for power prediction
,”
Energies
9
,
741
(
2016
).
22.
T.
Duc
,
O.
Coupiac
,
N.
Girard
,
G.
Giebel
, and
T.
Göçmen
, “
Local turbulence parameterization improves the Jensen wake model and its implementation for power optimization of an operating wind farm
,”
Wind Energy Sci.
4
,
287
302
(
2019
).
23.
J.
Pedersen
,
E.
Svensson
,
L.
Poulsen
, and
N.
Nygaard
, “
Turbulence optimized park model with Gaussian wake profile
,”
J. Phys.: Conf. Ser.
2265
,
022063
(
2022
).
24.
J. C.
Risco
,
M.
van der Laan
,
M.
Pedersen
,
A. M.
Forsting
, and
P.-E.
Réthoré
, “
A rans-based surrogate model for simulating wind turbine interaction
,”
J. Phys.: Conf. Ser.
2505
,
012016
(
2023
).
25.
L. A.
Martínez-Tossas
,
J.
Annoni
,
P. A.
Fleming
, and
M. J.
Churchfield
, “
The aerodynamics of the curled wake: A simplified model in view of flow control
,”
Wind Energy Sci.
4
,
127
138
(
2019
).
26.
L. A.
Martínez-Tossas
,
J.
King
,
E.
Quon
,
C. J.
Bay
,
R.
Mudafort
,
N.
Hamilton
,
M. F.
Howland
, and
P. A.
Fleming
, “
The curled wake model: A three-dimensional and extremely fast steady-state wake solver for wind plant flows
,”
Wind Energy Sci.
6
,
555
570
(
2021
).
27.
R.
Scott
,
L.
Martínez-Tossas
,
J.
Bossuyt
,
N.
Hamilton
, and
R. B.
Cal
, “
Evolution of eddy viscosity in the wake of a wind turbine
,”
Wind Energy Sci.
8
,
449
463
(
2023
).
28.
K. S.
Heck
and
M. F.
Howland
, “
Coriolis effects on wind turbine wakes across neutral atmospheric boundary layer regimes
,”
J. Fluid Mech.
1008
,
A7
(
2025
).
29.
K.
Devesse
,
L.
Lanzilao
, and
J.
Meyers
, “
A meso–micro atmospheric perturbation model for wind farm blockage
,”
J. Fluid Mech.
998
,
A63
(
2024
).
30.
O.
Ndindayino
,
A.
Puel
, and
J.
Meyers
, “
Effect of blockage on wind turbine power and wake development
,”
Wind Energy Sci.
(submitted, 2025).
31.
A.
Kirby
,
T.
Nishino
,
L.
Lanzilao
,
T. D.
Dunstan
, and
J.
Meyers
, “
Turbine-and farm-scale power losses in wind farms: An alternative to wake and farm blockage losses
,”
Wind Energy Sci.
10
,
435
450
(
2025
).
32.
M. F.
Howland
,
A. S.
Ghate
, and
S. K.
Lele
, “
Influence of the geostrophic wind direction on the atmospheric boundary layer flow
,”
J. Fluid Mech.
883
,
A39
(
2020
).
33.
J.
Liew
,
K. S.
Heck
, and
M. F.
Howland
, “
Unified momentum model for rotor aerodynamics across operating regimes
,”
Nat. Commun.
15
,
6658
(
2024
).
34.
M. V.
Lingad
,
M.
Rodrigues
,
S.
Leonardi
, and
A.
Zare
, “
Three-dimensional stochastic dynamical modeling for wind farm flow estimation
,”
J. Phys. Conf. Ser.
2767
,
052065
(
2024
).
35.
D. C.
Wilcox
,
Turbulence Modeling for CFD
(
DCW Industries
,
1998
).
36.
K. S.
Heck
,
H. M.
Johlas
, and
M. F.
Howland
, “
Modelling the induction, thrust and power of a yaw-misaligned actuator disk
,”
J. Fluid Mech.
959
,
A9
(
2023
).
37.
J. F.
Ainslie
, “
Calculating the flowfield in the wake of wind turbines
,”
J. Wind Eng. Ind. Aerodyn.
27
,
213
224
(
1988
).
38.
R.-E.
Keck
,
D.
Veldkamp
,
H. A.
Madsen
, and
G.
Larsen
, “
Implementation of a mixing length turbulence formulation into the dynamic wake meandering model
,”
J. Sol. Energy Eng.
134
,
021012
(
2012
).
39.
C. R.
Shapiro
,
D. F.
Gayme
, and
C.
Meneveau
, “
Generation and decay of counter-rotating vortices downstream of yawed wind turbines in the atmospheric boundary layer
,”
J. Fluid Mech.
903
,
R2
(
2020
).
40.
P.
Hulsman
,
L. A.
Martínez-Tossas
,
N.
Hamilton
, and
M.
Kühn
, “
Modelling and assessing the near-wake representation and turbulence behaviour of control-oriented wake models
,”
J. Phys.: Conf. Ser.
1618
,
022056
(
2020
).
41.
I.
Lopez-Gomez
,
Y.
Cohen
,
J.
He
,
A.
Jaruga
, and
T.
Schneider
, “
A generalized mixing length closure for eddy-diffusivity mass-flux schemes of turbulence and convection
,”
J. Adv. Model. Earth Syst.
12
,
e2020MS002161
(
2020
).
42.
NREL
(
2024
), “
FLORIS wake modeling and wind farm controls software
,” GitHub. https://github.com/NREL/floris.
43.
M. J.
LoCascio
,
C.
Gorle
, and
M. F.
Howland
, “
Data-driven wake model parameter estimation to analyze effects of wake superposition
,”
J. Renewable Sustainable Energy
15
,
063304
(
2023
).
44.
A.
Peña
,
P.-E.
Réthoré
, and
M. P.
van der Laan
, “
On the application of the Jensen wake model using a turbulence-dependent wake decay coefficient: The Sexbierum case
,”
Wind Energy
19
,
763
776
(
2016
).
45.
R. B.
Stull
,
An Introduction to Boundary Layer Meteorology
(
Springer Science & Business Media
,
2012
).
46.
A. S.
Ghate
and
S. K.
Lele
, “
Subfilter-scale enrichment of planetary boundary layer large eddy simulation using discrete Fourier–Gabor modes
,”
J. Fluid Mech.
819
,
494
539
(
2017
).
47.
S.
Nagarajan
,
S. K.
Lele
, and
J. H.
Ferziger
, “
A robust high-order compact method for large eddy simulation
,”
J. Comput. Phys.
191
,
392
419
(
2003
).
48.
S.
Gottlieb
,
C.-W.
Shu
, and
E.
Tadmor
, “
Strong stability-preserving high-order time discretization methods
,”
SIAM Rev.
43
,
89
112
(
2001
).
49.
F.
Nicoud
,
H. B.
Toda
,
O.
Cabrit
,
S.
Bose
, and
J.
Lee
, “
Using singular values to build a subgrid-scale model for large eddy simulations
,”
Phys. Fluids
23
,
085106
(
2011
).
50.
J.
Nordström
,
N.
Nordin
, and
D.
Henningson
, “
The fringe region technique and the Fourier method used in the direct numerical simulation of spatially evolving viscous flows
,”
SIAM J. Sci. Comput.
20
,
1365
1393
(
1999
).
51.
R. J. A. M.
Stevens
,
J.
Graham
, and
C.
Meneveau
, “
A concurrent precursor inflow method for Large Eddy Simulations and applications to finite length wind farms
,”
Renewable Energy
68
,
46
50
(
2014
).
52.
M.
Calaf
,
C.
Meneveau
, and
J.
Meyers
, “
Large eddy simulation study of fully developed wind-turbine array boundary layers
,”
Phys. Fluids
22
,
015110
(
2010
).
53.
S.
Basu
,
A. A.
Holtslag
,
B. J.
Van De Wiel
,
A. F.
Moene
, and
G.-J.
Steeneveld
, “
An inconvenient “truth” about using sensible heat flux as a surface boundary condition in models under stably stratified regimes
,”
Acta Geophys.
56
,
88
99
(
2008
).
54.
R.
Wagner
,
I.
Antoniou
,
S. M.
Pedersen
,
M. S.
Courtney
, and
H. E.
Jørgensen
, “
The influence of the wind speed profile on wind turbine performance measurements
,”
Wind Energy
12
,
348
362
(
2009
).
55.
A.
Choukulkar
,
Y.
Pichugina
,
C. T. M.
Clack
,
R.
Calhoun
,
R.
Banta
,
A.
Brewer
, and
M.
Hardesty
, “
A new formulation for rotor equivalent wind speed for wind resource assessment and wind power forecasting
,”
Wind Energy
19
,
1439
1452
(
2016
).
56.
S. A.
Mata
,
J. J.
Pena Martínez
,
J.
Bas Quesada
,
F.
Palou Larrañaga
,
N.
Yadav
,
J. S.
Chawla
,
V.
Sivaram
, and
M. F.
Howland
, “
Modeling the effect of wind speed and direction shear on utility-scale wind turbine power production
,”
Wind Energy
27
,
873
899
(
2024
).
57.
M. P.
van der Laan
,
N. N.
Sørensen
,
P.-E.
Réthoré
,
J.
Mann
,
M. C.
Kelly
,
N.
Troldborg
,
J. G.
Schepers
, and
E.
Machefaux
, “
An improved k-ε model applied to a wind turbine wake in atmospheric turbulence
,”
Wind Energy
18
,
889
907
(
2015
).
58.
R. J.
Stevens
,
D. F.
Gayme
, and
C.
Meneveau
, “
Coupled wake boundary layer model of wind-farms
,”
J. Renewable Sustainable Energy
7
,
023115
(
2015
).
Published open access through an agreement withMassachusetts Institute of Technology