Wake dynamics of wind turbines in unsteady streamwise flow conditions

Nathaniel J. Wei\aff1,2,3 \corresp [email protected]    Adnan El Makdah\aff4    JiaCheng Hu\aff4    Frieder Kaiser\aff4    David E. Rival\aff4,5       John O. Dabiri\aff3,6 \aff1Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, USA \aff2Andlinger Center for Energy and the Environment, Princeton University, Princeton, NJ 08544, USA \aff3Graduate Aerospace Laboratories, California Institute of Technology, Pasadena, CA 91125, USA \aff4Mechanical and Materials Engineering, Queen’s University, Kingston, ON, Canada K7L 3N6 \aff5Institute of Fluid Mechanics, Technische Universität Braunschweig, 38108 Braunschweig, Germany \aff6Mechanical and Civil Engineering, California Institute of Technology, Pasadena, CA 91125, USA
Abstract

The unsteady flow physics of wind-turbine wakes under dynamic forcing conditions are critical to the modeling and control of wind farms for optimal power density. Unsteady forcing in the streamwise direction may be generated by unsteady inflow conditions in the atmospheric boundary layer, dynamic induction control of the turbine, or streamwise surge motions of a floating offshore wind turbine due to floating-platform oscillations. This study seeks to identify the dominant flow mechanisms in unsteady wakes forced by a periodic upstream inflow condition. A theoretical framework for the problem is derived, which describes undulations in the wake radius and streamwise velocity that propagate as traveling waves downstream in the wake. These dynamics encourage the aggregation of tip vortices into large structures that are advected along in the wake. Flow measurements in the wake of a periodically surging turbine were obtained in an optically accessible towing-tank facility, with an average diameter-based Reynolds number of 300,000 and with surge-velocity amplitudes of up to 40% of the mean inflow velocity. Qualitative agreement between trends in the measurements and model predictions is observed, supporting the validity of the theoretical analyses. The experiments also demonstrate large enhancements in the recovery of the wake relative to the steady-flow case, with wake-length reductions of up to 46.5% and improvements in the available power at 10 diameters downstream of up to 15.7%. These results provide fundamental insights into the dynamics of unsteady wakes and serve as additional evidence that unsteady fluid mechanics can be leveraged to increase the power density of wind farms.

keywords:
Aerodynamics, vortex dynamics, wakes

1 Introduction

As the deployment of renewable energy continues to gain momentum worldwide, it is increasingly necessary to consider ways to optimize the power-generation capacity of large aggregations of these energy-harvesting systems. For this task, increasing the power density, or power generated per unit occupied land or sea area, is a critical objective for maximizing the contribution of renewable-energy sources to global energy demands while minimizing space and infrastructure requirements. For wind turbines, a major limiting factor to the power density of a wind farm is the wake regions downstream of each turbine. Downstream turbines that operate in these regions of low-speed and highly turbulent flow suffer drastic losses in power generation, thereby decreasing the overall power density of the array on the order of ten to twenty percent (Barthelmie et al., 2009).

Several methods for reducing these wake losses have been explored in recent years, including layout optimization using wake models (cf. Stevens & Meneveau, 2017), wake steering by turbine yaw misalignment (e.g. Howland et al., 2019), and flow control (cf. Meyers et al., 2022). Many of these approaches operate under the assumption that the incoming flow and its interactions with the turbines in an array behave in steady or quasi-steady manners. By contrast, flow conditions in the atmospheric boundary layer are inherently unsteady, and fluctuations across a wide range of length and time scales affect the power generation and wake dynamics of real wind farms. Therefore, including the effects of unsteady dynamics in the design, optimization, and analysis of wind farms can potentially uncover new strategies for maximizing power density (Dabiri, 2020).

Accordingly, significant efforts have recently been invested into leveraging unsteady fluid mechanics for improved turbine performance and reduced wake losses on downstream turbines. Goit & Meyers (2015), Munters & Meyers (2017), and Munters & Meyers (2018) introduced the idea of dynamic induction control, in which oscillations in the thrust force of a turbine are generated to excite wake instabilities, increase the mixing of high-momentum fluid in the free-stream flow into the wake, and thereby achieve accelerated wake recovery relative to steady-flow turbine operation. Wind-tunnel experiments by Frederik et al. (2020b) and van der Hoek et al. (2022) have demonstrated the effectiveness of this approach in enhancing the wake recovery downstream of a single turbine. A similar dynamic control strategy, involving individual pitch control of the turbine blades to create helical disturbances in the wake, has recently been developed by Frederik et al. (2020a) and tested in wind-tunnel experiments (van der Hoek et al., 2024). Other forms of dynamic wake control that introduce oscillations in the turbine rotation rate to excite tip-vortex pairing instabilities have been explored by Brown et al. (2022). Meyers et al. (2022) provide a review of many recent studies of these kinds of wake-control approaches.

For floating offshore wind turbines (FOWTs), the possibility of periodic turbine motions as a function of unsteady wind gusts, wave forcing, and floating-platform hydrodynamics represents an additional opportunity for leveraging unsteady flows for increased power density in offshore wind farms. Building on the work of Wen et al. (2018), Johlas et al. (2021), and others, Wei & Dabiri (2022) and Wei & Dabiri (2023) demonstrated using wind-tunnel experiments and analytical models that periodically surging turbines (i.e. turbines moving in linear oscillations along the direction of the incoming wind) can generate over six percent more power in certain loading conditions than equivalent stationary turbines. El Makdah et al. (2019) also measured large increases in power generation for turbines in axial ramp-up gusts, which could suggest that fixed-bottom turbines in unsteady streamwise flows can achieve time-averaged power-generation enhancements as well. The effects of unsteady turbine motions on the turbine wake and power generation of downstream turbines, however, are less well understood.

Several studies have investigated the wake dynamics of floating offshore wind turbines moving in streamwise rocking or linear-surge motions (cf. Cioni et al., 2023). Fontanella et al. (2022) observed traveling-wave oscillations in the streamwise velocity downstream of a periodically surging wind turbine in a wind tunnel, but at a measurement distance of 2.3 turbine diameters into the wake, no changes in wake recovery were found. In wind-tunnel experiments, Rockel et al. (2016) found that the rocking motions of a FOWT lead to a suppression in the entrainment of kinetic energy into the wake and therefore a decrease in the wake-recovery rate. Conversely, recent measurements by Messmer et al. (2024) using a turbine mounted on an actuated Stewart platform showed increases in wake recovery in excess of 20% relative to the fixed-turbine case, as a function of streamwise surge and transverse sway motions. Similarly, experiments by Bossuyt et al. (2023) involving arrays of scaled FOWT models undergoing simultaneous wind and wave forcing demonstrated improvements in power density with increased wave-induced platform oscillations. Additionally, van den Broek et al. (2023) and van den Berg et al. (2023) have investigated the coupled effects of dynamic induction control and turbine motions using free-vortex wake simulations, finding that improvements in wake recovery due to periodic turbine-thrust oscillations can in some cases be mitigated by the motions of the turbine, leading to a reduced overall effect on wake recovery. Based on the mixed results of these studies, it is apparent that a better understanding of the fluid mechanics underlying unsteady wake behaviors in FOWTs is needed.

The main purpose of the present work is to theoretically and experimentally investigate the unsteady wake dynamics generated by streamwise forcing from a turbine, in order to identify the dominant flow mechanisms and to determine their effects on wake dynamics and recovery. The current approach focuses on streamwise forcing because of the salience of these disturbances in FOWT platform dynamics (Johlas et al., 2019; Bossuyt et al., 2023), the direct relevance to the dynamic induction control literature, and the potential for power-generation enhancements in streamwise unsteady flows (Dabiri, 2020; Wei & Dabiri, 2023). The theoretical analysis is designed to be agnostic to the source of wake unsteadiness, and is therefore equally applicable to stationary turbines with dynamic induction control and FOWTs undergoing streamwise surge motions. While the experiments presented in this work use a periodically surging turbine to generate unsteady wakes, the results and conclusions should in principle apply to dynamic induction control scenarios as well. Lastly, this study identifies fundamental flow mechanisms for unsteady wakes that may find broader applications outside of wind-energy contexts, including hydrokinetic turbines in tidal flows (e.g. Scarlett & Viola, 2020), streamwise-oscillating cylinders (e.g. Currie & Turnbull, 1987), bio-inspired propulsors in free- or intermittent-swimming conditions (cf. Smits, 2019), and vehicles propelled by oscillatory jets (e.g. Ruiz et al., 2011).

The work is structured as follows. In Section 2, a theoretical framework for wakes with oscillatory streamwise inflow conditions is described, and its implications for wake properties and vortex dynamics are discussed. The assumptions, strengths, and limitations of the analysis approach are also considered. In Section 3, an experiment to obtain flow measurements in the wake of a periodically surging turbine using an optical towing-tank facility is described. Results from the experiments are presented in Section 4 and are compared qualitatively with predictions from the theoretical model to demonstrate that the modeling approach captures the dominant physics of the unsteady-wake problem. Implications of the findings for wind-energy applications are surveyed at the end of Section 4, and conclusions are given in Section 5.

2 Theoretical considerations

In this section, we derive a system of coupled partial differential equations as a phenomenological model for the unsteady dynamics of a turbine wake with an oscillatory upstream boundary condition. This general formulation can be applied to dynamic induction control for stationary turbines and streamwise periodic surge motions in FOWTs. The approach relies on several simplifying abstractions from real turbine wakes, and is therefore not intended to be fully quantitative. However, the theoretical framework can still provide useful insights into the dominant flow physics of the unsteady-wake problem, and it can also be applied to a kinematic description of vortex dynamics in the near wake.

In this work, we denote time averages with overbars, phase averages with tildes, and amplitudes with circumflexes. Quantities referring to steady-flow or quasi-steady measurements are written with a zero subscript (e.g. λ0subscript𝜆0\lambda_{0}italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT). For velocities, lowercase letters represent velocity fields that vary in space and time, e.g. u(x,r,t)𝑢𝑥𝑟𝑡u(x,r,t)italic_u ( italic_x , italic_r , italic_t ), where r𝑟ritalic_r represents the radial coordinate. The uppercase letter U𝑈Uitalic_U denotes the radially averaged streamwise velocity, i.e. u(r)𝑢𝑟u(r)italic_u ( italic_r ) averaged in space from r=0𝑟0r=0italic_r = 0 to the wake radius R𝑅Ritalic_R. The velocity of the turbine surge motions is defined as 𝒰(t)𝒰𝑡\mathcal{U}(t)caligraphic_U ( italic_t ) to distinguish it from these other velocities, and it is assumed to be periodic with the form

𝒰(t)U=usin(2πtT),𝒰𝑡subscript𝑈superscript𝑢2𝜋𝑡𝑇\frac{\mathcal{U}(t)}{U_{\infty}}=u^{*}\sin\left(2\pi\frac{t}{T}\right)\,,divide start_ARG caligraphic_U ( italic_t ) end_ARG start_ARG italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG = italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT roman_sin ( 2 italic_π divide start_ARG italic_t end_ARG start_ARG italic_T end_ARG ) , (1)

where T𝑇Titalic_T is the surge period and usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is the surge-velocity amplitude. The reduced frequency associated with these motions is

k=2πDTU,𝑘2𝜋𝐷𝑇subscript𝑈k=\frac{2\pi D}{TU_{\infty}}\,,italic_k = divide start_ARG 2 italic_π italic_D end_ARG start_ARG italic_T italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG , (2)

where D𝐷Ditalic_D is the turbine diameter. This is referred to as the Strouhal number in several recent studies on dynamic induction control for wind turbines (e.g. Munters & Meyers, 2018; Frederik et al., 2020b; Messmer et al., 2024).

2.1 Governing equations for streamwise-unsteady wake dynamics

To derive the governing equations for a turbine wake with streamwise unsteadiness, we define a control volume with a variable wake radius R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ), as shown in Figure 1. The flow inside the control volume is treated as incompressible and quasi one-dimensional, such that all flow occurs in the streamwise direction and the streamwise velocity U(x,t)𝑈𝑥𝑡U(x,t)italic_U ( italic_x , italic_t ) does not vary in the radial direction. The flow is additionally assumed to be inviscid in the near wake, where tip vortices dominate and viscous contributions to the kinetic-energy budget are comparatively small. The turbine itself is modeled as an actuator disc, which generates a thrust force on the flow and defines the inlet boundary condition to the control volume at x=0𝑥0x=0italic_x = 0. Variations in the thrust force due to unsteady inflow conditions or streamwise surge motions will create an oscillatory inlet condition Ui(x=0,t)subscript𝑈𝑖𝑥0𝑡U_{i}(x=0,t)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ) that will dictate the dynamics in the near wake. We do not seek to model the coupling between the turbine dynamics and this near-wake inflow condition here, assuming for simplicity a periodic inflow condition of the form Ui(x=0,t)=U¯+U^sin(ft)subscript𝑈𝑖𝑥0𝑡¯𝑈^𝑈𝑓𝑡U_{i}(x=0,t)=\overline{U}+\hat{U}\sin(ft)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ) = over¯ start_ARG italic_U end_ARG + over^ start_ARG italic_U end_ARG roman_sin ( italic_f italic_t ).

Refer to caption
Figure 1: Sketch of the quasi-1D axisymmetric problem formulation for an unsteady turbine wake.

Given these assumptions, the equations for conservation of mass and momentum within the wake control volume can be derived. (A detailed derivation is provided in Appendix A.) The relation for conservation of mass is given as a first-order nonlinear partial differential equation in terms of R𝑅Ritalic_R and U𝑈Uitalic_U:

Rt+12RUx+URx=0.𝑅𝑡12𝑅𝑈𝑥𝑈𝑅𝑥0\frac{\partial R}{\partial t}+\frac{1}{2}R\frac{\partial U}{\partial x}+U\frac% {\partial R}{\partial x}=0\,.divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG + italic_U divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG = 0 . (3)

If r𝑟ritalic_r is held constant, Equation 3 reduces to the more familiar form Ux=0𝑈𝑥0\frac{\partial U}{\partial x}=0divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG = 0. The equation therefore captures the effects of changes in the flow velocity on the local size of the wake. For a quasi-1D incompressible flow, a decrease in the flow velocity across a given distance ΔxΔ𝑥\Delta xroman_Δ italic_x must result in a corresponding increase in the cross-sectional area of the control volume so that the mass flux remains constant. Similarly, an increase in the flow velocity will result in a decrease in the cross-sectional area. Hence, if U(x,t)𝑈𝑥𝑡U(x,t)italic_U ( italic_x , italic_t ) takes the form of a traveling wave, Equation 3 dictates that the wake radius will also form traveling waves so that the mass flux at every streamwise location in the control volume is conserved.

The relation for conservation of momentum is the 1D Euler equation,

Ut+UUx=1ρpx.𝑈𝑡𝑈𝑈𝑥1𝜌𝑝𝑥\frac{\partial U}{\partial t}+U\frac{\partial U}{\partial x}=-\frac{1}{\rho}% \frac{\partial p}{\partial x}\,.divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + italic_U divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG = - divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ italic_p end_ARG start_ARG ∂ italic_x end_ARG . (4)

Here, we have assumed that the pressure is constant in the radial direction, both inside and outside of the wake. The pressure term represents a heterogeneous forcing on a Burgers-type partial differential equation. In kee** with the assumptions of many turbine wake models (e.g. Bastankhah & Porté-Agel, 2014), we choose to neglect this term, though in reality the pressure recovery in the near-wake region will be non-negligible.

For the remaining terms in the momentum equation, we decompose the flow velocity into phase-averaged and fluctuating components, in the style of a Reynolds decomposition:

U(x,t)=U~(x,t)+U(x,t).𝑈𝑥𝑡~𝑈𝑥𝑡superscript𝑈𝑥𝑡U(x,t)=\tilde{U}(x,t)+U^{\prime}(x,t)\,.italic_U ( italic_x , italic_t ) = over~ start_ARG italic_U end_ARG ( italic_x , italic_t ) + italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x , italic_t ) . (5)

Phase-averaging Equation 4 using this decomposition yields

U~t+U~U~x=12xU2~.~𝑈𝑡~𝑈~𝑈𝑥12𝑥~superscriptsuperscript𝑈2\frac{\partial\tilde{U}}{\partial t}+\tilde{U}\frac{\partial\tilde{U}}{% \partial x}=-\frac{1}{2}\frac{\partial}{\partial x}\widetilde{{U^{\prime}}^{2}% }\,.divide start_ARG ∂ over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_t end_ARG + over~ start_ARG italic_U end_ARG divide start_ARG ∂ over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_x end_ARG = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_x end_ARG over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (6)

To model the normal-stress term on the right-hand side of the equation, we combine two scaling relationships for wind-turbine wakes in steady inflow conditions. Quarton & Ainslie (1990) show that the turbulence intensity u/Usuperscript𝑢subscript𝑈u^{\prime}/U_{\infty}italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT / italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT decays monotonically toward zero with increasing streamwise distance. It is also well-known that the velocity in a turbine wake recovers monotonically toward the free-stream wind speed in a power-law relationship with streamwise distance (cf. Porté-Agel et al., 2020). Thus, one can typically assume that the magnitude of the streamwise velocity gradient ux𝑢𝑥\frac{\partial u}{\partial x}divide start_ARG ∂ italic_u end_ARG start_ARG ∂ italic_x end_ARG will decay monotonically toward zero as well. Combining these two observations, we argue that

U2~|dU~dx|,similar-to~superscriptsuperscript𝑈2𝑑~𝑈𝑑𝑥\widetilde{{U^{\prime}}^{2}}\sim\left|\frac{d\tilde{U}}{dx}\right|\,,over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∼ | divide start_ARG italic_d over~ start_ARG italic_U end_ARG end_ARG start_ARG italic_d italic_x end_ARG | , (7)

or in the form of a turbulent-viscosity hypothesis,

UU~=νu|U~x|.~superscript𝑈superscript𝑈subscript𝜈𝑢~𝑈𝑥-\widetilde{U^{\prime}U^{\prime}}=\nu_{u}\left|\frac{\partial\tilde{U}}{% \partial x}\right|\,.- over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG = italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT | divide start_ARG ∂ over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_x end_ARG | . (8)

The proportionality parameter νusubscript𝜈𝑢\nu_{u}italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT may vary as a function of x𝑥xitalic_x; we will obtain a rudimentary model from experimental measurements in Section 4 of the form

νu=κxsubscript𝜈𝑢𝜅𝑥\nu_{u}=\kappa xitalic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = italic_κ italic_x (9)

(cf. Figure 13). It is important to note that νusubscript𝜈𝑢\nu_{u}italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT is fundamentally different from a typical eddy viscosity based on the Reynolds shear stress, since it represents the interactions of streamwise quantities and has nothing to do with shear. Equation 8 thus suggests that stronger streamwise gradients are associated with stronger velocity fluctuations, and a linear model enforces a stronger coupling between the two quantities with increasing downstream distance.

Applying this model to Equation 6 yields

U~t+U~U~x+νu(x)2U~x2=0.~𝑈𝑡~𝑈~𝑈𝑥subscript𝜈𝑢𝑥superscript2~𝑈superscript𝑥20\frac{\partial\tilde{U}}{\partial t}+\tilde{U}\frac{\partial\tilde{U}}{% \partial x}+\nu_{u}(x)\frac{\partial^{2}\tilde{U}}{\partial x^{2}}=0\,.divide start_ARG ∂ over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_t end_ARG + over~ start_ARG italic_U end_ARG divide start_ARG ∂ over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_x end_ARG + italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ) divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over~ start_ARG italic_U end_ARG end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 0 . (10)

This first-order nonlinear partial differential equation is a viscous Burgers equation, which describes the growth, propagation, and steepening of nonlinear traveling waves. The proportionality parameter νu(x)subscript𝜈𝑢𝑥\nu_{u}(x)italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ) therefore represents a dam** term that inhibits the steepening of the waves and the formation of unphysical shock discontinuities, which would occur in the inviscid form of the Burgers equation (νu=0subscript𝜈𝑢0\nu_{u}=0italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = 0). This is physically consistent with the underlying UU~~superscript𝑈superscript𝑈\widetilde{U^{\prime}U^{\prime}}over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG quantity that the dam** term represents, which is related to turbulent convection and transport in the turbulent kinetic energy budget for axisymmetric wakes (Uberoi & Freymuth, 1970). This term thus models the transfer of energy from the phase-averaged base flow to turbulence via streamwise velocity gradients, which are created in this system by the time-varying inflow condition Uisubscript𝑈𝑖U_{i}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and the nonlinear wave steepening generated by the first two terms in Equation 10 (i.e. the inviscid Burgers equation).

In summary, the dynamics of the unsteady wake can be written as a system of two first-order nonlinear partial differential equations,

Rt+12RUx+URx=0and𝑅𝑡12𝑅𝑈𝑥𝑈𝑅𝑥0and\displaystyle\frac{\partial R}{\partial t}+\frac{1}{2}R\frac{\partial U}{% \partial x}+U\frac{\partial R}{\partial x}=0\;\rm{and}divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG + italic_U divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG = 0 roman_and (11a)
Ut+UUx+νu(x)2Ux2=0,𝑈𝑡𝑈𝑈𝑥subscript𝜈𝑢𝑥superscript2𝑈superscript𝑥20\displaystyle\frac{\partial U}{\partial t}+U\frac{\partial U}{\partial x}+\nu_% {u}(x)\frac{\partial^{2}U}{\partial x^{2}}=0\,,divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + italic_U divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG + italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ) divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_U end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 0 , (11b)

where the tildes denoting phase-averaging have been removed for clarity and will not be used for the remainder of this work. The system is subject to the boundary conditions

Rit|x=0,tevaluated-atsubscript𝑅𝑖𝑡𝑥0𝑡\displaystyle\left.\frac{\partial R_{i}}{\partial t}\right|_{x=0,\,t}divide start_ARG ∂ italic_R start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG | start_POSTSUBSCRIPT italic_x = 0 , italic_t end_POSTSUBSCRIPT =0andabsent0and\displaystyle=0\;\rm{and}= 0 roman_and (12a)
Ui|x=0,tevaluated-atsubscript𝑈𝑖𝑥0𝑡\displaystyle\left.U_{i}\right|_{x=0,\,t}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | start_POSTSUBSCRIPT italic_x = 0 , italic_t end_POSTSUBSCRIPT =U¯+U^sin(ft).absent¯𝑈^𝑈𝑓𝑡\displaystyle=\overline{U}+\hat{U}\sin(ft)\,.= over¯ start_ARG italic_U end_ARG + over^ start_ARG italic_U end_ARG roman_sin ( italic_f italic_t ) . (12b)

The equations exhibit one-way coupling, as the momentum equation only depends on U𝑈Uitalic_U and serves as a forcing on R𝑅Ritalic_R through the continuity equation. The time-varying boundary condition Uisubscript𝑈𝑖U_{i}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and dam** term in Equation 11b preclude the straightforward derivation of explicit analytical solutions, but it is well-known that the viscous Burgers equation generates damped nonlinear traveling waves and can be solved numerically. For Equation 11a, the method of characteristics can be applied by defining a characteristic variable ξ=xUt𝜉𝑥𝑈𝑡\xi=x-Utitalic_ξ = italic_x - italic_U italic_t, such that along lines of constant ξ𝜉\xiitalic_ξ, the partial differential equation reduces to a pair of ordinary differential equations,

dxdt|ξevaluated-at𝑑𝑥𝑑𝑡𝜉\displaystyle\left.\frac{dx}{dt}\right|_{\xi}divide start_ARG italic_d italic_x end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT =Uandabsent𝑈and\displaystyle=U\;\rm{and}= italic_U roman_and (13a)
dRdt|ξevaluated-at𝑑𝑅𝑑𝑡𝜉\displaystyle\left.\frac{dR}{dt}\right|_{\xi}divide start_ARG italic_d italic_R end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT =12UxR.absent12𝑈𝑥𝑅\displaystyle=-\frac{1}{2}\frac{\partial U}{\partial x}R\,.= - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG italic_R . (13b)

The first equation describes the advection of solutions for R𝑅Ritalic_R according to the traveling-wave velocity U𝑈Uitalic_U, while the second governs the growth or decay of the wake-radius amplitude independent of advection. Given the wave steepening that is inherent to solutions of Burgers-type equations for U𝑈Uitalic_U, we expect that the magnitude of Ux𝑈𝑥\frac{\partial U}{\partial x}divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG will be greater on the downstream side of the wave (where Ux<0𝑈𝑥0\frac{\partial U}{\partial x}<0divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG < 0) than on the upstream side of the wave. Thus, from Equation 13b, we anticipate that the period-averaged amplitude of R𝑅Ritalic_R will grow under the forcing provided by Equation 11b until streamwise velocity gradients are damped out, at which point the amplitude of R𝑅Ritalic_R will saturate and waves in the wake radius will simply advect downstream.

To demonstrate the dynamics of the system, results from numerical integrations are shown in Figure 2. The upstream boundary conditions were set based on measured values from experiments (cf. Figure 12 in Section 4.2). A third-order upwind scheme was used to discretize the Ux𝑈𝑥\frac{\partial U}{\partial x}divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG term, while a second-order central-difference scheme was used for the 2Ux2superscript2𝑈superscript𝑥2\frac{\partial^{2}U}{\partial x^{2}}divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_U end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG term. As we have anticipated in our analysis, the model produces damped nonlinear traveling waves in the velocity, which in turn result in waves in the wake radius that grow and saturate in amplitude as they propagate downstream.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Figure 2: Numerical solutions to the coupled-PDE modeling framework for the wake radius (left) and streamwise velocity (right), for three example cases.

2.2 Vortex dynamics

The kinematics of the control volume described by the system of equations derived in the previous section also allow predictions regarding the vortex dynamics in the wake to be made. Tip vortices shed by the turbine blades typically bound the wake of the turbine until they break down in the intermediate wake. This is a three-dimensional helical structure, but given the axisymmetric nature of the turbine wake, the successive appearances of the helical vortex line at a single azimuthal orientation are often treated as a discrete series of 2D point vortices (e.g. de Vaal et al., 2014b; van den Broek et al., 2023). In steady conditions, these vortex elements are generally arranged in relatively straight lines extending downstream from the blade tips. For an unsteady wake with a time-varying wake radius, the tip vortices will experience radial displacements along with the wake radius. We can therefore consider the dynamics of a series of discrete point vortices arranged in non-collinear patterns, as a representation of a helical tip vortex undergoing spatial and temporal changes in its radius. For more detailed stability analyses of helical tip vortices under dynamic conditions, we refer the reader to the work of Kleine et al. (2022) on the wakes of moving FOWTs and Rodriguez et al. (2021) on the wakes of turbines with flexible blades.

Consider an infinitely long line of point vortices with equal circulation ΓΓ\Gammaroman_Γ and equal spacing ΔxΔ𝑥\Delta xroman_Δ italic_x. The induced velocity from the i𝑖iitalic_ith vortex on a given vortex j𝑗jitalic_j is given by

[uv]i,j=Γ2π((xjxi)2+(yjyi)2)[(yjyi)xjxi],subscriptmatrix𝑢𝑣𝑖𝑗Γ2𝜋superscriptsubscript𝑥𝑗subscript𝑥𝑖2superscriptsubscript𝑦𝑗subscript𝑦𝑖2matrixsubscript𝑦𝑗subscript𝑦𝑖subscript𝑥𝑗subscript𝑥𝑖\begin{bmatrix}u\\ v\end{bmatrix}_{i,j}=\frac{\Gamma}{2\pi\left((x_{j}-x_{i})^{2}+(y_{j}-y_{i})^{% 2}\right)}\begin{bmatrix}-(y_{j}-y_{i})\\ x_{j}-x_{i}\end{bmatrix}\,,[ start_ARG start_ROW start_CELL italic_u end_CELL end_ROW start_ROW start_CELL italic_v end_CELL end_ROW end_ARG ] start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT = divide start_ARG roman_Γ end_ARG start_ARG 2 italic_π ( ( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG [ start_ARG start_ROW start_CELL - ( italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_CELL end_ROW start_ROW start_CELL italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] , (14)

where we assume that radial perturbations (ΔrΔyΔ𝑟Δ𝑦\Delta r\equiv\Delta yroman_Δ italic_r ≡ roman_Δ italic_y) are small such that ΓΓ\Gammaroman_Γ is approximately constant. Due to the symmetries of the interactions, the induced velocity on a given vortex by its left neighbor will be canceled out by the induced velocity by its right neighbor, and the line will not deform over time.

Next, consider a similar system of point vortices with circulations ΓΓ-\Gamma- roman_Γ, now arranged on a parabola of constant concavity given by y=αx2𝑦𝛼superscript𝑥2y=\alpha x^{2}italic_y = italic_α italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For the vortex initially at (xj,yj)=(0,0)subscript𝑥𝑗subscript𝑦𝑗00(x_{j},y_{j})=(0,0)( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = ( 0 , 0 ), the total induced velocity from its left and right neighbors at j1𝑗1j-1italic_j - 1 and j+1𝑗1j+1italic_j + 1 is

[uv]j±1,j=Γπ(1+(αΔx)2)[α0].subscriptmatrix𝑢𝑣plus-or-minus𝑗1𝑗Γ𝜋1superscript𝛼Δ𝑥2matrix𝛼0\begin{bmatrix}u\\ v\end{bmatrix}_{j\pm 1,j}=-\frac{\Gamma}{\pi\left(1+(\alpha\Delta x)^{2}\right% )}\begin{bmatrix}\alpha\\ 0\end{bmatrix}\,.[ start_ARG start_ROW start_CELL italic_u end_CELL end_ROW start_ROW start_CELL italic_v end_CELL end_ROW end_ARG ] start_POSTSUBSCRIPT italic_j ± 1 , italic_j end_POSTSUBSCRIPT = - divide start_ARG roman_Γ end_ARG start_ARG italic_π ( 1 + ( italic_α roman_Δ italic_x ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG [ start_ARG start_ROW start_CELL italic_α end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARG ] . (15)

A point-vortex distribution with positive concavity will thus induce a negative tangential velocity in a given vortex, and the magnitude of this induced motion scales with the concavity α𝛼\alphaitalic_α. Conversely, a distribution with negative concavity will induce a positive tangential velocity in a vortex on the curve.

This principle can be demonstrated for a system of point vortices initialized along a sine wave with no background flow and numerically integrated forward in time using Equation 14, shown in Figure 3. Vortices initially located on the wave crest move in the positive direction, while vortices initially located on the wave trough move in the negative direction. In both cases, vortices traveling in the same direction are pushed into closer proximity with each other as they move, rolling up into a vortex aggregate.

Refer to caption
Figure 3: 2D point-vortex simulation. Vortices with equal and negative (clockwise) circulation are initialized at the open circles on the grey dotted line. Their trajectories are shown in solid lines, and their final positions in the simulation are given by closed circles. Vortex locations and trajectories are colored by the concavity of the curve at the vortex’s initial position.

Applying this analysis to the sample model solutions for the wake radius shown in Figure 2, we can predict the evolution of tip vortices in the turbine wake as they are advected downstream. Assuming the initial distribution of the tip vortices follows the model solutions for r(x,t)𝑟𝑥𝑡r(x,t)italic_r ( italic_x , italic_t ), we would expect to see a similar aggregation behavior as that observed in Figure 3. Furthermore, since the nonlinear steepening of the wake-radius waves leads to higher concavities at the wave crests relative to the wave troughs, the tip vortices should tend to aggregate most strongly downstream of the wave crests and “surf” on these waves as they travel downstream.

A complementary mechanism for tip-vortex aggregation can be found by considering the effect of streamwise-velocity gradients. Consider again a line of point vortices with initial spacing ΔxΔ𝑥\Delta xroman_Δ italic_x, advected by a flow with a positive streamwise gradient ΔUΔxΔ𝑈Δ𝑥\frac{\Delta U}{\Delta x}divide start_ARG roman_Δ italic_U end_ARG start_ARG roman_Δ italic_x end_ARG. If we follow one point vortex at its advection velocity U𝑈Uitalic_U, after some time ΔtΔ𝑡\Delta troman_Δ italic_t, the distance between it and its neighbors will have increased to Δx+ΔUΔtΔ𝑥Δ𝑈Δ𝑡\Delta x+\Delta U\Delta troman_Δ italic_x + roman_Δ italic_U roman_Δ italic_t. By Equation 14, this increase in separation will result in weaker interactions between neighboring vortices. Conversely, in flows with a negative streamwise gradient, the separation between vortices will decrease as a function of time, thereby increasing the magnitude of vortex interactions by mutual induction.

Returning to the simulated traces of R𝑅Ritalic_R and U𝑈Uitalic_U in Figure 2, we observe that regions where Ux<0𝑈𝑥0\frac{\partial U}{\partial x}<0divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG < 0 coincide with peaks in the wake radius where the concavity of R𝑅Ritalic_R is most negative. The negative streamwise velocity gradients should therefore supplement the effect of negative concavity described previously and augment the tendency of tip-vortex elements to aggregate just downstream of peaks in the wake-radius waveform. By contrast, positive streamwise velocity gradients align with troughs in R𝑅Ritalic_R, which will discourage the roll-up of individual vortices in these regions. These analyses suggest that tip vortices will primarily aggregate in a single structure just downstream of peaks in the wake-radius waves. The strength of this behavior will depend on the amplitude and frequency of the unsteady wake forcing Ui(x=0,t)subscript𝑈𝑖𝑥0𝑡U_{i}(x=0,t)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ), which dictate both the magnitude of the streamwise velocity gradients in the wake and the sharpness of the peaks in the wake-radius waveform.

The proposed mechanisms for vortex aggregation in unsteady wakes are similar in principle to that of tip-vortex pairing, a phenomenon that is related to the breakdown of turbine wakes in steady inflow conditions (Okulov & Sørensen, 2007; Felli et al., 2011; Sarmast et al., 2014; Quaranta et al., 2015; Lignarolo et al., 2015). However, while the mutual-induction instabilities considered in canonical turbine wakes typically involve two or three vortex elements, the present analysis involves larger systems of vortices. The number of tip vortices shed into the wake over a single unsteady forcing period can be estimated as

Nv=2Nbλk,subscript𝑁𝑣2subscript𝑁𝑏𝜆𝑘N_{v}=2N_{b}\frac{\lambda}{k}\,,italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = 2 italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT divide start_ARG italic_λ end_ARG start_ARG italic_k end_ARG , (16)

where Nbsubscript𝑁𝑏N_{b}italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the number of turbine blades, λ𝜆\lambdaitalic_λ is the tip-speed ratio, and k𝑘kitalic_k is the reduced frequency. For the experiments that will be detailed in this work (cf. Table 1 in Section 3), 11<Nv<3311subscript𝑁𝑣3311<N_{v}<3311 < italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT < 33. The unsteady vortex-interaction mechanisms described above occur across length scales on the order of half the wavelength of the wake-radius perturbations, thus involving about Nv/2subscript𝑁𝑣2N_{v}/2italic_N start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT / 2 vortices per period. Therefore, while vortex-pairing instabilities may be present in periodically forced unsteady wakes, it is expected that the additional mechanisms described in this section will play a non-negligible role in the vortex dynamics of the wake, due to the number and distribution of the vortices present in the interactions.

In summary, the modeling approach outlined in this section provides an interpretable theoretical description of the dynamics of the wake radius, streamwise velocity, and tip-vortex aggregations in a periodically unsteady turbine wake. The framework offers insights into the dominant physics of the problem and can be used for qualitative predictions of trends in the dynamics, including that

1) Nonlinear traveling waves in the streamwise velocity U𝑈Uitalic_U are generated by the oscillatory boundary condition, undergo steepening, and are damped in amplitude as they advect downstream;

2) Traveling wakes in the wake radius R𝑅Ritalic_R are generated by the streamwise-velocity waves, grow with increasingly prominent crests and broad troughs, and saturate as velocity gradients dissipate;

3) Tip-vortex elements are encouraged by the unsteady wake dynamics to aggregate ahead of crests in the wake radius and advect downstream with these crests; and

4) Increasing the strength of the forcing in the boundary condition Ui(x=0,t)subscript𝑈𝑖𝑥0𝑡U_{i}(x=0,t)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ) increases the amplitudes of U𝑈Uitalic_U and R𝑅Ritalic_R in the wake, as well as the tendency toward vortex aggregation.

The modeling approach relies on several assumptions that limit its quantitative accuracy and restrict its applicability to the prediction of perturbations in the near wake. However, in the following sections we will demonstrate that it still captures many of the key dynamical features of unsteady turbine wakes.

3 Experimental methods

3.1 Experimental apparatus

Wake measurements downstream of a periodically surging rotor were conducted in an optically accessible towing-tank facility at Queen’s University. The apparatus is described in detail by El Makdah et al. (2019); a brief overview is given here, and a schematic is provided in Figure 4.

Refer to caption
Figure 4: Schematic of the experimental apparatus in the optical towing tank, including the turbine, turbine sensors, traverse, and the fields of view of the four high-speed cameras. A sample velocity waveform 𝒰(t)𝒰𝑡\mathcal{U}(t)caligraphic_U ( italic_t ) for the traverse is shown in an inset.

A 3D-printed turbine with a diameter of D=0.3𝐷0.3D=0.3italic_D = 0.3 m was used in the experiments. The turbine blades were designed with SD7003 airfoil profiles of constant chord and a spanwise twist that targeted a constant angle of attack of 10superscript1010^{\circ}10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT along the blade at a tip-speed ratio of λ=4𝜆4\lambda=4italic_λ = 4. The turbine had a maximum measured coefficient of power of Cp=0.29subscript𝐶𝑝0.29C_{p}=0.29italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0.29 at a tip-speed ratio of λ=3.89𝜆3.89\lambda=3.89italic_λ = 3.89. The turbine was mounted on a sting at the center of the towing-tank test section, and the turbine shaft was connected to a rotational encoder (Baumer ITD69H00), torque sensor (HBM T22), and frictional brake by means of a chain drive. The turbine system had a low rotational inertia relative to the hydrodynamic torques on the blades (El Makdah et al., 2021). The estimated blockage of the turbine based on swept area was 7.1%.

The test section of the towing tank was 15 m long and had a 1 m ×\times× 1 m cross-section. This was filled with water and enclosed by a ceiling that served to mitigate free-surface waves. The turbine was suspended from a traverse through a 50-mm wide opening in the ceiling. A rotary encoder on the traverse enabled its linear position to be recorded. The traverse was driven at a mean speed of U=1subscript𝑈1U_{\infty}=1italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for steady-flow measurements. For unsteady surge cases, periodic oscillations in the traverse velocity of up to ±0.4plus-or-minus0.4\pm 0.4± 0.4 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT were superimposed on this mean speed. The average magnitude of the error between the desired velocity profile and the measured traverse velocity was 0.023 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Each traverse run was initiated with a constant-acceleration ramp-up profile that transitioned smoothly into the test motion waveform, and a similar ramp-down profile was used to bring the traverse to rest after the experiment. A light sensor mounted 3.84 m from the starting position of the traverse served as a trigger to start flow measurements. This position was chosen such that the traverse would reach steady-state operation before data recording was initiated. The towing tank was allowed to settle for at least four minutes between individual runs, as it was determined that longer settling durations did not further reduce the average particle displacements significantly. Each phase-averaged ensemble for a given case was compiled from 20 separate runs of the corresponding traverse-motion profile.

Four high-speed cameras (Photron SA4) with a resolution of 1024×1024102410241024\times 10241024 × 1024 pixels were arranged in a line outside of the towing tank to capture a combined field of view of approximately 1.15 m ×\times× 0.39 m. The individual fields of view overlapped by approximately 30%. The cameras were triggered simultaneously by the aforementioned light sensor and recorded at a frame rate of 500 Hz. A high-speed laser (Photonics DM40-527) was used to illuminate the measurement domain along the rotational axis of the turbine. The flow was seeded by neutrally buoyant polyamide spherical particles (LaVision) with a diameter of 60 μm𝜇m\rm{\mu m}italic_μ roman_m. The turbine and sting were spray-painted black to minimize reflections.

3.2 Data analysis

2D particle-image velocimetry (PIV) was performed using the raw image data recorded by the cameras. An automated masking routine was written to mask the turbine, sting, and shadow cast by the turbine blades, based on the identified location of the nose of the turbine in the images. Only images recorded between the start trigger signal and the start of the traverse ramp-down motion were processed. Velocity vectors were computed using multi-pass cross-correlation with a final interrogation-window size of 32×32323232\times 3232 × 32 pixels with 50% overlap, using the open-source MATLAB package PIVlab (Thielicke & Stamhuis, 2014). A high-pass filter was applied to the images before correlation, and standard-deviation and median thresholds were applied to the vector fields after processing.

The individual vector fields from each camera were stitched together by interpolating onto a common spatial grid and averaging the velocities in the overlap regions. These composite lab-fixed velocity fields were then transferred into a reference frame moving at the mean speed of the turbine by identifying their locations in space and time within a single turbine surge period, interpolating onto a common spatial grid in the new reference frame, and stitching overlap** regions via linearly weighted blending. For the steady-flow reference cases, the representative period was set as T=1𝑇1T=1italic_T = 1 s (for U1subscript𝑈1U_{\infty}\leq 1italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ≤ 1) or T=0.5𝑇0.5T=0.5italic_T = 0.5 s (for U>1subscript𝑈1U_{\infty}>1italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT > 1). Vorticity fields were then calculated using Gaussian-filtered velocity fields to smooth out spurious results from numerical differentiation. Finally, the fields were phase-averaged across all 20 traverse runs, yielding time-resolved 2D velocity fields over a single period that covered at least 12 turbine diameters of the streamwise extent of the wake and over 1 turbine diameter in the radial direction. In some cases, over 18D18𝐷18D18 italic_D of the wake was measured.

To serve as quantities for comparison with the flow model described in Section 2, the wake radius and streamwise velocity were also computed from the PIV data. The wake radius was defined at a given location x𝑥xitalic_x as the radial location r𝑟ritalic_r at which the streamwise velocity reached 95% of the free-stream velocity, i.e. R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ) such that u(x,R,t)=0.95U𝑢𝑥𝑅𝑡0.95subscript𝑈u(x,R,t)=0.95U_{\infty}italic_u ( italic_x , italic_R , italic_t ) = 0.95 italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. These profiles of R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ) were smoothed using a moving-average filter with a width of 3 samples. The radially averaged streamwise velocity U(x,t)𝑈𝑥𝑡U(x,t)italic_U ( italic_x , italic_t ) was computed by averaging the local velocity u𝑢uitalic_u across a streamwise slice of the wake at x𝑥xitalic_x, from the center of the wake to the wake radius R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ).

3.3 Experimental cases

The parameter space investigated in these experiments is summarized in Table 1. Two loading conditions were applied to the turbine. In the first case, the frictional brake was manually tuned between runs to obtain a steady-flow tip-speed ratio of λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50. In the second case, the brake was released such that the only load on the turbine was due to shaft friction, yielding a steady-flow tip-speed ratio of λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07. Both of these scenarios result in higher tip-speed ratios than the power-maximizing loading condition. This was intended to mitigate the effects of flow separation and stall on the turbine blades, in accordance with the observations of Wei & Dabiri (2022).

At each loading condition, both steady-reference and unsteady surge-motion cases were investigated. Steady-flow reference cases were carried out for each loading condition at a constant inflow velocity of U=1subscript𝑈1U_{\infty}=1italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Additional constant-velocity cases at U=0.8subscript𝑈0.8U_{\infty}=0.8italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.8 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and U=1.2subscript𝑈1.2U_{\infty}=1.2italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1.2 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT were collected to represent a quasi-steady range of wake profiles for a surge-velocity amplitude of u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2, which will be compared against unsteady cases with the same surge-velocity amplitude to highlight differences between quasi-steady and unsteady wake dynamics.

Four unsteady cases for each loading condition spanned a range of surge-velocity amplitudes usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and reduced frequencies k𝑘kitalic_k: a baseline unsteady case with u=0.1superscript𝑢0.1u^{*}=0.1italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.1 and k=1.89𝑘1.89k=1.89italic_k = 1.89, a case with the same usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and higher k𝑘kitalic_k, a case with the same k𝑘kitalic_k as the baseline case and higher usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, and finally a case with as high of a value of usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT as could be achieved with the apparatus, which required k𝑘kitalic_k to be halved. For the case with u=0.4superscript𝑢0.4u^{*}=0.4italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.4, 20 additional runs were carried out in which the turbine motion waveform was offset by 1 m from its usual starting location, so that the combined set of 40 runs covered the entire wake over a single surge period. The surge-velocity amplitudes and reduced frequencies investigated in this study are relatively high and represent significant motions when applied to full-scale FOWTs, which in certain conditions experience motions with u>0.25superscript𝑢0.25u^{*}>0.25italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT > 0.25 (Wayman, 2006; Larsen & Hanson, 2007; de Vaal et al., 2014a). The range of reduced frequencies is somewhat higher than the range expected for FOWTs as a function of wave and platform frequencies, which Messmer et al. (2024) estimate to be k[0,1.5]𝑘01.5k\in[0,1.5]italic_k ∈ [ 0 , 1.5 ]. This was unavoidable due to limitations in the traverse velocity and length of the towing tank. Still, based on the findings of Messmer et al. (2024), we expect the dynamics observed at higher reduced frequencies to apply to somewhat lower reduced frequencies (k0.5greater-than-or-equivalent-to𝑘0.5k\gtrsim 0.5italic_k ≳ 0.5) as well.

λ0subscript𝜆0\lambda_{0}italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Cp,0subscript𝐶𝑝0C_{p,0}italic_C start_POSTSUBSCRIPT italic_p , 0 end_POSTSUBSCRIPT T𝑇Titalic_T [s] usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT k𝑘kitalic_k λ¯¯𝜆\overline{\lambda}over¯ start_ARG italic_λ end_ARG Cp¯¯subscript𝐶𝑝\overline{C_{p}}over¯ start_ARG italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG
4.50±0.193plus-or-minus4.500.1934.50\pm 0.1934.50 ± 0.193 0.144±0.045plus-or-minus0.1440.0450.144\pm 0.0450.144 ± 0.045 1 0.1 1.89 4.50±0.229plus-or-minus4.500.2294.50\pm 0.2294.50 ± 0.229 0.144±0.051plus-or-minus0.1440.0510.144\pm 0.0510.144 ± 0.051
0.8 0.1 2.36 4.51±0.226plus-or-minus4.510.2264.51\pm 0.2264.51 ± 0.226 0.139±0.047plus-or-minus0.1390.0470.139\pm 0.0470.139 ± 0.047
1 0.2 1.89 4.52±0.225plus-or-minus4.520.2254.52\pm 0.2254.52 ± 0.225 0.142±0.050plus-or-minus0.1420.0500.142\pm 0.0500.142 ± 0.050
2 0.4 0.94 4.25±0.495plus-or-minus4.250.4954.25\pm 0.4954.25 ± 0.495 0.123±0.043plus-or-minus0.1230.0430.123\pm 0.0430.123 ± 0.043
5.07±0.173plus-or-minus5.070.1735.07\pm 0.1735.07 ± 0.173 0.033±0.025plus-or-minus0.0330.0250.033\pm 0.0250.033 ± 0.025 1 0.1 1.89 5.05±0.196plus-or-minus5.050.1965.05\pm 0.1965.05 ± 0.196 0.034±0.024plus-or-minus0.0340.0240.034\pm 0.0240.034 ± 0.024
0.8 0.1 2.36 5.05±0.207plus-or-minus5.050.2075.05\pm 0.2075.05 ± 0.207 0.035±0.025plus-or-minus0.0350.0250.035\pm 0.0250.035 ± 0.025
1 0.2 1.89 5.06±0.180plus-or-minus5.060.1805.06\pm 0.1805.06 ± 0.180 0.034±0.022plus-or-minus0.0340.0220.034\pm 0.0220.034 ± 0.022
2 0.4 0.94 5.04±0.185plus-or-minus5.040.1855.04\pm 0.1855.04 ± 0.185 0.032±0.020plus-or-minus0.0320.0200.032\pm 0.0200.032 ± 0.020
Table 1: Operational parameters for the turbine used in these experiments. Two loading conditions were applied to the turbine, and the steady-flow tip-speed ratio λ0subscript𝜆0\lambda_{0}italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and coefficient of power Cp,0subscript𝐶𝑝0C_{p,0}italic_C start_POSTSUBSCRIPT italic_p , 0 end_POSTSUBSCRIPT are given as reference conditions in the leftmost two columns. For the unsteady cases, surge-motion parameters (including the period T𝑇Titalic_T in seconds, surge-velocity amplitude usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, and reduced frequency k𝑘kitalic_k) and the resulting time-averaged values of λ𝜆\lambdaitalic_λ and Cpsubscript𝐶𝑝C_{p}italic_C start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT are given on the right side of the table. For all cases, U=1subscript𝑈1U_{\infty}=1italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

The phase-averaged tip-speed ratios for all unsteady cases, referenced to the apparent inflow velocity in the rotor frame U𝒰(t)subscript𝑈𝒰𝑡U_{\infty}-\mathcal{U}(t)italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - caligraphic_U ( italic_t ), are shown in Figure 5. The fact that these signals are almost perfectly in phase with the surge-velocity waveform 𝒰(t)𝒰𝑡\mathcal{U}(t)caligraphic_U ( italic_t ) confirms that the inertia of the turbine is very low compared to the surge dynamics. We also note that, for most of the unsteady cases, the time-averaged tip-speed ratio and coefficient of power (shown in the rightmost columns of Table 1) remained relatively constant with changes in usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and k𝑘kitalic_k. For one case (u=0.4superscript𝑢0.4u^{*}=0.4italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.4 and λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50), the time-averaged tip-speed ratio and power dropped due to the onset of stall in the turbine at low instantaneous inflow velocities, as evidenced by the drop in rotation rate observed around t/T=0.8𝑡𝑇0.8t/T=0.8italic_t / italic_T = 0.8 for this case in Figure 5(a).

Refer to caption
(a)
Refer to caption
(b)
Figure 5: Phase-averaged tip-speed ratio data, defined by the inflow velocity in the rotor frame, as a function of time for the two loading conditions covered in this study (a, b). The steady reference values are shown as black dashed lines.

4 Results

In this section, velocity and vorticity fields from the towing-tank experiments are shown to highlight the dominant features of the unsteady turbine wake. These dynamics are compared with the qualitative predictions of the modeling framework from Section 2.1 to connect the previously discussed physical mechanisms with the observations. Finally, connections between the unsteady dynamics in the wake and enhancements in wake recovery observed in the unsteady cases are explored.

4.1 Velocity and vorticity fields

First, to demonstrate key differences between the wakes from the steady-flow and unsteady cases, streamwise-velocity and out-of-plane vorticity fields are shown in Figures 6 and 7 for sample cases at the higher reference tip-speed ratio, λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07. The top plot in each figure shows a snapshot from the steady-flow wake, while the bottom four plots show four instantaneous snapshots from the unsteady wake (u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89) at evenly spaced time steps. The steady wake shows typical wake features, such as gradual spreading in the velocity-deficit region, a monotonic recovery in the streamwise velocity with increasing downstream distance, and tip-vortex shedding in the near wake (x/D3less-than-or-similar-to𝑥𝐷3x/D\lesssim 3italic_x / italic_D ≲ 3) that breaks down in the intermediate wake. It is important to note that the azimuthal positions of the turbine blades were not synchronized across runs, so the fields shown in these and following figures are not phase-locked with respect to the turbine rotation. Despite this limitation, the signatures of tip vortices can still be identified in the vorticity fields. In the steady case shown in Figure 7, tip-vortex pairing instabilities are readily discernible, as the vortex elements collect into groups of three that then break down in the intermediate wake.

The unsteady wake, by contrast, exhibits strong departures from steady-flow wake behaviors. In Figure 6, a pulsatile streamwise velocity is visible in the near wake around x/D1𝑥𝐷1x/D\approx 1italic_x / italic_D ≈ 1, as a result of the time-varying thrust and power of the turbine. This oscillatory inflow condition propagates downstream as a traveling wave and creates periodic peaks in the velocity-deficit region where the wake radius extends out past the steady-flow wake boundary. In the corresponding vorticity fields in Figure 7, the tip-vortex elements roll up into a larger aggregate structure at around x/D2𝑥𝐷2x/D\approx 2italic_x / italic_D ≈ 2, which is then advected downstream as a coherent vortex packet. While the signatures of tip-vortex pairing are still visible in these unsteady snapshots, the smaller aggregates that form on account of the mutual-inductance instability are subsumed into the larger aggregate structure. Additionally, the large vortex aggregate is shed with a periodicity matching that of the turbine surge waveform, suggesting that its dynamics are governed more strongly by the periodic wake forcing than by conventional vortex-pairing mechanisms. All of these observations align well with the theoretical analyses that were presented in Section 2, as well as the numerical simulations of Kleine et al. (2022).

Refer to caption
Figure 6: Streamwise-velocity fields for a steady-flow case (top) and four time-steps of an unsteady case with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89. For both cases, λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07.
Refer to caption
Figure 7: Out-of-plane vorticity fields for a steady-flow case (top) and four time-steps of an unsteady case with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89. For both cases, λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07.

To demonstrate the effects of varying the unsteady surge-motion parameters usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and k𝑘kitalic_k, instantaneous snapshots at a fixed time instance t/T=0𝑡𝑇0t/T=0italic_t / italic_T = 0 are shown for all four unsteady cases at the lower tip-speed ratio (λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50) for the streamwise velocity (Figure 8) and out-of-plane vorticity (Figure 9). The dynamics observed in the example case discussed above are also visible in these instances. The degree of unsteadiness in the wake appears to increase with increasing surge-velocity amplitude, while the wavelength of the traveling waves in the wake scales inversely with the reduced frequency k𝑘kitalic_k. In all cases, the traveling-wave dynamics and vortex aggregates appear to persist well into the far wake (x/D10greater-than-or-equivalent-to𝑥𝐷10x/D\gtrsim 10italic_x / italic_D ≳ 10). The vortex aggregates that survive into the far wake are those that originally rolled up at the crests in the wake-radius wave, whereas no corresponding structures from the wake-radius wave troughs are visible past x/D8𝑥𝐷8x/D\approx 8italic_x / italic_D ≈ 8. Again, these observations are well in accordance with the theoretical conjectures advanced in Section 2.

Refer to caption
Figure 8: Streamwise-velocity fields for four unsteady cases at t/T=0𝑡𝑇0t/T=0italic_t / italic_T = 0, with T𝑇Titalic_T and usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT increasing from top to bottom. All cases have λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50.
Refer to caption
Figure 9: Vorticity fields for four unsteady cases at t/T=0𝑡𝑇0t/T=0italic_t / italic_T = 0, with T𝑇Titalic_T and usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT increasing from top to bottom. All cases have λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07.

4.2 Wake features and dynamics

4.2.1 Trends in the wake radius and radially averaged streamwise velocity

To quantify the dynamics observed in the velocity and vorticity fields, we use the definitions of the wake radius and radially averaged streamwise velocity given in Section 3.2 to produce experimental analogues to the model parameters R𝑅Ritalic_R and U𝑈Uitalic_U from Equations 11a and 11b. Examples of the calculated wake radius R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ) are shown as grey lines on two vorticity snapshots in Figure 10. This visualization highlights the undulatory behavior of the wake in the unsteady cases.

Refer to caption
Figure 10: Calculated wake radii (gray line), superimposed on vorticity fields for two cases with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89. The top and bottom plots show reference loading conditions of λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 and λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07, respectively. Both cases are shown at t/T=0.75𝑡𝑇0.75t/T=0.75italic_t / italic_T = 0.75.

The calculated wake radius and radially averaged streamwise velocity are then shown for all of the steady and unsteady cases at λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 for a single time instance t/T=0.75𝑡𝑇0.75t/T=0.75italic_t / italic_T = 0.75 in Figure 11. The wake-radius data in Figure 11(a) show evidence of nonlinear traveling waves, with wave steepening on the downwind sides of the waves in the far wake for the higher-amplitude cases. Corresponding perturbations in the streamwise velocity are visible in Figure 11(b), and the magnitude of these perturbations decreases with increasing streamwise distance into the wake. These data confirm the salience of the dynamics captured by the modeling framework, but also demonstrate the limitations of the theoretical analysis. Both R𝑅Ritalic_R and U𝑈Uitalic_U demonstrate strong evidence of nonlinear traveling waves, as predicted by the model. However, the model assumes a constant time-averaged base-flow velocity U¯¯𝑈\overline{U}over¯ start_ARG italic_U end_ARG that does not change as a function of downstream distance. The data in Figure 11(b), by contrast, show almost immediate signs of wake recovery at relatively short downstream distances. This therefore limits the scope of the model to perturbations about time-averaged quantities. These considerations should be kept in mind as we now turn to compare the model with the data.

Refer to caption
(a)
Refer to caption
(b)
Figure 11: Wake radius R𝑅Ritalic_R (a) and radially averaged streamwise velocity U𝑈Uitalic_U (b) as a function of streamwise distance at t/T=0.75𝑡𝑇0.75t/T=0.75italic_t / italic_T = 0.75. All cases with λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 are shown.

4.2.2 Comparisons between model solutions and measured data

While the modeling framework has limitations that may preclude its use for quantitatively accurate predictions, a comparison of its outputs with experimental data can still demonstrate that it is parameterizing the dominant physics of the unsteady-wake problem. To this end, we use the PIV data to extract representative inflow boundary conditions and a scaling for the turbulent fluctuations, and integrate the model with these parameters for direct comparison with the data.

To define the upstream boundary condition Ui(x=0,t)subscript𝑈𝑖𝑥0𝑡U_{i}(x=0,t)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ) for the wake, the time average U¯¯𝑈\overline{U}over¯ start_ARG italic_U end_ARG and amplitude U^^𝑈\hat{U}over^ start_ARG italic_U end_ARG of the radially averaged streamwise velocity were extracted from each unsteady dataset at x/D=1𝑥𝐷1x/D=1italic_x / italic_D = 1 and are shown in Figure 12. Notably, the amplitude of the streamwise velocity in the near wake appears to scale linearly with the surge-velocity amplitude in Figure 12(b). These quantities were applied directly to the boundary condition in Equation 12b.

Refer to caption
(a)
Refer to caption
(b)
Figure 12: Time-averaged streamwise velocity (a) and amplitude of the streamwise velocity (b), both spatially averaged across the wake at x/D=1𝑥𝐷1x/D=1italic_x / italic_D = 1. These data define the initial conditions for Ui(x=0,t)subscript𝑈𝑖𝑥0𝑡U_{i}(x=0,t)italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = 0 , italic_t ) in the analytical model, given in Equation 12b.

To estimate the coupling between streamwise-velocity fluctuations and streamwise-velocity gradients, as modeled by Equation 8, the variances of the individual streamwise velocities u(x,r,t)superscript𝑢𝑥𝑟𝑡u^{\prime}(x,r,t)italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x , italic_r , italic_t ) across all 20 ensembles were computed. These values were then averaged across the wake at every streamwise location to estimate the streamwise-velocity autocovariance UU~~superscript𝑈superscript𝑈\widetilde{U^{\prime}U^{\prime}}over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG at each location. Dividing these quantities by the magnitude of the velocity gradient at each location gave estimates for the proportionality parameter νu(x)subscript𝜈𝑢𝑥\nu_{u}(x)italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ). These empirical values are shown in Figure 13, and since they were observed to scale approximately linearly with streamwise distance in the intermediate wake (3x/D9less-than-or-similar-to3𝑥𝐷less-than-or-similar-to93\lesssim x/D\lesssim 93 ≲ italic_x / italic_D ≲ 9), a linear fit (cf. Equation 9) over all of the steady and unsteady estimates was employed as the model for νu(x)subscript𝜈𝑢𝑥\nu_{u}(x)italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ) in Equation 11b. The applicability of the linear fit breaks down in the far wake (x/D9greater-than-or-equivalent-to𝑥𝐷9x/D\gtrsim 9italic_x / italic_D ≳ 9), as the streamwise velocity gradients decrease in magnitude and the estimates for νusubscript𝜈𝑢\nu_{u}italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT thus become more susceptible to noise. The overarching theoretical framework is not expected to apply in this region anyway since it neglects wake-recovery effects due to turbulent momentum entrainment, so the linear approximation was deemed acceptable for the purposes of this study. With this model for νusubscript𝜈𝑢\nu_{u}italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT in place, Equations 11a and 11b could then be integrated numerically, as described in Section 2.1.

Refer to caption
Figure 13: Empirical estimates of the proportionality parameter νusubscript𝜈𝑢\nu_{u}italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT, derived from Equation 8. The fit to all of the data in the figure, shown as a red dashed line, was used as a rudimentary model for νu(x)subscript𝜈𝑢𝑥\nu_{u}(x)italic_ν start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_x ) in Equation 11b. The fit best matches the data for 3x/D9less-than-or-similar-to3𝑥𝐷less-than-or-similar-to93\lesssim x/D\lesssim 93 ≲ italic_x / italic_D ≲ 9, whereas the data become noisier further downstream.

For a more direct comparison with the model results, the envelope of the wake-radius perturbation was computed from the data. This represents the minimum and maximum values observed in the quantity R(x,t)R¯(x)𝑅𝑥𝑡¯𝑅𝑥R(x,t)-\overline{R}(x)italic_R ( italic_x , italic_t ) - over¯ start_ARG italic_R end_ARG ( italic_x ) over t/T[0,1)𝑡𝑇01t/T\in[0,1)italic_t / italic_T ∈ [ 0 , 1 ). The envelope representation removes the effects of wake recovery in terms of streamwise changes in both the time-averaged radius R¯(x)¯𝑅𝑥\overline{R}(x)over¯ start_ARG italic_R end_ARG ( italic_x ) and the wave advection velocity U¯(x)¯𝑈𝑥\overline{U}(x)over¯ start_ARG italic_U end_ARG ( italic_x ), neither of which are captured in the model. Plots of these envelopes along with the model solutions at two representative time instances are shown in Figure 14 for all four unsteady cases with λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50. Given the simplifying assumptions of the modeling framework, the agreement between the model solutions and experimental results is remarkable. For the lower-amplitude cases (u0.2superscript𝑢0.2u^{*}\leq 0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≤ 0.2), the model shows a very similar rate of growth in the wake-radius amplitude as in the envelope of the data. The saturated amplitude of the wake-radius waves downstream of x/D5𝑥𝐷5x/D\approx 5italic_x / italic_D ≈ 5 also corresponds well between the model and data. The model overestimates the final amplitude of the wake radius in the case with the highest forcing amplitude (u=0.4superscript𝑢0.4u^{*}=0.4italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.4), suggesting that the physics neglected in the modeling approach become more significant as the unsteady forcing amplitude increases.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 14: Excursions of the wake radius from the local temporal mean, R(x,t)R¯(x)𝑅𝑥𝑡¯𝑅𝑥R(x,t)-\overline{R}(x)italic_R ( italic_x , italic_t ) - over¯ start_ARG italic_R end_ARG ( italic_x ), comparing the envelope of the PIV data (grey region) with numerical solutions to the model (lines). The data envelope spans the minimum and maximum wake radii in the data over t/T[0,1)𝑡𝑇01t/T\in[0,1)italic_t / italic_T ∈ [ 0 , 1 ) at each streamwise location. Good agreement between the model solutions and measured data is observed for u0.2superscript𝑢0.2u^{*}\leq 0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≤ 0.2.

To further investigate the correspondence between the model results and experimental findings, we compute the amplitude of the wake-radius perturbations for both the model and the data for all cases. These are shown for both loading conditions in Figure 15 for three streamwise locations: x/D=2𝑥𝐷2x/D=2italic_x / italic_D = 2, 5, and 10, which represent the near-, intermediate-, and far-wake regions, respectively. The effective wake-radius amplitudes from the quasi-steady measurements at U=0.8subscript𝑈0.8U_{\infty}=0.8italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.8 and 1.2 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT are shown as open colored markers at u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2. It is apparent that the quasi-steady approximation of the wake-radius amplitude does not capture the unsteady dynamics of the wake, as there are no traveling waves present in the quasi-steady approximation. The model results, shown as darker-colored open markers (with dashed lines representing linear interpolations between solutions), follow the trends in the data (closed markers) more closely. At the lower surge-velocity amplitudes (u0.2superscript𝑢0.2u^{*}\leq 0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≤ 0.2) and in the near- and intermediate-wake regions, the model results align well with the data. The model solutions diverge from the data at the highest surge-velocity amplitude and in the far wake, as expected given the breakdown of its assumptions in these regimes. Still, even in these situations the model correctly predicts that the wake-radius amplitude will increase with increasing usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and with increasing x/D𝑥𝐷x/Ditalic_x / italic_D, suggesting that the underlying physics parameterized by the model are still relevant.

Refer to caption
(a)
Refer to caption
(b)
Figure 15: Comparison between theoretical and experimental results for the wake-radius amplitude R^/D^𝑅𝐷\hat{R}/Dover^ start_ARG italic_R end_ARG / italic_D, plotted as a function of surge-velocity amplitude usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Experimental data are shown as colored markers, model solutions are shown as darker-colored open markers, and results obtained from quasi-steady measurements are given as colored open markers. Linear interpolations between the model solutions are shown as dashed lines. The colors and shapes of the markers correspond to their streamwise locations in the wake. Relatively good agreement between the model and data is observed for x/D5𝑥𝐷5x/D\leq 5italic_x / italic_D ≤ 5 and u0.2superscript𝑢0.2u^{*}\leq 0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≤ 0.2.

Results from a similar analysis of the amplitude of the streamwise velocity are shown in Figure 22 in Appendix B. The model solutions strongly underpredicted these amplitudes but still captured the trends in the data, especially compared to the quasi-steady approximations. While the results suggest that the model in its current form is not a quantitatively accurate predictive tool for all wake properties, it is still able to represent key dynamical phenomena of these unsteady wake scenarios.

4.3 Wake-recovery enhancements via unsteady flow mechanisms

The preceding sections have demonstrated that the theoretical considerations in Section 2 are able to qualitatively describe the growth and propagation of traveling-wave perturbations in unsteady turbine wakes, but are not able to directly address the streamwise evolution of time-averaged quantities. Using the experimental results, however, it is possible to observe the effects of unsteady wake forcing on these time-averaged quantities and consider their underlying physics.

The radial- and time-averaged streamwise velocity U¯(x)¯𝑈𝑥\overline{U}(x)over¯ start_ARG italic_U end_ARG ( italic_x ) is shown for all cases in Figure 16. The steady-flow cases, shown as black dashed lines, lie below the unsteady cases from x/D2𝑥𝐷2x/D\approx 2italic_x / italic_D ≈ 2 and deep into the far wake (x/D>14𝑥𝐷14x/D>14italic_x / italic_D > 14). The gap between the steady and unsteady wake profiles is most apparent in the intermediate wake, around 2x/D6less-than-or-similar-to2𝑥𝐷less-than-or-similar-to62\lesssim x/D\lesssim 62 ≲ italic_x / italic_D ≲ 6, which is precisely where traveling-wave growth and tip-vortex aggregation occur. The extent to which the wake recovery is enhanced also increases with increasing surge-velocity amplitude as well as reduced frequency.

Refer to caption
(a)
Refer to caption
(b)
Figure 16: Radial- and time-averaged streamwise velocity U¯¯𝑈\overline{U}over¯ start_ARG italic_U end_ARG in the wake of the turbine for λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 (a) and λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07 (b). The steady reference cases are shown as black dashed lines.

The velocity profiles in Figure 16 can be used to further quantify the degree of wake enhancement observed in these experiments. First, a reduction in the streamwise extent of the wake can be calculated by defining the time-averaged streamwise distance x=L¯𝑥¯𝐿x=\overline{L}italic_x = over¯ start_ARG italic_L end_ARG required for the wake velocity to recover to U(x)=0.85U𝑈𝑥0.85subscript𝑈U(x)=0.85U_{\infty}italic_U ( italic_x ) = 0.85 italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. The percent reduction in this wake-length variable is shown in Figure 17(a). All unsteady cases showed reductions in L¯¯𝐿\overline{L}over¯ start_ARG italic_L end_ARG by at least 19%, with improvements in excess of 45% achieved at the highest amplitudes. Second, a similar quantity can be computed for the increase in wake velocity at a fixed streamwise distance of x/D=10𝑥𝐷10x/D=10italic_x / italic_D = 10, a reasonable inter-turbine spacing for a moderate-sized wind farm (Stevens & Meneveau, 2017). As shown in Figure 17(b), the wake velocities were increased at this particular location in all unsteady cases by 2 to 5% relative to the steady case. Taking the cube of these improvements gives a measure of the increase in the available power in the flow at the same streamwise location, shown in Figure 17(c). In these experiments, enhancements in the available power in the flow at x/D=10𝑥𝐷10x/D=10italic_x / italic_D = 10 ranged from 6.7 to 15.7%, which could represent a large potential increase in power generation for a downstream turbine placed at this location in an array. These findings are in line with the recent experimental results of Bossuyt et al. (2023) and Messmer et al. (2024) for FOWTs, as well as those of Frederik et al. (2020b) and van der Hoek et al. (2022) for fixed turbines under dynamic induction control.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 17: Measures of wake-recovery enhancement, reported as percentages. (a) shows the reduction in the streamwise distance L¯¯𝐿\overline{L}over¯ start_ARG italic_L end_ARG required for the wake to recover to U=0.85U𝑈0.85subscript𝑈U=0.85U_{\infty}italic_U = 0.85 italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, relative to the distance in the steady case L0subscript𝐿0L_{0}italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (b) shows the enhancement in the streamwise wake velocity U¯¯𝑈\overline{U}over¯ start_ARG italic_U end_ARG at x/D=10𝑥𝐷10x/D=10italic_x / italic_D = 10, again relative to the corresponding steady-flow quantity U0(x/D=10)subscript𝑈0𝑥𝐷10U_{0}(x/D=10)italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x / italic_D = 10 ). (c) shows the enhancement in the power available in the wake at x/D=10𝑥𝐷10x/D=10italic_x / italic_D = 10 relative to the steady case, (𝒫a¯𝒫a,0)/𝒫a,0¯subscript𝒫𝑎subscript𝒫𝑎0subscript𝒫𝑎0\left(\overline{\mathcal{P}_{a}}-\mathcal{P}_{a,0}\right)/\mathcal{P}_{a,0}( over¯ start_ARG caligraphic_P start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG - caligraphic_P start_POSTSUBSCRIPT italic_a , 0 end_POSTSUBSCRIPT ) / caligraphic_P start_POSTSUBSCRIPT italic_a , 0 end_POSTSUBSCRIPT.

The various measures of wake-recovery enhancement in Figure 17 are plotted against the surge-acceleration amplitude ku𝑘superscript𝑢ku^{*}italic_k italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. This choice leads to a better collapse in the data than plotting against usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT or k𝑘kitalic_k alone. The data from the lower tip-speed ratio cases seem to scale in direct proportionality with the surge acceleration, while the data from the higher tip-speed ratio cases appear to saturate toward higher acceleration amplitudes. This scaling may be a result of the time-varying thrust force from the surging turbine that generates the unsteady wake dynamics. The thrust force exerted by the turbine on the flow can be approximated as the sum of quasi-steady and added-mass forces,

FTπ8ρD2CT(U𝒰(t))2+12ρVdCa𝒰t,similar-tosubscript𝐹𝑇𝜋8𝜌superscript𝐷2subscript𝐶𝑇superscriptsubscript𝑈𝒰𝑡212𝜌subscript𝑉𝑑subscript𝐶𝑎𝒰𝑡F_{T}\sim\frac{\pi}{8}\rho D^{2}C_{T}(U_{\infty}-\mathcal{U}(t))^{2}+\frac{1}{% 2}\rho V_{d}C_{a}\frac{\partial\mathcal{U}}{\partial t}\,,italic_F start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∼ divide start_ARG italic_π end_ARG start_ARG 8 end_ARG italic_ρ italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ( italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - caligraphic_U ( italic_t ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ρ italic_V start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT divide start_ARG ∂ caligraphic_U end_ARG start_ARG ∂ italic_t end_ARG , (17)

where CTsubscript𝐶𝑇C_{T}italic_C start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the thrust coefficient, Vdsubscript𝑉𝑑V_{d}italic_V start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the effective volume of fluid displaced by the rotor disc, and Casubscript𝐶𝑎C_{a}italic_C start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is an added-mass coefficient. Assuming that CTsubscript𝐶𝑇C_{T}italic_C start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and Casubscript𝐶𝑎C_{a}italic_C start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT do not vary appreciably with time, the nondimensional force amplitude applied by the turbine on the flow in the wake scales as

F^F0u+kusimilar-to^𝐹subscript𝐹0superscript𝑢𝑘superscript𝑢\frac{\hat{F}}{F_{0}}\sim u^{*}+ku^{*}divide start_ARG over^ start_ARG italic_F end_ARG end_ARG start_ARG italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ∼ italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT + italic_k italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (18)

to first order. The fact that the data in Figure 17 appear to scale with ku𝑘superscript𝑢ku^{*}italic_k italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT may suggest that the flow accelerations associated with the effective added mass of the surging turbine represent an additional unsteady forcing on the wake that aids in the recovery of kinetic energy. In effect, the surge motions of the turbine may be pum** the flow in the wake, thereby stimulating faster wake recovery.

Complementary wake-recovery mechanisms may be postulated to explain the observed enhancements. For instance, the vortex aggregates visible in the intermediate and far wake in Figures 7 and 9 likely enhance mixing and the entrainment of free-stream momentum into the wake. To examine this hypothesis, the centroids of these structures were identified in the vorticity fields, the circulations were computed on a circular path of radius D/4𝐷4D/4italic_D / 4 around the centroids, and the circulations were bin-averaged by streamwise distance over all time instances for each case. These circulations are shown in Figure 18 for both loading conditions, and the data show that increasing the surge-velocity amplitude and surge frequency increase the circulations of the vortex aggregates throughout the wake. This aligns well with the idea that the surge motions of the turbine are applying a pulsatile forcing on the wake, since the roll-up of these structures can be described by the 1D modeling framework, and those dynamics are driven by the forcing amplitude of the wake velocity at the upstream boundary.

Refer to caption
(a)
Refer to caption
(b)
Figure 18: Circulations of vortex-aggregate structures in the wake, averaged over a surge period in streamwise bins of width D/4𝐷4D/4italic_D / 4. The steady reference case is shown in black.

The geometry of the wake in the unsteady cases may also play a role in wake-recovery enhancement. The wavy wake shape seen in Figures 10 and 11(a) implies that the unsteady wakes have a larger surface area than their steady-flow counterparts, which may give rise to additional turbulent momentum transport due to the larger interface. Similarly, the radial deformations in the wake themselves may encourage momentum transport in the radial direction. Though the 1D modeling framework ignores the radial velocity component in the wake, in reality an increase in the wake radius corresponds to a net momentum flux in the positive radial direction that scales with URt𝑈𝑅𝑡U\frac{\partial R}{\partial t}italic_U divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG. This local bulk flow in the radial direction may encourage sweeps of high-momentum fluid into the wake or ejections of low-momentum fluid out of the wake, depending on the direction of the local radial flow.

In practice, the thrust-amplitude, vortex-aggregate, and wake-geometry contributions to wake recovery are most likely all coupled, as the thrust-force variations drive both traveling waves in the wake and tip-vortex roll-up. A more detailed analysis of momentum transport across the wake is not possible here due to insufficient ensemble sizes for resolving turbulent fluxes. Future work could investigate these mechanisms more directly and parameterize their effects on wake recovery.

4.4 Discussion and implications

The experimental results described in the previous sections highlight several strengths of the modeling approach, as the theoretical hypotheses listed at the end of Section 2 are confirmed in the data. Traveling waves are observed in the streamwise velocity and undergo dam** as they propagate downstream, though the rate of this decay is not quantitatively captured by the model. The model also predicts the traveling-wave dynamics evident in the wake-radius measurements, including wave steepening and amplitude saturation. The tip-vortex aggregation explored in Section 2.2 is observed in the measured vorticity fields as well. Finally, the strengths of these unsteady dynamical features all increase with increasing forcing amplitude, represented by higher-amplitude oscillations at the inlet to the wake that scale with increasing surge-velocity amplitudes of the turbine.

The experimental results also highlight the assumptions and limitations of the theoretical analysis. Since the modeling approach in its current form neglects pressure gradients and wake-recovery effects, it is unable to capture changes in time-averaged quantities as a function of downstream distance. Mean-flow effects, such as pressure recovery and turbulent entrainment of momentum from the free stream into the wake, will alter the propagation velocity of traveling waves in the wake, the evolution of streamwise-velocity gradients, and ultimately the growth and saturation of the wake radius. In a strict sense, then, the model can only be used to predict perturbations about time-averaged quantities, and an extension would be required to capture the effects of wake recovery on the streamwise development of the time-averaged quantities themselves. These constraints are especially evident in the far-wake region where turbulent momentum entrainment contributes more significantly to wake recovery, as well as at high forcing amplitudes (e.g. u=0.4superscript𝑢0.4u^{*}=0.4italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.4) where additional nonlinearities not included in the model become more significant. Despite these limiting assumptions, the dynamics described by the model are observed to correlate with enhancements in wake recovery, suggesting that these unsteady mechanisms strongly influence momentum transport into the wake.

Additional limitations of the present approach suggest possibilities for future theoretical work. First, the relationship between the turbine dynamics (including rotor inertia, induction, thrust, and power) and the inlet boundary condition of the model is not explicitly defined. This was done intentionally to preserve the general applicability of the framework, but an analytical connection could be derived by extending the framework of Wei & Dabiri (2023) to include near-wake quantities, as done by Heck et al. (2023) for yaw-misaligned turbines. Such an approach could allow other effects to be included as well, including yaw and tilt misalignment or aeroelastic turbine-blade deformations (cf. Rodriguez & Jaworski, 2020). Also, a more physically motivated model for the streamwise-velocity autocovariance UU~~superscript𝑈superscript𝑈\widetilde{U^{\prime}U^{\prime}}over~ start_ARG italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG than the empirical linear fit used in this work could be implemented. Furthermore, the model could account for the pressure oscillations that occur downstream of turbines with time-varying thrust-force profiles. Finally, the vortex-dynamics analysis in Section 2.2 only considers one-way coupling from the wake properties to the induced motions of the vortex elements. The effects of the vortices on the flow field in the wake have not been considered, primarily because the treatment of the wake dynamics has been limited to 1D while point vortices generate velocity components in two dimensions.

The experimental approach taken in this study also involves necessary abstractions that limit the direct applicability of these findings to utility-scale wind farms. The Reynolds number regime in the experiments is two orders of magnitude lower than that of full-scale wind turbines, which affects both the turbine aerodynamics (Miller et al., 2019) and the wake properties (Piqué et al., 2022). However, since the theoretical framework neglects viscous stresses in the wake, the mechanisms identified in the study should still be representative of those that would dominate the flow physics of full-scale rotors. The experiments also only considered a narrow range of tip-speed ratios and did not measure thrust forces on the turbine, which limit the conclusions that can be drawn from these data regarding turbine aerodynamics and power generation.

The most significant limitation of this study regarding its applicability to real-world flow scenarios is the lack of inflow turbulence in the towing tank. Since the flow in the tank was nominally quiescent before each run of the traverse, background turbulence in the facility was minimal and wake recovery in the baseline case was slow relative to turbine wakes in flows with higher turbulence intensities (cf. Duckworth & Barthelmie, 2008). This could suggest that the relatively large enhancements in wake recovery due to the surge motions of the turbine may not be as significant in atmospheric boundary-layer conditions, where the additional mixing provided by the unsteady wake dynamics may be overshadowed by the mixing from the highly turbulent background flow. Experimental results in agreement with this hypothesis have been reported by Kadum et al. (2019).

With these considerations in mind, the theoretical framework and experimental findings presented here still have important implications for the dynamics and control of utility-scale wind turbines and wind farms. For onshore and fixed-bottom offshore wind turbines, the unsteady wake dynamics explored in this study describe the effects of dynamic induction control on the wake. Changing the collective pitch angle of the turbine blades or the tip-speed ratio of the turbine in a periodic manner will produce a periodic variation in the turbine thrust force, thus exciting oscillations in streamwise velocity that lead to traveling-wave propagation, tip-vortex aggregation, and accelerated wake recovery. In fact, similar dynamics have been reported in previous studies on dynamic induction control. For instance, the vortical structures observed in the large-eddy simulations of Munters & Meyers (2018) may be analogous to the vortex aggregates found in the present study. Similarly, the 3D flow measurements of van der Hoek et al. (2022) in the wake of a turbine with periodic variations in collective blade pitch also show propagating oscillations in the streamwise velocity in the near wake, as well as undulatory deformations in the wake radius and changes in the tip-vortex dynamics. It is noteworthy that both of these studies involve background flows with relatively high turbulence intensities and still show qualitatively similar dynamics to the present findings. This supports the broader applicability of the present work to utility-scale wind farms in more realistic atmospheric conditions.

The results of this study also highlight the potential benefits of unsteady turbine motion for wake-recovery enhancement. The typical methods for dynamic induction control on stationary turbines are limited by the thrust variations that can be achieved by changing the blade pitch angle and rotation rate. Surging the turbine back and forth can lead to much larger thrust-force variations, since the thrust force scales with the square of the rotor-frame velocity U𝒰(t)subscript𝑈𝒰𝑡U_{\infty}-\mathcal{U}(t)italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT - caligraphic_U ( italic_t ) (Johlas et al., 2021; Wei & Dabiri, 2023). The unsteady platform motions of floating offshore wind turbines could thus be leveraged to achieve faster wake recovery. This has recently been corroborated in the findings of Messmer et al. (2024). These same unsteady motions can also be exploited to enhance the power generation of the surging turbine itself (Wei & Dabiri, 2022, 2023). Therefore, the unsteady dynamics associated with the natural rocking motions of floating offshore wind turbines can be harnessed to achieve higher power densities in floating offshore wind farms relative to their fixed-bottom counterparts. The enhancements in available power shown in Figure 17(c), taken together with the power-generation improvements of up to 6.4% above the stationary-turbine case reported by Wei & Dabiri (2022), suggest that these increases in power density could be on the order of ten percent over conventional wind farms. It is also important to note that the power-density benefits in the current and aforementioned experimental studies were achieved without any active turbine control or array-scale optimization. Thus, even greater enhancements may be possible with the addition of physics-informed control strategies, though attention must be paid to the dynamic coupling between the controller, turbine thrust force, and floating-platform hydrodynamics (van den Berg et al., 2023).

5 Conclusions

This study demonstrates the effects of unsteady streamwise motions of a turbine on wake dynamics and recovery. A theoretical model of the wake radius and streamwise velocity in a wake with an oscillatory upstream boundary condition describes mechanisms for the growth and propagation of traveling waves in the wake, which encourage the roll-up of tip vortices into periodically shed vortical aggregates that advect into the far wake. These dynamics scale with the amplitude of the turbine motions. Experimental measurements in a towing-tank facility confirm the qualitative trends parameterized by the theoretical approach, and additionally show enhancements in wake recovery and the corresponding power available in the flow downstream of the turbine. These results support the findings of prior investigations involving dynamic induction control strategies for improving wind-farm power density. Taken together with power-generation enhancements observed in periodically surging turbines, these flow physics could yield improvements on the order of ten percent in floating offshore wind farms, relative to their fixed-bottom counterparts. The theoretical approach and findings of this study can also be applied to dynamic induction control strategies for traditional onshore and fixed-bottom offshore turbines in unsteady streamwise flow conditions such as axial gusts, as well as the dynamics and control of hydrokinetic turbine arrays in streamwise oscillatory tidal flows. Future work will investigate the coupled effects of floating offshore platform motions and turbine control schemes on wake recovery, as well as the unsteady loading effect of the traveling-wave disturbances on downstream turbines in an array.

\backsection

[Acknowledgements]The authors would like to thank Pengming Guo for help with setting up the experiments. The derivation of the theoretical framework was aided by insightful discussions with Omkar B. Shende, Young R. (Paul) Yi, and Lena Sabidussi. A colormap from the cmocean library (Thyng et al., 2016) was used for Figures 6 and 8.

\backsection

[Funding]This work was supported by the National Science Foundation (grant number CBET-2038071) as well as the Natural Sciences and Engineering Research Council of Canada (grant number RGPIN-2023-03525). N.J.W. was supported by a National Science Foundation Graduate Research Fellowship and a Distinguished Postdoctoral Fellowship from the Andlinger Center for Energy and the Environment at Princeton University.

\backsection

[Declaration of interests]The authors report no conflict of interest.

\backsection

[Data availability statement]The data that support the findings of this study are available upon reasonable request.

\backsection

[Author contributions]N.J.W. derived the theory and analyzed the data. N.J.W., A.E.M., J.C.H., and F.K. performed the experiments. D.E.R. and J.O.D. procured funding and provided guidance on the project direction. All authors contributed to reaching conclusions, as well as writing and revising the paper.

Appendix A Derivation of the governing equations

We consider incompressible and inviscid flow through a differential streamwise slice of the larger wake control volume shown in Figure 1, with streamwise width 2Δx2Δ𝑥2\Delta x2 roman_Δ italic_x and central radius R(x,t)𝑅𝑥𝑡R(x,t)italic_R ( italic_x , italic_t ). The wake radius is assumed to vary linearly with streamwise position and time, with spatial variation ΔxRsubscriptΔ𝑥𝑅\Delta_{x}Rroman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R over a streamwise distance of ΔxΔ𝑥\Delta xroman_Δ italic_x and a temporal variation ΔtRsubscriptΔ𝑡𝑅\Delta_{t}Rroman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R over a time step of ΔtΔ𝑡\Delta troman_Δ italic_t. The streamwise velocity through the differential control volume is U|xΔxevaluated-at𝑈𝑥Δ𝑥U|_{x-\Delta x}italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT at the inlet and U|x+Δxevaluated-at𝑈𝑥Δ𝑥U|_{x+\Delta x}italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT at the outlet. The pressure outside the radius of the control volume is denoted psubscript𝑝p_{\infty}italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT and may vary with space and time. Radial velocities are neglected, as these will generally be smaller than the streamwise velocities through the control volume. The outer radius of the control volume has zero flow across its surface. A sketch of this differential control volume is given in Figure 19.

Refer to caption
Figure 19: Sketch of the differential control volume used to derive conservation relations for an unsteady, radially deforming wake.

First, we derive the relation for conservation of mass, given in final form as Equation 3. The integral form of mass conservation for the differential control volume is

tcvρ𝑑V+csρU𝑑A=0,𝑡subscript𝑐𝑣𝜌differential-d𝑉subscript𝑐𝑠𝜌𝑈differential-d𝐴0\frac{\partial}{\partial t}\int_{cv}\rho\,dV+\int_{cs}\rho U\,dA=0\,,divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∫ start_POSTSUBSCRIPT italic_c italic_v end_POSTSUBSCRIPT italic_ρ italic_d italic_V + ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_ρ italic_U italic_d italic_A = 0 , (19)

where cv𝑑Vsubscript𝑐𝑣differential-d𝑉\int_{cv}dV∫ start_POSTSUBSCRIPT italic_c italic_v end_POSTSUBSCRIPT italic_d italic_V denotes integration over the control volume, cs𝑑Asubscript𝑐𝑠differential-d𝐴\int_{cs}dA∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_d italic_A is a flux integral over the control surface, and ρ𝜌\rhoitalic_ρ is the fluid density. We use a second-order finite-difference approximation for the growth in the control volume as a function of time,

VtV|t+ΔtV|tΔt2Δt.𝑉𝑡evaluated-at𝑉𝑡Δ𝑡evaluated-at𝑉𝑡Δ𝑡2Δ𝑡\frac{\partial V}{\partial t}\approx\frac{V|_{t+\Delta t}-V|_{t-\Delta t}}{2% \Delta t}\,.divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_t end_ARG ≈ divide start_ARG italic_V | start_POSTSUBSCRIPT italic_t + roman_Δ italic_t end_POSTSUBSCRIPT - italic_V | start_POSTSUBSCRIPT italic_t - roman_Δ italic_t end_POSTSUBSCRIPT end_ARG start_ARG 2 roman_Δ italic_t end_ARG . (20)

The volume of the control volume can be calculated from the volume of a partial cone with height hhitalic_h and radii R1subscript𝑅1R_{1}italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and R2subscript𝑅2R_{2}italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT,

V=13πh(R12+R1R2+R22).𝑉13𝜋superscriptsubscript𝑅12subscript𝑅1subscript𝑅2superscriptsubscript𝑅22V=\frac{1}{3}\pi h\left(R_{1}^{2}+R_{1}R_{2}+R_{2}^{2}\right)\,.italic_V = divide start_ARG 1 end_ARG start_ARG 3 end_ARG italic_π italic_h ( italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (21)

We therefore obtain the unsteady volumetric growth in the control volume as

Vt2πΔx(R2+2RΔtR)2πΔx(R22RΔtR)2Δt=4πRΔxΔtRΔt,𝑉𝑡2𝜋Δ𝑥superscript𝑅22𝑅subscriptΔ𝑡𝑅2𝜋Δ𝑥superscript𝑅22𝑅subscriptΔ𝑡𝑅2Δ𝑡4𝜋𝑅Δ𝑥subscriptΔ𝑡𝑅Δ𝑡\frac{\partial V}{\partial t}\approx\frac{2\pi\Delta x\left(R^{2}+2R\Delta_{t}% R\right)-2\pi\Delta x\left(R^{2}-2R\Delta_{t}R\right)}{2\Delta t}=4\pi R\Delta x% \frac{\Delta_{t}R}{\Delta t}\,,divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_t end_ARG ≈ divide start_ARG 2 italic_π roman_Δ italic_x ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_R roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R ) - 2 italic_π roman_Δ italic_x ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_R roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R ) end_ARG start_ARG 2 roman_Δ italic_t end_ARG = 4 italic_π italic_R roman_Δ italic_x divide start_ARG roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R end_ARG start_ARG roman_Δ italic_t end_ARG , (22)

where we have neglected terms of order Δ2superscriptΔ2\Delta^{2}roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and above. Including the flux terms, Equation 19 becomes

tcv𝑑V+csU𝑑A=4πRΔxΔtRΔt+πU|x+Δx(R+ΔxR)2πU|xΔx(RΔxR)24πRΔxΔtRΔt+πU|x+Δx(R2+ΔxR)πU|xΔx(R2ΔxR)=0,𝑡subscript𝑐𝑣differential-d𝑉subscript𝑐𝑠𝑈differential-d𝐴4𝜋𝑅Δ𝑥subscriptΔ𝑡𝑅Δ𝑡evaluated-at𝜋𝑈𝑥Δ𝑥superscript𝑅subscriptΔ𝑥𝑅2evaluated-at𝜋𝑈𝑥Δ𝑥superscript𝑅subscriptΔ𝑥𝑅24𝜋𝑅Δ𝑥subscriptΔ𝑡𝑅Δ𝑡evaluated-at𝜋𝑈𝑥Δ𝑥superscript𝑅2subscriptΔ𝑥𝑅evaluated-at𝜋𝑈𝑥Δ𝑥superscript𝑅2subscriptΔ𝑥𝑅0\frac{\partial}{\partial t}\int_{cv}\,dV+\int_{cs}U\,dA\\ =4\pi R\Delta x\frac{\Delta_{t}R}{\Delta t}+\pi U|_{x+\Delta x}\left(R+\Delta_% {x}R\right)^{2}-\pi U|_{x-\Delta x}\left(R-\Delta_{x}R\right)^{2}\\ \approx 4\pi R\Delta x\frac{\Delta_{t}R}{\Delta t}+\pi U|_{x+\Delta x}\left(R^% {2}+\Delta_{x}R\right)-\pi U|_{x-\Delta x}\left(R^{2}-\Delta_{x}R\right)=0\,,start_ROW start_CELL divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∫ start_POSTSUBSCRIPT italic_c italic_v end_POSTSUBSCRIPT italic_d italic_V + ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_U italic_d italic_A end_CELL end_ROW start_ROW start_CELL = 4 italic_π italic_R roman_Δ italic_x divide start_ARG roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R end_ARG start_ARG roman_Δ italic_t end_ARG + italic_π italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT ( italic_R + roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_π italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ( italic_R - roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL ≈ 4 italic_π italic_R roman_Δ italic_x divide start_ARG roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R end_ARG start_ARG roman_Δ italic_t end_ARG + italic_π italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) - italic_π italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) = 0 , end_CELL end_ROW (23)

again neglecting higher-order products of ΔΔ\Deltaroman_Δ. Assuming U𝑈Uitalic_U varies linearly with x𝑥xitalic_x, we can write U|x+Δx+U|xΔx=2Uevaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥2𝑈U|_{x+\Delta x}+U|_{x-\Delta x}=2Uitalic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT + italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT = 2 italic_U and U|x+ΔxU|xΔx=2ΔxUevaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥2subscriptΔ𝑥𝑈U|_{x+\Delta x}-U|_{x-\Delta x}=2\Delta_{x}Uitalic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT - italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT = 2 roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_U. Simplifying the relation for continuity, we obtain

4RΔxΔtRΔt+2R2ΔxU+4RUΔxR=0.4𝑅Δ𝑥subscriptΔ𝑡𝑅Δ𝑡2superscript𝑅2subscriptΔ𝑥𝑈4𝑅𝑈subscriptΔ𝑥𝑅04R\Delta x\frac{\Delta_{t}R}{\Delta t}+2R^{2}\Delta_{x}U+4RU\Delta_{x}R=0\,.4 italic_R roman_Δ italic_x divide start_ARG roman_Δ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_R end_ARG start_ARG roman_Δ italic_t end_ARG + 2 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_U + 4 italic_R italic_U roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R = 0 . (24)

Dividing by 4rΔx4𝑟Δ𝑥4r\Delta x4 italic_r roman_Δ italic_x and taking the limits as Δx0Δ𝑥0\Delta x\rightarrow 0roman_Δ italic_x → 0 and Δt0Δ𝑡0\Delta t\rightarrow 0roman_Δ italic_t → 0, we arrive at the differential form of the continuity equation,

Rt+12RUx+uRx=0,𝑅𝑡12𝑅𝑈𝑥𝑢𝑅𝑥0\frac{\partial R}{\partial t}+\frac{1}{2}R\frac{\partial U}{\partial x}+u\frac% {\partial R}{\partial x}=0\,,divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG + italic_u divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG = 0 , (25)

which corresponds to Equation 3.

A similar process can be undertaken to derive the differential form of the momentum equation (Equation 4). The integral form of conservation of momentum in the streamwise direction is

tcvρU𝑑V+csρU2𝑑A=csp𝑑A.𝑡subscript𝑐𝑣𝜌𝑈differential-d𝑉subscript𝑐𝑠𝜌superscript𝑈2differential-d𝐴subscript𝑐𝑠𝑝differential-d𝐴\frac{\partial}{\partial t}\int_{cv}\rho U\,dV+\int_{cs}\rho U^{2}\,dA=-\int_{% cs}p\,dA\,.divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∫ start_POSTSUBSCRIPT italic_c italic_v end_POSTSUBSCRIPT italic_ρ italic_U italic_d italic_V + ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_ρ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_A = - ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_p italic_d italic_A . (26)

The unsteady term can be integrated over the differential control volume using cylindrical coordinates:

tcvU𝑑V=2πΔxΔxt0R(t)Ur𝑑r𝑑x=2πΔxΔx[URRt+0R(t)tUr𝑑r]𝑑x2πΔx(2URRt+R2Ut),𝑡subscript𝑐𝑣𝑈differential-d𝑉2𝜋superscriptsubscriptΔ𝑥Δ𝑥𝑡superscriptsubscript0𝑅𝑡𝑈superscript𝑟differential-dsuperscript𝑟differential-dsuperscript𝑥2𝜋superscriptsubscriptΔ𝑥Δ𝑥delimited-[]𝑈𝑅𝑅𝑡superscriptsubscript0𝑅𝑡𝑡𝑈superscript𝑟differential-dsuperscript𝑟differential-dsuperscript𝑥2𝜋Δ𝑥2𝑈𝑅𝑅𝑡superscript𝑅2𝑈𝑡\frac{\partial}{\partial t}\int_{cv}U\,dV=2\pi\int_{-\Delta x}^{\Delta x}\frac% {\partial}{\partial t}\int_{0}^{R(t)}Ur^{\prime}\,dr^{\prime}dx^{\prime}=2\pi% \int_{-\Delta x}^{\Delta x}\left[UR\frac{\partial R}{\partial t}+\int_{0}^{R(t% )}\frac{\partial}{\partial t}Ur^{\prime}\,dr^{\prime}\right]dx^{\prime}\\ \approx 2\pi\Delta x\left(2UR\frac{\partial R}{\partial t}+R^{2}\frac{\partial U% }{\partial t}\right)\,,start_ROW start_CELL divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∫ start_POSTSUBSCRIPT italic_c italic_v end_POSTSUBSCRIPT italic_U italic_d italic_V = 2 italic_π ∫ start_POSTSUBSCRIPT - roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_x end_POSTSUPERSCRIPT divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R ( italic_t ) end_POSTSUPERSCRIPT italic_U italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_d italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 2 italic_π ∫ start_POSTSUBSCRIPT - roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_x end_POSTSUPERSCRIPT [ italic_U italic_R divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R ( italic_t ) end_POSTSUPERSCRIPT divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG italic_U italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ] italic_d italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL ≈ 2 italic_π roman_Δ italic_x ( 2 italic_U italic_R divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG ) , end_CELL end_ROW (27)

employing the Leibniz integral rule and neglecting higher-order terms in the integration over xsuperscript𝑥x^{\prime}italic_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. For the advection term, we have

csU2𝑑A=U|x+Δx2π(R+ΔxR)2U|xΔx2π(RΔxR)2πR2(U|x+Δx2U|xΔx2)+2πRΔxR(U|x+Δx2+U|xΔx2)=πR2(U|x+Δx+U|xΔx)(U|x+ΔxU|xΔx)+2πRΔxR((U|x+Δx+U|xΔx)22U|x+ΔxU|xΔx)4πRU(RΔxU+UΔxR),subscript𝑐𝑠superscript𝑈2differential-d𝐴evaluated-at𝑈𝑥Δ𝑥2𝜋superscript𝑅subscriptΔ𝑥𝑅2evaluated-at𝑈𝑥Δ𝑥2𝜋superscript𝑅subscriptΔ𝑥𝑅2𝜋superscript𝑅2evaluated-at𝑈𝑥Δ𝑥2evaluated-at𝑈𝑥Δ𝑥22𝜋𝑅subscriptΔ𝑥𝑅evaluated-at𝑈𝑥Δ𝑥2evaluated-at𝑈𝑥Δ𝑥2𝜋superscript𝑅2evaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥2𝜋𝑅subscriptΔ𝑥𝑅superscriptevaluated-at𝑈𝑥Δ𝑥evaluated-at𝑈𝑥Δ𝑥2evaluated-atevaluated-at2𝑈𝑥Δ𝑥𝑈𝑥Δ𝑥4𝜋𝑅𝑈𝑅subscriptΔ𝑥𝑈𝑈subscriptΔ𝑥𝑅\int_{cs}U^{2}\,dA={U|_{x+\Delta x}}^{2}\pi\left(R+\Delta_{x}R\right)^{2}-{U|_% {x-\Delta x}}^{2}\pi\left(R-\Delta_{x}R\right)^{2}\\ \approx\pi R^{2}\left({U|_{x+\Delta x}}^{2}-{U|_{x-\Delta x}}^{2}\right)+2\pi R% \Delta_{x}R\left({U|_{x+\Delta x}}^{2}+{U|_{x-\Delta x}}^{2}\right)\\ =\pi R^{2}\left({U|_{x+\Delta x}}+{U|_{x-\Delta x}}\right)\left({U|_{x+\Delta x% }}-{U|_{x-\Delta x}}\right)\\ +2\pi R\Delta_{x}R\left(\left({U|_{x+\Delta x}}+{U|_{x-\Delta x}}\right)^{2}-2% U|_{x+\Delta x}U|_{x-\Delta x}\right)\\ \approx 4\pi RU\left(R\Delta_{x}U+U\Delta_{x}R\right)\,,start_ROW start_CELL ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_A = italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π ( italic_R + roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π ( italic_R - roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL ≈ italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + 2 italic_π italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ( italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_CELL end_ROW start_ROW start_CELL = italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT + italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ) ( italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT - italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ) end_CELL end_ROW start_ROW start_CELL + 2 italic_π italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ( ( italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT + italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT ) end_CELL end_ROW start_ROW start_CELL ≈ 4 italic_π italic_R italic_U ( italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_U + italic_U roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) , end_CELL end_ROW (28)

where we have neglected higher-order terms, completed the square, and noted that U|x+ΔxU|xΔx=(U+ΔU)(UΔU)U2evaluated-atevaluated-at𝑈𝑥Δ𝑥𝑈𝑥Δ𝑥𝑈Δ𝑈𝑈Δ𝑈superscript𝑈2U|_{x+\Delta x}U|_{x-\Delta x}=(U+\Delta U)(U-\Delta U)\approx U^{2}italic_U | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT italic_U | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT = ( italic_U + roman_Δ italic_U ) ( italic_U - roman_Δ italic_U ) ≈ italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For a pressure distribution outside of the wake psubscript𝑝p_{\infty}italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT that does not vary in the radial direction, the pressure contribution to the momentum balance may be written as

csp𝑑A=π(R+ΔxR)2p|x+Δxπ(RΔxR)2p|xΔx2πRΔxRR+ΔxRrp𝑑rπ(R2+2rΔxR)p|x+Δxπ(R22rΔxR)p|xΔx4πpRΔxR=2πR(RΔxp+2(pp)ΔxR).subscript𝑐𝑠𝑝differential-d𝐴evaluated-at𝜋superscript𝑅subscriptΔ𝑥𝑅2𝑝𝑥Δ𝑥evaluated-at𝜋superscript𝑅subscriptΔ𝑥𝑅2𝑝𝑥Δ𝑥2𝜋superscriptsubscript𝑅subscriptΔ𝑥𝑅𝑅subscriptΔ𝑥𝑅superscript𝑟subscript𝑝differential-dsuperscript𝑟evaluated-at𝜋superscript𝑅22𝑟subscriptΔ𝑥𝑅𝑝𝑥Δ𝑥evaluated-at𝜋superscript𝑅22𝑟subscriptΔ𝑥𝑅𝑝𝑥Δ𝑥4𝜋subscript𝑝𝑅subscriptΔ𝑥𝑅2𝜋𝑅𝑅subscriptΔ𝑥𝑝2𝑝subscript𝑝subscriptΔ𝑥𝑅\int_{cs}p\,dA=\pi\left(R+\Delta_{x}R\right)^{2}p|_{x+\Delta x}-\pi\left(R-% \Delta_{x}R\right)^{2}p|_{x-\Delta x}-2\pi\int_{R-\Delta_{x}R}^{R+\Delta_{x}R}% r^{\prime}p_{\infty}\,dr^{\prime}\\ \approx\pi\left(R^{2}+2r\Delta_{x}R\right)p|_{x+\Delta x}-\pi\left(R^{2}-2r% \Delta_{x}R\right)p|_{x-\Delta x}-4\pi p_{\infty}R\Delta_{x}R\\ =2\pi R\left(R\Delta_{x}p+2(p-p_{\infty})\Delta_{x}R\right)\,.start_ROW start_CELL ∫ start_POSTSUBSCRIPT italic_c italic_s end_POSTSUBSCRIPT italic_p italic_d italic_A = italic_π ( italic_R + roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT - italic_π ( italic_R - roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_p | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT - 2 italic_π ∫ start_POSTSUBSCRIPT italic_R - roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R + roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL ≈ italic_π ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_r roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) italic_p | start_POSTSUBSCRIPT italic_x + roman_Δ italic_x end_POSTSUBSCRIPT - italic_π ( italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_r roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) italic_p | start_POSTSUBSCRIPT italic_x - roman_Δ italic_x end_POSTSUBSCRIPT - 4 italic_π italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R end_CELL end_ROW start_ROW start_CELL = 2 italic_π italic_R ( italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_p + 2 ( italic_p - italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) . end_CELL end_ROW (29)

Finally, putting all of these terms together, the momentum equation over the differential control volume is given by

2πΔx(2URRt+R2Ut)+4πRU(RΔxU+UΔxR)=2πR(RΔxp+2(pp)ΔxR)2𝜋Δ𝑥2𝑈𝑅𝑅𝑡superscript𝑅2𝑈𝑡4𝜋𝑅𝑈𝑅subscriptΔ𝑥𝑈𝑈subscriptΔ𝑥𝑅2𝜋𝑅𝑅subscriptΔ𝑥𝑝2𝑝subscript𝑝subscriptΔ𝑥𝑅2\pi\Delta x\left(2UR\frac{\partial R}{\partial t}+R^{2}\frac{\partial U}{% \partial t}\right)+4\pi RU\left(R\Delta_{x}U+U\Delta_{x}R\right)=2\pi R\left(R% \Delta_{x}p+2(p-p_{\infty})\Delta_{x}R\right)\,2 italic_π roman_Δ italic_x ( 2 italic_U italic_R divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG ) + 4 italic_π italic_R italic_U ( italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_U + italic_U roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) = 2 italic_π italic_R ( italic_R roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_p + 2 ( italic_p - italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) roman_Δ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_R ) (30)

which, after dividing by 2πRΔx2𝜋𝑅Δ𝑥2\pi R\Delta x2 italic_π italic_R roman_Δ italic_x and taking the limit as Δx0Δ𝑥0\Delta x\rightarrow 0roman_Δ italic_x → 0, yields

RUt+2U(Rt+RUx+URx)=1ρ(Rpx+2Rx(pp)).𝑅𝑈𝑡2𝑈𝑅𝑡𝑅𝑈𝑥𝑈𝑅𝑥1𝜌𝑅𝑝𝑥2𝑅𝑥𝑝subscript𝑝R\frac{\partial U}{\partial t}+2U\left(\frac{\partial R}{\partial t}+R\frac{% \partial U}{\partial x}+U\frac{\partial R}{\partial x}\right)=-\frac{1}{\rho}% \left(R\frac{\partial p}{\partial x}+2\frac{\partial R}{\partial x}(p-p_{% \infty})\right)\,.italic_R divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + 2 italic_U ( divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_t end_ARG + italic_R divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG + italic_U divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG ) = - divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG ( italic_R divide start_ARG ∂ italic_p end_ARG start_ARG ∂ italic_x end_ARG + 2 divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG ( italic_p - italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) ) . (31)

Subtracting Equation 25 from this equation gives

Ut+UUx=1ρpx2ρRRx(pp).𝑈𝑡𝑈𝑈𝑥1𝜌𝑝𝑥2𝜌𝑅𝑅𝑥𝑝subscript𝑝\frac{\partial U}{\partial t}+U\frac{\partial U}{\partial x}=-\frac{1}{\rho}% \frac{\partial p}{\partial x}-\frac{2}{\rho R}\frac{\partial R}{\partial x}% \left(p-p_{\infty}\right)\,.divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + italic_U divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG = - divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ italic_p end_ARG start_ARG ∂ italic_x end_ARG - divide start_ARG 2 end_ARG start_ARG italic_ρ italic_R end_ARG divide start_ARG ∂ italic_R end_ARG start_ARG ∂ italic_x end_ARG ( italic_p - italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) . (32)

If we assume that the pressure outside the wake is radially homogeneous, i.e. p=p(x)subscript𝑝𝑝𝑥p_{\infty}=p(x)italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = italic_p ( italic_x ), then the extra pressure term becomes zero and we arrive at the 1D Euler equation,

Ut+UUx=1ρpx,𝑈𝑡𝑈𝑈𝑥1𝜌𝑝𝑥\frac{\partial U}{\partial t}+U\frac{\partial U}{\partial x}=-\frac{1}{\rho}% \frac{\partial p}{\partial x}\,,divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + italic_U divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_x end_ARG = - divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ italic_p end_ARG start_ARG ∂ italic_x end_ARG , (33)

which corresponds to Equation 4.

Appendix B Further discussion of the quasi-steady wake approximation

The results presented in this study highlight the importance of unsteady fluid dynamics in the wakes of turbines with time-varying thrust profiles. The traveling-wave solutions and tip-vortex roll-up behaviors are not phenomena that can be obtained from quasi-steady flow assumptions. To demonstrate this even more clearly, the streamwise variations in the wake-radius amplitude r^^𝑟\hat{r}over^ start_ARG italic_r end_ARG and streamwise-velocity amplitude U^^𝑈\hat{U}over^ start_ARG italic_U end_ARG are shown in Figures 20 and 21 for both loading conditions. The dark-colored lines are taken from unsteady cases with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89, while the lines with shaded uncertainty intervals are calculated by subtracting the wake radius and streamwise velocity across steady-flow cases with U=0.8subscript𝑈0.8U_{\infty}=0.8italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.8 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and U=1.2subscript𝑈1.2U_{\infty}=1.2italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1.2 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT to represent an effective surge-velocity amplitude of u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 (with k=0𝑘0k=0italic_k = 0). The differences between the representative quasi-steady amplitudes and the measured unsteady amplitudes are clear. For the wake radius in the unsteady case, the amplitude increases with downstream distance. The differences in wake radius across the quasi-steady flow representations are small in comparison and do not display the same growth and saturation as seen in the measurements. Likewise, the streamwise-velocity amplitude shows opposite trends between the unsteady and quasi-steady measurements. While U^^𝑈\hat{U}over^ start_ARG italic_U end_ARG decays with downstream distance in the unsteady case, in accordance with the modeling framework from Section 2.1, the quasi-steady amplitude grows. This is due to differences in the steady-flow velocity deficit between the two quasi-steady datasets. These comparisons further underscore the need for dynamic models of the flows in unsteady turbine wakes.

Refer to caption
(a)
Refer to caption
(b)
Figure 20: Amplitude of the wake radius, calculated from unsteady measurements with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89 (darker lines) and quasi-steady flow measurements at U=0.8subscript𝑈0.8U_{\infty}=0.8italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.8 and 1.2 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (lines with uncertainty bounds). Cases with λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 (a) and λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07 (b) are shown.
Refer to caption
(a)
Refer to caption
(b)
Figure 21: Amplitude of the streamwise velocity, calculated from unsteady measurements with u=0.2superscript𝑢0.2u^{*}=0.2italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0.2 and k=1.89𝑘1.89k=1.89italic_k = 1.89 (darker lines) and quasi-steady flow measurements at U=0.8subscript𝑈0.8U_{\infty}=0.8italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.8 and 1.2 ms1superscriptms1\rm{ms^{-1}}roman_ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (lines with uncertainty bounds). Cases with λ0=4.50subscript𝜆04.50\lambda_{0}=4.50italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.50 (a) and λ0=5.07subscript𝜆05.07\lambda_{0}=5.07italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5.07 (b) are shown.

For the sake of completeness, we also plot the amplitudes of the radially averaged streamwise velocity as a function of surge-velocity amplitude, shown for both loading conditions in Figure 22. Unlike the wake-radius amplitudes shown previously in Figure 15, the streamwise-velocity amplitudes computed by the model strongly underpredict the measured data for most cases. This is likely due to the effects of wake recovery, which more directly impact the streamwise velocity than the wake radius. Despite this lack of quantitative agreement, the model still captures the trends in the data: the streamwise-velocity amplitude is shown to increase with increasing surge-velocity amplitude, and decrease with increasing streamwise distance. In accordance with the comparisons shown above in Figure 21, the quasi-steady amplitudes show the opposite trend with streamwise distance. Though for this quantity the modeling approach does not yield accurate quantitative predictions, these results suggest that the unsteady flow physics it parameterizes are to some extent reflective of the full dynamics of unsteady turbine wakes.

Refer to caption
(a)
Refer to caption
(b)
Figure 22: Comparison between theoretical and experimental results for the amplitude of the streamwise velocity U^/U^𝑈subscript𝑈\hat{U}/U_{\infty}over^ start_ARG italic_U end_ARG / italic_U start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, plotted as a function of surge-velocity amplitude usuperscript𝑢u^{*}italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Experimental data are shown as colored markers, model solutions are shown as darker-colored open markers, and results obtained from quasi-steady measurements are given as colored open markers. Linear interpolations between the model solutions are shown as dashed lines. The colors and shapes of the markers correspond to their streamwise locations in the wake. The model shows some qualitative agreement with the trends observed in the data, but is not quantitatively accurate.

References

  • Barthelmie et al. (2009) Barthelmie, R. J., Hansen, K., Frandsen, S. T., Rathmann, O., Schepers, J. G., Schlez, W., Phillips, J., Rados, K., Zervos, A., Politis, E. S. & Chaviaropoulos, P. K. 2009 Modelling and measuring flow and wind turbine wakes in large wind farms offshore. Wind Energy 12 (5), 431–444.
  • Bastankhah & Porté-Agel (2014) Bastankhah, Majid & Porté-Agel, Fernando 2014 A new analytical model for wind-turbine wakes. Renewable Energy 70, 116–123.
  • van den Berg et al. (2023) van den Berg, Daniel, De Tavernier, Delphine & van Wingerden, Jan-Willem 2023 The dynamic coupling between the pulse wake mixing strategy and floating wind turbines. Wind Energy Science 8 (5), 849–864.
  • Bossuyt et al. (2023) Bossuyt, Juliaan, Ferčák, Ondřej, Sadek, Zein, Meneveau, Charles, Gayme, Dennice F. & Cal, Raúl Bayoán 2023 Floating wind farm experiments through scaling for wake characterization, power extraction, and turbine dynamics. Physical Review Fluids 8 (12), 120501.
  • van den Broek et al. (2023) van den Broek, Maarten J., Van den Berg, Daniel, Sanderse, Benjamin & van Wingerden, Jan-Willem Van 2023 Optimal Control for Wind Turbine Wake Mixing on Floating Platforms. IFAC-PapersOnLine 56 (2), 7656–7661.
  • Brown et al. (2022) Brown, Kenneth, Houck, Daniel, Maniaci, David, Westergaard, Carsten & Kelley, Christopher 2022 Accelerated Wind-Turbine Wake Recovery Through Actuation of the Tip-Vortex Instability. AIAA Journal 60 (5), 3298–3310.
  • Cioni et al. (2023) Cioni, Stefano, Papi, Francesco, Pagamonci, Leonardo, Bianchini, Alessandro, Ramos-García, Néstor, Pirrung, Georg, Corniglion, Rémi, Lovera, Anaïs, Galván, Josean, Boisard, Ronan, Fontanella, Alessandro, Schito, Paolo, Zasso, Alberto, Belloli, Marco, Sanvito, Andrea, Persico, Giacomo, Zhang, Lijun, Li, Ye, Zhou, Yarong, Mancini, Simone, Boorsma, Koen, Amaral, Ricardo, Viré, Axelle, Schulz, Christian W., Netzband, Stefan, Soto-Valle, Rodrigo, Marten, David, Martín-San-Román, Raquel, Trubat, Pau, Molins, Climent, Bergua, Roger, Branlard, Emmanuel, Jonkman, Jason & Robertson, Amy 2023 On the characteristics of the wake of a wind turbine undergoing large motions caused by a floating structure: an insight based on experiments and multi-fidelity simulations from the OC6 project Phase III. Wind Energy Science 8 (11), 1659–1691.
  • Currie & Turnbull (1987) Currie, I.G. & Turnbull, D.H. 1987 Streamwise oscillations of cylinders near the critical Reynolds number. Journal of Fluids and Structures 1 (2), 185–196.
  • Dabiri (2020) Dabiri, John O. 2020 Theoretical framework to surpass the Betz limit using unsteady fluid mechanics. Physical Review Fluids 5 (2), 022501, publisher: American Physical Society.
  • Duckworth & Barthelmie (2008) Duckworth, A. & Barthelmie, R.J. 2008 Investigation and Validation of Wind Turbine Wake Models. Wind Engineering 32 (5), 459–475.
  • El Makdah et al. (2019) El Makdah, Adnan M., Ruzzante, Sacha, Zhang, Kai & Rival, David E. 2019 The influence of axial gusts on the output of low-inertia rotors. Journal of Fluids and Structures 88, 71–82.
  • El Makdah et al. (2021) El Makdah, Adnan M., Zhang, Kai & Rival, David E. 2021 The scaling of rotor inertia under dynamic inflow conditions. Journal of Fluids and Structures 106, 103357.
  • Felli et al. (2011) Felli, M., Camussi, R. & Di Felice, F. 2011 Mechanisms of evolution of the propeller wake in the transition and far fields. Journal of Fluid Mechanics 682, 5–53.
  • Fontanella et al. (2022) Fontanella, A., Zasso, A. & Belloli, M. 2022 Wind tunnel investigation of the wake-flow response for a floating turbine subjected to surge motion. Journal of Physics: Conference Series 2265 (4), 042023, publisher: IOP Publishing.
  • Frederik et al. (2020a) Frederik, Joeri A., Doekemeijer, Bart M., Mulders, Sebastiaan P. & van Wingerden, Jan-Willem 2020a The helix approach: Using dynamic individual pitch control to enhance wake mixing in wind farms. Wind Energy 23 (8), 1739–1751.
  • Frederik et al. (2020b) Frederik, Joeri Alexis, Weber, Robin, Cacciola, Stefano, Campagnolo, Filippo, Croce, Alessandro, Bottasso, Carlo & van Wingerden, Jan-Willem 2020b Periodic dynamic induction control of wind farms: proving the potential in simulations and wind tunnel experiments. Wind Energy Science 5 (1), 245–257.
  • Goit & Meyers (2015) Goit, Jay P. & Meyers, Johan 2015 Optimal control of energy extraction in wind-farm boundary layers. Journal of Fluid Mechanics 768, 5–50.
  • Heck et al. (2023) Heck, K.S., Johlas, H.M. & Howland, M.F. 2023 Modelling the induction, thrust and power of a yaw-misaligned actuator disk. Journal of Fluid Mechanics 959, A9.
  • van der Hoek et al. (2024) van der Hoek, Daan, van den Abbeele, Bert, Simao Ferreira, Carlos & van Wingerden, Jan-Willem 2024 Maximizing wind farm power output with the helix approach: Experimental validation and wake analysis using tomographic particle image velocimetry. Wind Energy 27 (5), 463–482.
  • van der Hoek et al. (2022) van der Hoek, Daan, Frederik, Joeri, Huang, Ming, Scarano, Fulvio, Simao Ferreira, Carlos & van Wingerden, Jan-Willem 2022 Experimental analysis of the effect of dynamic induction control on a wind turbine wake. Wind Energy Science 7 (3), 1305–1320.
  • Howland et al. (2019) Howland, Michael F., Lele, Sanjiva K. & Dabiri, John O. 2019 Wind farm power optimization through wake steering. Proceedings of the National Academy of Sciences 116 (29), 14495–14500.
  • Johlas et al. (2021) Johlas, Hannah M., Martínez-Tossas, Luis A., Churchfield, Matthew J., Lackner, Matthew A. & Schmidt, David P. 2021 Floating platform effects on power generation in spar and semisubmersible wind turbines. Wind Energy 24 (8), 901–916.
  • Johlas et al. (2019) Johlas, Hannah M., Martínez-Tossas, Luis A., Schmidt, David P., Lackner, Matthew A. & Churchfield, Matthew J. 2019 Large eddy simulations of floating offshore wind turbine wakes with coupled platform motion. Journal of Physics: Conference Series 1256 (1), 012018, publisher: IOP Publishing.
  • Kadum et al. (2019) Kadum, Hawwa, Rockel, Stanislav, Hölling, Michael, Peinke, Joachim & Cal, Raúl Bayoán 2019 Wind turbine wake intermittency dependence on turbulence intensity and pitch motion. Journal of Renewable and Sustainable Energy 11 (5), 053302, publisher: American Institute of Physics.
  • Kleine et al. (2022) Kleine, V. G., Franceschini, L., Carmo, B. S., Hanifi, A. & Henningson, D. S. 2022 The stability of wakes of floating wind turbines. Physics of Fluids 34 (7), 074106, publisher: American Institute of Physics.
  • Larsen & Hanson (2007) Larsen, T. J. & Hanson, T. D. 2007 A method to avoid negative damped low frequent tower vibrations for a floating, pitch controlled wind turbine. Journal of Physics: Conference Series 75, 012073, publisher: IOP Publishing.
  • Lignarolo et al. (2015) Lignarolo, L. E. M., Ragni, D., Scarano, F., Simão Ferreira, C. J. & Van Bussel, G. J. W. 2015 Tip-vortex instability and turbulent mixing in wind-turbine wakes. Journal of Fluid Mechanics 781, 467–493.
  • Messmer et al. (2024) Messmer, Thomas, Hölling, Michael & Peinke, Joachim 2024 Enhanced recovery caused by nonlinear dynamics in the wake of a floating offshore wind turbine. Journal of Fluid Mechanics 984, A66.
  • Meyers et al. (2022) Meyers, Johan, Bottasso, Carlo, Dykes, Katherine, Fleming, Paul, Gebraad, Pieter, Giebel, Gregor, Göçmen, Tuhfe & van Wingerden, Jan-Willem 2022 Wind farm flow control: prospects and challenges. Wind Energy Science 7 (6), 2271–2306.
  • Miller et al. (2019) Miller, Mark A., Kiefer, Janik, Westergaard, Carsten, Hansen, Martin O. L. & Hultmark, Marcus 2019 Horizontal axis wind turbine testing at high Reynolds numbers. Physical Review Fluids 4 (11), 110504.
  • Munters & Meyers (2017) Munters, W. & Meyers, J. 2017 An optimal control framework for dynamic induction control of wind farms and their interaction with the atmospheric boundary layer. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 375 (2091), 20160100.
  • Munters & Meyers (2018) Munters, Wim & Meyers, Johan 2018 Towards practical dynamic induction control of wind farms: analysis of optimally controlled wind-farm boundary layers and sinusoidal induction control of first-row turbines. Wind Energy Science 3 (1), 409–425.
  • Okulov & Sørensen (2007) Okulov, V. L. & Sørensen, J. N. 2007 Stability of helical tip vortices in a rotor far wake. Journal of Fluid Mechanics 576, 1–25, publisher: Cambridge University Press.
  • Piqué et al. (2022) Piqué, Alexander, Miller, Mark A. & Hultmark, Marcus 2022 Laboratory investigation of the near and intermediate wake of a wind turbine at very high Reynolds numbers. Experiments in Fluids 63 (6), 106.
  • Porté-Agel et al. (2020) Porté-Agel, Fernando, Bastankhah, Majid & Shamsoddin, Sina 2020 Wind-Turbine and Wind-Farm Flows: A Review. Boundary-Layer Meteorology 174 (1), 1–59.
  • Quaranta et al. (2015) Quaranta, Hugo Umberto, Bolnot, Hadrien & Leweke, Thomas 2015 Long-wave instability of a helical vortex. Journal of Fluid Mechanics 780, 687–716.
  • Quarton & Ainslie (1990) Quarton, D. C. & Ainslie, J. F. 1990 Turbulence in Wind Turbine Wakes. Wind Engineering 14 (1), 15–23.
  • Rockel et al. (2016) Rockel, Stanislav, Peinke, Joachim, Hölling, Michael & Cal, Raúl Bayoán 2016 Wake to wake interaction of floating wind turbine models in free pitch motion: An eddy viscosity and mixing length approach. Renewable Energy 85, 666–676.
  • Rodriguez & Jaworski (2020) Rodriguez, Steven N. & Jaworski, Justin W. 2020 Strongly-coupled aeroelastic free-vortex wake framework for floating offshore wind turbine rotors. Part 2: Application. Renewable Energy 149, 1018–1031.
  • Rodriguez et al. (2021) Rodriguez, Steven N., Jaworski, Justin W. & Michopoulos, John G. 2021 Stability of helical vortex structures shed from flexible rotors. Journal of Fluids and Structures 104, 103279.
  • Ruiz et al. (2011) Ruiz, Lydia A., Whittlesey, Robert W. & Dabiri, John O. 2011 Vortex-enhanced propulsion. Journal of Fluid Mechanics 668, 5–32.
  • Sarmast et al. (2014) Sarmast, Sasan, Dadfar, Reza, Mikkelsen, Robert F., Schlatter, Philipp, Ivanell, Stefan, Sørensen, Jens N. & Henningson, Dan S. 2014 Mutual inductance instability of the tip vortices behind a wind turbine. Journal of Fluid Mechanics 755, 705–731.
  • Scarlett & Viola (2020) Scarlett, Gabriel Thomas & Viola, Ignazio Maria 2020 Unsteady hydrodynamics of tidal turbine blades. Renewable Energy 146, 843–855.
  • Smits (2019) Smits, Alexander J. 2019 Undulatory and oscillatory swimming. Journal of Fluid Mechanics 874, P1.
  • Stevens & Meneveau (2017) Stevens, Richard J.A.M. & Meneveau, Charles 2017 Flow Structure and Turbulence in Wind Farms. Annual Review of Fluid Mechanics 49 (1), 311–339.
  • Thielicke & Stamhuis (2014) Thielicke, William & Stamhuis, Eize J. 2014 PIVlab – Towards User-friendly, Affordable and Accurate Digital Particle Image Velocimetry in MATLAB. Journal of Open Research Software 2.
  • Thyng et al. (2016) Thyng, Kristen, Greene, Chad, Hetland, Robert, Zimmerle, Heather & DiMarco, Steven 2016 True Colors of Oceanography: Guidelines for Effective and Accurate Colormap Selection. Oceanography 29 (3), 9–13.
  • Uberoi & Freymuth (1970) Uberoi, Mahinder S. & Freymuth, Peter 1970 Turbulent Energy Balance and Spectra of the Axisymmetric Wake. The Physics of Fluids 13 (9), 2205–2210.
  • de Vaal et al. (2014a) de Vaal, J. B., Hansen, M. O. L. & Moan, T. 2014a Effect of wind turbine surge motion on rotor thrust and induced velocity. Wind Energy 17 (1), 105–121.
  • de Vaal et al. (2014b) de Vaal, J. B., Hansen, M. O. L. & Moan, T. 2014b Validation of a vortex ring wake model suited for aeroelastic simulations of floating wind turbines. Journal of Physics: Conference Series 555 (1), 012025, publisher: IOP Publishing.
  • Wayman (2006) Wayman, Elizabeth N. 2006 Coupled dynamics and economic analysis of floating wind turbine systems. Thesis, Massachusetts Institute of Technology, accepted: 2007-01-10T16:56:13Z.
  • Wei & Dabiri (2022) Wei, Nathaniel J. & Dabiri, John O. 2022 Phase-averaged dynamics of a periodically surging wind turbine. Journal of Renewable and Sustainable Energy 14 (1), 013305, publisher: American Institute of Physics.
  • Wei & Dabiri (2023) Wei, Nathaniel J. & Dabiri, John O. 2023 Power-generation enhancements and upstream flow properties of turbines in unsteady inflow conditions. Journal of Fluid Mechanics 966, A30.
  • Wen et al. (2018) Wen, Binrong, Dong, Xingjian, Tian, Xinliang, Peng, Zhike, Zhang, Wenming & Wei, Kexiang 2018 The power performance of an offshore floating wind turbine in platform pitching motion. Energy 154, 508–521.