Hostname: page-component-cb9f654ff-hn9fh Total loading time: 0 Render date: 2025-09-05T09:00:56.005Z Has data issue: false hasContentIssue false

The role of motion-excited coherent structures in improved wake recovery of a floating wind turbine

Published online by Cambridge University Press:  01 September 2025

Thomas Messmer*
Affiliation:
Carl von Ossietzky Universität Oldenburg, School of Mathematics and Science, Institute of Physics, Oldenburg, Germany ForWind - Center for Wind Energy Research, Küpkersweg 70, Oldenburg 26129, Germany
Joachim Peinke
Affiliation:
Carl von Ossietzky Universität Oldenburg, School of Mathematics and Science, Institute of Physics, Oldenburg, Germany ForWind - Center for Wind Energy Research, Küpkersweg 70, Oldenburg 26129, Germany
Alessandro Croce
Affiliation:
Department of Aerospace Science and Technology, Politecnico di Milano, Milan, Italy
Michael Hölling
Affiliation:
Carl von Ossietzky Universität Oldenburg, School of Mathematics and Science, Institute of Physics, Oldenburg, Germany ForWind - Center for Wind Energy Research, Küpkersweg 70, Oldenburg 26129, Germany
*
Corresponding author: Thomas Messmer, thomas.messmer@uni-oldenburg.de

Abstract

This study experimentally investigates wake recovery mechanisms behind a floating wind turbine subjected to imposed fore-aft (surge) and side-to-side (sway) motions. Wind tunnel experiments with varying free-stream turbulence intensities ($\textit{TI}_{\infty } \in [1.1, 5.8]\,\%$) are presented. Rotor motion induces large-scale coherent structures – pulsating for surge and meandering for sway – whose development critically depends on the energy ratio between the incoming turbulence and the platform motion. The results provide direct evidence supporting the role of these structures in enhancing wake recovery, as previously speculated by Messmer, Peinke & Hölling (J. Fluid Mech., vol. 984, 2024, A66). These periodic structures significantly increase Reynolds shear stress gradients, particularly in the streamwise–lateral direction, which are key drivers of wake recovery. However, their influence diminishes with increasing $\textit{TI}_{\infty }$: higher background turbulence weakens the coherent flow patterns, reducing their contribution to recovery. Beyond a threshold turbulence level – determined by the energy, frequency and direction of motion – rotor-induced structures no longer contribute meaningfully to recovery, which becomes primarily driven by the free-stream turbulence. Finally, we show that the meandering structures generated by sway motion are more resilient in turbulent backgrounds than the pulsating modes from surge, making sway more effective for promoting enhanced wake recovery.

Information

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

Early work on idealised free shear flows (jet, mixing layer, wake) showed the existence of ‘large-scale orderly patterns’ forming due to the amplification of small excitation (Crow & Champagne Reference Crow and Champagne1971; Oster & Wygnanski Reference Oster and Wygnanski1982; Cantwell & Coles Reference Cantwell and Coles1983; Ho & Huerre Reference Ho and Huerre1984). These structures play a central role in the development of the flow. In the context of floating offshore wind turbines (FOWT), Li, Dong & Yang (Reference Li, Dong and Yang2022), Messmer, Peinke & Hölling (Reference Messmer, Peinke and Hölling2024a ), Li, Yu & Sarlak (Reference Li, Yu and Sarlak2024), Li & Yang (Reference Li and Yang2024), Hubert, Conan & Aubrun (Reference Hubert, Conan and Aubrun2025) found organised flow structures in the wake of the turbine that arise from the harmonic motion of the rotor. Numerical studies by Kleine et al. (Reference Kleine, Franceschini, Carmo, Hanifi and Henningson2022) and Ramos-García et al. (Reference Ramos‐García, Kontos, Pegalajar‐Jurado, González Horcas and Bredmose2022) examined the influence of platform motion on wake dynamics under laminar wind and found that rotor oscillations lead to near-wake instabilities (Kleine et al. Reference Kleine, Franceschini, Carmo, Hanifi and Henningson2022), while also enhancing turbulence, mixing and energy distribution within the wake (Ramos-García et al. Reference Ramos‐García, Kontos, Pegalajar‐Jurado, González Horcas and Bredmose2022). Experimental works by Rockel et al. (Reference Rockel, Peinke, Hölling and Cal2017), Fu et al. (Reference Fu, Jin, Zheng and Chamorro2019) and Fontanella et al. (Reference Fontanella, Zasso and Belloli2022, Reference Fontanella, Fusetti, Cioni, Papi, Muggiasca, Persico, Dossena, Bianchini and Belloli2025) further support these findings. Under more realistic conditions, the intricate interactions between platform motion, inflow conditions and rotor aerodynamics result in complex wake dynamics, as explored and discussed in recent research by Bossuyt et al. (Reference Bossuyt, Ferčák, Sadek, Meneveau, Gayme and Cal2023), Schliffke, Conan & Aubrun (Reference Schliffke, Conan and Aubrun2024), Pagamonci et al. (Reference Pagamonci, Papi, Cojocaru, Belloli and Bianchini2025).

Wake recovery is a key aspect in wind energy, as it determines how the flow behind a turbine can regain lost energy through interaction with its surroundings. Li et al. (Reference Li, Dong and Yang2022), Messmer et al. (Reference Messmer, Hölling and Peinke2024a ) showed in idealised conditions that rotor motion can enhance wake recovery by up to $25\,\%$ more than that of a fixed turbine. Especially for motion frequencies characterised by a platform Strouhal number $St = f_p D / U_{\infty } \in [0.2, 0.6]$ , with $f_p$ the platform’s motion frequency, $D$ the rotor diameter and $U_{\infty }$ the incoming wind speed. They observed the formation of amplified meandering-like coherent structures from sideways rotor motion and pulsating structures from fore-aft motion, synchronised with the movements of the rotor. It remains to be seen what role these large-scale coherent structures play in the enhanced wake recovery.

Interestingly, similar wake dynamics of synchronisation and enhanced recovery are found for quite different types of harmonic excitation (actuation) on a fixed wind turbine with helix blade pitching control strategy (Frederik et al. Reference Frederik, Doekemeijer, Mulders and van Wingerden2020; Korb, Asmuth & Ivanell Reference Korb, Asmuth and Ivanell2023; Van der Hoek et al. Reference van der Hoek, den Abbeele, Simao Ferreira and van Wingerden2024), periodic variation of inflow structures (Mao & Sørensen Reference Mao and Sørensen2018; Hodgson, Madsen & Andersen Reference Hodgson, Madsen and Andersen2023; Wei et al. Reference Wei, El Makdah, Hu, Kaiser, Rival and Dabiri2024), harmonic yaw motion (Mühle et al. Reference Mühle, Tamaro, Klinger, Campagnolo and Bottasso2024) or with dynamic variation of a turbine’s thrust (Munters & Meyers Reference Munters and Meyers2018; Yılmaz & Meyers Reference Yılmaz and Meyers2018; Croce et al. Reference Croce, Cacciola, Montero Montenegro, Stipa and Praticó2023). Combined floating platform motion and rotor excitation strategies were simulated by van den Berg et al. (Reference van den Berg, De Tavernier and van Wingerden2023), showing promising results for the coupling of wake flow mixing strategies. Several studies suggest that an optimal actuation frequency, $ f_a$ , corresponding to $ St_a = ({f_a D}/{U_{\infty }}) \approx 0.3$ , leads to the most enhanced wake recovery and the emergence of large coherent structures (Cheung et al. Reference Cheung, Brown, Houck and Develder2024). This aligns with the findings of Crow & Champagne (Reference Crow and Champagne1971), who reported that $ St_a \approx 0.3$ induces the strongest instabilities in jets. The similarity suggests that common dynamics may govern both jet flows and wind turbine wakes, as both are characterised by strong shear layers.

The papers mentioned primarily investigated harmonic excitations with idealised flow conditions (i.e. laminar). Previous studies on the wake of a fixed cylinder (Kankanwadi & Buxton Reference Kankanwadi and Buxton2020) and fixed wind turbine (Neunaber et al. Reference Neunaber, Hölling, Stevens, Schepers and Peinke2020; Gambuzza & Ganapathisubramani Reference Gambuzza and Ganapathisubramani2023) analysed the impact of free-stream turbulence, characterised by the turbulence intensity ( $\textit{TI}_{\infty } =\sigma /U_{\infty }$ ) and integral length scale ( $\mathcal{L}_0$ ) on wake mixing and development. These studies show that with increasing $\textit{TI}_{\infty }$ , wake recovery is enhanced, due to improved mixing. The interaction between inflow turbulence and turbine harmonic excitation on wake dynamics has been studied numerically by Yılmaz & Meyers (2018), Korb et al. (Reference Korb, Asmuth and Ivanell2023), who found that periodic control strategies have a reduced impact on wake behaviour at higher turbulence intensities. Back to floating wind, the computational fluid dynamics (CFD) results from Li et al. (Reference Li, Dong and Yang2022), Li & Yang (Reference Li and Yang2024) further emphasise the sensitivity of wake dynamics and recovery to varying $\textit{TI}_{\infty }$ levels. The findings align with the observations of Yılmaz & Meyers (Reference Yılmaz and Meyers2018), Korb et al. (Reference Korb, Asmuth and Ivanell2023); namely, the effect of rotor motion on wake behaviour decreases relative to that of a fixed turbine as the intensity of the turbulence increases.

From the review of the literature, two key fluid dynamics questions for wind energy remain open: How is wake recovery driven by the large-scale flow structures in the wake of a periodically excited wind turbine and what is the interplay with increasing free-stream turbulence? In this paper we address these questions by means of wind tunnel testings in two facilities (Milan and Oldenburg). We conducted experiments with a model floating wind turbine, varying the rotor motion’s amplitude, frequency and direction of motion together with different inflow turbulence intensities ( $\textit{TI}_{\infty } \in [1.1,5.8] \,\%$ ) modulated by an active grid. Based on the rich literature from previous studies on canonical flows and the wake of fixed wind turbines, we provide an experimental proof of the speculated role of the excited coherent structures on wake recovery. We analyse their development and impact on recovery with different levels of turbulence in the free stream. The paper is organised as follows: § 2 details the experiments used and provides a theoretical basis for the analysis of coherent structures and wake recovery, § 3 presents the results in terms of wake recovery rate budget, recovery with increasing turbulence and motion-induced coherent structures shape and energy depending on the direction and amplitude of motion, § 4 summarises the main findings and § 5 discusses the results, draws parallels with previous work and concludes with their implications for wind energy.

2. Methodology

First, we describe the experimental set-up used for the tests in Milan and Oldenburg and detail the generation and type of inflows considered (§ 2.1). Second, we depict the platform motion cases investigated and wind turbine operating parameters (§ 2.2). Third, we provide a theoretical basis for the analysis of the wake data (§ 2.3).

2.1. Experimental set-up and inflows

In this study we carried out complementary measurements in two wind tunnel facilities: Milan (M), Italy ( $13.8\,\text{m} \times 3.8\,\text{m}$ section, length of $ 35\,\text{m}$ ) and Oldenburg (O), Germany ( $ 3\,\text{m} \times 3\,\text{m}$ section, length of $ 30\,\text{m}$ ) with a simplified model ‘floating’ wind turbine. The lab-scale FOWT used is composed of the model wind turbine Oldenburg, MoWiTO 0.6 with a diameter $D = 2R = 0.58$ m mounted on a motorised Stewart platform (see figure 1); further details on the model turbine and platform can be found in Messmer et al. (Reference Messmer, Brigden, Peinke and Hölling2022, Reference Messmer, Hölling and Peinke2024a ).

In Milan (figure 1 a) the model turbine was placed 25 m downstream of the inlet. The natural flow of the boundary layer test section is uniform at the rotor area, from bottom tip to top blade tip about $0.67\!-\!1.25$ m from the ground. The inflow is characterised by a background turbulence intensity $\textit{TI}_{\infty } = \sqrt {\overline {u^ {\prime 2}_{\infty }}} / U_{\infty } = 1.5 \pm 0.2 \,\%$ and an integral length scale of $\mathcal{L}_0/D = 0.2 \pm 0.05$ for $U_{\infty } = 5$ m s–1, referred to as inflow case M1.5. We computed the integral length scale in the $x$ direction, $\mathcal{L}_0$ as follows: $\mathcal{L}_0 = U_{\infty } \int _{0}^{\tau _0} R_{xx}(\tau ) {\rm d}\tau$ with $R_{xx}$ the auto-correlation function of $u'(t)$ in the free stream and $\tau _0$ the time lag at which $R_{xx}(\tau _0) = 0$ as done by Gambuzza & Ganapathisubramani (Reference Gambuzza and Ganapathisubramani2023).

We measured the wake of the wind turbine with two three-dimensional (3-D) hot wires, getting the three components of the wind speed, expressed in a Cartesian frame of reference (see figure 1 a): $\boldsymbol{u} = u_i \boldsymbol{e_i} = u \boldsymbol{e_x} + v \boldsymbol{e_y} + w \boldsymbol{e_z}$ (with $x$ pointing downstream). The two probes were installed on a motorised mounting that allows translation in the $y$ $z$ plane and that can be manually moved to different downstream positions, $x$ . With this measurement set-up, we recorded wake data behind the turbine at $x \in \{2, 4, 6, 8, 10\}D$ and different $y$ and $z$ positions (see the blue crosses in figure 2 a) with a $\Delta y$ and $\Delta z$ spacing between points of 0.34 $R$ . Data were recorded with two triaxial fibre-film Dantec Dynamics 55R91 probes. A temperature correction was applied on the calibration to account for temperature drift, as described in the sensor data sheet (temperature coefficient of resistance, TCR, about $0.5\,\%/{}^\circ\text{C}$ ). For each point, the measurement lasted 60 s and was sampled with a frequency $f_s$ of 6 kHz. In Milan, three motion cases were investigated; more details are given below.

Figure 1. Experimental set-up: MoWiTO 0.6 mounted on the 6-degrees of freedom (DoF) motorised platform. (a) Wind tunnel in Milan. (b) Wind tunnel in Oldenburg (with active grid).

Figure 2. (a) Measurement points $y$ - $z$ positions. (b) Picture of surge motion (fore-aft to streamwise). (c) Picture of sway motion (sideways to streamwise).

In Oldenburg (figure 1 b) the model turbine was placed in the closed test section of the wind tunnel behind an active grid mounted at the inlet. The model was placed at $d_{\textit{grid}\rightarrow \textit{turbine}}=$ 10 m and 15 m downstream of the inlet depending on the inflow cases (see table 1). Such long distances enable the development of a uniform inflow with a low level of turbulence. The active grid used for the experiments is described in more detail in Neuhaus et al. (Reference Neuhaus, Berger, Peinke and Hölling2021). It is composed of 80 independent motorised shafts on which flaps are mounted. Depending on the mean angle of the shaft, $\overline {\alpha }$ , rotational speed and the type of shaft motion, the free flow can be tailored to different $\textit{TI}_{\infty }$ and $\mathcal{L}_0$ . Motivated by real applications for offshore conditions where low to moderate turbulent inflows ( $\textit{TI}_{\infty } \leqslant 6 \,\%$ ) are found sufficiently frequently to be of relevance (Platis et al. Reference Platis, Hundhausen, Lampert, Emeis and Bange2021; Angelou, Mann & Dubreuil-Boisclair Reference Angelou, Mann and Dubreuil-Boisclair2023), we investigated cases with $\textit{TI}_{\infty } \in \sim [1, 6] \,\%$ and $\mathcal{L}_0 \leqslant 1.2D$ . However, in this study we do not analyse and discuss further the role of $\mathcal{L}_0$ on the results.

Table 1. Inflow cases investigated. Here M stands for Milan and O for Oldenburg.

The active grid was operated in passive mode, noted P (i.e no shaft motion but different mean shaft angles), and active mode, noted A (i.e with shaft motion). For the passive mode, we set a mean angle $\overline {\alpha }$ between $0^\circ \text{ and } 25^\circ$ . For the active mode, we used a coupled shaft motion protocol ensuring a constant global blockage of the grid at any time, with a mean shaft angle, $\overline {\alpha }$ of $30^\circ$ and different levels of fluctuations around the mean, characterised by $\text{std}(\alpha )$ . We characterised the inflows without the wind turbine model placed by using the same hot-wire array as for the wake measurements, described below. The passive grid modes enabled us to generate inflows with $\textit{TI}_{\infty } \lt 3\,\%$ and the active grid modes, inflows with $\textit{TI}_{\infty } \geqslant 3\,\%$ . All the inflows analysed in this paper are reported in table 1. Most of the tests were done at $U_{\infty }$ of 3 m s–1 except for the O1.5 case for which $U_{\infty } = 5$ m s–1, and is a duplicate of the flow from Milan experiments. Important to note is that the inflows generated were far enough from the grid to be fully developed, decaying and in a region where the decay is low so that between $x=0D$ and $x=10D$ a maximum of $20 \,\%$ of decay was observed, meaning that $\textit{TI}_{\infty }$ remain sufficiently constant while moving downstream.

We plot in figure 3 the $y$ profile of $\textit{TI}_{\infty }$ at hub height (figure 3 a) and the power spectrum of $u'_{\infty }$ , noted $\varPhi _x$ of the different inflows with $U_{\infty } = 3$ m s–1 (figure 3 b). The turbulence within the measurement area is homogeneous (see the low deviation of the different points to the averaged lines in figure 3 a) with a standard deviation, $\Delta TI_{\infty }$ of maximum $0.2 \,\%$ . The lowest $\textit{TI}_{\infty }$ generated is $1.1 \,\%$ (case O1.1) with the grid fully open and the highest is $5.8\,\%$ (case O5.8) obtained with a large amplitude of shaft motions, characterised by $\text{std}(\alpha ) = 15 ^{\circ }$ . The power spectra in figure 3(b) further describe the inflows with regard to the energy distribution across the various flow scales expressed in the frequency domain, $fD/U_{\infty }$ . All inflows feature a typical –5/3 Kolmogorov inertia subrange (Pope Reference Pope2001), i.e $\varPhi _x \propto f^{-5/3}$ for $f \in f_{\textit{sub}\hbox{-}\textit{range}}$ with different ranges depending on the case (see the dashed line in figure 3 b for the –5/3 slope). Case O1.1, for instance, has the shortest inertia subrange with $f_{\textit{sub}\hbox{-}\textit{range}}D/U_{\infty } \in \sim [2,5]$ whereas case O5.8 has the largest with $f_{\textit{sub}\hbox{-}\textit{range}}D/U_{\infty } \in \sim [0.2,10]$ . Since $\overline {u^{{\prime}2}_{\infty }} = \int _0^{\infty } \varPhi _x(f){\rm d}f$ , the higher $\textit{TI}_{\infty }$ cases naturally feature a larger level of energy. Last but not least, whereas passive modes generate inflow with a low level of turbulence intensity and low $\mathcal{L}_0$ , active modes lead to larger $\mathcal{L}_0$ up to 1.2D (see table 1).

Figure 3. (a) Horizontal profile of turbulence intensity at $x=0D$ (measurement points and mean $\textit{TI}_{\infty }$ plotted with solid lines). (b) Power spectrum of $u^{\prime}_{\infty }$ , the turbulent fluctuations of the inflow cases O1.1 to O5.8 in table 1 (with $U_{\infty } = 3$ m s–1).

In Oldenburg the wake measurements were carried out using 21 one-dimensional (1-D) hot wires (grey points in figure 2 a) aligned horizontally at hub height, measuring a good estimate of $u$ , the streamwise wind speed component. They were mounted on a motorised array enabling them to travel downstream and measure at different $x$ positions. The hot wire covered $y \in [-2.8, 2.8]R$ with a spacing of $\Delta y \approx 0.3R$ and a finer spacing of $\Delta y \approx 0.25R$ within $y \in [-R,R]$ . Measurements were taken at $x \in \{2,4, 6, 8, 10\}D$ for all cases except cases (fixed-5-O, surge-5-O, sway-5-O) hereafter in table 2 for which $x \in [1.5,10]D$ with $\Delta x = 0.5D$ were measured. Similar to the experiments discussed in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ), data were acquired with multichannel constant temperature anemometer (54N80-CTA) modules from Dantec Dynamics, with 2-minute measurements per case ( $T_{\textit{Meas}} = 120$ s) and $f_s = 6$ kHz. The hot wires were calibrated twice a day, with a Prandtl tube placed in the free stream as reference for $U_{\infty }$ . Similarly, a Prandtl tube was used to record $U_{\infty }$ during the wake measurements.

Table 2. Matrix of motion cases and inflows. Here DoF stands for degree of freedom.

2.2. Motion cases investigated and turbine parameter

The review of the literature (§ 1) outlined that rotor excitation, characterised by a Strouhal number $St_a = ({f_a D}/{U_{\infty }}) \approx 0.3$ , is often found to be the optimum excitation for wake mixing. Interestingly, when considering realistic motions of floating wind turbines, such as semi-submersibles or spar platforms, the natural period of rotational motion (fore-aft and sideways) often falls within a range of $ f_p \approx {1}/{40}$ Hz (Robertson et al. Reference Robertson, Jonkman, Masciola, Song, Goupee, Coulling and Luan2014). When this frequency is normalised using a typical offshore wind turbine diameter of $ D \approx 150$ m and an inflow velocity of $ U_{\infty } \approx 12$ m s–1, it results in a platform Strouhal number, $St$ of $\sim 0.3$ . This suggests that the ‘optimum’ wake excitation frequency naturally arises in real offshore conditions, reinforcing the relevance of the wake-mixing mechanism for floating wind turbines. The amplitude of platform motion, $ A_p$ , varies with environmental conditions but generally remains low, typically below a few degrees (Robertson et al. Reference Robertson, Jonkman, Masciola, Song, Goupee, Coulling and Luan2014). For small amplitudes ( $ A_p \leqslant 3^\circ$ ), we demonstrated in Messmer et al. (Reference Messmer, Peinke and Hölling2024a ) that rotational motions are equivalent to translational ones in terms of wake behaviour. Therefore, in this study we focus on single translational surge (fore-aft) and sway (sideways) motions. Figures 2(b) and 2(c) illustrate surge and sway motions, respectively. We consider idealised harmonic motion, characterised by a single reduced amplitude, $A^* = A_p/D$ , and Strouhal, $St$ , giving a motion signal, $ \xi (t)$ :

(2.1) \begin{equation} \xi (t) = A^* D \sin \left (2\pi \left (\frac {St U_{\infty }}{D}\right )t\right)\!. \end{equation}

In Milan we measured the wake for three motion cases: fixed turbine ( $St = 0$ , used as reference), and surge and sway motions with $St = 0.38$ , $A^* = 0.01$ , with a free-stream velocity of $U_{\infty } = 5$ m s–1. As concluded by Yılmaz & Meyers (Reference Yılmaz and Meyers2018), a rotor excitation frequency corresponding to $St = 0.38$ is effective for enhancing wake mixing, providing further motivation for the selected cases. The turbine was operated with a constant blade pitch and a tip speed ratio of $\lambda = ({R \omega }/{U_{\infty }}) = 6 \pm 0.1$ ( $\omega$ is the rotor rotational speed), close to the turbine optimum. The thrust coefficient was measured using a strain gauge mounted at the bottom of the tower, yielding $\overline {C_T} = 0.86 \pm 0.05$ . These tests are referred to as fixed-5-M, surge-5-M and sway-5-M in table 2.

In Oldenburg we repeated the same motion cases as in Milan under equivalent inflow conditions using the active grid to reproduce the free-stream turbulence ( $\textit{TI}_{\infty } = 1.5 \,\%$ ), and maintained the same turbine operating parameters.

In addition, we conducted further tests with $St = 0.3$ – identified in the literature as optimal for wake mixing – with three low motion amplitudes: $A^* \in \{0.01, 0.017, 0.024\}$ , with a wind speed $U_{\infty }$ of 3 m s–1. The choice of lower inflow velocity enabled the testing of higher $A^*$ values (bounded by the platform’s maximum physical amplitude $A_p$ at a given motion frequency $f_p$ , as discussed in Messmer et al. (Reference Messmer, Brigden, Peinke and Hölling2022)) while preserving comparable wake dynamics and recovery behaviour, as shown in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). This is explained by the sufficiently high Reynolds number, $\textit{Re} = D U_{\infty } / \nu = 1.4 \times 10^5$ (for $U_{\infty } = 3$ m s–1), which ensures that the wake behaviour is independent of $\textit{Re}$ for $U_{\infty } \geqslant 3$ m s–1.

For the $St = 0.3$ cases, we varied the inflow turbulence intensity, $\textit{TI}_{\infty }$ , from $1.1 \,\%$ (case O1.1) to $5.8 \,\%$ (case O5.8), to explore the combined effects of motion and turbulence on wake behaviour. The turbine was also operated at $\lambda = 6 \pm 0.1$ , and we measured a thrust coefficient of $\overline {C_T} = 0.80 \pm 0.05$ . All investigated cases are summarised in table 2.

In Kleine et al. (Reference Kleine, Franceschini, Carmo, Hanifi and Henningson2022) the authors analysed the effect of platform motion on the interaction between blade-tip flow structures and platform motion. They define the ratio between the platform motion frequency, $ f_p$ , and the rotor rotational frequency, $ \omega / 2\pi$ , as $ \omega ^* = ({2\pi f_p}/{\omega }) = ({\pi St}/{\lambda })$ . In our experiments, since $ \lambda = 6$ , we have $ \omega ^* \in \{0.16, 0.2\}$ for $ St \in \{0.3, 0.38\}$ . This indicates that the platform motion’s frequency is about five times slower than the rotor rotation frequency and lies below the range where the strongest interaction between motion and vortex structures in the near wake is expected, as described in Kleine et al. (Reference Kleine, Franceschini, Carmo, Hanifi and Henningson2022), namely $ \omega ^* \in \{0.5, 1.5\}$ .

The data collected in the two wind tunnels are complementary. The experiments conducted in Milan provide 3-D wake measurements, which are crucial for assessing the role of coherent structures in wake recovery mechanisms as depicted later on. However, these measurements are limited to a smaller set of motion and inflow conditions, as well as fewer downstream positions.

In contrast, the Oldenburg experiments offer 1-D wake field data at a greater number of downstream positions than those available from the Milan tests, for the three shared cases (fixed-5, surge-5, sway-5). This extended spatial coverage is essential for computing streamwise gradient terms (e.g. $\partial /\partial x$ ). Additionally, the Oldenburg data include a wider range of motion and inflow conditions, with simultaneous measurements enabling the reconstruction of the motion-induced large-scale wake structures.

2.3. Mathematical concepts on wake recovery and coherent structures

In this work we are interested in wake recovery as defined in Boudreau & Dumas (Reference Boudreau and Dumas2017), i.e the average wind speed in the wake on the rotor area. Since we mainly measure on a horizontal line at hub height, we define the recovery as

(2.2) \begin{equation} R_w(x) = \int _{y_c-R}^{y_c+R} \overline {u(x,y,t)} {\rm d}y / (D U_{\infty }) = \left \langle \overline {u} \right \rangle / U_{\infty }, \quad{\rm for}\quad y_c \approx 0,\end{equation}

where $y_c$ is the wake centre identified as the region where $\overline {u}$ is the lowest, in this paper for most cases $y_c \approx 0$ . Here $\left \langle \boldsymbol{\cdot} \right \rangle$ is the integrated value in the rotor area normalised by $D$ ; $R_w$ provides a good estimate of the wind available for a potential downstream wind turbine aligned with the one upfront generating the wake. The rate, $\partial R_w / \partial x$ , at which the wake recovers can be expressed directly by the Reynolds-averaged Navier–Stokes equation (RANS), as done for a fixed wind turbine by Bastankhah & Porté-Agel (Reference Bastankhah and Porté-Agel2016), Boudreau & Dumas (Reference Boudreau and Dumas2017), van der Laan, Baungaard & Kelly (Reference van der Laan, Baungaard and Kelly2023). The local RANS equation projected in $\boldsymbol{e_x}$ gives

(2.3) \begin{equation} \begin{aligned} \frac {D}{U_{\infty }} \frac {\partial \overline {u}}{\partial x} &= -\frac {D}{U_{\infty }} \frac {1}{\overline {u}} \left [ \frac {1}{\rho } \frac {\partial \overline {p}}{\partial x} + \overline {v} \frac {\partial \overline {u}}{\partial y} + \overline {w} \frac {\partial \overline {u}}{\partial z} + \frac {\partial \overline {u^{\prime } u^{\prime } }}{\partial x} + \frac {\partial \overline {u^{\prime } v^{\prime } }}{\partial y} + \frac {\partial \overline {u^{\prime } w^{\prime } }}{\partial z} \right ] \ \end{aligned} \end{equation}

after Boudreau & Dumas (Reference Boudreau and Dumas2017), we neglected the viscous term $1 / {\textit{Re}} \boldsymbol{\cdot}{{\nabla}}^2 \boldsymbol{u_i}\boldsymbol{\cdot}\boldsymbol{e_x}$ since $\textit{Re} \gg 1$ . The integration of (2.3) provides a decomposition of the contribution of the different terms to wake recovery. In this paper and based on our previous work (Messmer et al. Reference Messmer, Hölling and Peinke2024a ), we investigate platform motion cases that lead to the formation of a synchronised coherent structure at the motion frequency characterised by $St$ . In such cases, the wake flow field can be decomposed as (Reynolds & Hussain Reference Reynolds and Hussain1972; Baj, Bruce & Buxton Reference Baj, Bruce and Buxton2015; Lignarolo et al. Reference Lignarolo, Ragni, Scarano, Simão Ferreira and van Bussel2015)

(2.4) \begin{equation} u_i = \overline {u_i} + u^{\prime}_i = \overline {u_i} + \widetilde {u_i} + u^{\prime \prime }_i ,\end{equation}

where the velocity fluctuations $u^{\prime}_i$ are split up into $\widetilde {u_i}$ , the contribution from the coherent structure of $St$ (with a frequency $f_p$ ) and $u^{\prime \prime }_i$ , the stochastic turbulent contribution, i.e the velocity fluctuations without any periodic coherence. As done in Cantwell & Coles (Reference Cantwell and Coles1983), $\widetilde {u_{i}}$ is calculated by phase averaging the measured wind speed signals so that $\widetilde {u_{i}}(\phi (t)) = \widetilde {u_i}(\phi (t + 1/f_p))$ . Here, $\phi (t)$ is the instantaneous phase of the periodic wake motion, typically defined based on the known excitation frequency $f_p$ ,

(2.5) \begin{equation} \phi (t) = 2\pi f_p t + \phi _0, \end{equation}

where $\phi _0$ denotes the initial random phase shift. Throughout this paper, we use the normalised phase defined as $\phi ^* = (\phi (t) - \phi _0) \bmod 2 \pi$ , which maps the phase into a dimensionless range $\phi ^* \in [0, 2 \pi ]$ . Further details on the methodology used to compute $\widetilde {u_i}$ are provided in Appendix A. Per definition, $\overline {\widetilde {u_i} u^{\prime \prime }_i} = 0$ . Accounting for the decomposition of (2.4) in (2.3) writes as

(2.6) \begin{align} \frac {D}{U_{\infty }} \frac {\partial \overline {u}}{\partial x} &= -\frac {D}{U_{\infty }} \frac {1}{\overline {u}} \left [ \frac {1}{\rho } \frac {\partial \overline {p}}{\partial x} + \overline {v} \frac {\partial \overline {u}}{\partial y} + \overline {w} \frac {\partial \overline {u}}{\partial z} + \frac {\partial \overline {u^{\prime \prime } u^{\prime \prime } }}{\partial x} + \frac {\partial \overline {\widetilde {u} \widetilde {u}}}{\partial x} + \frac {\partial \overline {u^{\prime \prime } v^{\prime \prime } }}{\partial y} + \frac {\partial \overline {\widetilde {u} \widetilde {v}}}{\partial y}\right .\nonumber\\&\left . \quad + \frac {\partial \overline {u^{\prime \prime } w^{\prime \prime } }}{\partial z} + \frac {\partial \overline {\widetilde {u} \widetilde {w}}}{\partial z} \right] \!. \end{align}

The recovery rate $D \partial R_w / \partial x := D/U_{\infty } \langle \partial \overline {u} / \partial x \rangle$ is obtained from (2.6) by spatial integration of each term as done in (2.2). The recovery rate budget, i.e. the different physical mechanisms that contribute to the restoration of the mean flow velocity in the wake, is composed of the following terms.

  1. (i) The mean pressure gradient ( $-\partial \overline {p} / \partial x \lt 0$ ): the turbine induces a large pressure drop that restores primarily to momentum (Boudreau & Dumas Reference Boudreau and Dumas2017). This term can thus be viewed as a sink for the ‘recovery’ process.

  2. (ii) The mean speed transverse transport terms ( $-\overline {v} \partial \overline { u} / \partial y$ and $-\overline {w} \partial \overline {u} / \partial z$ ): wake rotation generates horizontal and vertical mean wind speed components, $\overline {v}$ and $\overline {w}$ , involved in momentum transport (Boudreau & Dumas Reference Boudreau and Dumas2017).

  3. (iii) The gradient of the Reynolds shear stress terms from incoherent fluctuations ( $-\partial \overline {u^{\prime \prime } v^{\prime \prime } / }\partial y$ etc.): incoherent turbulent fluctuations transport momentum and are the primary contributors to wake recovery (van der Laan et al. Reference van der Laan, Baungaard and Kelly2023; Gambuzza & Ganapathisubramani Reference Gambuzza and Ganapathisubramani2023).

  4. (iv) The gradient of the Reynolds shear stress terms from the coherent fluctuations ( $-\partial \overline {\widetilde {u} \widetilde {v}} / \partial y$ etc.): the organised structures may also contribute to wake recovery, though their role has not been extensively studied in previous research on wind turbine wakes, and is therefore addressed in this paper.

3. Results

First, in § 3.1 we present results from the shared cases between Milan and Oldenburg at $ \textit{TI}_{\infty } = 1.5\,\%$ (M1.5 and O1.5 inflow cases in table 1). Using these detailed datasets, we compute the terms of the recovery rate budget (2.6) and provide quantitative evidence for the role of coherent structures in enhancing wake recovery. Second, in § 3.2 we show how wake profiles and recovery vary with inflow turbulence intensity and platform motion amplitude, at a fixed Strouhal number of $ St = 0.3$ . Third, § 3.3 focuses on the coherent structures formed in the wake. We analyse their shape, strength and dependence on both $ \textit{TI}_{\infty }$ and the type of rotor motion. Fourth, based on these structures, § 3.4 explores their influence on phase-resolved wake recovery. Finally, § 3.5 examines the energy content of the coherent structures and analyses the dependency on inflow turbulence and platform motion energy and relation with increased recovery compared with the fixed case.

3.1. Wake recovery mechanisms

We begin the results section by presenting findings from the shared cases of the Milan and Oldenburg experiments. In Appendix B we detail the close agreement between the two datasets, as well as the methodology used to compute the various terms of (2.6), which supports the combined use of both datasets. The 3-D wake fields from Milan enable the computation of the $\partial /\partial y$ and $\partial /\partial z$ terms, while the 1-D measurements from Oldenburg, with their higher resolution in the streamwise direction, are used to compute the $\partial /\partial x$ terms. Details regarding the estimation of uncertainty in the derivative calculations are also provided in Appendix B.

For the three shared motion cases (fixed-5, surge-5 and sway-5; see table 2), figure 4(a) shows the evolution of wake recovery and figure 4(b) presents the normalised recovery rate, $D \partial R_w / \partial x$ . The wake of the moving turbine recovers more rapidly – with approximately $14 \,\%$ higher recovery for the sway case and $3 \,\%$ for surge compared with the fixed case at $x = 6D$ . Correspondingly, the recovery rate (figure 4 b) is higher for both sway and surge motions up to $x \leqslant 5D$ . This enhancement in wake recovery is less pronounced than in cases with laminar inflow, as reported by Li et al. (Reference Li, Dong and Yang2022) and Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). Compared with surge motion, sway leads to a $11 \,\%$ higher recovery, even though both cases share the same values of $St$ and $A^*$ . This result provides an initial indication that sway motion has a larger impact on wake recovery enhancement than surge motion – a point we investigate in more detail later in § 3.2.

Figure 4. (a) Wake recovery (2.2) for the fixed-5-O, surge-5-O and sway-5-O cases ( $St = 0.38$ , $A^* = 0.01$ ). Error bars represent the average discrepancy between the Milan and Oldenburg datasets (Appendix B). (b) Wake recovery rate. Error bars represent the estimated uncertainty based on a Monte Carlo approach (Appendix B). Both quantities are derived from the Oldenburg dataset.

In the following, we examine the mechanisms behind the enhanced wake recovery found with rotor motion. Figure 5 presents horizontal ( $y$ ) profiles of the individual terms in the recovery rate budget. At $x = 4D$ (figure 5 ac), the black dotted line highlights a larger value of $\partial \overline {u}/\partial x$ for the sway case (figure 5 c) compared with the fixed case (figure 5 a), particularly within $y \in [-0.5D, 0.5D]$ . While the fixed and surge cases still exhibit characteristics of a developing wake – evidenced by a double-Gaussian-like shape in $\partial \overline {u}/\partial x$ – the sway case presents a developed wake profile, already resembling the single-Gaussian distribution typical of the far wake (Neunaber et al. Reference Neunaber, Hölling, Stevens, Schepers and Peinke2020). This suggests that sway motion promotes a faster transition to the far-wake regime. By $x = 8D$ (figure 5 df), all cases display Gaussian-like $\partial \overline {u}/\partial x$ profiles, confirming that the far wake has been reached for all cases. These findings are consistent with observations under laminar inflow conditions reported by Messmer et al. (Reference Messmer, Hölling and Peinke2024a ), where platform motion was shown to accelerate wake development.

Returning to the profiles at $x = 4D$ , the primary contributor to the positive $\partial \overline {u}/\partial x$ appears to be the term $-(1/\overline {u}) \partial \overline {u'v'} / \partial y$ , i.e. the gradient of Reynolds shear stress in the $x$ $y$ plane. The gradient of $u'w'$ (the $x$ $z$ component) also contributes positively to wake recovery. As discussed in Boudreau & Dumas (Reference Boudreau and Dumas2017), van der Laan et al. (Reference van der Laan, Baungaard and Kelly2023), Gambuzza & Ganapathisubramani (Reference Gambuzza and Ganapathisubramani2023), these terms are known to dominate wake recovery processes. The negative regions of $\partial \overline {u}/\partial x$ observed at the edges of the wake ( $|y| \gt 0.5D$ ) act as sinks of recovery, supplying momentum toward the wake centre where $\partial \overline {u}/\partial x \gt 0$ , consistent with the interpretation of van der Laan et al. (Reference van der Laan, Baungaard and Kelly2023). The $y$ -rotational term, $-(1/\overline {u}) \overline {v} \partial \overline {u}/\partial y$ , is positive and, thus, contributes mathematically to $\partial \overline {u}/\partial x \gt 0$ ; whereas the other rotational term, $-(1/\overline {u}) \overline {w}\partial \overline {u}/\partial z$ is negative. The sum of the two appears to be counterbalanced by the remaining residual, which is mathematically equivalent to the pressure gradient term. Finally, figure 5(df) shows that at $x = 8D$ the recovery rate is higher for the fixed case (figure 5 d) than for the sway case (figure 5 f), in agreement with the results shown previously in figure 4(b). At this downstream location, all cases have transitioned to the far-wake regime, where recovery is primarily driven by gradients in Reynolds shear stresses in both the $x$ $y$ and $x$ $z$ planes.

Figure 5. The $y$ profiles of the terms in the recovery rate budget (2.6) for the fixed-5 (a,d), surge-5 (b,e) and sway-5 (c,f) cases. Panels (ac) correspond to $x = 4D$ and panels (df) to $x = 8D$ . In the legend, (M) indicates terms computed using Milan data, while (O) indicates those based on Oldenburg data.

To provide a broader view, we integrate all terms of the recovery rate budget (2.6) over the rotor diameter, following the same approach used in (2.2) for $\overline {u}$ . This integration yields a quantitative assessment of each term’s contribution (rotation, pressure, Reynolds stresses) to the overall recovery rate. Figure 6 presents the downstream evolution of these integrated contributions. Sway motion leads to a substantial increase in the Reynolds stress gradient terms in both the streamwise – spanwise and streamwise–vertical planes – up to $65 \,\%$ higher than in the fixed case (blue $\blacklozenge$ symbols in figure 6(a,c). In contrast, surge motion shows only a modest enhancement of these terms. Meanwhile, mean flow transport terms (associated with $\overline {v}$ and $\overline {w}$ ) decay more rapidly for sway – see the yellow $\blacksquare$ symbols at $x = 6D$ in figure 6(c) – compared with fixed and surge cases. The turbulent transport in the $x$ direction (grey $\times$ ) acts as a mild sink of recovery in all cases.

Notably, the residual term (green $\blacktriangle$ ) emerges as the dominant sink, suggesting that a large portion of the momentum rate budget is associated with pressure restoration or rotation-induced pressure effects. Noriega & Mazellier (Reference Fuentes Noriega and Mazellier2025) showed, using a porous disk with tilted blades, that wake rotation can induce a low-pressure core in the wake. This low-pressure region likely counteracts the potential benefits of rotational effects on wake recovery. In fact, between $x = 4D$ and $8D$ , the rotation contribution (yellow $\blacksquare$ ) appears nearly balanced by the residual, potentially indicating a negligible net impact on recovery.

Most importantly, for $x \in [2D, 6D)$ , the enhanced recovery observed for sway primarily originates from the increased Reynolds stress gradient terms: $-\langle (1/\overline {u}) \partial \overline {u' u^{\prime}_i} /\partial x_i\rangle$ with $i \in {y,z}$ . These terms represent turbulent transport in the $x$ $y$ and $x$ $z$ planes, and have been consistently identified as key contributors to wake recovery in previous studies on fixed wind turbines (Bastankhah & Porté-Agel 2016; Boudreau & Dumas Reference Boudreau and Dumas2017; Gambuzza & Ganapathisubramani Reference Gambuzza and Ganapathisubramani2023; van der Laan et al. Reference van der Laan, Baungaard and Kelly2023). Finally, for $x \gtrsim 8D$ , the far-wake regime is reached, where the recovery rate $D \partial R_w / \partial x$ is well approximated by the combined Reynolds shear stress gradients, i.e. $D \partial R_w / \partial x \approx -\langle (1/\overline {u}) (\partial \overline {u' v'} / \partial y + \partial \overline {u' w'} / \partial z ) \rangle$ where rotation and pressure terms are zero (Gambuzza & Ganapathisubramani Reference Gambuzza and Ganapathisubramani2023).

Figure 6. Downstream evolution of recovery equation terms (2.6) integrated over the rotor diameter for fixed-5 (a), surge-5 (b) and sway-5 (c). Panels (df) show $x$ $y$ Reynolds shear stress gradients; panels (gi) show $x$ $z$ gradients for the same cases.

For a clearer insight into the underlying processes, we next explore the decomposition of the Reynolds shear stresses. In the $x{-}y$ plane the total term is expressed as $-\langle 1/\overline {u} \partial \overline {u' v'}/ \partial y \rangle = - \langle 1/\overline {u} \partial \overline {\widetilde {u} \widetilde {v}} / \partial y \rangle - \langle 1/\overline {u} \partial \overline {u'' v''} / \partial y \rangle$ (see figure 6 d–f) and in the $x{-}z$ plane it is written as $- \langle 1/\overline {u} \partial \overline {u' w'}/ \partial z \rangle = - \langle 1/\overline {u} \partial \overline {\widetilde {u} \widetilde {w}} / \partial z \rangle - \langle 1/\overline {u} \partial \overline {u'' w''} / \partial z \rangle$ (see figure 6 g–i). Here, the tilde terms $\overline {\widetilde {u} \widetilde {v}}$ and $\overline {\widetilde {u} \widetilde {w}}$ represent the contributions from the coherent structures induced by platform motion at a given $St$ , while the double-prime terms $\overline {u'' v''}$ and $\overline {u'' w''}$ capture the contributions from purely stochastic turbulence.

Figure 6(f) shows that the total gradient of the Reynolds shear stress in the $x{-}y$ plane is significantly enhanced for the sway case compared with the fixed case (compare the blue $\blacklozenge$ symbols in figures 6 d and 6 f). Surprisingly, for sway and $3D \lesssim x \lesssim 10D$ , more than half of the total $x{-}y$ term is covered by $\left \langle 1/\overline {u} \partial \overline {\widetilde {u} \widetilde {v}} / \partial y \right \rangle$ , i.e the coherent structure contribution. In contrast, the $x{-}z$ plane (figure 6 i) shows both a lower overall Reynolds shear stress gradient and a relatively smaller contribution from the coherent structure subterm. Although surge motion (figures 6 e and 6 h) also exhibits contributions from coherent structures in both the $x{-}y$ and $x{-}z$ planes, these are much lower than those observed for sway, reflecting its reduced impact on wake recovery.

These results conclusively show that the coherent structures induced by platform motion play a critical role in enhancing wake recovery. Consistent with previous findings on fixed wind turbines (Boudreau & Dumas Reference Boudreau and Dumas2017; van der Laan et al. Reference van der Laan, Baungaard and Kelly2023), our study shows that the gradients of Reynolds shear stress in the $x{-}y$ and $x{-}z$ planes are central to the recovery process. In the sway case, the improved recovery – evidenced by the higher $D \partial R_w / \partial x$ in figure 4(a) – is directly linked to increased Reynolds stress gradient terms: $- \langle 1/\overline {u} \partial \overline {u' v'}/ \partial y \rangle - \langle 1/\overline {u} \partial \overline {u' w'}/ \partial z \rangle$ .

Notably, a substantial portion of these gradients is attributable to the contributions of coherent structures – an observation that supports our earlier speculation in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). Previous studies, such as Yılmaz & Meyers (Reference Yılmaz and Meyers2018), Korb et al. (Reference Korb, Asmuth and Ivanell2023), Van der Hoek et al. (Reference van der Hoek, den Abbeele, Simao Ferreira and van Wingerden2024), have similarly linked the presence of coherent structures from periodic excitations to mechanisms like wake deflection and enhanced momentum entrainment. In our work, the effect of these mechanisms on recovery are quantitatively represented by the two terms: $-\langle 1/\overline {u} \partial \overline {\widetilde {u} \widetilde {v}}/ \partial y \rangle - \langle 1/\overline {u} \partial \overline {\widetilde {u} \widetilde {w}}/ \partial z \rangle$ .

Later in the paper we further decompose wake recovery by distinguishing the effects of periodic wake deflection and enhanced momentum entrainment, both of which are influenced by periodic structures, as discussed in Korb et al. (Reference Korb, Asmuth and Ivanell2023). These factors are, to some extent, hidden in our recovery rate budget analysis.

3.2. Wake profiles and recovery with increasing free-stream turbulence

So far, we have investigated wake recovery for a specific inflow condition with $\textit{TI}_{\infty } = 1.5 \,\%$ and fixed values of $St$ and $A^*$ . Next, we present the results on wake profiles and recovery for a range of inflows with different $\textit{TI}_{\infty }$ ( $\textrm{O}1.1\!-\!\textrm{O}5.8$ in table 1) and for various motion cases with $St = 0.3$ and $A^* \in \{0.01,0.017,0.024\}$ (see table 2). As detailed in § 2.1, the active grid at the University of Oldenburg was used to achieve $\textit{TI}_{\infty }$ levels in the range $[1.1, 5.8] \,\%$ . These experiments allow us to assess the combined effects of increasing inflow turbulence and platform motion on wake development and recovery.

Figure 7. The $y$ profiles of normalised wake deficit (a) and turbulence intensity (b) for fixed, surge- and sway- 3–0.01 and 0.024, i.e $St = 0.3, A^* = 0.01,\,0.024$ (see table 2) for O1.1 (a–0)–(a–4), O3.0 (b–0)–(b–4), O5.8 (c–0)–(c–4) and $x \in [2,10]D$ (0,4). Inflows in table 1.

We present in figure 7 the lateral ( $y$ ) profiles of the normalised wake deficit $\Delta u^* = (U_{\infty } - \overline {u}) / U_{\infty }$ (figure 7 a) and turbulence intensity $ TI = \sqrt {\overline {u^{\prime 2}}}/\overline {u}$ (figure 7 b). The data are organised as follows.

  1. (i) Three inflow conditions, with $\textit{TI}_{\infty } \in \{1.1, 3.0, 5.8 \} \,\%$ , corresponding to cases O1.1, O3.0 and O5.8 from table 1), shown respectively in (a–0)–(a–4), (b–0)–(b–4), (c–0)–(c–4).

  2. (ii) Two motion amplitudes per motion type ( $A^*=0.01,\,0.024$ ) for both surge and sway, with $St = 0.3$ .

  3. (iii) Five downstream locations, $x \in [2,10]D$ denoted (.–0)–(.–4), where ‘.’ stands for a, b or c.

Figure 7(a) (a–0), (b–0) and (c–0) reveal that, in the near wake at $x = 2D$ , the wake deficit profiles are nearly identical across all motion cases for a given inflow condition. This observation, consistent with earlier findings by Fontanella, Zasso & Belloli (Reference Fontanella, Zasso and Belloli2022) and Messmer et al. (Reference Messmer, Hölling and Peinke2024a ) (in low turbulent inflows, $\textit{TI}_{\infty } \lt 2 \,\%$ ), suggests that the mean induction, i.e. the average slowing of the flow by the turbine, is effectively unchanged between cases. In other words, $\overline {C_T}$ is constant across motion cases. This equivalence is crucial as it ensures that any observed differences in wake evolution further downstream can be attributed to platform motion effects rather than discrepancies in the initial near-wake, rotor-induction conditions.

As we move downstream, figure 7(a) (a–0)–(a–4) illustrate the pronounced impact of platform motion on the wake deficit under low inflow turbulence ( $\textit{TI}_{\infty } = 1.1 \,\%$ ). Both surge and sway cases show enhanced wake recovery relative to the fixed case, with the most significant effect observed for sway motion at the highest amplitude ( $A^* = 0.024$ , green $\times$ ). This case produces a much broader and more uniform velocity deficit profile compared with the fixed case (black dots), indicating that sway motion promotes a wider wake and accelerates lateral wake expansion. Notably, by $x = 10D$ , the wake profile for this sway case deviates from the classic Gaussian shape, potentially challenging conventional analytical wake models (e.g. the Bastankhah & Porté-Agel (Reference Bastankhah and Porté-Agel2014) wake model). In contrast, the lower amplitude case (green triangle) yields a wake profile more similar to the fixed case, suggesting a weaker influence on wake growth. Comparing the two motion directions, sway induces more noticeable changes in the wake mean structure than surge, which remains closer to the fixed reference across all downstream positions.

As inflow turbulence increases, the influence of motion diminishes. At $\textit{TI}_{\infty } = 3.0 \,\%$ , the sway-induced differences are still visible (figure 7 a(b–2)–(b–4)), but to a lesser extent. For surge motion, however, this turbulence level appears to mark a threshold – no discernible difference is observed between the surge and fixed cases, even at the highest motion amplitude. Finally, under higher turbulence ( $\textit{TI}_{\infty } = 5.8 \,\%$ ), all wake deficit profiles converge, indicating that inflow turbulence dominates the wake mean profiles and suppresses any motion-induced effects.

Concordantly, the differences in the turbulence intensity profiles between the motion cases and the fixed case are most pronounced under low inflow turbulence ( $\textit{TI}_{\infty } = 1.1\,\%$ ), as shown in figure 7(b) (a–0)–(a–4).

Surge motion primarily increases the level of turbulence intensity within the rotor region ( $|y| \lt 0.5D$ ); for instance, at $x = 4D$ , the surge case with $A^* = 0.024$ (blue squares) shows a pronounced increase in turbulence intensity compared with the fixed case (black dots), as seen in figure 7(b) (a–1). This elevated turbulence accelerates the transition to the far-wake regime, as detailed in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). The influence of motion amplitude is clear: the lower amplitude case ( $A^* = 0.01$ ) induces only a slight increase in turbulence intensity, consistent with its weaker effect on the wake deficit seen in figure 7(a) (a–0)–(a–4).

Sway motion, by contrast, predominantly enhances turbulence intensity at the edges of the wake ( $|y| \gt 0.5D$ ), particularly at downstream locations ( $x \in [8D, 10D]$ ); see green crosses in figures 7(b) (a–3) and 7(b) (a–4). Similar to the trends observed for $\Delta u^*$ , lateral platform motion results in a broader and more uniform turbulence intensity distribution, with a lower intensity in the wake centre and a higher intensity at the edges – indicating a wider and more dispersed wake for the larger amplitude sway case.

As $\textit{TI}_{\infty }$ increases, the influence of platform motion on wake turbulence intensity diminishes. At $\textit{TI}_{\infty } = 3.0\,\%$ , surge motion with $A^* = 0.024$ leads to a moderate increase in near-wake turbulence for $x \leqslant 4D$ (figure 7 b (b-0)–(b-1)). Sway motion at this turbulence level still causes slight wake broadening in the far wake ( $x \geqslant 8D$ ), though the effect is noticeably reduced compared with the lower $\textit{TI}_{\infty }$ case (figure 7 b (b–3)–(b–4)). For both surge and sway, the lower amplitude case ( $A^* = 0.01$ ) shows no discernible effect on turbulence intensity profiles when $\textit{TI}_{\infty } \geqslant 3\,\%$ .

Finally, at the highest inflow turbulence ( $\textit{TI}_{\infty } = 5.8\,\%$ ), platform motion has negligible impact on turbulence intensity across all downstream positions beyond $x = 2D$ (figure 7 b (c–1)–(c–4)).

This analysis of the turbulence intensity profiles, in agreement with the wake deficit observations, confirms that the effect of platform motion weakens as inflow turbulence increases. This trend aligns well with previous CFD studies by Yılmaz & Meyers (Reference Yılmaz and Meyers2018) and Li et al. (Reference Li, Dong and Yang2022).

We next quantify the differences between the fixed case and the motion cases for multiple inflow turbulence levels $\textit{TI}_{\infty }$ and all amplitudes of motion. To do so, we compute the recovery using (2.2) for the fixed case ( $St = 0$ ), as well as for surge and sway motions at $St = 0.3$ with $A^* \in \{0.01, 0.017, 0.024\}$ (table 2). Figure 8(af) shows the results for the different inflow turbulence conditions $\textit{TI}_{\infty } \in \{1.1, 2.2, 3.0, 4.1, 5.0, 5.8\}\,\,\%$ .

As noted earlier in the wake deficit profiles (figure 7 a), for a given inflow condition, all cases – fixed or with platform motion – exhibit nearly identical wake recovery at $x = 2D$ . The effect of platform motion on the wake develops progressively farther downstream, driven by nonlinear interactions (Messmer et al. Reference Messmer, Hölling and Peinke2024a ). Between the different inflow cases, $R_w(x=2D)$ varies slightly but remains close to $R_w \approx 0.5$ .

The clearest differences between motion cases appear for the lowest free-stream turbulence level $\textit{TI}_{\infty } = 1.1\,\,\%$ (figure 8 a), consistent with the wake profiles in figure 7(a). Sway motion with $A^* = 0.024$ improves recovery by up to $21 \,\%$ compared with the fixed case. For both surge and sway, the impact is less with decreasing motion amplitude; for example, sway with $A^* = 0.01$ yields a $12 \,\%$ increase in recovery over the fixed case. Interestingly, even the largest surge amplitude results in a weaker recovery enhancement than sway motion at the lowest amplitude.

As $\textit{TI}_{\infty }$ increases (figure 8 bf), the influence of platform motion diminishes. At $\textit{TI}_{\infty } = 2.2\,\,\%$ (figure 8 b), surge with $A^* = 0.024$ enhances recovery by only $3.5 \,\%$ , while sway still yields an $11 \,\%$ increase. Sway at $A^* = 0.01$ continues to improve recovery by $4 \,\%$ , whereas surge at the same amplitude has no observable effect. At $\textit{TI}_{\infty } = 3.0\,\,\%$ (figure 8 c), surge and fixed recovery curves overlap entirely, while sway continues to provide a $ 5 \,\%$ improvement. The effect of sway becomes almost negligible around $\textit{TI}_{\infty } = 5.0\,\,\%$ , and at $\textit{TI}_{\infty } = 5.8\,\,\%$ all cases, fixed, surge and sway, converge.

Figure 8 highlights three important trends. First, sway motion has a stronger influence on wake recovery than surge, a result we further quantify and discuss later in this paper. Second, larger motion amplitudes lead to greater recovery improvements, as also pointed out by Li et al. (Reference Li, Dong and Yang2022), Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). Third, the effect of platform motion does not add linearly to the effect of inflow turbulence; rather, the two blend. To highlight this point, we present in figure 9 the downstream evolution of offset-adjusted wake recovery, denoted as $ R^*_w$ , for the fixed case (figure 9 a) and for the sway case with $ A^* = 0.024$ (figure 9 b), across all levels of inflow turbulence, $ \textit{TI}_{\infty } \in [1.1, 5.8]\,\%$ .

We define the adjusted recovery as

(3.1) \begin{align} R^*_w(x) = R_w(x) - R_w(x=2D) + 0.5 \end{align}

in order to align all curves at a common starting point of 0.5 at $ x = 2D$ , thereby removing small variations in $ R_w(x=2D)$ that arise from minor dependencies on $ \textit{TI}_{\infty }$ .

Figure 8. Wake recovery from (2.2) for different free-stream turbulence levels: $\textit{TI}_{\infty } = 1.1\,\% \!-\!5.8\,\%$ (cases O1.1–O5.8 in table 1). Results include fixed ( $St = 0$ ), surge and sway motions at $St = 0.3$ with $A^* \in \{0.01, 0.017, 0.024\}$ (table 2).

Figure 9. Offset-adjusted wake recovery: $R^*_w(x) = (R_w(x) - R_w(x=2D))+0.5$ ( $R_w$ defined in (2.2) for different free-stream turbulence levels: $\textit{TI}_{\infty } = 1.1\,\%\!-\!5.8\,\%$ (cases O1.1–O5.8 in table 1). Panel (a) shows the fixed case ( $St = 0$ ) and panel (b) sway with $St = 0.3$ and $A^* = 0.024$ (table 2).

In the fixed case, wake recovery increases with inflow turbulence. For example, $ R_w(x=10D)$ increases from approximately 0.7 at $ \textit{TI}_{\infty } = 1.1\,\%$ (dark blue in figure 9 a) to about 0.83 at $ \textit{TI}_{\infty } = 5.8\,\%$ (light yellow in the same figure), in agreement with trends reported in the literature, such as Bastankhah & Porté-Agel (Reference Bastankhah and Porté-Agel2014). This confirms the general observation that higher levels of free-stream turbulence lead to faster wake recovery.

In contrast, for the sway case with $ A^* = 0.024$ (figure 9 b), the recovery appears largely independent of inflow turbulence: $ R_w(x=10D) = 0.83 \pm 0.01$ across all $ \textit{TI}_{\infty }$ levels. A slight effect of turbulence is visible around $ x = 4D$ , where higher turbulence modestly enhances recovery; however, this difference diminishes farther downstream. This suggests that while higher $ \textit{TI}_{\infty }$ can slightly accelerate the onset of wake recovery, the influence of platform motion is substantially stronger. Remarkably, platform motion at low $ \textit{TI}_{\infty }$ produces recovery levels equivalent to those achieved with high turbulence in the fixed case, emphasising the dominant role of platform movements in accelerating wake recovery. This effect is examined next in relation to motion-induced coherent structures.

3.3. Motion-induced coherent structures with increasing $\textit{TI}_{\infty }$

In figure 6 we found that coherent structures formed in the wake due to platform motion contribute to enhanced recovery by increasing the gradients of Reynolds shear stresses in the $x{-}y$ and $x{-}z$ planes. However, as discussed in figures 8 and 9, this improvement diminishes with increasing $\textit{TI}_{\infty }$ . To investigate the connection between this trend and the evolution of coherent structures at higher turbulence intensities, we examine the phase-resolved wake dynamics. The Oldenburg experiments, which use an array of 21 hot wires to simultaneously measure the streamwise velocity along a vertical line, enable the computation of $\widetilde {u}(\phi ^*)$ at fixed streamwise positions. The measurements span $y \in [-1.4, 1.4]D$ and are synchronised – meaning that for a given $x$ , all probes share the same initial phase $\phi _0$ .

Figure 10. (a) Sum of the mean wake flow and the surge-induced coherent fluctuation. (b) Surge-induced coherent fluctuation alone. All velocities are normalised by the free-stream velocity $U_{\infty }$ . Columns (- .0)–(- .4) correspond to streamwise positions $x/D \in \{2, 4, 6, 8, 10\}$ . Each panel represents one full period of the motion, shown over the normalised phase $\phi ^* \in [0, 2\pi ]$ . Rows correspond to different inflow turbulence levels: (a.x) $\textit{TI}_{\infty } = 1.1\,\%$ , (b.x) $\textit{TI}_{\infty } = 3.0\,\%$ and (c.x) $\textit{TI}_{\infty } = 5.8\,\%$ , corresponding to cases O1.1, O3.0 and O5.8 in table 1. The motion parameters are $St = 0.3$ and $A^* = 0.024$ (case surge-3-0.024 in table 2).

We begin by analysing surge-induced periodic structures, illustrated in figure 10 for platform motion characterised by $St = 0.3$ and $A^* = 0.024$ (case surge-3-0.024 in table 2). Figure 10(a) shows contour plots of the combined mean and coherent streamwise velocity $(\overline {u} + \widetilde {u})/U_{\infty }$ , while figure 10(b) presents the coherent fluctuation component $\widetilde {u}/U_{\infty }$ alone. Each panel (a.-), (b.-) and (c.-) corresponds to a different inflow turbulence intensity, $\textit{TI}_{\infty } \in \{1.1, 3.0, 5.8\}\,\%$ – denoted as O1.1, O3.0 and O5.8 in table 1 – and spans one full oscillation cycle ( $\phi ^* \in [0, 2\pi ]$ ) at downstream positions $x \in [2,10]D$ . At low turbulence ( $\textit{TI}_{\infty } = 1.1\,\%$ ), the combined velocity field shows a distinct periodic pattern of alternating bulges and constrictions, especially between $x \in [6,8]D$ (see panels a.2–a.3 in figure 10 a). The corresponding fluctuation fields in figure 10(b) (a.0)–(a.4) display in-phase red and blue bands – indicative of in-phase alternating speed-up and slow-down – characterising a pulsating wake. This structure is analogous to the varicose mode observed in a planar wake excited periodically in Wygnanski, Champagne & Marasli (Reference Wygnanski, Champagne and Marasli1986). The pulsating structure is likely resulting from symmetric vortex roll-up in the shear layer due to the platform’s surging motion. Similar patterns have been reported for thrust-excited and surging floating turbines in the studies of Yılmaz & Meyers (Reference Yılmaz and Meyers2018), Messmer et al. (Reference Messmer, Hölling and Peinke2024a ), Li & Yang (Reference Li and Yang2024), Li et al. (Reference Li, Yu and Sarlak2024), Wei et al. (Reference Wei, El Makdah, Hu, Kaiser, Rival and Dabiri2024), Hubert et al. (Reference Hubert, Conan and Aubrun2025).

For $\textit{TI}_{\infty } = 1.1\,\%$ , coherent fluctuations reach amplitudes up to $15 \,\%$ of the free-stream wind speed, underscoring the strong influence of platform motion under low turbulence conditions. The downstream evolution of the mode shows an initial low amplitude near $x = 2D$ , which increases up to a peak around $x = 6D$ before decaying further downstream. This trend is consistent with observations at a fixed lateral position, which we reported in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). As $\textit{TI}_{\infty }$ increases, the amplitude of coherent fluctuations decreases substantially. At $\textit{TI}_{\infty } = 3.0\,\%$ , the pulsating pattern remains visible (see figure 10 b (b.0)–(b.4)) but is less pronounced, with amplitudes reduced by a factor of approximately three. At the highest turbulence level ( $\textit{TI}_{\infty } = 5.8\,\%$ ), coherent structures are barely discernible, with amplitudes nearly six times lower than in the low turbulence case. At this level of inflow turbulence, the development of coherent wake structures is significantly inhibited, consistent with the CFD analysis of Yılmaz & Meyers (Reference Yılmaz and Meyers2018).

Figure 11. (a) Sum of the mean wake flow and the sway-induced coherent fluctuation. (b) Sway-induced coherent fluctuation alone. All velocities are normalised by the free-stream velocity $U_{\infty }$ . Columns (-.0)–(-.4) correspond to streamwise positions $x/D \in \{2, 4, 6, 8, 10\}$ . Each panel represents one full period of the motion, shown over the normalised phase $\phi ^* \in [0, 2\pi ]$ . Rows correspond to different inflow turbulence levels: (a.x) $\textit{TI}_{\infty } = 1.1\,\%$ , (b.x) $\textit{TI}_{\infty } = 3.0\,\%$ and (c.x) $\textit{TI}_{\infty } = 5.8\,\%$ , corresponding to cases O1.1, O3.0 and O5.8 in table 1. The motion parameters are $St = 0.3$ and $A^* = 0.024$ (case sway-3-0.024 in table 2).

An equivalent analysis can be performed for sway motion. Figure 11 presents results for sideways platform motion characterised by $St = 0.3$ and $A^* = 0.024$ (case sway-3-0.024 in table 2), showing the same inflow conditions as for surge in figure 10. In contrast to surge, sway-induced platform motion causes strong lateral displacement of the wake, especially pronounced at $\textit{TI}_{\infty } = 1.1\,\%$ (see figure 11 b (a.0)–(a.4)). The phase-averaged streamwise wake velocity, combined with the mean flow, exhibits a sinusoidal shape that causes the entire wake to oscillate laterally. This lateral motion contributes to the increased wake spreading observed in figure 7. For example, at $x = 8D$ , the side-to-side amplitude of wake motion reaches approximately one rotor diameter, as shown by the $y$ displacement between the wake centre at $\phi ^* = 0$ and $\phi ^* = \pi$ in figure 11(a) (a.3). This illustrates how the relatively small lateral rotor motion ( $A_p = 0.024D$ ) is strongly amplified in the wake – a phenomenon we previously discussed in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ) in terms of nonlinear dynamics.

Interestingly, the shape of the phase-averaged wake bears a resemblance to the flow past a cylinder undergoing vortex shedding, as described by Cantwell & Coles (Reference Cantwell and Coles1983), and also to the flow behind high-solidity porous disks (see the smoke visualisations in Cannon, Champagne & Glezer (Reference Cannon, Champagne and Glezer1993)). The periodic lateral wake motion shares similarities with the sinuous mode identified by Wygnanski et al. (Reference Wygnanski, Champagne and Marasli1986) in planar wakes and the coherent structures found in excited jets (Crow & Champagne Reference Crow and Champagne1971). These analogies suggest that many shear flows exhibit an intrinsic sinuous mode, which, when excited at the appropriate frequency or under favourable conditions (e.g. low turbulence and strong shear layers), can develop into large-scale coherent structures due to shear layer instabilities – even in the presence of background turbulence, as observed in our case. In the context of wind energy, this phenomenon is referred to as coherent wake meandering and studied in detail in Mao & Sørensen (Reference Mao and Sørensen2018), Li et al. (Reference Li, Dong and Yang2022). In our case, the meandering motions are triggered by the lateral turbine movements.

Accordingly – and in contrast to the surge-induced structures – the meandering patterns in figure 11(b) (a.0)–(a.4) exhibit anti-phase velocity fluctuations: $\widetilde {u}(\phi ^* = \pi , y = -R) \approx -\widetilde {u}(\phi ^* = \pi , y = R)$ . The amplitude of these streamwise meandering fluctuations reaches up to 20 % of the free-stream velocity. As with surge, increasing inflow turbulence reduces the magnitude of these coherent structures: at $\textit{TI}_{\infty } = 3.0\,\%$ amplitudes are roughly halved, while at $\textit{TI}_{\infty } = 5.8\,\%$ they are reduced by a factor of six. Similar to the results obtained for surge motion, the shape of the mode remains consistent across different turbulence intensities. Its magnitude increases with downstream distance, reaching a maximum between $x \in [4D, 6D]$ , before gradually decaying further downstream.

Later in the paper, we quantify the energy contained in these coherent structures and examine the relationship between their energy content, wake recovery enhancement and the energy input from platform excitation. A direct comparison between surge and sway motions is presented to highlight the differences in their respective impacts.

Prior to this, we address and quantify a frequently debated issue in wind energy: the effect of coherent fluctuations on the time evolution of phase-resolved wake recovery.

3.4. Coherent structures and phase-resolved wake recovery

It is often argued that the observed increase in wake recovery – computed using (2.2) and shown in figure 8 – in response to platform motion or rotor excitation could be an artefact of periodic wake displacement, as discussed by Korb et al. (Reference Korb, Asmuth and Ivanell2023). This is particularly relevant in the case of sway-induced coherent structures (figure 11 a), where the entire wake undergoes large-amplitude lateral oscillations. When such motion is averaged over time and across the rotor area, it can artificially smooth the wake profile, giving a feigned increased recovery. This raises an important question: Is the enhanced recovery a result of actual momentum exchange, or is it merely a consequence of averaging over periodic lateral displacements? In the following, we clarify this distinction and present evidence to evaluate the true contribution of coherent structures to enhance momentum exchange.

Figure 12(a) displays horizontal wind speed profiles in the wake at $x = 6D$ . Panels 12(a)(cg) show both the time-averaged wake profile and phase-resolved wake profiles for the surging turbine operating at $St = 0.3$ , $A^* = 0.024$ and $\textit{TI}_\infty = 1.1\,\%$ . For comparison, the wake of the fixed (non-oscillating) turbine is also included. Similarly, the wake profiles for the swaying turbine under identical operating conditions are shown in figure 12(a)(hl). We manually adjusted the phase origin $\phi _0$ such that, for $\phi ^* = 0$ , the phase-resolved wake aligns most closely with the time-averaged profile.

Figure 12. Comparison of time-averaged and phase-averaged wake recovery. (a) Time-averaged and local (phase-averaged) streamwise wind speed $y$ profiles at $x = 6D$ for selected phases $\phi ^* \in \{0, 2\pi , 4\pi , 6\pi , 8\pi \}/5$ , for both surge (ae) and sway (fj) motions, with $St = 0.3$ , $A^* = 0.024$ and $\textit{TI}_{\infty } = 1.1\,\%$ . (b) Local wake recovery $\widetilde {R_w}(\phi ^*)$ computed across the phase range $\phi ^*/2\pi \in [0,1]$ . The plot includes the mean of the phase-averaged recovery $\overline {\widetilde {R_w}(\phi ^*)}$ , the time-averaged recovery $R_w$ (also for the fixed reference case), shown for both surge (1) and sway (2) cases.

As $\phi ^*$ increases ( $\phi ^* \in [0, \pi ]$ ), the surging turbine’s wake initially contracts and exhibits higher wind speeds than the time-averaged case (see figure 12 a(ce)). To better illustrate the connection between the wake profiles and the coherent dynamics, we display in figure 12(a)(ce) the structure previously shown in figure 10(b)(a.2), and mark with a red thick line the position of $\phi ^*$ . Beyond $\phi ^* \gt \pi$ , the streamwise component of the coherent structure becomes negative, the wake expands, and the wind speed decreases below the time-averaged profile. Notably, at $\phi ^* = 8\pi /5$ , the wake speed is even lower than in the fixed case, i.e. $\overline {u}_{\textit{fixed}} \gt (\overline {u} + \widetilde {u})_{\textit{surge}}$ .

To quantify the effects of motion on recovery at different phases, we define the phase-resolved wake recovery $\widetilde {R_w}(\phi ^*)$ as

(3.2) \begin{equation} \widetilde {R_w}(\phi ^*) = \frac {1}{D U_\infty } \int _{-R + y_c(\phi ^*)}^{R + y_c(\phi ^*)} \big(\overline {u} + \widetilde {u}(\phi ^*) \big) \, {\rm d}y .\end{equation}

Here, $y_c(\phi ^*)$ is the lateral location of the wake centre at a given phase, defined as the position where $\overline {u} + \widetilde {u}$ reaches its minimum. For surge motion, this position remains approximately constant and centred: $y_c(\phi ^*) \approx 0$ .

Figure 12(b)(1) shows the evolution of $\widetilde {R_w}$ over a full motion cycle ( $\phi ^*/2\pi \in [0,\sim 1]$ ), alongside the average, $\overline {\widetilde {R_w}}$ , and the conventional time-averaged wake recovery from (2.2), both for the surging and fixed cases. In agreement with the profiles in figure 12(a)(cg), the phase-resolved recovery initially increases (red x-. curves) when $\widetilde {u} \gt 0$ for $\phi ^*/2\pi \in [0, 0.5]$ . Then, as the wake expands for $\phi ^*/2\pi \gt 0.5$ , the recovery decreases, even falling below that of the fixed case. Overall, $\overline {\widetilde {R_w}}$ (red dashed line) is only slightly higher (by $\sim 1 \,\%$ ) than the conventional time-averaged value (blue solid line). This indicates that, for surge motion, the coherent structures do not significantly alter the mean of the phase-averaged recoveries, although they do introduce large fluctuations depending on $\phi ^*$ . As a result, downstream turbines in the wake would experience substantial power variability.

In contrast, the wake of the swaying turbine behaves differently. As seen in figure 12(a)(hl), the wake undergoes significant lateral movement, with the wake centre $y_c(\phi ^*)$ varying between $-R$ and $R$ . This behaviour, driven by the meandering mode illustrated in figure 11(b)(a.2) and reproduced in figure 12(a)(hl), causes the phase-averaged wake profile to shift side to side rather than contract or expand. Unlike the surge case, the minimum of $\overline {u} + \widetilde {u}$ remains almost constant across phases, but is slightly lower than that of the time-averaged wake alone. This suggests that the smoothed time-averaged profile (green solid line in figure 12(a)(hl)) also includes a contribution from the phase displacements rather than solely from enhanced momentum transport.

Another important difference lies in the shape of the phase-resolved wake. For the surging turbine, the wake retains a near-Gaussian profile even locally, whereas the swaying turbine’s wake becomes asymmetric at different phases – indicating a distortion caused by lateral motion. This further highlights the distinct nature of wake modulation between meandering and pulsating modes.

In figure 12(b)(2) we display the phase-averaged recovery for sway motion at various $\phi ^*$ (orange squares), along with the average over all the phases, the time-averaged recovery and the fixed case for reference. Compared with the surge case, the variation in phase-resolved recovery is smaller ( ${\sim}6\,\%$ around the mean versus ${\sim}16\,\%$ for surge). Interestingly, the time-averaged recovery (green solid line) is slightly greater than the mean of the phase-resolved wake recovery, however, by only about $3\,\%$ , indicating that the enhanced recovery for sway is primarily due to increased momentum transport rather than smoothing artefacts – a result also analysed by Korb et al. (Reference Korb, Asmuth and Ivanell2023). However, the approach of Korb et al. (Reference Korb, Asmuth and Ivanell2023) differs: using CFD, they examined energy differences between the time-averaged wake and the wake in a meandering frame of reference, integrated over the full wake area at a given downstream position. They reported differences up to $15 \,\%$ , whereas our wind-speed-based line profiles show only a $3\,\%$ deviation. While the cases and metrics differ (they examined energy; we analyse streamwise velocity), both studies highlight that averaging over periodic wake motion introduces discrepancies in evaluating true recovery. If other metrics like $\overline {u^2}$ (kinetic energy) or $\overline {u^3}$ (wind power) were considered, these discrepancies would likely be more pronounced in our case. Nonetheless, we find that the dominant source of enhanced recovery remains the increased exchange between the wake and the outer flow – a point we return to in the conclusion.

3.5. Coherent structures: energy perspectives

We conclude the results section with an analysis of the energy of the coherent structures observed in the wake of the moving wind turbine, focusing on their relationship with free-stream turbulence intensity, energy of platform motion and enhanced wake recovery.

Reynolds & Hussain (Reference Reynolds and Hussain1972) provide a mathematical framework for the energy dynamics of coherent structures, describing the transfer of energy between the mean flow, stochastic fluctuations and coherent motions. They define the local kinetic energy of a coherent structure in a flow as

(3.3) \begin{equation} e_{\textit{CS}} = \frac {1}{2} \overline {\widetilde {u_i}^2} = \frac {1}{2} \big ( \overline {\widetilde {u}^2} + \overline {\widetilde {v}^2} + \overline {\widetilde {w}^2} \big ) .\end{equation}

In their work, Reynolds & Hussain (Reference Reynolds and Hussain1972) derive a transport equation for $e_{\textit{CS}}$ from the phase-averaged Navier–Stokes equations. While such an analysis would be of interest in the present context, it falls beyond the scope of this study. However, the concept of energy remains relevant, as we can compute $e_{\textit{CS}}$ from the full 3-D dataset available from Milan, and evaluate $\overline {\widetilde {u}^2}$ (a subset of $e_{\textit{CS}}$ ) across all cases with varying degrees of freedom (DoF), Strouhal number $ St$ , motion reduced-amplitude $ A^*$ and free-stream turbulence intensity $ \textit{TI}_{\infty }$ .

Hussain (Reference Hussain1986) emphasises the 3-D nature of coherent structures, which often occupy a substantial volume of a sheared flow. This motivates extending the concept of local energy to a more spatially integrated quantity, particularly across the dimensions where measurements are available. Previous CFD studies from Yılmaz & Meyers (Reference Yılmaz and Meyers2018) and Li et al. (Reference Li, Dong and Yang2022) provide visual evidence of the large-scale extent of these structures in 3-D space.

In our case, measurements are taken along a horizontal line at hub height. As observed in figures 10 and 11, the motion-induced coherent structures span a significant portion of the lateral direction. Most of the energy is concentrated in the region $ y \in [-D, D]$ , as illustrated, for example, in figure 11(b)(a.2). Based on this, and following a similar approach to Yılmaz & Meyers (Reference Yılmaz and Meyers2018), we define the spanwise-integrated kinetic energy of the coherent structures, normalised by the inflow mean kinetic energy:

(3.4) \begin{equation} E^*_{\textit{CS}} = \frac {1}{2D} \int _{-D}^{D} \frac {1}{2} \big ( \overline {\widetilde {u}^2} + \overline {\widetilde {v}^2} + \overline {\widetilde {w}^2} \big ) {\rm d}y \Big / \left ( \frac {1}{2} U_{\infty }^2 \right ) = \frac {1}{2D U^2_{\infty }} \int _{-D}^{D} e_{\textit{CS}} \, {\rm d}y. \end{equation}

To enable comparison across all datasets, we also define a streamwise-only version, $ E^*_{x,CS}$ , which uses only the streamwise component $ \overline {\widetilde {u}^2}$ :

(3.5) \begin{equation} E^*_{x,CS} = \frac {1}{2D U^2_{\infty }} \int _{-D}^{D} \overline {\widetilde {u}^2} \, {\rm d}y .\end{equation}

Similar to (3.5), we define $ E^*_{y,CS}$ and $ E^*_{z,CS}$ as the normalised energy contributions from the spanwise ( $ \overline {\widetilde {v}^2}$ ) and vertical ( $ \overline {\widetilde {w}^2}$ ) components of the coherent structures, respectively.

We begin by examining the shared dataset between Milan and Oldenburg for the cases surge-5 and sway-5 (see table 2), under inflow conditions M1.5 and O1.5 with $ \textit{TI}_{\infty } = 1.5\,\%$ (see table 1). Figure 13(a) presents the downstream evolution of $ E^*_{x,CS}$ from both the Milan and Oldenburg datasets, while figure 13(b) shows the evolution of $ E^*_{i,CS}$ , where $ i \in \{x, y, z\}$ , and the total energy $ E^*_{\textit{CS}}$ , using the 3-D data from Milan. The good agreement (i.e. a mean deviation of less than $10 \,\%$ ) between the curves marked with empty symbols (Oldenburg) in figure 13(a) and those with filled symbols (Milan) confirms the consistency of the results from the two experiments. A more detailed comparison is provided in Appendix B.

Figure 13. Integrated energy of surge- and sway-induced coherent structures for cases surge-5 and sway-5 (see table 2). (a) Comparison of the integrated energy of the streamwise component of the coherent fluctuations $E^*_{x,CS}$ (3.5) for surge and sway with $St = 0.38$ , $A^* = 0.01$ , using Milan and Oldenburg datasets ( $\textit{TI}_{\infty } = 1.5\,\%$ ). (b) Integrated energy of each component of the coherent structure: streamwise $\overline {\widetilde {u}^2}$ , lateral $\overline {\widetilde {v}^2}$ and vertical $\overline {\widetilde {w}^2}$ velocities, as well as their total noted $E^*_{x,CS},\,E^*_{y,CS}, E^*_{z,CS}$ and $E^*_{\textit{CS}}$ , respectively, for surge (b.1) and sway (b.2). Panels (b.3) and (b.4) show the same energy components normalised by their respective maximum values.

In figure 13(a) the normalised, spanwise-integrated energy of $ \overline {\widetilde {u}^2}$ for both surge- and sway-induced structures initially increases for $ x \lesssim 4D$ . This indicates that, in the near-wake to the transition region, the coherent structures are gaining energy. The energy associated with the meandering mode (sway, green curve) is approximately $50 \,\%$ larger than that of the pulsating mode (surge), correlating with the higher wake recovery for sway observed in figure 4(a). Further downstream, for $ 3.5 \lesssim x/D \lesssim 5.5$ , the energy reaches a peak and then decays. This decline indicates that the coherent structures are dissipating energy, consistent with our pointwise analysis in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). By $ x = 10D$ , the energy has dropped to about one-fifth of its peak value, showing strong dissipation and potential breakdown into smaller-scale turbulence, consistent with the classic energy cascade described by Richardson/Kolmogorov (Pope Reference Pope2001).

Figure 13(b)(1–2) provide additional insights into the downstream development of $ E^*_{y,CS}$ , $ E^*_{z,CS}$ and the total $ E^*_{\textit{CS}}$ for the Milan dataset. The differences between the surge- and sway-induced structures are much more pronounced when considering the full 3-D energy components. In particular, the spanwise component $ \overline {\widetilde {v}^2}$ (triangular markers with dashed lines) shows almost four times more energy for sway than for surge. As a result, the total energy $ E^*_{\textit{CS}}$ is nearly three times higher for sway, emphasising the energetic dominance of the meandering mode under identical excitation conditions ( $ St = 0.38,\, A^* = 0.01$ ).

Despite these differences, the overall trends are similar: $ E^*_{\textit{CS}}$ increases to a peak before decreasing downstream. This full 3-D analysis shows that a significant portion of the meandering mode’s energy resides in the spanwise direction ( $ \widetilde {v}$ ), as expected from its lateral motion characteristics. Figure 13(b).(3–4) show the same quantities as in b.(1–2), but normalised by their respective maximum values. While the normalised total energy $ E^*_{\textit{CS}} / \max E^*_{\textit{CS}}$ does not exactly match the normalised streamwise component $ E^*_{x,CS} / \max E^*_{x,CS}$ , the trends remain consistent. This supports the decision to investigate $ E^*_{x,CS}$ as a proxy for total energy across all cases – while acknowledging that it only represents a portion of the total energy. In particular, for the meandering mode (sway), a large fraction of the total energy is concentrated in the spanwise component $ E^*_{y,CS}$ , meaning that differences between surge and sway are more significant than suggested by $ E^*_{x,CS}$ alone. Nonetheless, it is reasonable to assume that a lower $ E^*_{x,CS}$ generally implies a lower total energy $ E^*_{\textit{CS}}$ .

We next analyse the $E^*_{x,CS}$ dependency to inflow conditions and motion parameters. We start with figure 14 for the normalised energy evolution of the pulsating mode from surge motion (figure 14 a) and meandering mode from sway (figure 14 b) both with $St = 0.3,\,A^*=0.024$ for the different inflows: $\textit{TI}_{\infty } \in [1.1, 5.8]\,\%$ . On both panels, we added red symbols that represent the normalised energy of the free-stream fluctuations, $\epsilon ^*_{\infty }$ , defined as

(3.6) \begin{align} \epsilon ^*_{\infty } = \epsilon _{\infty }/\big(1/2 U^2_{\infty }\big) = \big(1/2 \overline {u^{{\prime}2}_{\infty }}\big) /\big(1/2 U^2_{\infty }\big) = \overline {u^{{\prime}2}_{\infty }}/U^2_{\infty } = TI^2_{\infty } . \end{align}

Figure 14. Integrated energy of the streamwise component of the coherent structures induced by motions $E^*_{x,CS}$ for surge (a) and sway (b) both with $St = 0.3,\,A^*=0.024$ (surge- and sway-3–0.024 in table 2) for $\textit{TI}_{\infty } \in [1.1, 5.8]\,\%$ (O1.1–O5.8 in table 1). The red symbols represent the normalised energy of the free-stream fluctuations $\epsilon ^*_{\infty }$ (3.6) and the dashed line shows $\epsilon ^*_{p}$ , the normalised energy of the platform motion (3.7).

From (3.6), the higher the inflow $\textit{TI}_{\infty }$ , the higher the energy of the fluctuations. Here $\epsilon ^*_{\infty }$ can be compared with $E^*_{x,CS}$ as well as the normalised energy of the platform motion, $\epsilon ^*_p$ , which we defined and discussed already in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ), as

(3.7) \begin{equation} \epsilon ^*_{p} = 2 (\pi St A^*)^2 ,\end{equation}

which quantifies the normalised energy of the platform motion. Figure 14 illustrates that, for both surge (figure 14 a) and sway (figure 14 b), increasing inflow turbulence intensity $ \textit{TI}_{\infty }$ (i.e. higher $ \epsilon ^*_{\infty }$ ) leads to a reduction in the energy of the coherent structures. This trend is also visible in figures 10 and 11, based on the spatial maps of $ \widetilde {u}$ .

For example, in the surge case, the normalised streamwise energy of the coherent structure, $ E^*_{x,CS}$ , is approximately eight times larger for $ \textit{TI}_{\infty } = 1.1\,\%$ than for $ \textit{TI}_{\infty } = 5.8\,\%$ . A similar trend is observed for sway, with the energy about five times higher at low turbulence. At a fixed inflow condition, sway-induced coherent structures exhibit higher energy levels than those induced by surge, consistent with the observations in figure 13(a).

Both surge and sway exhibit a similar evolution in coherent structure energy: an initial growth up to $ x \approx 4{-}6D$ , followed by a downstream decay. At $ x = 2D$ , the energy levels are nearly identical across all inflow turbulence intensities $ \textit{TI}_{\infty }$ , and closely match $ \epsilon ^*_p$ – the energy associated with platform motion. This suggests that at this early position, the coherent structures have not yet developed and their energy is primarily determined by the imposed motion. Further downstream, however, the growth and peak level of coherent structure energy become strongly dependent on $ \textit{TI}_{\infty }$ . For low turbulence intensities ( $ \textit{TI}_{\infty } \leqslant 2.2\,\%$ ), the structures grow substantially in energy. In contrast, for higher turbulence levels ( $ \textit{TI}_{\infty } \gt 3\,\%$ ), their growth is significantly dampened. The location of the peak energy $ E^*_{x,{CS}}$ , as well as the rate of energy increase and decay, vary depending on both the direction of motion and $ \textit{TI}_{\infty }$ . This indicates that the formation and evolution of coherent structures are highly sensitive to inflow conditions.

Under surge excitation, when the inflow turbulence intensity exceeds approximately $ \textit{TI}_{\infty } \gtrsim 3\,\%$ , the energy of the streamwise coherent structure, $ E^*_{x,{CS}}$ , generally falls below the platform motion energy threshold $ \epsilon ^*_p$ . While a slight increase above $ \epsilon ^*_p$ is observed close to the source, it quickly diminishes downstream, suggesting that higher turbulence intensity inhibits the formation and persistence of coherent structures. For sway, by contrast, coherent structure energy remains above the threshold until turbulence levels exceed approximately $ 5 \,\%$ , demonstrating that the sway-induced structures are more resilient to free-stream turbulence.

Interestingly, this energy threshold seems to correlate with the wake recovery trends shown in figure 8. In the surge case, no significant improvement in recovery is observed for $ \textit{TI}_{\infty } \gtrsim 3\,\%$ , whereas for the sway case, this limit rises to $ \textit{TI}_{\infty } \gtrsim 5\,\%$ . This suggests that $ \epsilon ^*_{p}$ may act as a critical threshold – below which the coherent structure’s energy is too weak to affect wake recovery meaningfully. Figure 14 also highlights that sway-induced structures not only carry more energy but also better resist the damping effects of increased inflow turbulence compared with surge-induced structures.

So far, we have investigated only the highest amplitude of motion ( $ A^* = 0.024$ ). We now extend the analysis in figure 15 to include all three tested amplitudes: $ A^* \in \{0.01,\,0.017,\,0.024\}$ . To facilitate the analysis of how coherent structure energy depends on $ A^*$ and inflow turbulence intensity $ \textit{TI}_{\infty }$ , we introduce a streamwise-integrated metric.

As shown in figure 14, increasing levels of $ \textit{TI}_{\infty }$ lead to a clear reduction in the overall energy of the coherent structures. On average, this energy is well above the platform motion energy threshold $ \epsilon ^*_p$ for low turbulence levels ( $ \textit{TI}_{\infty } \lt 2\,\%$ ), approximately equal to it around $ \textit{TI}_{\infty } \approx 3\,\%$ , and falls below it for higher turbulence intensities ( $ \textit{TI}_{\infty } \gt 3\,\%$ ) in the surge case (figure 14 a). For sway, the thresholds are at higher $ \textit{TI}_{\infty }$ values (figure 14 b). These observations motivate the use of a spatially integrated metric that captures the overall strength of the structure and allows direct comparison with $ \epsilon ^*_p$ . Inspired by the approach of Yılmaz & Meyers (Reference Yılmaz and Meyers2018), we define the streamwise-integrated coherent structure energy as

(3.8) \begin{align} \int E^*_{x,{CS}} = \frac {1}{8D} \int _{x=2D}^{x=10D} E^*_{x,{CS}}(x)\,{\rm d}x. \end{align}

This quantity represents a spatially averaged estimate of the coherent structure energy in the $ x$ direction, computed over the region where the structures are observed ( $ x \times y \in [2D, 10D] \times [-D, D]$ ). Figure 15 presents $ \int E^*_{x,{CS}}$ as a function of inflow turbulence intensity $ \textit{TI}_{\infty }$ for surge (a) and sway (b) excitation, at fixed Strouhal number $ St = 0.3$ , and for all amplitudes and inflow conditions considered (O1.1–O5.8 in table 1, corresponding to $ \textit{TI}_{\infty } \in [1.1\,\%, 5.8\,\%]$ ).

Figure 15. Streamwise-integrated energy of the coherent structures in the $ x$ direction, defined as $ \int E^*_{x,CS} = ({1}/{8D}) \int _{x=2D}^{x=10D} E^*_{x,CS}(x) \, {\rm d}x$ , plotted against the inflow turbulence intensity $ \textit{TI}_{\infty }$ for (a) surge and (b) sway. All cases use $ St = 0.3$ and $ A^* \in \{0.01,\,0.017,\,0.024\}$ in table 2 and inflows O1.1–O5.8 in table 1. Here red :, –,- lines represent $\epsilon ^*_p$ defined in (3.7).

A consistent trend is observed for both motion directions: for a given inflow condition, increasing $ A^*$ leads to a higher coherent structure energy. For example, at $ \textit{TI}_{\infty } = 1.1\,\%$ , the integrated energy for surge at $ A^* = 0.024$ is approximately three times higher than at $ A^* = 0.01$ (compare blue $\bullet$ and x in figure 15 a). Similarly, for sway, the increase is about 1.8 times (compare green $\bullet$ and x in figure 15 b).

A higher $ A^*$ naturally results in a higher excitation energy $ \epsilon ^*_p$ , which scales with $ {A^*}^2$ (3.7). Thus, the platform motion at $ A^* = 0.024$ carries approximately 5.8 times more energy than at $ A^* = 0.01$ . Interestingly, the relative response of the wake structures (i.e. the energy of the structures relative to $ \epsilon ^*_p$ ) is larger for smaller $ A^*$ , which aligns with classical findings in periodic excitation and synchronisation theory. As shown by Pikovsky et al. (Reference Pikovsky, Rosenblum and Kurths2001), weakly forced nonlinear systems often exhibit enhanced sensitivity, leading to more efficient energy amplification at lower excitation amplitudes. Nonetheless, in terms of absolute energy levels, our results show clearly that increasing the excitation amplitude leads to more energetic coherent structures. Since $ \epsilon ^*_p \ll 1$ , even the largest tested amplitudes still fall in the range of what can be considered infinitesimal forcing.

The further analysis of figure 15 confirms, as previously analysed in figure 14, that sway motion generates coherent structures with higher energy and greater resilience to increasing inflow turbulence intensity $ \textit{TI}_{\infty }$ , compared with surge. The connection between the energy of the coherent structures and wake recovery enhancement becomes apparent: as $ \textit{TI}_{\infty }$ increases, the energy of the structures decreases, and so does the impact of the structures on recovery. For the largest amplitude $ A^* = 0.024$ , the energy drops below the corresponding platform motion energy $ \epsilon ^*_p(A^* = 0.024)$ at approximately $ \textit{TI}_{\infty } \approx 3.0\,\%$ for surge and $ \textit{TI}_{\infty } \approx 5.0\,\%$ for sway. For the lower amplitudes ( $ A^* \leqslant 0.017$ ), $ \int E^*_{x,CS}$ remains above the respective $ \epsilon ^*_p(A^*)$ for all inflow conditions, but the absolute energy remains significantly lower than that of the structures induced by $ A^* = 0.024$ . By linking figures 8 and 15, we observe that, for any $ A^*$ , when $ \int E^*_{x,CS} \lesssim \epsilon ^*_p(A^* = 0.024)$ , the impact of platform motion on wake recovery improvement becomes negligible. Therefore, the higher energy of the meandering structures induced by sway, along with their greater robustness to increasing $ \textit{TI}_{\infty }$ , leads to an extended range of inflow turbulence conditions where sway motion improves recovery compared with surge. This observation is further quantified in figure 16.

Figure 16. Relative increase in wake recovery, $ \Delta \mathcal{R} / \mathcal{R}$ (3.10), plotted against the ratio $ \epsilon ^*_{\infty } / \epsilon ^*_p$ , which compares the normalised energy of free-stream fluctuations to that of platform motion. Data includes all motion cases with $ St = 0.3$ and inflow conditions from O1.1 to O5.8 (see tables 1 and 2). A threshold of $ 1.5 \,\%$ is shown as a red dashed line.

Figures 14 and 15 have shown that both $ \epsilon ^*_{\infty }$ (the normalised energy of the free-stream fluctuations) and $ \epsilon ^*_p$ (the normalised energy of the platform motion) are key parameters governing the energy of coherent structures and, consequently, the wake recovery enhancement. An increase in $ \epsilon ^*_p$ leads to more energetic structures for a fixed $ \epsilon ^*_{\infty }$ , while an increase in $ \epsilon ^*_{\infty }$ results in weaker structures, as they struggle to emerge and persist in a more turbulent background at constant $ \epsilon ^*_p$ . As the energy of the coherent structures decreases, so does their impact on wake recovery improvement.

This observation motivates the study of the relative increase in wake recovery, $ \Delta \mathcal{R} / \mathcal{R}$ , between a case with platform motion and the corresponding fixed-platform reference, for different energy levels of inflow and platform motion. The streamwise-integrated wake recovery $ \mathcal{R}$ is defined as

(3.9) \begin{equation} \mathcal{R} = \frac {1}{8D} \int _{x=2D}^{x=10D} \frac {1}{D} \int _{y=-R}^{y=R} \frac {\overline {u}(x,y)}{U_{\infty }} \, {\rm d}y \, {\rm d}x = \frac {1}{8D} \int _{x=2D}^{x=10D} R_w(x) \, {\rm d}x.\end{equation}

Here, $ \mathcal{R}$ represents the average recovery over the measurement region $ x \in [2D, 10D]$ . The relative increase in recovery is given by

(3.10) \begin{equation} \frac {\Delta \mathcal{R}}{\mathcal{R}} = \frac {\mathcal{R}(St, A^*, \text{Dof}, TI_{\infty }) - \mathcal{R}(St=0, TI_{\infty })}{\mathcal{R}(St, A^*, \text{Dof}, TI_{\infty })}. \end{equation}

This quantity quantifies the relative improvement in wake recovery due to platform motion, averaged over the streamwise measurement domain. The final figure 16 shows the evolution of $ \Delta \mathcal{R} / \mathcal{R}$ (3.10), plotted against the ratio $ \epsilon ^*_{\infty } / \epsilon ^*_p$ for all motion cases with $ St = 0.3$ and inflows ranging from O1.1 to O5.8. This figure complies some of the outcomes of figures 8, 14 and 15 combined.

For $ \textit{TI}_{\infty } = 1.1\,\%$ , $ \epsilon ^*_\infty = 1.2 \times 10^{-4}$ , which is lower than $ \epsilon ^*_p$ for all $ A^*$ . For example, for $ A^* = 0.024$ , $ \epsilon ^*_\infty / \epsilon ^*_p \approx 0.1$ , indicating that there is an order of magnitude more energy in the platform motion than in the free-stream turbulence. As a result, the energy of the coherent structure found in the wake and induced by the platform is significant (the largest) for both surge and sway DoFs (see the blue and green $ \bullet$ in figure 15(a) and 15(b) for quantification). Consequently, the impact on recovery is also large: about $ 16 \,\%$ more averaged recovery is observed for sway compared with $7 \,\%$ for surge, relative to the fixed case’s recovery (see the $ \bullet$ points for $ \epsilon ^*_\infty / \epsilon ^*_p \approx 0.1$ in figure 16). As $ \epsilon ^*_{\infty } / \epsilon ^*_p$ increases, the inflow turbulence either increases (if $ \epsilon ^*_p$ is fixed) or $ \epsilon ^*_p$ decreases (with fixed $ \epsilon ^*_{\infty }$ ). Regardless of the scenario, as $ \epsilon ^*_{\infty } / \epsilon ^*_p$ increases, the relative importance of the motion diminishes because the inflow turbulence becomes more dominant. For both surge and sway, small discrepancies in the impact of motion on recovery improvement can be observed at different amplitudes for a given $ \epsilon ^*_{\infty } / \epsilon ^*_p$ ratio. Specifically, the lowest amplitude ( x in figure 16) leads to larger recovery improvements compared with the highest amplitude, consistent with the statement that ‘the lower the excitation amplitude, the greater the relative impact’. Nevertheless, the overall trend remains clear: as $ \epsilon ^*_{\infty } / \epsilon ^*_p$ increases, the influence of platform motion on wake recovery diminishes, eventually falling below $ 1.5 \,\%$ – a level at which the effect becomes indistinguishable from measurement uncertainty here (Messmer et al. Reference Messmer, Peinke and Hölling2024b ).

Notably, the difference between surge and sway, discussed throughout the paper, becomes evident here. For $ \epsilon ^*_{\infty } / \epsilon ^*_p \gtrsim 1$ (marked with the transparent blue rectangle in figure 16), the impact of surge motion on wake recovery becomes negligible. In contrast, the threshold for sway is about 3.5 times higher, with no measurable impact for $ \epsilon ^*_{\infty } / \epsilon ^*_p \gtrsim 3.5$ . This key result demonstrates that surge-induced coherent structures, responsible for enhanced recovery, are more sensitive to inflow turbulence than sway-induced structures, with a factor of about 3.5. This is further discussed in the conclusion.

4. Summary of the main findings

In this paper we investigated wake recovery mechanisms and the interplay between inflow turbulence and motion-induced coherent structures in the wake of a model floating wind turbine. We considered harmonic surge and sway motions (figure 2 b,c), with motion parameters $ St \in \{0.3, 0.38\}$ and $ A^* \in \{0.01, 0.017, 0.024\}$ . Wind tunnel experiments (figure 1) were conducted in two facilities under various inflow conditions, with turbulence intensities $ \textit{TI}_{\infty } \in [1.1, 5.8]\,\%$ , controlled using an active grid. Hereafter, we summarise the main findings.

The new insights of this paper are the connection between motion-induced coherent structures and enhanced wake recovery, as well as the interplay with increasing $ \textit{TI}_{\infty }$ .

First, at $ \textit{TI}_{\infty } = 1.5\,\%$ , we found that the increased wake recovery with platform motion, most pronounced for sway (figure 4 a), is directly linked to an increase (up to $ 65 \,\%$ more than the fixed reference case) in the Reynolds shear stress gradient term in the $ x$ $ y$ plane (i.e. from streamwise-lateral fluctuations), and secondly in the $ x$ $ z$ plane. Previous work by Bastankhah & Porté-Agel (Reference Bastankhah and Porté-Agel2016), Boudreau & Dumas (Reference Boudreau and Dumas2017), van der Laan et al. (Reference van der Laan, Baungaard and Kelly2023), Gambuzza & Ganapathisubramani (Reference Gambuzza and Ganapathisubramani2023) identified these terms as the main drivers of recovery. In this paper we further decompose the terms $- \langle (1/\overline {u}) (\partial \overline {u' v'} / \partial y + \partial \overline {u' w'} / \partial z ) \rangle$ as the sum of the contributions from the coherent structure ( $ \overline {\widetilde {u} \widetilde {v}}$ , $ \overline {\widetilde {u} \widetilde {w}}$ ) and purely stochastic turbulence ( $ \overline {u''v''}$ , $ \overline {u''w''}$ ). We showed in figure 6 the large contribution (more than $ 50 \,\%$ ) from $-\left \langle (1/\overline {u})\partial \overline {\widetilde {u} \widetilde {v}} / \partial y \right \rangle$ , i.e. from the coherent structure induced by the rotor motion of the floating platform to the improved $-\langle (1/\overline {u}) \partial \overline {u' v'} / \partial y \rangle$ . This term is found in the wake recovery rate budget (2.6) and contributes to the higher recovery rate, $D \partial R_w/\partial x$ . This new result confirms our speculation in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ) and provides clear evidence as well as quantification of the role of large-scale coherent structures in improving wake recovery.

Second, as $ \textit{TI}_{\infty }$ increases, the wake recovery of the moving turbine approaches that of the fixed case, indicating that higher inflow turbulence reduces the influence of motion on recovery enhancement (figure 8). This trend is consistent with CFD results for sway motion reported by Li et al. (Reference Li, Dong and Yang2022) and surge motion from Li et al. (Reference Li, Yu and Sarlak2024), as well as with other rotor excitation techniques investigated in Yılmaz & Meyers (Reference Yılmaz and Meyers2018), Korb et al. (Reference Korb, Asmuth and Ivanell2023). We considered three amplitudes of motion and observed that higher amplitudes result in a more pronounced impact on recovery. Comparing fore-aft and sideways motion, we saw in figure 8 that sway motion leads to a larger increase in recovery than surge for equivalent cases. Furthermore, this study identifies the inflow turbulence levels, $ \textit{TI}_{\infty }$ , beyond which the impact of platform motion on wake recovery becomes negligible. Specifically, for the maximum amplitude tested ( $ A^* = 0.024$ at $ St = 0.3$ ), the critical threshold is approximately $ \textit{TI}_{\infty } \approx 3\,\%$ for surge and $ \textit{TI}_{\infty } \approx 5\,\%$ for sway motion.

Third, the measurements from Oldenburg enabled the computation of the shape of the coherent structures and allowed us to distinguish two different modes induced by surge (fore-aft) and sway (sideways) motion, as well as their development under increasing free-stream turbulence. Fore-aft motion leads to the formation of pulsating coherent structures (figure 10), characterised by alternating shrinking and expansion of the wake, also reported in Yılmaz & Meyers (Reference Yılmaz and Meyers2018), Li & Yang (Reference Li and Yang2024), Li et al. (Reference Li, Yu and Sarlak2024), Wei et al. (Reference Wei, El Makdah, Hu, Kaiser, Rival and Dabiri2024), Hubert et al. (Reference Hubert, Conan and Aubrun2025), and resembling the varicose mode observed in planar wakes described by Wygnanski et al. (Reference Wygnanski, Champagne and Marasli1986). In contrast, sway motion generates structures that oscillate side to side, a typical meandering mode, similar to those found and discussed in Li et al. (Reference Li, Dong and Yang2022), Li & Yang (Reference Li and Yang2024). This structure shares characteristics with vortex shedding behind bluff bodies (see figure 11) and with the sinous mode described by Wygnanski et al. (Reference Wygnanski, Champagne and Marasli1986). Both types of structures (pulsating and meandering) contribute to increased Reynolds shear stresses, $ \partial \overline {u'v'} / \partial y$ and $ \partial \overline {u'w'}/ \partial z$ , as discussed earlier and shown in figure 6. For both DoF, increasing the level of inflow turbulence, $ \textit{TI}_{\infty }$ , leads to less pronounced coherent structures, which become barely visible at $ \textit{TI}_{\infty } = 5.8\,\%$ .

Fourth, we investigated the influence of coherent structures on phase-resolved wake recovery, by examining changes in the phase-resolved wake profiles (figure 12). Surge motion induces significant phase-to-phase variations in wake recovery, but these fluctuations average out over a cycle to result in little to no difference in the average of the phase-averaged recoveries compared with the time-averaged recovery defined in (2.2). Sway motion, in contrast, causes smaller phase-to-phase fluctuations, yet the mean phase-resolved recovery is lower than the classic time-averaged recovery. This suggests, as also noted by Korb et al. (Reference Korb, Asmuth and Ivanell2023), that part of the enhanced wake recovery observed with sway motion, reflected in the increased term $ \partial \overline {u'v'} / \partial y$ , is due to lateral smearing of the wake profile. However, this smearing mechanism accounts for only about $ 10 \,\% $ of the total recovery enhancement, indicating in our case that the dominant contribution arises from increased momentum entrainment.

Finally, we linked enhanced recovery and coherent structures by examining the energy of the large-scale organised patterns. We focused on the streamwise component of the coherent structure energy, $ E^*_{x,\textit{CS}}$ (3.5), and provided rationale for this choice in figure 13. Results show that meandering structures exhibit higher energy and are more resilient to increasing inflow turbulence than surge-induced structures (figure 14). Moreover, higher motion amplitudes, $ A^*$ , generate structures with greater energy compared with lower amplitudes, directly linking to the lower recovery increase under smaller motion amplitudes (figure 15). Last but not least, we quantified the relationship between the energy of free-stream fluctuations, $ \epsilon ^*_\infty$ (3.6), and that of the platform motion, $ \epsilon ^*_p$ (3.7), in relation to the relative increase in wake recovery (figure 16). For $ \epsilon ^*_\infty / \epsilon ^*_p \lt 0.5$ , i.e. when motion energy dominates over free-stream fluctuations, both surge and sway motions enhance wake recovery, with sway leading to up to $16 \,\%$ improvement. However, as $ \epsilon ^*_\infty$ increases, the impact of motion on recovery diminishes. Our results clearly show that sway consistently leads to larger recovery enhancement than surge and that sway-induced structures are more resilient to inflow turbulence. We found that, for $ \epsilon ^*_\infty / \epsilon ^*_p \gtrsim 1$ , the effect of surge motion on wake recovery improvement becomes negligible, while for sway, this threshold is about $ \epsilon ^*_\infty / \epsilon ^*_p \gtrsim 3.5$ , indicating the greater robustness of sway-induced structures to free-stream turbulence.

5. Discussion and conclusion

Last, we discuss and draw conclusions based on our results regarding the similarity with canonical flows and local wake recovery mechanisms (§ 5.1), the formation of coherent structures in turbulent backgrounds (§ 5.2) and the implications for wind energy (§ 5.3).

5.1. Similarities with canonical flows and local wake recovery mechanisms

Despite the complexity of wind turbine wakes and the high Reynolds numbers involved, it is remarkable to observe wake dynamics similar of those found in canonical shear flows with laminar backgrounds, as described in foundational studies from the 1970s and 1980s (Crow & Champagne Reference Crow and Champagne1971; Ho & Huerre Reference Ho and Huerre1984; Wygnanski et al. Reference Wygnanski, Champagne and Marasli1986). These structures arise from shear-flow instabilities, where periodic disturbances are strongly amplified (Gupta & Wan Reference Gupta and Wan2019), even in turbulent environments where excitations are less distinct. In the wake of a fixed turbine, we observed inherent sinuous (Wygnanski et al. Reference Wygnanski, Champagne and Marasli1986) (i.e. meandering) modes, as evidenced by the erratic meandering reported in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). However, these modes are typically smeared out, resulting in disorganised motion similar to that observed behind porous disks with high porosity (Cannon et al. Reference Cannon, Champagne and Glezer1993). This behaviour occurs within a characteristic meandering frequency range ( $f_m$ ), defined by the Strouhal number $ St_m = ({f_m D}/{U_{\infty }}) \in [0.1, 0.5]$ , as previously discussed in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). When such a shear flow is excited at an appropriate frequency (e.g. $f_p D/U_{\infty } \approx 0.3$ ), via harmonic platform motion or other methods, the convectively unstable wake responds strongly, even to small perturbations (Ho & Huerre Reference Ho and Huerre1984). The dynamics then become dominated by the excited coherent structures, similar to vortex shedding behind a cylinder. As noted by Cantwell & Coles (Reference Cantwell and Coles1983), ‘the relatively large stress associated with deep incursions of free-stream fluid into the regions between the (coherent) vortices’ is the ‘most interesting feature’ of such flows. This observation raises questions about the local mechanisms driving wake recovery, which we discuss in the following.

Our findings demonstrate that, in a time-averaged sense, large-scale periodic structures enhance wake recovery by increasing the gradient of the Reynolds shear stresses. Instantaneously, two key mechanisms appear to be involved: periodic wake deflection (Korb et al. Reference Korb, Asmuth and Ivanell2023) and enhanced entrainment – achieved via engulfment and nibbling processes, as discussed by Kankanwadi & Buxton (Reference Kankanwadi and Buxton2020) in the context of cylinder wakes. As shown in figure 12, the majority of the recovery enhancement originates from increased entrainment, while wake deflection plays a secondary role.

An open question remains: Which local entrainment mechanism is primarily enhanced by the presence of large-scale structures – nibbling, engulfment, or both? On one hand, the organised structures likely increase the length of the interface between the wake and the outer flow (i.e. the turbulent–turbulent interface), potentially enabling enhanced momentum transfer via small-scale nibbling (da Silva et al. Reference da Silva, Hunt, Eames and Westerweel2014). On the other hand, following the observations of Cantwell & Coles (Reference Cantwell and Coles1983), large-scale structures may promote engulfment by drawing outer fluid into the wake core, thereby accelerating momentum recovery.

At this stage, these remain hypotheses, but they represent promising directions for future investigation aimed at better understanding the very local mechanisms by which large-scale coherence influences wake recovery.

5.2. Formation of excited coherent structures in a turbulent background

In this paper we turned to analytical methods originally used to study coherent structures in canonical flows (Reynolds & Hussain Reference Reynolds and Hussain1972), applying them here to measure structures’ energy and its dependence on free-stream turbulence to a problem motivated by wind energy. We observed that increased inflow turbulence ( $\textit{TI}_\infty$ ) leads to less energetic motion-excited coherent structures. This suggests that higher turbulence levels either smear out the periodic excitation, preventing the formation of coherent structures, or disrupt the structures rapidly, or both. We found that reduced coherent structure energy corresponds to weaker enhancements in the Reynolds stress gradients, as evidenced by the comparison between surge- and sway-induced structures in figure 13(b.1,b.2). The reduced surge-induced structure energy is further linked to the diminished $\partial \overline {u'v'} / \partial y,\,\partial \overline {u'w'} / \partial z$ observed in figure 6(e) compared with figure 6(f) for sway, highlighting the limited capacity of low energy structures to improve wake recovery.

Hussain (Reference Hussain1986) wrote: ‘the excitation amplitude should be small but sufficiently above the background disturbance or free stream turbulence’ for organised structures to emerge from periodic excitation. This highlights a key requirement: although platform-induced excitations may be of low amplitude, they must remain distinguishable from the background turbulence to effectively influence wake dynamics. In this context, the energy ratio between free-stream turbulence and platform motion, $\epsilon ^*_\infty / \epsilon ^*_p$ , emerges as a useful metric for assessing the conditions under which coherent structures can form and grow sufficiently to enhance recovery. Our results, illustrated in figures 13, 14, 15 and 16, reveal pronounced differences between the effects of surge (fore-aft) and sway (side-to-side) excitations on wake dynamics and recovery. For surge, when the energy of inflow turbulence approaches that of the platform motion, the wake fails to sustain coherent structures with sufficient energy to influence recovery. In contrast, sway-induced structures remain effective up to significantly higher turbulence levels, with $\epsilon ^*_\infty / \epsilon ^*_p \approx 3.5$ . This suggests that meandering modes triggered by sway motion are either more easily excited, more robust against turbulent noise or both – making sway-induced excitations more persistent than their surge-induced (pulsating) counterparts under similar inflow conditions. To enhance practical applicability, it becomes essential to identify excitation methods that retain their influence on wake dynamics even under high $\epsilon ^*_\infty / \epsilon ^*_p$ where background turbulence dominates over platform motion, and to understand further the underlying flow mechanisms.

The fundamental differences between surge- and sway-induced wake structures, along with their varying resilience to inflow turbulence, raise important questions about the underlying mechanisms driving their formation. Previous studies by Kleine et al. (Reference Kleine, Franceschini, Carmo, Hanifi and Henningson2022) and Wei et al. (Reference Wei, El Makdah, Hu, Kaiser, Rival and Dabiri2024) suggest that blade-tip vortices play a key role in the development of pulsating wake structures. As $\textit{TI}_{\infty }$ increases, however, tip vortices become destabilised by the background fluctuations, leading to their early breakdown in the wake (Neunaber et al. Reference Neunaber, Hölling, Stevens, Schepers and Peinke2020; Porté-Agel et al. Reference Porté-Agel, Bastankhah and Shamsoddin2020). This raises the question: Could the breakdown of these disturbed tip vortices limit the ability of pulsating structures to form? In contrast, sway-induced meandering structures (as seen in figure 11) are more resilient in turbulent conditions, suggesting that their formation is less dependent on near-wake tip vortices. However, Biswas & Buxton (Reference Biswas and Buxton2024a ) found that the hub vortex may also influence the development of meandering structures. These findings emphasise the need for further investigation into the local flow mechanisms behind the formation and stability of large-scale wake structures due to periodic excitations, and their relationship with near-wake flow structures, such as tip and hub vortices.

Furthermore, the origin of the energy feeding the excited large-scale structures remains a key question. Portela, Papadakis & Vassilicos (Reference Alves Portela, Papadakis and Vassilicos2020) showed that in the wake of a square obstacle, coherent structures are primarily energised by the mean flow. Analogous results were observed in the wake of a fixed wind turbine by Biswas & Buxton (Reference Biswas and Buxton2024b ), who investigated energy exchanges in the near wake and identified the mean flow as the primary source of erratic wake meandering. A similar mechanism is likely at play in our case. Other contributions – such as from tip vortices, hub vortices or even inflow turbulence with the hypothesis of a backward energy cascade – may also be found. While a full analysis of these energy exchanges lies beyond the present scope, these aspects merit further investigation.

Our line measurements are limited in providing a full 3-D picture of the coherent structures, especially the meandering mode. It remains uncertain whether this mode forms as a helical structure or resembles a more von Kármán-like vortex, primarily driven by lateral motion with minimal vertical displacement. Additionally, it remains unclear whether meandering and pulsating wake modes emerge from a common system of interacting helical vortex structures, or if they represent fundamentally different vortex dynamics.

In the broader fluid dynamic context of transition to turbulence, our study shows similarities with other means of transition to turbulence in a flow. Figure 14 suggests a behaviour qualitatively similar to that observed downstream of passive fractal grids (Valente & Vassilicos Reference Valente and Vassilicos2011; Hearst & Lavoie Reference Hearst and Lavoie2014) and active grids (Mora, Obligado & Vassilicos Reference Mora, Muñiz Pladellorens, Riera Turró, Lagauzere and Obligado2019; Neuhaus et al. Reference Neuhaus, Hölling, Bos and Peinke2020). In both cases, flow structures (periodic in our case) initially gain energy at a rate determined by factors such as the direction of motion (which influences the type of flow structure), excitation energy (e.g. forcing amplitude or grid motion) and the inflow turbulence intensity $\textit{TI}_{\infty }$ . As these structures accumulate energy while being advected downstream, they eventually become unstable, roughly at the point where the local turbulence level is highest, as observed in Messmer et al. (Reference Messmer, Hölling and Peinke2024a ). The destabilised, high-energy structures then decay, likely to small-scale eddies, feeding the turbulent cascade, similar to the mechanisms described by Mora et al. (Reference Mora, Muñiz Pladellorens, Riera Turró, Lagauzere and Obligado2019) behind an active grid. Although grid-generated turbulence typically begins with a non-turbulent inflow, the underlying phenomena remain qualitatively similar to what is observed in turbine wake dynamics. From this perspective, a row of floating wind turbines within a wind farm can be seen as an array of independently actuated elements – analogous to shafts in an active grid – each capable of modulating the flow and potentially accelerating the transition to turbulence in the whole wake, offering potential applications for cluster wake control, as outlined by Gutknecht et al. (Reference Gutknecht, Becker, Taschner, Stipa, Allaerts, Viré and Van Wingerden2024).

5.3. Implication for wind energy

The wake profiles and recovery results presented in figures 7 and 8 demonstrate that platform motion – or, more generally, periodic excitation strategies applied at the rotor level – can enhance wake recovery, particularly under low to moderate inflow turbulence conditions ( $ \textit{TI}_{\infty } \leqslant 3{-}4\,\%$ ). From a wind energy perspective, this implies that downstream turbines can benefit from increased mean flow velocities when placed behind an upstream rotor undergoing periodic excitation at a Strouhal number $ St \approx 0.3$ . For instance, a $ 3 \,\%$ increase in mean streamwise velocity ( $ \overline {u}$ ) was observed for sway excitation at $ \textit{TI}_{\infty } = 4.1\,\%$ (see figure 8 d). This corresponds to a $ 9 \,\%$ increase in available power ( $ \overline {u}^3$ ), eventually converted into electrical power by a potential downstream wind turbine, highlighting the industrial relevance of wake control strategies even under moderate turbulence intensity conditions.

Future studies should focus on optimising wake mixing at the wind farm scale. The current findings suggest that promoting wake meandering may be particularly effective in enhancing momentum recovery. This is especially relevant for floating wind turbines, which could be designed to naturally exhibit motion dynamics favouring wake mixing; e.g. oscillations near $ St \approx 0.3$ . However, the potential benefits must be carefully balanced against possible increases in structural loads. More broadly, the effectiveness of wake control strategies appears to diminish at higher inflow turbulence intensities ( $\textit{TI}_{\infty } \gt 5 \,\%$ ), where natural turbulent mixing already promotes sufficient wake recovery. In such conditions, active control may either be energetically inefficient or require derating to achieve meaningful gains.

Last, this study did not analyse the influence of the largest-scale structures in the inflow, characterised by the integral length scale $ \mathcal{L}_0$ , on wake dynamics and recovery. These structures have been shown to be an important additional parameter influencing wake behaviour (Gambuzza & Ganapathisubramani Reference Gambuzza and Ganapathisubramani2023; Bourhis & Buxton Reference Bourhis and Buxton2024; Bourhis et al. 2025; Hodgson, Troldborg & Andersen Reference Hodgson, Troldborg and Andersen2025). Given the variability of large-scale structures that wind farms might encounter (Hodgson et al. Reference Hodgson, Troldborg and Andersen2025), the interaction between $ \mathcal{L}_0$ and rotor-induced motion is of particular interest, especially in the context of wake meandering modelling. As discussed by Li et al. (Reference Li, Dong and Yang2022), two primary mechanisms are typically considered for wake meandering. The first is passive advection, where large-scale inflow eddies passing through the rotor (typically having a size $ \mathcal{L} \gt 2D$ , corresponding to $ St_{\mathcal{L}} = D / \mathcal{L} \lt 0.5$ ) displace the wake laterally and vertically, as described in the well-known dynamic wake meandering (DWM) model from Larsen et al. (Reference Larsen, Madsen, Thomsen and Larsen2008). The second is shear-flow-induced meandering, where natural instabilities in the wake give rise to organised large-scale motions. These instabilities may arise naturally within the shear layer or be externally excited, as demonstrated in the present study, where platform sway induces large coherent meandering structures (figure 11) at the platform motion frequency, $ f_p$ .

The forced meandering observed here is highly amplified: platform motions with amplitudes of only about $ 2\,\%D$ generate lateral structures of span $ \sim D$ , and the associated energy amplification is substantial. For example, figure 13(b) (b.2) shows that the total energy in the coherent structure $ E^*_{\textit{CS}}$ is approximately 30 times greater than the input excitation energy $ \epsilon ^*_p$ . This indicates a strong nonlinear response that is not captured by classical passive-advection-based models like DWM.

Our results suggest that, for inflow turbulence intensities $ \textit{TI}_{\infty } \lesssim 4\,\%$ , shear-flow-induced meandering may play a dominant role in wake dynamics, especially when the rotor experiences forced excitations. Incorporating such nonlinear amplification effects into engineering models – for example, through low-order resolvent-based frameworks such as that proposed by Li & Yang (Reference Li and Yang2024, Reference Li and Yang2025), or other methods – could improve predictions of structural loads and performance for downstream turbines. Finally, an improved understanding of the interplay between $ \mathcal{L}_0$ -driven meandering (i.e. passive advection, a primarily linear phenomenon) and shear-flow meandering (a nonlinear phenomenon), recently discussed by Bourhis et al. (Reference Bourhis, Messmer, Hölling and Buxton2025), is of great interest. For $ \textit{TI}_{\infty } \lesssim 4\,\%$ , wake meandering is likely driven by shear-flow instabilities, whereas increasing levels of $ \textit{TI}_{\infty }$ lead to a reduced impact of shear instabilities and an increased impact of the linear effect of passive advection. The overlap, interaction and dependence on $ C_T$ , $ \mathcal{L}_0$ , $\textit{TI}_{\infty }$ and potentially the wake excitation strategy remain open questions that are of interest for improving models and predictions.

Acknowledgements

The authors would like to thank J. Jüchter, J. Maus, J. Puczylowski, K. Silwal, T. Wester, L. Ledoux and A. Hölling for their valuable help with the experiments in Oldenburg. Special thanks to S. Goedeke and L. Neuhaus for the support with the active grid. They also thank C. Messmer for the precious help with the transport of the set-up from Oldenburg to Milan. The authors are thankful to M. Montenegro Montero and K. Yilmazlar for their great help with the experiments. Many thanks to L. Riccobene and D. Grassi for their technical support during the experiments in Milan.

The authors would like to acknowledge the assistance of ChatGPT (OpenAI) in refining the clarity and English language style of this paper. As non-native English speakers, this support helped improve the overall readability of the paper.

The authors wish to express their appreciation to the reviewers for their constructive critique and suggestions, which contributed meaningfully to the refinement of this work.

Lastly, the authors thank Martin Bourhis and Sandrine Aubrun for the helpful conversations over the past years and for their thoughtful input, which contributed to analyses and discussions.

Funding

This work has received funding from (i) the EU’s Horizon 2020 research and innovation programme under the Marie Sklodowska–Curie grant agreement N 860879, as part of the FLOAWER consortium; (ii) the FLORIDA project, funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – N 529978479.

Declaration of interests

The authors report no conflict of interest.

Appendix A. Phase-averaging methodology

We briefly present the methodology used to compute $\widetilde {u_i}$ (see 2.4), the phase-averaged component of the wind speed signal with a period equal to the platform excitation, $T_p = 1/f_p$ . This method applies to both the Milan dataset (3-D field) and the Oldenburg dataset (1-D field).

The goal is to compute $\widetilde {u_i}(\phi ^*)$ , a periodic function of a coherent mode at frequency $f_k$ , identified in the wake. The phase satisfies $\phi (t) = \phi (t + 1/f_k)$ and is rescaled to a normalised phase via $\phi ^* = (\phi (t) - \phi _0) \bmod 2 \pi$ . In this study we focus on $T_p = 1/f_p = D / (U_\infty St)$ , but the methodology is applicable for any $f_k$ for which the signal exhibits coherent energy (i.e. a distinct spectral peak; see Messmer et al. (Reference Messmer, Peinke and Hölling2024a ) for details).

To compute $\widetilde {u_i}(\phi ^*)$ , we implement the following steps (which we display in figure 17) using a Python algorithm that takes as input the fluctuating velocity field

(A1) \begin{align} u^{\prime}_i(x, y, z, t) = u_i(x, y, z, t) - \overline {u_i(x, y, z, t)}, \quad \text{with } i = x, y, z .\end{align}

  1. (i) Step 1: apply a low-pass filter to $u^{\prime}_i$ with cutoff frequency $f_{{cut}} \approx 3 f_k$ to isolate coherent fluctuations. Denote the filtered signal as $u_i^{\textit{filt}}$ .

  2. (ii) Step 2: segment $u_i^{\textit{filt}}$ into bins of size $n_k = f_s / f_k$ , where $f_s$ is the sampling frequency. This results in approximately $N_{\textit{seg}} = T_{{meas}} \boldsymbol{\cdot }f_k$ segments, each corresponding to one period of the coherent mode. Since $N_{\textit{seg}} \gg 10$ in all cases, phase averaging is statistically meaningful. The segmented signal is (see as an example the 20 segments in figure 17 a):

    (A2) \begin{align} u_i^{\textit{filt}} =\ & \big [ u_i^{\textit{filt}} (1),\ u_i^{\textit{filt}} (2),\ \ldots ,\ u_i^{\textit{filt}} (n_k) \big ] + \end{align}
    (A3) \begin{align} & \big [ u_i^{\textit{filt}} (n_k + 1),\ \ldots ,\ u_i^{\textit{filt}} (2n_k) \big ] + \end{align}
    (A4) \begin{align} & {\cdots} + \end{align}
    (A5) \begin{align} & \big [ u_i^{\textit{filt}} (N_{\textit{seg}} \boldsymbol{\cdot }n_k),\ \ldots ,\ u_i^{\textit{filt}} ((N_{\textit{seg}} + 1) \boldsymbol{\cdot }n_k ) \big ] .\end{align}
  3. (iii) Step 3: compute the phase-averaged signal over all periods (figure 17 b):

    (A6) \begin{align} \widetilde {u_i^{\textit{filt}}} = \Bigg [ & \big ( u_i^{\textit{filt}} (1) + u_i^{\textit{filt}} (n_k + 1) + \cdots + u_i^{\textit{filt}} (N_{\textit{seg}} \boldsymbol{\cdot }n_k + 1) \big ) / N_{\textit{seg}}, \end{align}
    (A7) \begin{align} & \big ( u_i^{\textit{filt}} (2) + u_i^{\textit{filt}} (n_k + 2) + \cdots + u_i^{\textit{filt}} (N_{\textit{seg}} \boldsymbol{\cdot }n_k + 2) \big ) / N_{\textit{seg}}, \end{align}
    (A8) \begin{align} & \ \ldots , \end{align}
    (A9) \begin{align} & \big ( u_i^{\textit{filt}} (n_k) + u_i^{\textit{filt}} (2n_k) + \cdots + u_i^{\textit{filt}} ((N_{\textit{seg}} + 1) \boldsymbol{\cdot }n_k ) \big ) / N_{\textit{seg}} \Bigg ]. \end{align}
  4. (iv) Finally, we define the discrete phase-averaged signal as $\widetilde {u_i}(\phi _k)$ with $\phi _k \in [\![ 0, n_k ]\!]$ , corresponding to $\phi ^* \in [0, 2\pi ]$ sampled at $n_k$ points per cycle. The decomposition of (2.4) is displayed in figure 17(c).

Figure 17. Visualisation of phase-averaging methodology. (a) Wake wind speed time series of the floating wind turbine partitioned in bins of size $f_s/f_p$ , giving one bin per period $T_p = 1/f_p$ (step 2). (b) Superposition of the $N_{seg}$ bins and phase averaged signal (step 3). (c) Raw signal decomposed as the sum of the mean, phase-averaged and stochastic fluctuations (2.4).

Appendix B. Comparison of Milan and Oldenburg data and additional details about the computation of $\boldsymbol{\partial / \partial x_i}$ terms and error estimation

In § 3.1 we use two datasets (corresponding to the cases fixed-5, surge-5 and sway-5 in table 2) to compute the recovery rate budget terms of (2.6). These datasets were obtained from experiments conducted in two different wind tunnels – Milan and Oldenburg – but great care was taken to replicate identical inflow and operating conditions across both facilities. To validate the comparability of the datasets, we present in figure 18(a) comparison of wake profiles of the mean streamwise wind speed (figure 18 ae) and streamwise velocity fluctuations (figure 18 fj), both normalised by the free-stream wind speed. Figure 18 ae show good agreement between the mean wake velocities measured in Milan (solid lines) and Oldenburg (semi-transparent dashed lines) across all three motion cases. The inter-experimental discrepancy, defined as the relative difference between datasets, $\mid (\overline {u}_{{M}} - \overline {u}_{{O}}) / \overline {u}_{{M}} \mid$ , reaches a maximum of approximately $10 \,\%$ at $ x = 6D$ (figure 18 c), while remaining below $5 \,\%$ at most other downstream locations. For the velocity fluctuations (figures 18 fj), the relative difference between the two experiments is generally within $5 \,\%$ , indicating a satisfactory level of consistency between the independently acquired datasets. To reflect the inter-experimental uncertainty between the two independent datasets, error bars corresponding to $\pm 5\,\%$ were added to figure 4(a). Note that the intrinsic measurement uncertainty of hot-wire anemometry for an individual measurement, $\epsilon _{\textit{HW}}$ , is considerably lower, estimated at approximately $\epsilon _{\textit{HW}} \sim 2 \,\%$ in Neunaber (Reference Neunaber2019) for similar experiments.

Figure 18. Comparison of Milan and Oldenburg datasets (fixed-5, surge-5 and sway-5 cases in table 2). (ae) Horizontal profiles of the mean streamwise wind speed at $x \in [2D, 10D]$ . (fj) Horizontal profiles of the streamwise wind speed fluctuations at $x \in [2D, 10D]$ . Both quantities are normalised by the free-stream velocity.

It is worth noting that the experiments in Milan were conducted on different days and even in different months depending on the downstream position: measurements at $x = 2D$ , $6D$ and $10D$ were taken in March 2022, while those at $4D$ and $8D$ were performed in June 2022. This required resetting the experiment each time, resulting in slight variations in turbine and probe positioning as well as operating conditions. In contrast, the experiments in Oldenburg for all three cases were performed consecutively on the same day for $x \in [1.5D, 10D]$ under consistent operating conditions. As such, we expect the Oldenburg data to exhibit greater stability in the mean wind speed. The close agreement in velocity fluctuations between the Milan and Oldenburg datasets further validates our approach – namely, using Oldenburg data for recovery computation and Milan data for Reynolds stress terms – as the comparison of profiles in figure 18 fj indicates that the essential turbulent features are robust to minor experimental variations.

As a point of comparison, Aubrun et al. (Reference Aubrun2019) evaluated wake measurements of fixed porous and actuator disks across nine wind tunnel facilities worldwide under nominally identical inflow conditions. Despite the simplicity and standardisation of these tests, differences of up to 15 % were observed between facilities. These discrepancies are often attributed to numerous factors, including probe positioning, probe type and calibration, operating conditions (e.g. inflow turbulence intensity $\textit{TI}_\infty$ , thrust coefficient $C_T$ ), yaw misalignment and reference wind speed. Even slight differences in these parameters between two experiments can cumulatively lead to variations in the results.

Given these considerations and looking at our results, the Milan and Oldenburg datasets are sufficiently consistent to be jointly analysed in this study.

The different terms in (2.6) involve derivatives of the mean flow field and Reynolds stresses in the $x$ , $y$ and $z$ directions. To compute these terms, we used the Milan dataset for those involving $\partial / \partial y$ and $\partial / \partial z$ . Specifically, we applied the numpy.gradient (v2.1 at the time of the analysis) function in Python to the wake field, where the spatial resolution is $\Delta y, \Delta z = 0.34R$ – the spacing between measurement points (see figure 2 a). For derivatives in the streamwise direction, $\partial / \partial x$ , we used the Oldenburg dataset, which provides data points at intervals of half a rotor diameter, i.e. $\Delta x = 0.5D$ .

As mentioned in the manual of numpy.gradient: ‘the gradient is computed using second order accurate central differences in the interior points and either first or second order accurate one-sides (forward or backwards) differences at the boundaries’. For the interior points (i.e. $x \in ]2,10[D, \{y,z \} \in ]-1.34,1.34[R,$ ), the derivative of a given $\overline {g}(x,y,z)$ is computed as

(B1) \begin{align} \frac {\partial \overline {g}}{\partial x} &\approx \frac {\overline {g}(x + \Delta x, y, z) - \overline {g}(x - \Delta x, y, z)}{2 \Delta x} + \mathcal{O}(\Delta x^2),\quad \Delta x= 0.5D,\,\,\, \end{align}
(B2) \begin{align} \frac {\partial \overline {g}}{\partial y} &\approx \frac {\overline {g}(x, y + \Delta y, z) - \overline {g}(x, y - \Delta y, z)}{2\Delta y} + \mathcal{O}(\Delta y^2),\quad\Delta y=0.17D ,\end{align}
(B3) \begin{align} \frac {\partial \overline {g}}{\partial z} &\approx \frac {\overline {g}(x, y, z + \Delta z) - \overline {g}(x, y, z - \Delta z)}{2\Delta z} + \mathcal{O}(\Delta z^2), \quad\, \Delta z=0.17D. \end{align}

The resulting recovery rate budget shown in figure 6 balances well, especially in the far wake ( $x \geqslant 8D$ , where the pressure and rotation terms are null), supporting the assumption that the spatial resolution in all three directions is sufficient. Thus, we consider the finite-difference approximation $\partial \overline {g} / \partial x_i \approx \Delta \overline {g} / \Delta x_i$ to be appropriate for our computations.

To quantify the uncertainty in spatial derivatives, we employed a Monte Carlo uncertainty propagation method, similar to the approach of Rezaeiravesh et al. (Reference Rezaeiravesh, Vinuesa, Liefvendahl and Schlatter2018). A total of $ N_{\textit{MC}} \gt 10^3$ synthetic realisations of each measured quantity $ \overline {g}(x,y,z)$ were generated by perturbing the original data with Gaussian noise. The noise amplitude corresponded to a typical hot-wire measurement uncertainty of $ \epsilon _{{HW}} = 2\,\%$ , consistent with values reported in previous experimental studies (e.g. (Neunaber Reference Neunaber2019)). For each realisation, the gradient $ \partial \overline {g} / \partial x_i$ was computed using the numpy.gradient function. The standard deviation across the ensemble of synthetic gradients was then taken as the estimated uncertainty in the derivative. On average, the resulting uncertainty in the various mean flow and Reynolds stress derivative terms (e.g. $ \partial \overline {u} / \partial x$ , $ \partial \overline {u'v'} / \partial y$ , etc.) is approximately $ \epsilon _{\partial / \partial x_i} = 15\,\%$ , as shown in figure 4(b) for the example of $ D\langle \partial \overline {u} / \partial x \rangle$ . This Monte Carlo-based uncertainty estimate supports the interpretation of the derivative terms presented in figures 5 and 6, which can be assumed to carry an error margin of approximately $ \epsilon _{\partial / \partial x_i} = 15\,\%$ .

References

Alves Portela, F., Papadakis, G. & Vassilicos, J.C. 2020 The role of coherent structures and inhomogeneity in near-field interscale turbulent energy transfers. J. Fluid Mech. 896, A16.10.1017/jfm.2020.341CrossRefGoogle Scholar
Angelou, N., Mann, J. & Dubreuil-Boisclair, C. 2023 Revealing inflow and wake conditions of a 6 MW floating turbine. Wind Energy Sci. 8 (10), 15111531.10.5194/wes-8-1511-2023CrossRefGoogle Scholar
Aubrun, S. et al. 2019 Round-robin tests of porous disc models. J. Phys. Conf. Ser. 1256, 012004.10.1088/1742-6596/1256/1/012004CrossRefGoogle Scholar
Baj, P., Bruce, P.J.K. & Buxton, O.R.H. 2015 The triple decomposition of a fluctuating velocity field in a multiscale flow. Phys. Fluids 27 (7), 075104-1–075104-24.10.1063/1.4923744CrossRefGoogle Scholar
Bastankhah, M. & Porté-Agel, F. 2014 A new analytical model for wind-turbine wakes. Renew. Energy 70, 116123.10.1016/j.renene.2014.01.002CrossRefGoogle Scholar
Bastankhah, M. & Porté-Agel, F. 2016 Experimental and theoretical study of wind turbine wakes in yawed conditions. J. Fluid Mech. 806, 506541.10.1017/jfm.2016.595CrossRefGoogle Scholar
Biswas, N. & Buxton, O.R.H. 2024 a Effect of tip speed ratio on coherent dynamics in the near wake of a model wind turbine. J. Fluid Mech. 979, A34.10.1017/jfm.2023.1095CrossRefGoogle Scholar
Biswas, N. & Buxton, O.R.H. 2024 b Energy exchanges between coherent modes in the near wake of a wind turbine model at different tip speed ratios. J. Fluid Mech. 996, A8.10.1017/jfm.2024.581CrossRefGoogle Scholar
Bossuyt, J., Ferčák, O., Sadek, Z., Meneveau, C., Gayme, D. & Cal, R.B. 2023 Floating wind farm experiments through scaling for wake characterization, power extraction, and turbine dynamics. Phys. Rev. Fluids. 8 (12), 120501-1–120501-39.10.1103/PhysRevFluids.8.120501CrossRefGoogle Scholar
Boudreau, M. & Dumas, G. 2017 Comparison of the wake recovery of the axial-flow and cross-flow turbine concepts. J. Wind Eng. Ind. Aerodyn. 165, 137152.10.1016/j.jweia.2017.03.010CrossRefGoogle Scholar
Bourhis, M. & Buxton, O.R.H. 2024 Influence of freestream turbulence and porosity on porous disk-generated wakes. Phys. Rev. Fluids 9 (12), 124501.10.1103/PhysRevFluids.9.124501CrossRefGoogle Scholar
Bourhis, M., Messmer, T., Hölling, M. & Buxton, O.R.H. 2025 Impact of freestream turbulence and thrust coefficient on wind turbine-generated wakes. arXiv: 2503.21903.Google Scholar
Cannon, S., Champagne, F. & Glezer, A. 1993 Observations of large-scale structures in wakes behind axisymmetric bodies. Exp. Fluids 14 (6), 447450.10.1007/BF00190199CrossRefGoogle Scholar
Cantwell, B. & Coles, D. 1983 An experimental study of entrainment and transport in the turbulent near wake of a circular cylinder. J. Fluid Mech. 136, 321374.10.1017/S0022112083002189CrossRefGoogle Scholar
Cheung, L.C., Brown, K.A., Houck, D.R. & Develder, N.B. 2024 Fluid-dynamic mechanisms underlying wind turbine wake control with Strouhal-timed actuation. Energies 17 (4), 865.10.3390/en17040865CrossRefGoogle Scholar
Croce, A., Cacciola, S., Montero Montenegro, M., Stipa, S. & Praticó, R. 2023 A CFD‐based analysis of dynamic induction techniques for wind farm control applications. Wind Energy 26 (3), 325343.10.1002/we.2801CrossRefGoogle Scholar
Crow, S.Cj & Champagne, F.H. 1971 Orderly structure in jet turbulence. J. Fluid Mech. 48 (3), 547591.10.1017/S0022112071001745CrossRefGoogle Scholar
da Silva, C.B., Hunt, J.C.R., Eames, I. & Westerweel, J. 2014 Interfacial layers between regions of different turbulence intensity. Annu. Rev. Fluid Mech. 46, 567590.10.1146/annurev-fluid-010313-141357CrossRefGoogle Scholar
Fontanella, A., Fusetti, A., Cioni, S., Papi, F., Muggiasca, S., Persico, G., Dossena, V., Bianchini, A. & Belloli, M. 2025 Wake development in floating wind turbines: new insights and an open dataset from wind tunnel experiments. Wind Energy Sci. 10 (7), 13691387.10.5194/wes-10-1369-2025CrossRefGoogle Scholar
Fontanella, A., Zasso, A. & Belloli, M. 2022 Wind tunnel investigation of the wake-flow response for a floating turbine subjected to surge motion. J. Phys.: Conf. Ser. 2265, 042023.Google Scholar
Frederik, J., Doekemeijer, B., Mulders, S. & van Wingerden, J. 2020 The helix approach: using dynamic individual pitch control to enhance wake mixing in wind farms. Wind Energy 23 (8), 17391751.10.1002/we.2513CrossRefGoogle Scholar
Fu, S., Jin, Y., Zheng, Y. & Chamorro, L.P. 2019 Wake and power fluctuations of a model wind turbine subjected to pitch and roll oscillations. Appl. Energy 253, 113605.10.1016/j.apenergy.2019.113605CrossRefGoogle Scholar
Fuentes Noriega, E. & Mazellier, N. 2025 Scaling analysis of the swirling wake of a porous disc: application to wind turbines. J. Fluid Mech. 1003, A34.10.1017/jfm.2024.1215CrossRefGoogle Scholar
Gambuzza, S. & Ganapathisubramani, B. 2023 The influence of free stream turbulence on the development of a wind turbine wake. J. Fluid Mech. 963, A19.10.1017/jfm.2023.302CrossRefGoogle Scholar
Gupta, V. & Wan, M. 2019 Low-order modelling of wake meandering behind turbines. J. Fluid Mech. 877, 534560.10.1017/jfm.2019.619CrossRefGoogle Scholar
Gutknecht, J., Becker, M., Taschner, E., Stipa, S., Allaerts, D., Viré, A. & Van Wingerden, J.-W. 2024 Active cluster wake mixing. J. Phys.: Conf. Ser. 2767, 092052.Google Scholar
Hearst, R.J. & Lavoie, P. 2014 Decay of turbulence generated by a square-fractal-element grid. J. Fluid Mech. 741, 567584.10.1017/jfm.2013.684CrossRefGoogle Scholar
Ho, C.M. & Huerre, P. 1984 Perturbed free shear layers. Annu. Rev. Fluid Mech. 16, 365424.10.1146/annurev.fl.16.010184.002053CrossRefGoogle Scholar
Hodgson, E.L., Madsen, M. H.A. & Andersen, S.J. 2023 Effects of turbulent inflow time scales on wind turbine wake behavior and recovery. Phys. Fluids 35 (9), 095125-1–095125-21.10.1063/5.0162311CrossRefGoogle Scholar
Hodgson, E.L., Troldborg, N. & Andersen, S.J. 2025 Impact of freestream turbulence integral length scale on wind farm flows and power generation. Renew. Energy 238, 121804.10.1016/j.renene.2024.121804CrossRefGoogle Scholar
Hubert, A., Conan, B. & Aubrun, S. 2025 Spatiotemporal behavior of the far wake of a wind turbine model subjected to harmonic motions: phase averaging applied to stereo particle image velocimetry measurements. Wind Energy Sci. 10 (7), 13511368.10.5194/wes-10-1351-2025CrossRefGoogle Scholar
Hussain, A.K.M.F. 1986 Coherent structures and turbulence. J. Fluid Mech. 173, 303356.10.1017/S0022112086001192CrossRefGoogle Scholar
Kankanwadi, K.S. & Buxton, O.R.H. 2020 Turbulent entrainment into a cylinder wake from a turbulent background. J. Fluid Mech. 905, A35.10.1017/jfm.2020.755CrossRefGoogle Scholar
Kleine, V.G., Franceschini, L., Carmo, B.S., Hanifi, A. & Henningson, D.S. 2022 The stability of wakes of floating wind turbines. Phys. Fluids 34 (7), 074106-1–074106-21.10.1063/5.0092267CrossRefGoogle Scholar
Korb, H., Asmuth, H. & Ivanell, S. 2023 The characteristics of helically deflected wind turbine wakes. J. Fluid Mech. 965, A2.10.1017/jfm.2023.390CrossRefGoogle Scholar
Larsen, G.C., Madsen, H.A., Thomsen, K. & Larsen, T.J. 2008 Wake meandering: a pragmatic approach. Wind Energy, 11 (4), 377395.10.1002/we.267CrossRefGoogle Scholar
Li, Y.T., Yu, W. & Sarlak, H. 2024 Wake structures and performance of wind turbine rotor with harmonic surging motions under laminar and turbulent inflows. Wind Energy 27 (12), 14991525.10.1002/we.2949CrossRefGoogle Scholar
Li, Z., Dong, G. & Yang, X. 2022 Onset of wake meandering for a floating offshore wind turbine under side-to-side motion. J. Fluid Mech. 934, A29.10.1017/jfm.2021.1147CrossRefGoogle Scholar
Li, Z. & Yang, X. 2024 Resolvent-based motion-to-wake modelling of wind turbine wakes under dynamic rotor motion. J. Fluid Mech. 980, A48.10.1017/jfm.2023.1097CrossRefGoogle Scholar
Li, Z. & Yang, X. 2025 Self-consistent model for active control of wind turbine wakes. J. Fluid Mech. 1013, A36.10.1017/jfm.2025.10263CrossRefGoogle Scholar
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. J. Fluid Mech. 781, 467493.10.1017/jfm.2015.470CrossRefGoogle Scholar
Mao, X. & Sørensen, J.N. 2018 Far-wake meandering induced by atmospheric eddies in flow past a wind turbine. J. Fluid Mech. 846, 190209.10.1017/jfm.2018.275CrossRefGoogle Scholar
Messmer, T., Brigden, C., Peinke, J. & Hölling, M. 2022 A six degree-of-freedom set-up for wind tunnel testing of floating wind turbines. J. Phys.: Conf. Ser. 2265, 042015.Google Scholar
Messmer, T., Hölling, M. & Peinke, J. 2024 a Enhanced recovery caused by nonlinear dynamics in the wake of a floating offshore wind turbine. J. Fluid Mech. 984, A66.10.1017/jfm.2024.175CrossRefGoogle Scholar
Messmer, T., Peinke, J. & Hölling, M. 2024 b Wind tunnel investigation on the recovery and dynamics of the wake of a floating offshore wind turbine subjected to low inflow turbulence. J. Phys.; Conf. Ser 2767, 092083.10.1088/1742-6596/2767/9/092083CrossRefGoogle Scholar
Mora, D.O., Muñiz Pladellorens, E., Riera Turró, P., Lagauzere, M. & Obligado, M. 2019 Energy cascades in active-grid-generated turbulent flows. Phys. Rev. Fluids 4 (5), 054602.10.1103/PhysRevFluids.4.104601CrossRefGoogle Scholar
Mühle, F.V., Tamaro, S., Klinger, F., Campagnolo, F. & Bottasso, C.L. 2024 Experimental and numerical investigation on the potential of wake mixing by dynamic yaw for wind farm power optimization. J. Phys.: Conf. Ser. 2767, 092068.Google Scholar
Munters, W. & Meyers, J. 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 Sci. 3 (1), 409425.10.5194/wes-3-409-2018CrossRefGoogle Scholar
Neuhaus, L., Berger, F., Peinke, J. & Hölling, M. 2021 Exploring the capabilities of active grids. Exp. Fluids 62 (6), 130.10.1007/s00348-021-03224-5CrossRefGoogle Scholar
Neuhaus, L., Hölling, M., Bos, W.J. T. & Peinke, J. 2020 Generation of atmospheric turbulence with unprecedentedly large Reynolds number in a wind tunnel. Phys. Rev. Lett. 125 (15), 154503.10.1103/PhysRevLett.125.154503CrossRefGoogle Scholar
Neunaber, I. 2019 Stochastic investigation of the evolution of small-scale turbulence in the wake of a wind turbine exposed to different inflow conditions. PhD thesis, Universität Oldenburg, Germany.Google Scholar
Neunaber, I., Hölling, M., Stevens, R.J.A.M., Schepers, G. & Peinke, J. 2020 Distinct turbulent regions in the wake of a wind turbine and their inflow-dependent locations: the creation of a wake map. Energies 13 (20), 5392.10.3390/en13205392CrossRefGoogle Scholar
Oster, D. & Wygnanski, I. 1982 The forced mixing layer between parallel streams. J. Fluid Mech. 123, 91130.10.1017/S0022112082002973CrossRefGoogle Scholar
Pagamonci, L., Papi, F., Cojocaru, G., Belloli, M. & Bianchini, A. 2025 How does turbulence affect wake development in floating wind turbines? A critical assessment. Wind Energy Sci. Discuss. 2025, 136.Google Scholar
Pikovsky, A., Rosenblum, M. & Kurths, J. 2001, Synchronization. Cambridge University Press.10.1017/CBO9780511755743CrossRefGoogle Scholar
Platis, A., Hundhausen, M., Lampert, A., Emeis, S. & Bange, J. 2021 The role of atmospheric stability and turbulence in offshore wind-farm wakes in the German bight. Boundary Layer Meteorol. 129.10.1007/s10546-021-00668-4CrossRefGoogle Scholar
Pope, S.B. 2001 Turbulent flows. Meas. Sci. Technol. 12 (11), 20202021.10.1088/0957-0233/12/11/705CrossRefGoogle Scholar
Porté-Agel, F., Bastankhah, M. & Shamsoddin, S. 2020 Wind-turbine and wind-farm flows: a review. Boundary-Layer Meteorol. 174 (1), 159.10.1007/s10546-019-00473-0CrossRefGoogle ScholarPubMed
Ramos‐García, N., Kontos, S., Pegalajar‐Jurado, A., González Horcas, S. & Bredmose, H. 2022 Investigation of the floating IEA Wind 15 MW RWT using vortex methods. Part I: Flow regimes and wake recovery. Wind Energy 25 (3), 468504.10.1002/we.2682CrossRefGoogle Scholar
Reynolds, W.C. & Hussain, A.K.M.F. 1972 The mechanics of an organized wave in turbulent shear flow. Part 3. Theoretical models and comparisons with experiments. J. Fluid Mech. 54 (2), 263288.10.1017/S0022112072000679CrossRefGoogle Scholar
Rezaeiravesh, S., Vinuesa, R., Liefvendahl, M. & Schlatter, P. 2018 Assessment of uncertainties in hot-wire anemometry and oil-film interferometry measurements for wall-bounded turbulent flows. Eur. J. Mech. B/Fluids 72, 5773.10.1016/j.euromechflu.2018.04.012CrossRefGoogle Scholar
Robertson, A., Jonkman, J., Masciola, M., Song, H, Goupee, A., Coulling, A. & Luan, C. 2014 Definition of the semisubmersible floating system for phase II of OC4. Tech. Rep, National Renewable Energy Lab (NREL).10.2172/1155123CrossRefGoogle Scholar
Rockel, S., Peinke, J., Hölling, M. & Cal, R.A.B. 2017 Dynamic wake development of a floating wind turbine in free pitch motion subjected to turbulent inflow generated with an active grid. Renew. Energy 112, 116.10.1016/j.renene.2017.05.016CrossRefGoogle Scholar
Schliffke, B., Conan, B. & Aubrun, S. 2024 Floating wind turbine motion signature in the far-wake spectral content – a wind tunnel experiment. Wind Energy Sci. 9 (3), 519532.10.5194/wes-9-519-2024CrossRefGoogle Scholar
Valente, P.C. & Vassilicos, J.C. 2011 The decay of turbulence generated by a class of multiscale grids. J. Fluid Mech. 687, 300340.10.1017/jfm.2011.353CrossRefGoogle Scholar
van den Berg, D., De Tavernier, D. & van Wingerden, J.-W. 2023 The dynamic coupling between the pulse wake mixing strategy and floating wind turbines. Wind Energy Sci. 8 (5), 849864.10.5194/wes-8-849-2023CrossRefGoogle Scholar
van der Hoek, D., den Abbeele, B.V., Simao Ferreira, C. & van Wingerden, J‐W. 2024 Maximizing wind farm power output with the helix approach: experimental validation and wake analysis using tomographic particle image velocimetry. Wind Energy 27 (5), 463482.10.1002/we.2896CrossRefGoogle Scholar
van der Laan, M.P., Baungaard, M. & Kelly, M. 2023 Brief communication: a clarification of wake recovery mechanisms. Wind Energy Sci. 8 (2), 247254.10.5194/wes-8-247-2023CrossRefGoogle Scholar
Wei, N.J., El Makdah, A., Hu, J.C., Kaiser, F., Rival, D.E. & Dabiri, J.O. 2024 Wake dynamics of wind turbines in unsteady streamwise flow conditions. J. Fluid Mech. 1000, A66.10.1017/jfm.2024.999CrossRefGoogle Scholar
Wygnanski, I., Champagne, F. & Marasli, B. 1986 On the large-scale structures in two-dimensional, small-deficit, turbulent wakes. J. Fluid Mech. 168, 3171.10.1017/S0022112086000289CrossRefGoogle Scholar
Yılmaz, A.E. & Meyers, J. 2018 Optimal dynamic induction control of a pair of inline wind turbines. Phys. Fluids 30 (8), 085106-1–085106-20.10.1063/1.5038600CrossRefGoogle Scholar
Figure 0

Figure 1. Experimental set-up: MoWiTO 0.6 mounted on the 6-degrees of freedom (DoF) motorised platform. (a) Wind tunnel in Milan. (b) Wind tunnel in Oldenburg (with active grid).

Figure 1

Figure 2. (a) Measurement points $y$-$z$ positions. (b) Picture of surge motion (fore-aft to streamwise). (c) Picture of sway motion (sideways to streamwise).

Figure 2

Table 1. Inflow cases investigated. Here M stands for Milan and O for Oldenburg.

Figure 3

Figure 3. (a) Horizontal profile of turbulence intensity at $x=0D$ (measurement points and mean $\textit{TI}_{\infty }$ plotted with solid lines). (b) Power spectrum of $u^{\prime}_{\infty }$, the turbulent fluctuations of the inflow cases O1.1 to O5.8 in table 1 (with $U_{\infty } = 3$ m s–1).

Figure 4

Table 2. Matrix of motion cases and inflows. Here DoF stands for degree of freedom.

Figure 5

Figure 4. (a) Wake recovery (2.2) for the fixed-5-O, surge-5-O and sway-5-O cases ($St = 0.38$, $A^* = 0.01$). Error bars represent the average discrepancy between the Milan and Oldenburg datasets (Appendix B). (b) Wake recovery rate. Error bars represent the estimated uncertainty based on a Monte Carlo approach (Appendix B). Both quantities are derived from the Oldenburg dataset.

Figure 6

Figure 5. The $y$ profiles of the terms in the recovery rate budget (2.6) for the fixed-5 (a,d), surge-5 (b,e) and sway-5 (c,f) cases. Panels (ac) correspond to $x = 4D$ and panels (df) to $x = 8D$. In the legend, (M) indicates terms computed using Milan data, while (O) indicates those based on Oldenburg data.

Figure 7

Figure 6. Downstream evolution of recovery equation terms (2.6) integrated over the rotor diameter for fixed-5 (a), surge-5 (b) and sway-5 (c). Panels (df) show $x$$y$ Reynolds shear stress gradients; panels (gi) show $x$$z$ gradients for the same cases.

Figure 8

Figure 7. The $y$ profiles of normalised wake deficit (a) and turbulence intensity (b) for fixed, surge- and sway- 3–0.01 and 0.024, i.e $St = 0.3, A^* = 0.01,\,0.024$ (see table 2) for O1.1 (a–0)–(a–4), O3.0 (b–0)–(b–4), O5.8 (c–0)–(c–4) and $x \in [2,10]D$ (0,4). Inflows in table 1.

Figure 9

Figure 8. Wake recovery from (2.2) for different free-stream turbulence levels: $\textit{TI}_{\infty } = 1.1\,\% \!-\!5.8\,\%$ (cases O1.1–O5.8 in table 1). Results include fixed ($St = 0$), surge and sway motions at $St = 0.3$ with $A^* \in \{0.01, 0.017, 0.024\}$ (table 2).

Figure 10

Figure 9. Offset-adjusted wake recovery: $R^*_w(x) = (R_w(x) - R_w(x=2D))+0.5$ ($R_w$ defined in (2.2) for different free-stream turbulence levels: $\textit{TI}_{\infty } = 1.1\,\%\!-\!5.8\,\%$ (cases O1.1–O5.8 in table 1). Panel (a) shows the fixed case ($St = 0$) and panel (b) sway with $St = 0.3$ and $A^* = 0.024$ (table 2).

Figure 11

Figure 10. (a) Sum of the mean wake flow and the surge-induced coherent fluctuation. (b) Surge-induced coherent fluctuation alone. All velocities are normalised by the free-stream velocity $U_{\infty }$. Columns (-.0)–(-.4) correspond to streamwise positions $x/D \in \{2, 4, 6, 8, 10\}$. Each panel represents one full period of the motion, shown over the normalised phase $\phi ^* \in [0, 2\pi ]$. Rows correspond to different inflow turbulence levels: (a.x) $\textit{TI}_{\infty } = 1.1\,\%$, (b.x) $\textit{TI}_{\infty } = 3.0\,\%$ and (c.x) $\textit{TI}_{\infty } = 5.8\,\%$, corresponding to cases O1.1, O3.0 and O5.8 in table 1. The motion parameters are $St = 0.3$ and $A^* = 0.024$ (case surge-3-0.024 in table 2).

Figure 12

Figure 11. (a) Sum of the mean wake flow and the sway-induced coherent fluctuation. (b) Sway-induced coherent fluctuation alone. All velocities are normalised by the free-stream velocity $U_{\infty }$. Columns (-.0)–(-.4) correspond to streamwise positions $x/D \in \{2, 4, 6, 8, 10\}$. Each panel represents one full period of the motion, shown over the normalised phase $\phi ^* \in [0, 2\pi ]$. Rows correspond to different inflow turbulence levels: (a.x) $\textit{TI}_{\infty } = 1.1\,\%$, (b.x) $\textit{TI}_{\infty } = 3.0\,\%$ and (c.x) $\textit{TI}_{\infty } = 5.8\,\%$, corresponding to cases O1.1, O3.0 and O5.8 in table 1. The motion parameters are $St = 0.3$ and $A^* = 0.024$ (case sway-3-0.024 in table 2).

Figure 13

Figure 12. Comparison of time-averaged and phase-averaged wake recovery. (a) Time-averaged and local (phase-averaged) streamwise wind speed $y$ profiles at $x = 6D$ for selected phases $\phi ^* \in \{0, 2\pi , 4\pi , 6\pi , 8\pi \}/5$, for both surge (ae) and sway (fj) motions, with $St = 0.3$, $A^* = 0.024$ and $\textit{TI}_{\infty } = 1.1\,\%$. (b) Local wake recovery $\widetilde {R_w}(\phi ^*)$ computed across the phase range $\phi ^*/2\pi \in [0,1]$. The plot includes the mean of the phase-averaged recovery $\overline {\widetilde {R_w}(\phi ^*)}$, the time-averaged recovery $R_w$ (also for the fixed reference case), shown for both surge (1) and sway (2) cases.

Figure 14

Figure 13. Integrated energy of surge- and sway-induced coherent structures for cases surge-5 and sway-5 (see table 2). (a) Comparison of the integrated energy of the streamwise component of the coherent fluctuations $E^*_{x,CS}$ (3.5) for surge and sway with $St = 0.38$, $A^* = 0.01$, using Milan and Oldenburg datasets ($\textit{TI}_{\infty } = 1.5\,\%$). (b) Integrated energy of each component of the coherent structure: streamwise $\overline {\widetilde {u}^2}$, lateral $\overline {\widetilde {v}^2}$ and vertical $\overline {\widetilde {w}^2}$ velocities, as well as their total noted $E^*_{x,CS},\,E^*_{y,CS}, E^*_{z,CS}$ and $E^*_{\textit{CS}}$, respectively, for surge (b.1) and sway (b.2). Panels (b.3) and (b.4) show the same energy components normalised by their respective maximum values.

Figure 15

Figure 14. Integrated energy of the streamwise component of the coherent structures induced by motions $E^*_{x,CS}$ for surge (a) and sway (b) both with $St = 0.3,\,A^*=0.024$ (surge- and sway-3–0.024 in table 2) for $\textit{TI}_{\infty } \in [1.1, 5.8]\,\%$ (O1.1–O5.8 in table 1). The red symbols represent the normalised energy of the free-stream fluctuations $\epsilon ^*_{\infty }$ (3.6) and the dashed line shows $\epsilon ^*_{p}$, the normalised energy of the platform motion (3.7).

Figure 16

Figure 15. Streamwise-integrated energy of the coherent structures in the $ x$ direction, defined as $ \int E^*_{x,CS} = ({1}/{8D}) \int _{x=2D}^{x=10D} E^*_{x,CS}(x) \, {\rm d}x$, plotted against the inflow turbulence intensity $ \textit{TI}_{\infty }$ for (a) surge and (b) sway. All cases use $ St = 0.3$ and $ A^* \in \{0.01,\,0.017,\,0.024\}$ in table 2 and inflows O1.1–O5.8 in table 1. Here red :, –,- lines represent $\epsilon ^*_p$ defined in (3.7).

Figure 17

Figure 16. Relative increase in wake recovery, $ \Delta \mathcal{R} / \mathcal{R}$ (3.10), plotted against the ratio $ \epsilon ^*_{\infty } / \epsilon ^*_p$, which compares the normalised energy of free-stream fluctuations to that of platform motion. Data includes all motion cases with $ St = 0.3$ and inflow conditions from O1.1 to O5.8 (see tables 1 and 2). A threshold of $ 1.5 \,\%$ is shown as a red dashed line.

Figure 18

Figure 17. Visualisation of phase-averaging methodology. (a) Wake wind speed time series of the floating wind turbine partitioned in bins of size $f_s/f_p$, giving one bin per period $T_p = 1/f_p$ (step 2). (b) Superposition of the $N_{seg}$ bins and phase averaged signal (step 3). (c) Raw signal decomposed as the sum of the mean, phase-averaged and stochastic fluctuations (2.4).

Figure 19

Figure 18. Comparison of Milan and Oldenburg datasets (fixed-5, surge-5 and sway-5 cases in table 2). (ae) Horizontal profiles of the mean streamwise wind speed at $x \in [2D, 10D]$. (fj) Horizontal profiles of the streamwise wind speed fluctuations at $x \in [2D, 10D]$. Both quantities are normalised by the free-stream velocity.