Hostname: page-component-68c7f8b79f-rgmxm Total loading time: 0 Render date: 2025-12-19T16:08:08.779Z Has data issue: false hasContentIssue false

Wake turbulence of an inclined prolate spheroid

Published online by Cambridge University Press:  16 December 2025

Sanidhya Jain
Affiliation:
University of California, San Diego, CA, USA
Sheel Nidhan
Affiliation:
University of California, San Diego, CA, USA
Sutanu Sarkar*
Affiliation:
University of California, San Diego, CA, USA
*
Corresponding author: Sutanu Sarkar, ssarkar@ucsd.edu

Abstract

The flow past a $6:1$ prolate spheroid at a moderate pitch angle $\alpha =10^\circ$ is investigated with a focus on the turbulent wake in a high-fidelity large eddy simulation (LES) study. Two length-based Reynolds numbers, ${\textit{Re}}_L=3\times 10^4$ and $9\times 10^4$, and four Froude numbers, ${\textit{Fr}} = \infty \text{(unstratified)}, 6, 1.9 \text{ and }1$, are selected for the parametric study. Spectral proper orthogonal decomposition (SPOD) analysis of the flow reveals the leading coherent modes in the unsteady separated flow at the tail of the body. At the higher ${\textit{Re}}_L=9\times 10^4$, a high-frequency spanwise flapping of shear layers on either side of the body is observed in the separated boundary layer for all cases. The flapping does not perturb the lateral symmetry of the wake. At ${\textit{Fr}}=\infty$, a low-frequency oscillating laterally asymmetric mode, which is found in addition to the shear-layer mode, leads to a sidewise unsteady lateral load. All temporally averaged wakes at ${\textit{Re}}=9\times 10^4$ are found to be spanwise symmetric in the mean as opposed to the lower ${\textit{Re}}=3\times 10^4$, at which the ${\textit{Fr}}=\infty \text{ and }6$ wakes exhibit asymmetry. The turbulent kinetic energy (TKE) budget is compared among cases. Here, ${\textit{Fr}}=\infty$ exhibits higher production and dissipation compared with ${\textit{Fr}}=6 \text{ and }1.9$. The streamwise vortex pair in the wake induces a significant mean vertical velocity ($U_z$). Therefore, in contrast to straight-on flow, the terms involving gradients of $U_z$ matter to TKE production. Buoyancy reduces $U_z$ and also the Reynolds shear stresses involving $u^{\prime}_z$. Through this indirect mechanism, buoyancy exerts control on the wake TKE budget, albeit being small relative to production and dissipation. Buoyancy, through the baroclinic torque, is found to qualitatively affect the streamwise vorticity. In particular, the primary vortex pair is extinguished in the intermediate wake and two new vortex pairs form with opposite-sense circulation relative to the primary.

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

Rapid advancements in aerial and underwater exploration technologies have made the study of turbulent wakes – from near the body to the far wake – important to understand/model vehicle dynamics and its environmental impact. For such applications, slender bodies like a prolate spheroid are used quite extensively due to their low drag characteristics. Numerous experimental and numerical studies have looked at the near and intermediate wake of slender objects in the past (Chevray Reference Chevray1968; Jimenez, Hultmark & Smits Reference Jimenez, Hultmark and Smits2010; Kumar & Mahesh Reference Kumar and Mahesh2018). However, due to it’s expensive nature and experimental limitations, the far wake of a slender body has been studied only recently (Ortiz-Tarin et al. Reference Ortiz-Tarin, Nidhan and Sarkar2021, Reference Ortiz-Tarin, Nidhan and Sarkar2023).

At zero pitch angle configuration ( $\alpha =0^\circ$ ) with a trip, the near wake of a slender body is characterised by a small separation bubble with the absence of significant vortex shedding at a high ${\textit{Re}}_L=U_\infty L/\nu \sim O(10^5)-O(10^6)$ (Posa & Balaras Reference Posa and Balaras2016; Kumar & Mahesh Reference Kumar and Mahesh2018; Ortiz-Tarin et al. Reference Ortiz-Tarin, Nidhan and Sarkar2021, Reference Ortiz-Tarin, Nidhan and Sarkar2023). In contrast, at moderate values of ${\textit{Re}}_L\sim O(10^4)$ , Ortiz-Tarin, Chongsiripinyo & Sarkar (Reference Ortiz-Tarin, Chongsiripinyo and Sarkar2019) and Ohh & Spedding (Reference Ohh and Spedding2024) observed clear vortex shedding for a $4:1$ and $6:1$ prolate spheroid, respectively. In the same configuration, both a high- ${\textit{Re}}_L$ wake (Ortiz-Tarin, Nidhan & Sarkar Reference Ortiz-Tarin, Nidhan and Sarkar2021) and a moderate- ${\textit{Re}}_L$ wake (Ortiz-Tarin et al. Reference Ortiz-Tarin, Chongsiripinyo and Sarkar2019) exhibit self-similar evolution with a decay law of $U_d\sim x^{-1}$ , which is significantly different to the classical law $U_d\sim x^{-2/3}$ . In the literature, this difference has been attributed to the non-equilibrium dissipation scaling due to unsteadiness in the energy cascade (Nedić et al. Reference Nedić, Vassilicos and Ganapathisubramani2013; Dairay, Obligado & Vassilicos Reference Dairay, Obligado and Vassilicos2015).

On-body flow and wake dynamics change significantly when the pitch angle $\alpha$ is increased. The separation line moves upstream towards the leading edge of the body. The flow wraps around the body and forms streamwise vortices on both sides during separation, even at moderate $\alpha \in [10^\circ ,20^\circ ]$ (Han & Patel Reference Han and Patel1979; Fu et al. Reference Fu, Shekarriz, Katz and Huang1994; Patel & Kim Reference Patel and Kim1994; Jiang et al. Reference Jiang, Andersson, Gallardo and Okulov2016). Along with strong primary vortices, counter-rotating secondary vortices with weaker circulation often emerge from inside the boundary layer at these flow configurations (Fu et al. Reference Fu, Shekarriz, Katz and Huang1994; Chesnakas & Simpson Reference Chesnakas and Simpson1997; Guo, Kaiser & Rival Reference Guo, Kaiser and Rival2023).

Although the wake generator is an axisymmetric slender body, the vortex wake at non-zero pitch angle can be laterally asymmetric. Such asymmetry has been reported by many studies of sharp-nosed slender bodies at high angle of attack ( $\alpha \gt 40^\circ$ ), where the strength of one vortex is much higher than the other (Nelson & Pelletier Reference Nelson and Pelletier2003; Bridges Reference Bridges2006; Nelson, Corke & Matsuno Reference Nelson, Corke and Matsuno2006). Similarly, spheroid wakes at both moderate and high pitch are also known to have lateral asymmetry in flow separation and the resulting vortex wake (Jiang et al. Reference Jiang, Gallardo, Andersson and Zhang2015, Reference Jiang, Andersson, Gallardo and Okulov2016). In a study by Tezuka & Suzuki (Reference Tezuka and Suzuki2006), different regimes of lateral asymmetry were identified for the vortex wake behind a $4:1$ prolate spheroid at $\alpha =10^\circ$ . A steady, symmetric wake was found below a diameter based critical ${\textit{Re}}_{C1}\approx 4\times 10^3$ which transitioned into a steady asymmetric mode upon increasing ${\textit{Re}}$ . Beyond the second critical ${\textit{Re}}_{C2}\approx 6.5\times 10^3$ , the wake exhibits an oscillating asymmetric mode. Recently, Ohh & Spedding (Reference Ohh and Spedding2024) and Nidhan et al. (Reference Nidhan, Jain, Ortiz-Tarin and Sarkar2025) also reported steady asymmetry for the mean vortex wake for an inclined $6:1$ spheroid at ${\textit{Re}}_L\sim O(10^4)$ .

Stratification effects for bluff and slender body wakes at zero pitch have long been a subject of investigation (Lin et al. Reference Lin, Lindberg, Boyer and Fernando1992; Spedding, Browand & Fincham Reference Spedding, Browand and Fincham1996; Spedding Reference Spedding1997; Dommermuth et al. Reference Dommermuth, Rottman, Innis and Novikov2002; Meunier & Spedding Reference Meunier and Spedding2006; Brucker & Sarkar Reference Brucker and Sarkar2010; de Stadler & Sarkar Reference de Stadler and Sarkar2012; Chongsiripinyo, Pal & Sarkar Reference Chongsiripinyo, Pal and Sarkar2017; Pal et al. Reference Pal, Sarkar, Posa and Balaras2017; Zhou & Diamessis Reference Zhou and Diamessis2019; Chongsiripinyo & Sarkar Reference Chongsiripinyo and Sarkar2020; Rowe, Diamessis & Zhou Reference Rowe, Diamessis and Zhou2020; Madison, Xiang & Spedding Reference Madison, Xiang and Spedding2022; Ortiz-Tarin et al. Reference Ortiz-Tarin, Chongsiripinyo and Sarkar2019, Reference Ortiz-Tarin, Nidhan and Sarkar2023; Li, Yang & Kunz Reference Li, Yang and Kunz2024). In a stratified environment, buoyancy effects on the wake are governed by the Froude number ${\textit{Fr}}=U_\infty /ND$ , where $U_\infty$ is the free stream velocity of the fluid, $D$ is the characteristic length scale of the body and $N$ is the buoyancy frequency. The mean defect velocity ( $U_d$ ) in the stratified wake is known to display three regimes of wake decay (Spedding Reference Spedding1997). First, a three-dimensional (3-D) regime where the buoyancy does not affect the wake and it behaves as unstratified up to $Nt\sim O(1)$ . Subsequently, it enters a non-equilibrium (NEQ) regime ( $2\lesssim Nt\lesssim 50$ ), where the buoyancy force on the vertical motion is sufficiently strong to slow down the streamwise decay of $U_d$ . Beyond $Nt\approx 50$ , buoyancy forces dominate vertical overturning motions and the wake enter a quasi-two-dimensional (Q2-D) regime organising into thin pancake eddies. The streamwise decay of $U_d$ speeds up in the Q2-D regime.

Ohh & Spedding (Reference Ohh and Spedding2024) performed the first experimental study of stratified $6:1$ spheroid wakes for non-zero pitch angles ( $\alpha =\{10^\circ ,20^\circ \}$ ) and varying stratification levels. Nidhan et al. (Reference Nidhan, Jain, Ortiz-Tarin and Sarkar2025) recently performed high-fidelity large eddy simulation (LES) for the same body shape and $\alpha =10^\circ$ at ${\textit{Re}}_L=3\times 10^4$ , and a range of stratification levels ( ${\textit{Fr}}\in \{6,1.9,1\}$ ). In both studies, the wake exhibits an initial downward displacement due to the angle of attack, followed by an oscillatory response about the hydrostatic equilibrium with a wavelength of $2\pi /N$ – the lee wavelength. The downward displacement of the streamwise vortex pair was found to be inhibited immediately after the tail of the body even at large ${\textit{Fr}} =6$ and this strong buoyancy effect was linked to a lower value of vorticity-based Froude number ( ${\textit{Fr}}_\omega$ ). Here, ${\textit{Fr}}_\omega = W_c/Nd$ , where $W_c$ is the measured descent velocity, and $d$ is the distance between the centroids of the positive and the negative vortex. Furthermore, a reduction in $U_d$ decay rates was observed along with restricted growth of the wake height compared with the unstratified counterpart. For ${\textit{Fr}}\geqslant 6$ , lateral asymmetry was observed in the separation region and the post-separation vortex pair. However, under heavy stratification ( ${\textit{Fr}}=1.9 \text{ and }1$ ), Nidhan et al. (Reference Nidhan, Jain, Ortiz-Tarin and Sarkar2025) observed a loss of asymmetry, with the wake being characterised by strong layering effects.

Apart from altering wake dynamics and topology, stratification crucially affects wake turbulence. Wakes under stratification tend to have slower streamwise decay of mean $U_d$ and are said to ‘survive longer’ compared with their unstratified counterparts. This phenomenon is attributed to reduced transfer of mean kinetic energy (MKE) into turbulent kinetic energy (TKE) by production (Brucker & Sarkar Reference Brucker and Sarkar2010). Specifically, the component $P_{\textit{xz}}=-\langle u^{\prime}_xu^{\prime}_z\rangle \partial _zU_x$ is reduced due to a decrease in the magnitude of $\langle u^{\prime}_xu^{\prime}_z\rangle$ , owing to buoyancy related suppression of $u^{\prime}_z$ and its correlation coefficient with $u^{\prime}_x$ (Jacobitz, Sarkar & Van Atta Reference Jacobitz, Sarkar and Van Atta1997; Jacobitz & Sarkar Reference Jacobitz and Sarkar1998; Meunier, Diamessis & Spedding Reference Meunier, Diamessis and Spedding2006). Chongsiripinyo, Pal & Sarkar (Reference Chongsiripinyo, Pal and Sarkar2019) performed a detailed turbulence energetics study for a disk wake at ${\textit{Re}}=5\times 10^4$ , where they found a power-law decay $u^{\prime}_z\sim x^{-1}$ in the strongly stratified turbulence (SST) regime, much faster than $u^{\prime}_x,u^{\prime}_y,U_d \sim x^{-0.18}$ .

Other than Ohh & Spedding (Reference Ohh and Spedding2024) and Nidhan et al. (Reference Nidhan, Jain, Ortiz-Tarin and Sarkar2025) (hereafter referred to as NJOS25), there is little work on pitched slender body wakes in a stratified environment. A relevant question is that, given NJOS25 which reports on the flow at ${\textit{Re}}_L=3\times 10^4$ , what is the motivation of this paper? The answer lies in the qualitatively different features introduced into the wake by increasing ${\textit{Re}}_L$ to $9 \times 10^4$ in both homogeneous and density-stratified fluids. We highlight some of these features here and will elaborate in later sections of the manuscript. NJOS25 reported on the mean wake at ${\textit{Re}}_L=3\times 10^4$ and at various $\textit{Fr}$ , but could not investigate the coherent structures (as done in detail here) behind the body because of their absence. Investigation of turbulence (TKE and its budgets) was not undertaken by NJOS25 because of the weak turbulence in the low- ${\textit{Re}}$ stratified cases of NJOS25. Furthermore, NJOS25 reported unexpected new results regarding the mean wake at ${\textit{Re}}_L=3\times 10^4$ . For instance, the unstratified ( ${\textit{Fr}}=\infty$ ) and weakly stratified ( ${\textit{Fr}}=6$ ) mean wake exhibited spanwise lateral asymmetry, despite the axisymmetry of the wake generator, with the important consequence of a sideways force imparted to the body. Does this asymmetry persist when ${\textit{Re}}_L$ is increased to $9\times 10^4$ ? For ${\textit{Fr}}=1$ , NJOS25 found the wake to form a vertically arranged double-lobed structure. How does that structure change at higher ${\textit{Re}}$ ?

Thus motivated, we perform a high-resolution computational study, and conduct an extensive analysis of the TKE budget and component terms for ${\textit{Re}}_L=9\times 10^4$ wakes for different stratification levels and compare with ${\textit{Re}}_L=3\times 10^4$ . Furthermore, the dominant temporal modes of the wake are identified using pointwise spectra and spectral proper orthogonal decomposition (SPOD) is employed to identify and analyse the coherent structures in the turbulent wake.

The paper is organised as follows. A description of the solver and methodology is given in § 2. The unsteady flow and its dominant modes are discussed using single-point spectra and by analysis of two-point, two-time correlations using SPOD in § 3. The qualitative difference upon introducing pitch and by increasing ${\textit{Re}}$ on the evolution of mean wake flow-field is discussed in § 4. A detailed analysis of TKE evolution and budget terms is given in § 5. Finally, the results are summarised and conclusions are drawn in § 6.

Figure 1. Schematic of the flow configuration in a cylindrical computational domain (not to scale). $L_x^{-}$ , $L_x^{+}$ and $L_r$ refer to the upstream, downstream and radial domain distance, respectively. $\alpha$ is the pitch angle. The centre of the spheroid is at the origin of the coordinate system.

2. Methodology

2.1. Governing equations and numerical scheme

The near and intermediate wake of a 6 : 1 prolate spheroid at moderate pitch ( $\alpha =10^\circ$ ) is investigated using high-fidelity large eddy simulation (LES). These simulations are implemented with the set-up and solver used by NJOS25, see figure 1 for a detailed schematic. In stratified cases, the background is thermally stratified with a constant temperature gradient. Both studies solve the non-dimensional filtered Navier–Stokes equations under the Boussinesq approximation, in conjunction with the continuity and density transport equation, in a cylindrical coordinate system:

(i) continuity,

(2.1) \begin{equation} \frac {\partial u_{i}}{\partial x_{i}} = 0; \end{equation}

(ii) momentum,

(2.2) \begin{equation} \frac {\partial u_{i}}{\partial t} + \frac {\partial (u_{i}u_{j})}{\partial x_{j}} = -\frac {\partial p}{\partial x_{i}} + \frac {1}{\textit{Re}}\frac {\partial }{\partial x_{j}}\Big [\Big (1 + \frac {\nu _{s}}{\nu }\Big )\frac {\partial u_{i}}{\partial x_{j}}\Big ] - \frac {1}{{Fr}^{2}}{\rho _d} \delta _{i3}; \end{equation}

(iii) density,

(2.3) \begin{equation} \frac {\partial \rho }{\partial t} + \frac {\partial (\rho u_{j})}{\partial x_{j}} = \frac {1}{Re Pr}\frac {\partial }{\partial x_{j}}\Big [\Big (1 + \frac {\kappa _{s}}{\kappa }\Big )\frac {\partial \rho }{\partial x_{j}}\Big ]. \end{equation}

Following the Boussinesq approximation, $\rho$ is decomposed as $\rho (x_i, t) = \rho _0 + \rho _{b}(x_i) + \rho _d (x_i,t)$ , where $\rho _0$ is a density, $\rho _{b}(x_i)$ is the background density variation and $\rho _d (x_i,t)$ is the density deviation from the background. For linear stratification, $\partial \rho _{b}(x_i)/\partial z = C$ , where $C$ is a constant. Furthermore, it is assumed that $\rho _{b}/\rho _0 \lt \lt 1$ and $\rho _d/\rho _0 \lt \lt 1$ to obtain the simplified treatment of density variation in (2.1) and (2.2).

The non-dimensionalisation employs the following parameters: (i) free stream velocity $U_{\infty }$ for $u_{i}$ ; (ii) minor axis length $D$ for $x_{i}$ ; (iii) dynamic pressure $\rho _{0}U_{\infty }^{2}$ for $p$ ; (iv) $D/U_{\infty }$ for time $t$ ; (v) $-DC$ for density; (vi) kinematic viscosity $\nu$ for the subgrid kinematic viscosity $\nu _s$ and (vii) molecular diffusivity $\kappa$ for the subgrid molecular diffusivity $\kappa _s$ .

The stratification is set by a linear density profile characterised by the non-dimensional Froude number ${\textit{Fr}}=U_\infty /ND$ . Here, $N = \sqrt {-gC/\rho _0}$ is the buoyancy frequency and $C = {\rm d}\rho _{b}/{\rm d} z$ is a constant. In this study, three different levels of stratification are simulated: ${\textit{Fr}}=6,1.9\text{ and }1$ , ranging from weak to high stratification relative to the flow inverse time scale, respectively. Note that ${\textit{Fr}}=1.9$ is a critical case ( ${\textit{Fr}} = AR/\pi = 6/\pi \approx 1.9$ ), where the lee wavelength ( $\lambda /D=2\pi {Fr}$ ) is approximately equal to twice the major axis of the body.

Two length-based Reynolds numbers ${\textit{Re}}=U_\infty L/\nu =3\times 10^4\text{ and }9\times 10^4$ are considered. For analysing the ${\textit{Re}}=3\times 10^4$ case, the data from simulations of NJOS25 are used. New simulations are performed for the higher ${\textit{Re}}=9\times 10^4$ case.

An immersed boundary method (Balaras Reference Balaras2004; Yang & Balaras Reference Yang and Balaras2006) is used to resolve the flow past a 6 : 1 prolate spheroid at moderate pitch. The solver uses second-order, central finite difference schemes to calculate spatial derivatives on a staggered grid. The flow variables are temporally advanced with a fractional step method that combines the Crank–Nicolson method with a low-storage Runge–Kutta scheme (RKW3). The kinematic subgrid viscosity ( $\nu _{s}$ ) and density diffusivity ( $\kappa _{s}$ ) are determined using the dynamic Smagorinsky model (Germano et al. Reference Germano, Piomelli, Moin and Cabot1991). For the simulations, Dirichlet boundary conditions at the inlet, an Orlanski-type convective boundary condition (Orlanski Reference Orlanski1976) at the outlet and Neumann boundary conditions at the radial boundary are implemented. Similar to NJOS25, a sponge layer is added to all the boundaries to avoid spurious reflection of internal gravity waves. For a more detailed description of the stratified-flow solver employed in this study, interested readers are referred to NJOS25, Pal et al. (Reference Pal, Sarkar, Posa and Balaras2017) and Chongsiripinyo & Sarkar (Reference Chongsiripinyo and Sarkar2020).

2.2. Domain and grid parameters

A cylindrical structured grid is used and the grid parameters for each case is shown in table 1 along with the corresponding case labels. Here, $L_r$ is the radial extent of the domain, and $L_x^-$ and $L_x^+$ are the upstream and downstream extent measured from the body centre at $x =0$ . Furthermore, an ${\textit{Re}}=9\times 10^4$ spheroid without pitch ( $\alpha =0^\circ$ ) and in an unstratified ( ${\textit{Fr}}=\infty$ ) background is also simulated. The R30 ( ${\textit{Re}}=3\times 10^4$ ) series reported in NJOS25 was found to have DNS-like resolution (over-resolved for an LES) as summarised in NJOS25 and, therefore, the grid was not changed for the R90 ( ${\textit{Re}}=9\times 10^4$ ) series. A posteriori analysis of the R90F $\infty$ data shows $\text{max}(\Delta x/\eta ) = 7.5$ , $\text{max}(\Delta r/\eta ) = 1.8$ and $\text{max}(r\Delta \theta /\eta ) = 3.5$ after $x/D=5$ (intermediate wake). Similarly for R90F6, $\text{max}(\Delta x/\eta ) = 8.5$ , $\text{max}(\Delta r/\eta ) = 3.2$ and $\text{max}(r\Delta \theta /\eta ) = 5$ . Therefore, both these grids are sufficiently resolved for a high-fidelity LES. Since the flow is found to be less turbulent in the other stratified cases, the resolution in units of Kolmogrov scale is even better for ${\textit{Fr}}=1.9$ and ${\textit{Fr}} = 1$ , and are not shown here for brevity.

Table 1. Simulation parameters. $N_r, N_\theta , N_x$ correspond to the number of grid points in radial, azimuthal and streamwise directions, respectively. Here, ${\textit{Re}}_L$ and $\textit{Fr}$ are the major-axis Reynolds number and minor-axis-based Froude number, respectively. Case labels given as RxxFyy refers to a case with ${\textit{Re}}=$ xx $\times 10^3$ and ${\textit{Fr}}=$ yy. The label with A0 at the end refers to a zero angle of attack case

Similar to ${\textit{Re}}=3\times 10^4$ , the quality of the grid is good over the body for ${\textit{Re}}=9\times 10^4$ and the boundary layer (BL) in all cases is adequately resolved. At $x/D=0$ , for R90F6, there are approximately 80 points across the $\delta _{99}$ thickness of the BL. For R90F $\infty$ , there are 40 points in the BL. At the thinnest BL location on the windside of the body, there are 35 points across $\delta _{99}$ for the stratified cases and 15 for the unstratified case. The Ozmidov scale, $L_O=\sqrt {\varepsilon /N^3}$ , is well resolved with $\Delta r/L_O\lt 0.1$ inside the wake. For a detailed discussion of the grid quality, interested readers are referred to § 2.2 of NJOS25.

2.3. Statistics

All reported velocities and lengths are normalised with free stream velocity $U_\infty$ and the body minor axis $D$ , respectively. For stratified cases, the normalised streamwise distance $x$ from the centre of the body is also measured as a function of the buoyancy frequency ( $N$ ) and time ( $t$ ). The time in the $Nt$ axis refers to the time measured by an observer in the frame of reference of the free stream flow that sees the body moving at speed $-U_\infty$ . From a Galilean transform, we get $x/{Fr}=Nt$ .

After the flow achieves statistical stationarity, temporal averaging, denoted by $\langle . \rangle$ , is performed to obtain the statistics. Instantaneous quantities are written with lower case, mean with upper case and fluctuations with primes. For all cases, averaging is performed over a span of $100 D/U_\infty$ , approximately two flow-through times. This time interval is found, a posteriori, to significantly exceed that required for convergence of the turbulence statistics reported here. Apart from temporal averaging, some statistics are obtained by cross-wake area-integration denoted by $\{.\}$ . The integral is performed over a circular cross-section of radius $4D$ , unless otherwise stated. Wake turbulence becomes negligible well before reaching the lateral boundaries of the region of integration.

2.4. Spectral proper orthogonal decomposition (SPOD)

Spectral proper orthogonal decomposition (SPOD) is a modal decomposition technique that extracts coherent structures from turbulent flows in a statistically optimal manner (Towne, Schmidt & Colonius Reference Towne, Schmidt and Colonius2018). Unlike classical snapshot POD (Sirovich Reference Sirovich1987), which identifies spatial modes based solely on energy maximisation without considering temporal coherence, SPOD identifies dominant structures by leveraging temporal correlation among flow snapshots. Dynamic mode decomposition (DMD) also identifies structures with space–time coherence. However, unlike DMD, where the eigenmodes are typically non-orthogonal, SPOD offers a more precise basis set for the fluctuations due to the orthogonality of its eigenmodes. SPOD modes, as ensemble DMD modes that have been optimally averaged to provide the most accurate representation of the second-order space–time flow statistics, provide an excellent representation of coherent structures by providing rank-ordered spatial eigenmodes for the energetic frequencies in the wake. Nidhan, Schmidt & Sarkar (Reference Nidhan, Schmidt and Sarkar2022) demonstrated the effectiveness of this method in stratified disk wakes, where the leading SPOD eigenmode effectively captured the generation of internal waves and layering effects within the wake core. In the present study, we employ a similar approach to analyse the time-resolved data from the R90 series.

Consider a spatio-temporally evolving field $\boldsymbol{{\varLambda}}(\boldsymbol{\textit{x}},t) = [\boldsymbol{\textit{u}}'(\boldsymbol{\textit{x}},t),\rho '(\boldsymbol{\textit{x}},t)]^T$ , where $\rho '(\boldsymbol{\textit{x}},t)$ represents the density fluctuation field and $\boldsymbol{\textit{u}}'(\boldsymbol{\textit{x}},t) = [\textit{u}_r'(\boldsymbol{\textit{x}},t),\textit{u}_\theta '(\boldsymbol{\textit{x}},t),\textit{u}_x'(\boldsymbol{\textit{x}},t)]$ is the velocity fluctuation field. The SPOD method seeks to identify orthogonal modes $\boldsymbol{{\varPsi}}(\boldsymbol{\textit{x}},t)$ that maximise the ensemble-averaged projection of $\boldsymbol{{\varLambda}}(\boldsymbol{\textit{x}},t)$ on these modes (Lumley Reference Lumley1970).

For stratified flow, the eigenvalues and eigenmodes satisfy

(2.4) \begin{equation} \int _{{\varOmega}} S_{\textit{ij}}(\boldsymbol{x},\boldsymbol{x}', f) \boldsymbol{W}(\boldsymbol{x}'){\varPhi}_j^{(n)}(\boldsymbol{x}', f) \,{\rm d}\boldsymbol{x}' = \lambda ^{(n)}(f){\varPhi}_i^{(n)}(\boldsymbol{x}, f), \end{equation}

which is an eigenvalue problem that is solved at each frequency $f$ . Here, $\lambda ^{(n)}(f)$ is the $n{\textrm{th}}$ eigenvalue at $f$ and the modified eigenmodes are given by ${\varPhi}_i^{(n)}(\boldsymbol{x}, f) = {\varPsi}_i^{(n)}(\boldsymbol{x}, t)e^{-i2\pi ft}$ . The eigenvalues are sorted such that $\lambda ^{(1)}(f) \geqslant \lambda ^{(2)}(f) \ldots \geqslant \lambda ^{(n)}(f)$ . Furthermore, $S_{\textit{ij}}(\boldsymbol{x},\boldsymbol{x}', f)$ is the Fourier transform of the kernel $\langle {\varLambda}_i(\boldsymbol{x}, t) {\varLambda}_j(\boldsymbol{x}, t) \rangle$ and $\boldsymbol{W}(\boldsymbol{x})=\operatorname {diag} ( 1, 1, 1, ({g^2}/{\rho _0^2 N^2}) )$ is a positive-definite Hermitian matrix that represents the respective weights of each variable. The sum of all eigenvalues at $f$ is equivalent to twice the total fluctuation energy content, i.e. 2(TKE + TPE) at that frequency. The obtained eigenmodes at the same frequency are spatially orthogonal to one another.

For numerical implementation, the mean-subtracted dataset, comprising $N$ temporal snapshots, is divided into $N_{\textit{blk}}$ overlapping blocks, each containing $N_{\textit{ovlp}}$ shared snapshots. Each block comprises $N_{\textit{freq}}$ entries, represented as $\boldsymbol{Q} = [\boldsymbol{q}^{(1)}, \boldsymbol{q}^{(2)}, \boldsymbol{q}^{(3)}, \ldots , \boldsymbol{q}^{(N_{\textit{freq}})}]$ , where each entry is defined as $\boldsymbol{q}^{(i)} = [\boldsymbol{u}'^{(i)}, \rho '^{(i)}]^T$ . Next, a discrete Fourier transform (DFT) is applied to each block in the temporal direction. This results in a set of $N_{\textit{blk}}$ Fourier realisations at a given frequency $f$ . These realisations are collected as $\hat {\boldsymbol{Q}}_f = [\boldsymbol{q}^{(1)(f)}, \boldsymbol{q}^{(2)(f)}, \boldsymbol{q}^{(3)(f)}, \ldots , \boldsymbol{q}^{(N_{\textit{blk}})(f)}]$ . Once $\hat {\boldsymbol{Q}}_f$ is obtained, the SPOD eigenvalues and eigenvectors corresponding to frequency $f$ are determined through the following eigenvalue decomposition:

(2.5) \begin{equation} \hat {\boldsymbol{Q}}_f^* \boldsymbol{W} \hat {\boldsymbol{Q}}_f \boldsymbol{{\varGamma}}_{\!f} = \boldsymbol{{\varGamma}}_{\!f}\boldsymbol{{\lambda}}_f .\end{equation}

Here, the asterisk denotes the complex conjugate and $\boldsymbol{{\lambda}}_f$ is a diagonal matrix defined as $\boldsymbol{{\lambda}}_f = \operatorname {diag} (\lambda ^{(1)}_f, \lambda ^{(2)}_f, \ldots , \lambda ^{(N_{\textit{blk}})}_f)$ , where the eigenvalues are arranged in descending order based on their energy content, indexed from $i = 1$ to $N_{\textit{blk}}$ . The corresponding spatial eigenmodes, $\hat {\boldsymbol{{\varPhi}}}_f$ , are derived from the eigenvectors $\boldsymbol{{\varGamma}}_{\!f}$ as $\hat {\boldsymbol{{\varPhi}}}_f = \hat {\boldsymbol{Q}}_f \boldsymbol{{\varGamma}}_{\!f} \boldsymbol{{\lambda}}_f^{-1/2}$ . A total of $N_{\textit{blk}} = (N - N_{\textit{ovlp}})/(N_{\textit{freq}}-N_{\textit{ovlp}})$ SPOD modes are identified at each frequency. In (2.5), $\boldsymbol{W}$ is a diagonal matrix of size $4 N^{\textit{SPOD}}_r N_{\theta }$ , which contains numerical quadrature weights multiplied by coefficients required for computing the energy quantities. Here, $N^{\textit{SPOD}}_r$ is the number of points in the radial extent of the snapshot and $N_\theta$ in the azimuthal direction. For further insights into the theoretical aspects and numerical implementation of SPOD, readers are referred to the studies by Towne et al. (Reference Towne, Schmidt and Colonius2018) and Schmidt & Colonius (Reference Schmidt and Colonius2020).

3. Unsteadiness and coherent structures

The near wakes in the R90 series exhibit unsteadiness soon after the laminar boundary layer separates. The unsteadiness is related to coherent structures and also impacts the body loads as will be shown later. The R30 series has weaker unsteadiness after flow separation and will not be discussed in this section.

Figure 2. Contours of instantaneous defect velocity ( $u_d=U_\infty -u_x$ ) at $x/D = 3$ for (a–c) R90F1.9, (d–f) R90F6 and (g–i) R90F $\infty$ at three distinct timestamps for each case. Radial extent for each panel is $r/D=0.75$ .

3.1. Instantaneous wake behaviour

Figure 2(a–c) illustrates the instantaneous defect velocity ( $u_d = U_\infty -u_x$ ) at $x/D=3$ for R90F1.9 at three different time instances. The near wake exhibits two vertically elongated, ear-like lobes positioned atop a central structure. Specifically, at $tU_\infty /D=109.34$ (figure 2 a), both lobes are situated close to the centre-plane ( $y=0$ ). Subsequently, at $tU_\infty /D=109.76$ (figure 2 b), the lobes diverge in the spanwise direction, moving away from the centre-plane. Finally, at $tU_\infty /D=110.152$ (figure 2 c), the lobes converge, returning to a position closest to the centre-plane. This sequence demonstrates a spanwise flapping motion of the wake, characterised by a distinct temporal frequency, as will be shown later. The selected timestamps are chosen to accurately capture a complete cycle of this flapping motion. Notably, the wake structure maintains a configuration with predominantly lateral symmetry throughout the oscillation, resembling varicose modes with symmetric, in-phase deformations that have been found previously in shear flows, e.g. shear layers at the edge of a plane jet (Stanley & Sarkar Reference Stanley and Sarkar1997) and the wake of a cylindrical roughness element on a flat plate (Bucci et al. Reference Bucci, Puckert, Andriano, Loiseau, Cherubini, Robinet and Rist2018). For a more dynamic visualisation of this flapping motion, please refer to the supplementary movie 1 available at https://doi.org/10.1017/jfm.2025.10923.

At ${\textit{Fr}}=6$ , the wake exhibits a structural morphology qualitatively similar to that observed at ${\textit{Fr}}=1.9$ , characterised by two lobes connected to a central structure (figure 2 df and supplementary movie 2). A notable dissimilarity, however, lies in the reduced elongation of the lobes compared with R90F1.9. Nevertheless, an analogous flapping motion is evident, wherein the lobes undergo continuous inward and outward oscillations, maintaining a symmetric configuration.

The temporal evolution of the unstratified case (R90F $\infty$ ) deviates somewhat from the previously described stratified cases. Figure 2(gi) depicts $u_d$ contours at $x/D=3$ for three different time instances. While a double-lobed structure remains discernible, the lobes are considerably less differentiated from the central structure. Moreover, the instantaneous wake structure exhibits asymmetry about the centre-plane ( $y=0$ ). At $tU_\infty /D=72.24$ (figure 2 g), the left half of the wake displays a larger extent compared with the right half. Conversely, this polarity of asymmetry reverses as the simulation progresses to $tU_\infty /D=84.76$ (figure 2 h), where the right half manifests a larger extent. Subsequently, the asymmetry oscillates back to the left half at $tU_\infty /D=98.24$ (figure 2 i). These timestamps were carefully selected to fully capture a complete cycle of these oscillations. Based on the timestamps, it can be inferred that this oscillating asymmetry is characterised by a low frequency. The unstratified wake also exhibits a high-frequency spanwise flapping of the double lobes, analogous to the stratified cases, as demonstrated in supplementary movie 3. Consequently, both the oscillating asymmetry (low frequency) and flapping (high frequency) modes are dominant in the R90F $\infty$ flow, and these modes will be discussed in subsequent sections.

To gain further insight into the high frequency flapping mode, contours of instantaneous vertical vorticity ( $\omega _z$ ) for the R90F1.9 wake, chosen as an example, are shown on the horizontal centre-plane ( $z=0$ ) in figure 3. Shear layer separation is evident on each side of the body at $x/D\approx 1.5{-}2.5$ . These shear layers later become unstable and rollup into Kelvin–Helmholtz (K–H) billows after $x/D\approx 2.5$ before breaking up into small-scale structures further downstream. Looking closely at $x/D=3$ at $tU_\infty /D=109.34$ , the K–H rollups have not yet reached that streamwise location and, thus, the $\omega _z$ imprint is the narrowest in the spanwise direction. This is the same time at which the $u_d$ double-lobed structure observed in figure 2(a) is the closest to the centre-plane ( $y=0$ ). Furthermore, at $tU_\infty /D=109.76$ , the K–H billows are somewhat centred around $x/D=3$ giving the widest spanwise imprint of $\omega _z$ . This timestamp also corresponds to the diverging double-lobed structure in figure 2(b). Therefore, it can be surmised that the aforementioned flapping mode is related to the shear layers becoming unstable after leaving the inclined spheroid. For a more dynamic visualisation of the shear layer instability, please refer to supplementary movie 4.

Figure 3. Instantaneous vertical vorticity ( $\omega _z$ ) contours plotted on the centre horizontal plane ( $z=0$ ) for R90F1.9 at three distinct timestamps similar to figure 2(ac). The separated shear layers exhibit Kelvin–Helmholtz billows.

3.2. Point spectra

The spanwise fluctuation velocity ( $u^{\prime}_y$ ) is obtained from the computed $u^{\prime}_r, u^{\prime}_\theta$ components. Analysis of $u^{\prime}_y (t)$ with pointwise spectra in this subsection and $u^{\prime}_y (r, \theta , t)$ with SPOD spectra in the following subsection reveal dominant modes as elaborated here.

Figure 4. Point spectra of spanwise velocity fluctuation ( $u^{\prime}_y$ ) for (a) R90F1.9, (b) R90F6 and (c,d) R90F $\infty$ at different locations in the wake, where panel (d) is computed for a longer time series. Location of points A, B and C for each $\textit{Fr}$ case is marked atop the $u_d$ contours shown for $x/D=3$ . At $x/D=10$ , spectra shown in dashed red represent a point at the edge of the wake and dashed green represent a point on the vertical centre-plane.

Figure 4(a) shows temporal point spectra of the spanwise fluctuation velocity ( $u^{\prime}_y$ ) for R90F1.9 at different points on the cross-stream planes at $x/D=3$ and $10$ . At $x/D=3$ , point spectra show a peak at $St \approx 1.15$ for points at $\theta =0^\circ \text{(point A) and } 180^\circ$ (point B) and radial distance of $r/D=0.2$ from centre. Points A and B are located in the right and left shear layers. The spectrum at point C ( $\theta =270^\circ , r/D=0.5$ ) in the central structure at $x/D=3$ does not show any prominent peak. Hence, not much fluctuation activity is present in the central wake structure and the $St \approx 1.15$ peak is the dominant wake mode, as will be more rigorously established with SPOD analysis. At a downstream location of $x/D=10$ , the spectra peak at $St\approx 1$ for point $\theta =0^\circ \text{ and } r/D=0.2$ at the wake edge. However, this peak is weaker compared with that at $x/D=3$ , suggesting streamwise evolution of the coherent shear-layer mode. The spectra at further downstream locations show progressive diminishing of this peak until it disappears at $x/D \approx 20$ . The spectrum at a point on the centre-plane ( $\theta =270^\circ , r/D=0.2$ ) does not show any dominant frequency at $x/D=10$ .

Point spectra of $u^{\prime}_y$ for R90F6 (figure 4 b) show characteristics similar to R90F1.9. A peak at $St\approx 1.25$ is observed at $x/D=3$ for locations A ( $r/D=0.4,\theta =315^\circ$ ) and B ( $r/D=0.4,\theta =225^\circ$ ). Again, no peak is observed in the spectrum at point C ( $r/D=0.6,\theta =270^\circ$ ) in the centre-plane. At $x/D=10$ , a local peak is observed at $St\approx 0.75$ for a point at the wake edge ( $r/D=0.75,\theta =300^\circ$ ), whereas no dominant peak is observed in the vertical centre-plane ( $r/D=1,\theta =270^\circ$ ).

Spectra of $u^{\prime}_y$ for the R90F $\infty$ wake at $x/D=3$ show a peak at $St\approx 1.28$ (figure 4 c) for points A ( $r/D=0.5,\theta =315^\circ$ ) and B ( $r/D=0.5,\theta =225^\circ$ ) on the double-lobed structure, see figure 2(gi). This peak corresponds to the flapping mode. A significant power spectral density (PSD) level is also observed for low frequencies $St\sim O(0.01)$ at point C ( $r/D=0.5,\theta =270^\circ$ ) on the centre-plane. Taking a longer time series for the point spectra helped resolve those low frequencies, see figure 4(d). A peak is observed at $St\approx 0.04$ , which corresponds to the slow oscillating asymmetry exhibited by the unstratified wake, discussed in § 3.1. At $x/D = 10$ and beyond, no prominent peak is seen in the power spectra. Thus, shear-layer flapping and oscillating asymmetric modes are confined to the near wake.

3.3. SPOD eigenspectra and leading eigenmenodes

SPOD analysis was conducted for the flow on a 2-D plane at $x/D=3$ for the R90 series. For the unstratified case R90F $\infty$ , the analysed flow field spans $0\leqslant r/D \leqslant 10$ , comprising a total of $N_r^{\textit{SPOD}}=678$ radial grid points and $N_\theta =256$ points in the azimuthal direction. Stratified cases R90F6 and R90F1.9 contain $N_r^{\textit{SPOD}}=940$ grid points over $0\leqslant r/D \leqslant 11.5$ , and $N_\theta =128$ points in the azimuthal direction. Table 2 lists the parameters for the SPOD analysis across various $\textit{Fr}$ cases at ${\textit{Re}}=9\times 10^4$ . In this context, $N$ denotes the total number of snapshots, spaced by $\Delta t D/U_\infty$ between consecutive snapshots; $N_{\textit{freq}}$ represents the number of frequencies that are being resolved. The $N$ snapshots are divided into discrete sets of overlapping blocks with an overlap of $N_{\textit{ovlp}}$ snapshots. A total of $N_{\textit{blk}} = (N - N_{\textit{ovlp}})/(N_{\textit{freq}}-N_{\textit{ovlp}})$ SPOD modes are extracted at each frequency.

Table 2. SPOD parameters for ${\textit{Re}}=9\times 10^4$ wakes

Figure 5. (a) SPOD eigenspectra, $\lambda ^{(1)}, \lambda ^{(2)},\ldots , \lambda ^{(n)}$ , for all eigenmodes of R90F1.9 at a streamwise location of $x/D=3$ . (b) Real part of the leading SPOD mode for the spanwise velocity, ${\varPhi}_{v}^{(1)}(y,z,St;x/D)$ , corresponding to the peak in the eigenspectrum of $\lambda ^{(1)}$ .

Figure 5(a) presents the SPOD eigenspectra of the R90F1.9 wake at $x/D=3$ . The leading eigenmode ( $\lambda ^{(1)}$ ) is highlighted in blue. A prominent high-frequency peak is evident in the eigenspectrum of $\lambda ^{(1)}$ at $St\approx 1.15$ . Notably, the frequency of $St\approx 1.15$ is rarely associated with a vortex shedding (VS) mode in the literature, which typically has a frequency $St\sim \textit{O}(0.1)$ (Berger, Scholz & Schumm Reference Berger, Scholz and Schumm1990; El Khoury, Andersson & Pettersen Reference El Khoury, Andersson and Pettersen2012; Ohh & Spedding Reference Ohh and Spedding2024). To delve deeper into the nature of this high-frequency mode, the real part of the normalised (by $L_\infty$ norm) leading SPOD mode, ${\varPhi}_v^{(1)}(y,z,St;x/D)/{\Vert}{\varPhi}_v^{(1)}(y,z,St;x/D){\Vert}_\infty$ , where ${\varPhi}_v$ refers to the spanwise velocity $u_y$ , is presented in figure 5(b). From the plot, two distinct coherent regions/patches of opposite sign are discernible, situated on either side of the centre-plane ( $y=0$ ). The sign of these patches ensures that they are in opposition, meaning that if the right patch exhibits a positive spanwise velocity, the corresponding left patch will have a negative spanwise velocity of comparable magnitude. Consequently, both coherent patches simultaneously move outward or inward relative to the centre-plane. Furthermore, a striking similarity emerges between figures 2(ac) and 5(b), as the oscillating portion of the wake aligns with the coherent patches in the eigenmode. Moreover, the size of the eigenmode coherent patches is comparable to the swept area of the oscillating double-lobes in the instantaneous wake. Therefore, this eigenmode corresponds to the flapping motion of the shear layers observed in the instantaneous defect velocity field ( $u_d$ ). It is worth noting that high-frequency modes with $St \sim O(1)$ have been identified in the near wake of a disk and attributed to shear layer instability (Berger et al. Reference Berger, Scholz and Schumm1990; Nidhan et al. Reference Nidhan, Chongsiripinyo, Schmidt and Sarkar2020).

Figure 6. Same as figure 5 for R90F6.

Similarly, the SPOD eigenspectra for all eigenmodes of the R90F6 wake at $x/D=3$ are depicted in figure 6(a). A prominent high-frequency peak at $St\approx 1.21$ is evident in the leading mode ( $\lambda ^{(1)}$ ). Furthermore, examining the shape of the leading SPOD mode for spanwise velocity ( ${\varPhi}_v^{(1)}(y,z,St;x/D)$ ) at the peak frequency (figure 6 b), it becomes evident that two patches of opposite signs are present in a spanwise symmetric manner, reminiscent of ${\textit{Fr}}=1.9$ . By juxtaposing $u_d$ contours in figure 2(df) with the leading SPOD mode, it can be deduced that the leading mode corresponds to the high-frequency flapping of the two lobes/shear layers that were observed in § 3.1.

Figure 7. Same as figure 5 for R90F $\infty$ . Here, panel (b) corresponds to the leading eigenmode at $St\approx 1.28$ peak and panel (c) corresponds to eigenmode of low-frequency mode at $St\approx 0.04$ .

Eigenspectra for R90F $\infty$ differ somewhat from the stratified cases, as illustrated in figure 7(a). While a high-frequency local maximum is observed for $\lambda ^{(1)}$ at $St\approx 1.28$ , similar to the aforementioned cases, a substantial amount of energy persists in the low-frequency modes for $St\lt 0.1$ . Similar to point spectra (figure 4 d), eigenspectra also show a peak at $St\approx 0.04$ , therefore, the eigenmode for this frequency was also diagnosed along with that for the high-frequency peak. Figures 7(b) and 7(c) depict the shape of the leading eigenmodes for spanwise velocity ( ${\varPhi}_v^{(1)}(y,z,St;x/D)$ ) at $St\approx 1.28$ (high-frequency peak) and $St\approx 0.04$ (low-frequency mode), respectively. At $St\approx 1.28$ , the leading eigenmode again exhibits a two-patch structure with a sign change between $y \gt 0$ and $y\lt 0$ . Comparing the shape and location of the patches with $u_d$ contours in figure 2(gi), it can be inferred that this mode represents the flapping mode of the double-lobed structure. Notably, the low-frequency mode of $St\approx 0.04$ exhibits a prominent coherent structure on the centre-plane ( $y=0$ ). The sign of the $St \approx 0.04$ mode remains consistent across the structure, implying that the spanwise velocity in this region is uniformly in the same direction. This mode corresponds to the oscillating asymmetric mode, which modifies the size of both lobes simultaneously, increasing one and decreasing the other.

It is noted that there is significant energy at the low-frequency cutoff of the spectrum in figure 7(a). The capture of such a very low frequency (VLF) mode, which is often seen in laboratory wakes, presents a well-known challenge to numerical simulations because of the necessity of a very long simulation time. For example, Zhu & Morrison (Reference Zhu and Morrison2021) found increased energy at low frequencies ( $St\lt 0.01$ ) in the SPOD eigenspectra for the $m=\pm 1$ mode, but were unable to find a specific VLF peak because of the insufficient length of the time series. They inferred VLF dynamics – a multistable wake that exhibits a slow rotation of the vortex shedding plane (Rigas et al. Reference Rigas, Oxlade, Morgans and Morrison2014). Investigation of the VLF mode is beyond the scope of this paper.

3.4. Oscillating force on the body

Dominant modes of the flow, presented in previous subsections, impart significant forces on the wake generator. Force coefficients in this section are defined as $C_i=F_i/(1/2\rho U^2_\infty A)$ , where $A=\pi D^2/4$ and $F_i$ is the net force on the body in the $i$ -direction. Figure 8(a) shows the time evolution of the net spanwise force coefficient ( $C_y$ ) on the body for R90F $\infty$ . A cyclic forcing can be observed with an amplitude of $|C_y|\approx 0.025$ with a mean value $\langle C_y\rangle \approx 0$ . The amplitude of the sideways force is significant, ${\sim} 10\,\%$ of the mean drag ( $\langle C_x\rangle =0.238$ ) and lift ( $\langle C_z\rangle =0.264$ ). The sideways force (harmonic oscillation superposed on a mean) is significantly different from the ${\textit{Re}}=3\times 10^4$ wake, where the sideways force has a non-zero mean and no temporal variability.

Figure 8. (a) Time evolution of total spanwise force on the body, (b) Spectra of total force in the spanwise, vertical and streamwise directions for R90F $\infty$ .

The spectrum deduced from the $C_y$ time series exhibits a dominant peak at $St\approx 0.04$ (figure 8 b). This is the same frequency as the leading eigenmode observed for the asymmetric oscillatory mode. This clearly shows that the flow field imparts a cyclic spanwise force on the body through its leading eigenmode. Furthermore, no high-frequency peak is observed for any of the force coefficients. Since the high-frequency SPOD mode at $St \approx 1.28$ had two patches of opposite phase on either side of the body, their force contributions cancel out and this oscillatory mode imparts no net force to the body.

Since the R90F1.9 and R90F6 wakes are found to be spanwise symmetric at every time instant, no net oscillating spanwise force is observed. Also, no dominant peaks are exhibited in the spectra (not shown for brevity) of the time-varying force.

4. Mean flow evolution

Flow quantities, such as the mean defect velocity ( $U_d$ ), mean vertical velocity ( $U_z$ ) and mean streamwise vorticity ( $\langle \omega _x\rangle$ ), provide quantitative measures of the mean wake dynamics and are discussed in this section. The mean flow evolution for inclined spheroid wakes at ${\textit{Re}}=3\times 10^4$ and their buoyancy-induced changes upon varying stratification were discussed extensively in § 5 of NJOS25. Therefore, the discussion here is brief and serves as a juxtaposition of the R90 series with its R30 counterpart.

4.1. Mean defect velocity ( $U_d$ )

Figure 9. Contours of mean defect velocity ( $U_d$ ) for two streamwise locations (a,b,e, f,i,j) $x/D=3$ and (c,d,g,h,k,l) $x/D=10$ are shown for (a–d) ${\textit{Fr}}=\infty$ , (e–h) ${\textit{Fr}}=6$ and (i–l) ${\textit{Fr}}=1.9$ at (a,c,e,g,i,k) ${\textit{Re}}=9\times 10^4$ and (b,d, f,h,j,l) ${\textit{Re}}=3\times 10^4$ (adapted from NJOS25). Isopycnals are overlaid on the plots (red) for stratified cases. Radial extent ( $r/D$ ) of domain is $r/D=1$ unless explicitly mentioned.

The wake is characterised by a momentum deficit relative to an inbound flow and quantification of the mean defect velocity, defined as $U_d = U_\infty - U_x$ , is of importance. Figure 9 displays $U_d$ contours at two streamwise locations $x/D=3\text{ and }10$ for ${\textit{Fr}}=\infty$ , $6$ and $1.9$ wakes at ${\textit{Re}}=9\times 10^4$ and ${\textit{Re}}=3\times 10^4$ (adapted from NJOS25). For the stratified cases, isopycnals (lines of constant density) are superimposed on the $U_d$ contours.

In the post-separation region at $x/D=3$ , the time-averaged wakes for all $\textit{Fr}$ cases at ${\textit{Re}}=9\times 10^4$ exhibit symmetry about the centre-plane ( $y=0$ ), as depicted in figures 9(a), 9(e) and 9(i). It is worth emphasising that, although the instantaneous wake of R90F $\infty$ was found to be laterally asymmetric (figure 2 g–i), the mean stays symmetric at all streamwise locations. This behaviour contrasts sharply with the asymmetric wakes observed at R30F $\infty$ and R30F6 (figure 9 b, f). Furthermore, all wakes at ${\textit{Re}}=9\times 10^4$ display a double-lobed structure with a lobe emanating from each lateral side of the body. For R90F $\infty$ , the lobes are well rounded (figure 9 a). In contrast, the stratified cases R90F6 and R90F1.9 demonstrate elongation of the lobes in the vertical direction, with this stretching increasing as stratification intensifies (figure 9 e,i). Thus, similar to the lower- ${\textit{Re}}$ case of NJOS25, buoyancy inhibits rollup and streamwise vortex formation. At the same $x/D$ location, significant distortion of the isopycnals is also observed within the wake for both stratified cases. The wake edges are bounded by isopycnals.

At downstream locations of $x/D=10$ , the double-lobed structure appears to be preserved for the R90F $\infty$ case, as shown in figure 9(c). Two distinct local maxima for $U_d$ are also observed at these locations, similar to R30F $\infty$ (figure 9 d). These lobes constitute a streamwise vortex pair, discussed extensively by NJOS25. Furthermore, higher vertically downward drift of the wake structure is observed for R90F $\infty$ as compared with R30F $\infty$ . The stratified cases R90F6 and R90F1.9 exhibit a loss of the double-lobed structures by $x/D=10$ , as shown in figures 9(g) and 9(k). Both stratified wakes show a slight increase of the width relative to the height, demonstrating the anisotropic effects of stratification. The isopycnals tend to flatten out after $x/D=3$ for both cases. For ${\textit{Fr}}=6$ , the flattening process is slower compared with ${\textit{Fr}}=1.9$ due to weaker buoyancy, as R90F6 exhibits slightly higher distortions of isopycnals at $x/D=10$ . A key ${\textit{Re}}$ effect is that the wakes at ${\textit{Re}}=9\times 10^4$ are more mixed at $x/D =10$ relative to ${\textit{Re}}=3\times 10^4$ as is reflected by the spatial contours. Due to this increased mixing, the $U_d$ magnitude also decays faster at higher ${\textit{Re}}$ , as shown by the difference in contour levels at $x/D=10$ .

Figure 10. Contours of mean defect velocity ( $U_d$ ) for R90F1 at four streamwise locations: (a) $x/D=3$ ; (b) $x/D=5$ ; (c) $x/D=10$ and (d) $x/D=20$ . Isopycnals are overlaid on the plots (red). Radial extent of domain is $r/D=1$ for each panel.

At the strong stratification of ${\textit{Fr}}=1$ , buoyancy significantly alters the wake structure and dynamics. NJOS25 observed two distinct patches in the mean defect velocity ( $U_d$ ) contours for the R30F1 wake (see their figure 11) near separation ( $x/D=3$ ), with a smaller secondary defect region present above the larger primary region. This structure, characterised by two local maxima, evolved downstream into distinct layers. Figure 10 displays $U_d$ contours for the R90F1 wake at four streamwise locations: $x/D=3, 5, 10$ and $20$ . At $x/D=3$ , two distinct $U_d$ patches are observed, similar to R30F1. At both values of ${\textit{Re}}$ , the edges of the $U_d$ contours appear to align consistently with the isopycnals throughout the domain, resembling a jigsaw fit. It can be inferred that buoyancy strongly influences the flow topology in the vertical plane leading to a secondary wake above a primary wake. Progressing further downstream, this double-layered structure of the wake is maintained throughout the domain. Overall, the $\textit{Fr} = 1$ wake remains double-layered at ${\textit{Re}}=9\times 10^4$ although the boundaries of the two layers are not as sharp as at ${\textit{Re}}=3\times 10^4$ . Both R90F1 and R30F1 wakes are found to be transitional throughout the domain, i.e. they do not become fully turbulent, and therefore, are not discussed further in this study.

4.2. Vertical velocity for spheroid at pitch

When a slender object, a prolate spheroid in this study, is placed at a moderate pitch angle with respect to the oncoming flow, a substantial vertical velocity is imparted to the fluid by the body. Furthermore, a counter-rotating vortex pair is observed in the wake and it induces a significant vertical velocity between the vortices. Figure 11 compares the mean vertical velocity ( $U_z$ ) between a moderate value of pitch angle ( $\alpha = 10^\circ$ , R90F $\infty$ ) and zero pitch ( $\alpha = 0^\circ$ , R90F $\infty$ A0). In the pitched case R90F $\infty$ (figure 11 ac), a region of negative $U_z$ is observed in the centre plane ( $y=0$ ) at all streamwise locations. These regions of negative $U_z$ are flanked by two weaker positive $U_z$ regions on either side at $x/D=10$ and $20$ . This is consistent with a $U_z$ field generated by a pair of counter-rotating vortices, with a positive vortex on the right and a negative vortex on the left, see figure 12. Furthermore, when compared with the mean defect velocity $U_d$ (figure 9 a,c), $U_z$ exhibits a similar magnitude to $U_d$ at all streamwise locations and, therefore, becomes significant to the evolution of wake turbulence statistics.

Figure 11. Contours of mean vertical velocity ( $U_z$ ) at three streamwise locations $x/D=3$ , $x/D=10$ and $x/D=20$ . Panels (a)–(c) corresponds to the $\alpha =10^\circ$ case R90F $\infty$ and panels (d)–( f) to the $\alpha =0^\circ$ case R90F $\infty$ A0. Radial extent ( $r/D$ ) of the domain is shown on each panel.

The unpitched R90F $\infty$ A0 case in figure 11(df) exhibits a significantly different $U_z$ field compared to the moderate pitched case. When the flow leaves the body at $x/D=3$ , the magnitude of $U_z$ is approximately 10 times smaller than that in the pitched case. Due to the absence of a counter-rotating vortex pair, no region of negative $U_z$ is observed in the centre-plane. Furthermore, as the flow progresses downstream, $U_z$ remains 10 times smaller compared with the pitched case.

4.3. Buoyancy effects on streamwise vorticity

The shedding of a streamwise vortex pair from a body and its vertical drift has been well studied for an homogeneous fluid in the literature. Recently, Ohh & Spedding (Reference Ohh and Spedding2024) and NJOS25 studied the vertical drift of the vortex wake for an inclined 6 : 1 spheroid at $\alpha =10^\circ$ . Unlike the unequal-strength vortices formed by the asymmetric separation reported by NJOS25, the vortex pair at the present higher ${\textit{Re}} = 9\times 10^4$ (figure 12) is symmetric. Since the downward drift of the vortex pair and its modelling was discussed by NJOS25, it is not further considered here.

Figure 12. Contours of mean streamwise vorticity ( $\langle \omega _x\rangle$ ) for R90F $\infty$ at three streamwise locations: (a) $x/D=5$ ; (b) $x/D=10$ and (c) $x/D=20$ . Radial extent ( $r/D$ ) of domain is shown on each panel.

Figure 13. Contours of (a–d) mean streamwise vorticity ( $\langle \omega _x\rangle$ ) and (e–h) baroclinic torque ( $\omega _{BT}$ ) for R90F6 at four streamwise locations $x/D=5$ , $x/D=10$ , $x/D=20$ and $x/D=30$ . Isopycnals are overlaid on the $\omega _{BT}$ contours (black). Radial extent ( $r/D$ ) of the domain is shown on each panel.

Buoyancy is found to substantially affect the evolution of streamwise vorticity even for the relatively high $\textit{Fr} = 6$ . Figure 13(ad) shows the mean streamwise vorticity ( $\langle \omega _x\rangle$ ) for R90F6 at four locations. Additionally, figure 13(eh) shows the corresponding streamwise component of the mean baroclinic torque ( $-(1/{Fr})^2\partial \langle \rho \rangle /\partial y$ ) denoted succinctly by $\omega _{BT}$ . At $x/D=5$ , the primary vortex pair looks very similar to R90F $\infty$ (figure 12 a), with positive on the right and negative on the left. The induced velocity displaces the vortex pair vertically downwards. However, $\omega _{BT}$ has patches opposing the existing vorticity field with negative towards the right and positive towards the left. The opposing torque has two notable effects. First, the strength of the primary vortex pair is reduced so that it is substantially weaker at $x/D =10$ and is decimated by $x/D = 20$ . Second, its cumulative action generates two new secondary vortex pairs, one above and another below the centreline. For instance, compare $\omega _{BT}$ ( $x/D = 5)$ in the top half of figure 13(e) with $\langle \omega _x\rangle$ ( $x/D = 10$ ) in the top half of figure 13(b) and, similarly, the bottom halves of $\omega _{BT}$ ( $x/D = 10)$ and $\langle \omega _x\rangle$ ( $x/D = 20$ ).

Figure 14. Schematic for estimating the streamwise component of baroclinic torque ( $\omega _{BT}=-(1/{Fr})^2\partial \langle \rho \rangle /\partial y$ ).

In the near wake, the initial value of $\omega _{BT}$ not only depends on the stratification, but also on the geometry of the wake generator. We estimate the initial $\omega _{BT}$ as follows. Superscript $^*$ will distinguish dimensional quantities (when used) from their non-dimensional counterparts. Consider two fluid particles upstream of the spheroid (figure 14), where green has a vertical offset of $\delta z^*$ and a spanwise offset of $\delta y^*$ , and assume that the particles take the shown simplified paths. Initially, blue is vertically offset from green by $-\delta z^* = -6D \sin (\alpha )$ and is heavier than green by

(4.1) \begin{equation} \delta \rho ^* = -6D \sin (\alpha )C, \end{equation}

where $C$ is the dimensional background density gradient in the vertical. Reverting back to non-dimensional variables with $D$ for length and $CD$ for the density variation, it follows that

(4.2) \begin{equation} \delta \rho = -6 \sin (\alpha ){\textrm {sgn}}(C), \end{equation}

where $\textrm {sgn}(C) = -1$ . After travelling their simplified pathlines, both particles eventually reach the same vertical level, while maintaining the initial spanwise offset ( $\delta y^* = D/2$ ) of green relative to blue or $\delta y = 1/2$ , leading to a spanwise non-dimensional density gradient,

(4.3) \begin{equation} \frac {\partial \langle \rho \rangle }{\partial y} \sim \frac {-\delta \rho }{\delta y} = 12 \sin (\alpha ) \textrm {sgn}(C)\, . \end{equation}

The baroclinic torque $\omega _{BT}$ originating from the motion of the blue/green particle pair is

(4.4) \begin{equation} \omega _{BT} = \frac {-1}{{Fr}^2}\frac {\partial \langle \rho \rangle }{\partial y} = -\frac {12 \sin (\alpha )}{{Fr}^2} \textrm {sgn}(C) . \end{equation}

For the case R90F6, where $\alpha =10^\circ$ and ${\textit{Fr}}=6$ , the above-mentioned estimate leads to $\omega _{BT} = 0.06$ . Notwithstanding the assumption of simplified pathlines, we arrive at a reasonable estimate since the simulation result (figure 13 e) has $\omega _{BT} = O(0.1)$ . The sign is also correctly estimated for the $y\lt 0$ half-plane considered here. Since, $\omega _{BT}$ flips sign if a similar argument is followed for a blue particle that originates at $y^* = D/2$ instead of $y^* = - D/2$ , the final result for the initial value of non-dimensional baroclinic torque is written as

(4.5) \begin{equation} \omega _{BT} = \pm \frac {12 \sin (\alpha )}{{Fr}^2} . \end{equation}

By $x/D=10$ , a secondary vortex pair is formed directly above the primary vortex pair, but has the opposite sense of rotation (figure 13 b). Correspondingly, its motion reverses to vertically upward. By $x/D =20$ , the primary vortex has been replaced by two secondary vortex pairs, one above and the other below the centreline. Each secondary vortex pair drifts upward, $\omega _{BT}$ reverses sign too and acts to oppose the secondary vorticity. Thus, the vortex pair experiences a vertical oscillation.

It is worth noting that, at the lower ${\textit{Re}}$ of NJOS25 where the $\textit{Fr} = 6$ case exhibited asymmetry in the vortex pair formed after flow separation, buoyancy effects on the unequal-strength vortices resulted in intertwined patches of opposite-signed baroclinic torque that weakened the amplitude and coherence of the generated secondary vorticity. Consequently, the vertical oscillation of the wake found by Ohh & Spedding (Reference Ohh and Spedding2024) and NJOS25 is stronger in the present R90F6 case where the secondary vortex pair is more coherent.

5. Turbulent flow evolution

The wakes exhibit laminar boundary-layer separation from the body in both R30 and R90 series and eventually transition to turbulence downstream, except at ${\textit{Fr}} = 1$ . To analyse the turbulence level in the wake, the turbulent kinetic energy (TKE), defined as $E^T_K = ({1}/{2})\langle u^{\prime}_iu^{\prime}_i \rangle$ , is studied along with the component terms in the TKE transport equation, paying attention to buoyancy effects and ${\textit{Re}}$ -associated differences.

5.1. Streamwise evolution of TKE

Figure 15(a) depicts the streamwise evolution of area-integrated TKE ( $\{E^T_K\}=\int _A E^T_K \,{\rm d}A$ ) for the R90 series. The area integral is computed over a circular region of radius $r=4D$ . All $\textit{Fr}$ cases exhibit a peak in TKE immediately after leaving the body at $x/D\approx 3$ with comparable values. Beyond this location, a monotonic decay is observed. This location also corresponds to the breakup of wake into small-scale structures. The unstratified R90F $\infty$ wake displays a power-law decay of $\{E^T_K\}\sim x^{-2/3}$ , whereas the weakly stratified case R90F6 decays faster, following $\{E^T_K\}\sim x^{-1.1}$ . It is important to note that these fits are empirical, do not represent self-similar power laws and are employed here for differentiating among the various cases. The strongly stratified case R90F1.9 exhibits the lowest TKE of all $\textit{Fr}$ cases at ${\textit{Re}}=9\times 10^4$ throughout the evolution, with the decay slowing down towards the end of the domain. For this case, buoyancy-induced oscillations can also be observed in $\{E^T_K\}$ after $x/D\approx 10$ .

Figure 15. Area-integrated TKE evolution in the streamwise direction at (a) ${\textit{Re}}=9\times 10^4$ and (b) ${\textit{Re}}=3\times 10^4$ for $\textit{Fr} = \infty , 6$ and $1.9$ . All cases are for $\alpha =10^\circ$ . Dotted lines denote empirical curve fit.

Area-integrated TKE evolution for the R30 wakes is depicted in figure 15(b) for all $\textit{Fr}$ cases. R30F $\infty$ and R30F6 wakes are indistinguishable until approximately $x/D\approx 7$ or $Nt\approx 1.1$ , consistent with the similarities observed in mean defect velocity (figure 9 b, f). After $Nt \approx 1.1$ , buoyancy effects in the ${\textit{Fr}}=6$ wake become progressively important as will be elaborated within the framework of the TKE transport equation in the following subsections. The maxima of TKE for both cases occur at $x/D\approx 5$ and is slightly delayed when compared with ${\textit{Re}}=9\times 10^4$ . These peak values correspond to locations of transition to turbulence when the flow, after laminar boundary-layer separation, breaks up into small-scale structures (see figure 4 of NJOS25). At low ${\textit{Re}}$ , the relatively stronger viscous forces delay transition and, correspondingly, the peak in TKE. Notably, both $\textit{Fr}$ cases exhibit faster power-law decay of $\{E^T_K\}$ compared with their higher- ${\textit{Re}}$ counterparts. The evolution of the unstratified R30F $\infty$ wake is $\{E^T_K\}\sim x^{-1}$ instead of $x^{-2/3}$ at ${\textit{Re}} = 9 \times 10^4$ and, for the ${\textit{Fr}} = 6$ cases, $\{E^T_K\}\sim x^{-1.6}$ for R30F6 instead of $x^{-1.1}$ for R90F6.

Unlike all other cases in figure 15, TKE for R30F1.9 exhibits a slow and gradual increase after flow separation. For this case, TKE peaks quite far downstream from the body at $x/D\approx 12$ . Subsequently, TKE levels remain relatively low as the flow progresses downstream.

5.2. TKE budget terms

In the preceding subsection, it was noted that wakes at the two ${\textit{Re}}$ levels exhibit distinct TKE decay rates. Moreover, these decay rates are significantly influenced by the background stratification. To further understand these ${\textit{Re}}$ and $\textit{Fr}$ trends, we quantify the following TKE transport equation:

(5.1) \begin{equation} U_i \frac {\partial E^T_K}{\partial x_i} + \frac {\partial T_i}{\partial x_i} = P - \varepsilon + B , \end{equation}

where $T_i$ is the turbulent transport, $P$ is turbulent production, $\varepsilon$ is turbulent dissipation rate and $B$ is turbulent buoyancy flux defined as

(5.2) \begin{equation} P = -\big\langle u^{\prime}_{i}u^{\prime}_j\big\rangle \frac {\partial U_i}{\partial x_j}, \qquad \varepsilon = 2\nu \big\langle s^{\prime}_{\textit{ij}}s^{\prime}_{\textit{ij}}\big\rangle - \big\langle \tau ^{{\prime}s}_{\textit{ij}}s^{\prime}_{\textit{ij}} \big\rangle , \qquad B = -\frac {g}{\rho _0}\big\langle \rho ^{\prime}u^{\prime}_z \big\rangle . \end{equation}

Here, $s_{\textit{ij}}=(\partial _ju_i+\partial _iu_j)/2$ is the strain rate tensor and $\tau ^s_{\textit{ij}}=-2\nu _ss_{\textit{ij}}$ is the subgrid stress tensor. The subgrid term in the TKE transport equation is found to be relatively small in the wake at both ${\textit{Re}}$ – a consequence of the good grid resolution employed here.

Production $P$ acts as a source in the transport equation, extracting energy from the MKE reservoir and generating an equivalent amount of TKE. Dissipation $\varepsilon$ serves as a sink of TKE. Buoyancy $B$ facilitates energy exchange between TKE and TPE (turbulent potential energy), with a positive buoyancy flux converting TPE into TKE and vice versa. Unlike the aforementioned source and sink terms, turbulent transport ( $T_i$ ) is responsible for spatial redistribution of TKE and is given by

(5.3) \begin{equation} T_i = \frac {1}{2}\big\langle u^{\prime}_{i}u^{\prime}_{j}u^{\prime}_{j}\big\rangle + \big\langle u^{\prime}_{i}p^{\prime}\big\rangle - 2\nu \big\langle u^{\prime}_{j}s^{\prime}_{\textit{ij}}\big\rangle - \big\langle u^{\prime}_{j}\tau ^{{\prime}s}_{\textit{ij}}\big\rangle . \end{equation}

In the wake, turbulent transport redistributes energy, primarily in the $y{-}z$ plane. Consequently, the contribution of the area integral of $\partial T_i/ \partial x_i$ to the area-integrated TKE balance is negligible.

Figure 16. Area-integrated turbulent (a,c) production and (b,d) dissipation evolution in streamwise direction at (a,c) ${\textit{Re}}=9\times 10^4$ and (b,d) ${\textit{Re}}=3\times 10^4$ for $\textit{Fr} = \infty , 6$ and $1.9$ . All cases are for $\alpha =10^\circ$ . Dotted lines denote empirical curve fit.

The streamwise evolution of area-integrated production, $\{P\}$ , is depicted in figure 16(a) for the R90 series. The  $\{P\}$ has already reached its peak for all $\textit{Fr}$ cases by $x/D=3$ , followed by a consistent decline. Stratification generally tends to decrease $\{P\}$ . For example, $\{P\}\sim x^{-1.96}$ in the R90F $\infty$ wake, while the R90F6 wake exhibits $\{P\}\sim x^{-2.65}$ , which is a faster decay. For R90F1.9, production initially exhibits a rapid decay compared with other cases until $x/D\approx 7$ , beyond which, the decay rate slows down and $\{P\}$ becomes comparable to the other cases.

Area-integrated dissipation, $\{\varepsilon \}$ , in the R90 series also shows a decay after reaching its peak at $x/D\approx 3$ , as illustrated in figure 16(c). For the R90F $\infty$ wake, $\{\varepsilon \}\sim x^{-1.85}$ , with the decay rates increasing as the stratification level increases, indicating a damping effect due to buoyancy.

The R30 cases are shown in figures 16(b) and 16(d). Buoyancy affects the evolution of $\{P\}$ and $\{\varepsilon \}$ in R30F6 relative to R30F $\infty$ similar to its influence for their R90 equivalents. For example, $\{P\}$ for both values of $\textit{Fr}$ experiences a peak at $x/D\approx 4$ and subsequently decays with rates similar to the corresponding R90 cases. Moving to ${\textit{Re}}$ , its effect on the wake TKE is weak if the $\textit{Fr}$ value is fixed at $\infty$ or $6$ . Although the initial values of $\{P\}$ and $\{\varepsilon \}$ are smaller and their initial evolution is different for the ${\textit{Re}}=3\times 10^4$ wakes, their eventual decay rates are similar to their higher- ${\textit{Re}}$ counterparts.

The R30F1.9 wake stands out from all the other cases discussed in this section due to the absence of small-scale structures, resulting in a relatively small level of velocity fluctuations in the near wake. Both $\{P\}$ and $\{\varepsilon \}$ exhibit the lowest initial values among all cases and, after an initial increase, remain nearly constant until $x/D\approx 12$ or $Nt\approx 6$ . This region also corresponds to the gradual increase in TKE, as depicted in figure 15(b). Beyond this point, both quantities undergo a decay, with production values being comparable and dissipation slightly higher than that of the R90F1.9 wake.

Figure 17. Ratio of area-integrated production and dissipation with respect to streamwise coordinate at (a) ${\textit{Re}}=9\times 10^4$ and (b) ${\textit{Re}}=3\times 10^4$ . All cases are for $\alpha =10^\circ$ .

The ratio $\{P\}/\{\varepsilon \}$ (figure 17) is a direct measure of the relative significance of production and dissipation in the TKE balance. When $\{P\}/\{\varepsilon \}\lt 1$ , dissipation consumes more energy than production contributes, leading to a decrease in TKE. For all cases at both Reynolds numbers, $\{P\}/\{\varepsilon \}$ exceeds 1, while the flow leaves the body at $x/D=3$ , resulting in a streamwise increase of TKE. However, this trend is short-lived, as $\{P\}/\{\varepsilon \}$ soon dips below 1. This occurrence initiates the decay of TKE, which was previously seen in figure 15. For instance, all cases at ${\textit{Re}}=9\times 10^4$ exhibit $\{P\}/\{\varepsilon \}\lt 1$ from $x/D\approx 4.5$ onwards, accompanied by the onset of TKE decay at the same location. Another example is the R30F1.9 wake, which exhibits an extended region of $\{P\}/\{\varepsilon \}\gt 1$ until $x/D\approx 12$ , corresponding to the gradual TKE increase before the decay commences.

The relationship between the power laws of TKE and dissipation in some of the simulated wakes can be understood as follows. When the TKE budget is predominantly dominated by dissipation, the area-integrated TKE transport equation simplifies to

(5.4) \begin{equation} U_x \frac {\partial \{E^T_K\}}{\partial x} = - \{\varepsilon \} . \end{equation}

Assuming a power law for TKE as $\{E^T_K\}\sim x^{-p}$ and assuming $U_x\sim U_\infty$ remains relatively unchanged, (5.4) leads to $\{\varepsilon \} \sim x^{-p-1}$ . The above-mentioned relationship between TKE and dissipation power-law exponents is typically observed in this study when $\{P\}/\{\varepsilon \}$ falls below approximately 0.5. This correlation can be seen for the wake R90F6, where $\{E^T_K\}\sim x^{-1.1}, \{\varepsilon \}\sim x^{-2.1}$ and $p\approx -1.1$ . Similarly, it can be observed for R30F $\infty$ , where $\{E^T_K\}\sim x^{-1}, \{\varepsilon \}\sim x^{-1.85}$ and $p\approx -1$ , and also R30F6, where $\{E^T_K\}\sim x^{-1.6}$ , $\{\varepsilon \}\sim x^{-2.47}$ and $p\approx -1.6$ . However, $\{P\}/ \{\varepsilon \}$ does not become as small as 0.5 in the R90F $\infty$ wake and the dissipation does not follow the $x^{-p-1}$ scaling.

Figure 18. (a,b) Area-integrated turbulent buoyancy flux $\{B\}$ and (c,d) ratio of area-integrated buoyancy and dissipation for ${\textit{Re}}=9\times 10^4$ and ${\textit{Re}}=3\times 10^4$ , respectively. All cases are for $\alpha =10^\circ$ . Note that the $\{ B \}$ -scale in panel (b) is an order of magnitude smaller than that in panel (a).

We now move to a discussion of the direct contribution of the buoyancy flux to the TKE balance in the stratified wakes. The streamwise evolution of the area-integrated buoyancy flux, $\{B\}$ , is depicted in figure 18(a) for ${\textit{Re}}=9\times 10^4$ . It is evident that the R90F1.9 wake exhibits a stronger (larger amplitude) buoyancy flux compared with the R90F6 wake. For a given vertical displacement of the turbulent motions ( $\Delta z_t$ ) in a background with buoyancy frequency $N^2$ , the resulting buoyancy is $ B \sim g/\rho _0 \, ( \partial \rho _b/\partial z) \Delta z_t = -N^2 \Delta z_t$ . Since the initial downward $\Delta z_t$ at $x/D =3$ , just after flow separation, is set by the inclination of the spheroid and the Reynolds number, the difference among cases is small as is suggested by comparing among the rows of figure 2. Therefore, the higher $N^2$ case R90F1.9 has a higher buoyancy flux.

Moreover, $\{B\}$ for both wakes exhibits significant modulation by lee waves and oscillates with a wavelength equal to the wavelength ( $\lambda \approx 2\pi {Fr}$ ) that has been previously reported for body generated lee waves. Figure 18(c) shows the evolution of $\{B\}/\{\varepsilon \}$ for ${\textit{Re}}=9\times 10^4$ . For R90F6, the contribution of buoyancy is relatively small to the TKE budget until $x/D\approx 20$ or $Nt\approx 3$ , beyond which it contributes up to half of the dissipation value. Whereas, for R90F1.9, buoyancy becomes important close to the body, starting at $x/D\approx 6$ or $Nt\approx 3$ ( $Nt$ same as for R90F6). Along with high buoyancy induced fluctuations, buoyancy occasionally becomes comparable to dissipation, converting a significant amount of TKE into TPE and vice versa. Therefore, $B$ directly affects the R90F1.9 wake for most of the domain.

Turning to the lower- ${\textit{Re}}$ series (figure 18 b), the R30F6 wake displays a comparable level of $\{B\}$ as its higher ${\textit{Re}}$ counterpart R90F6 (note the smaller $\{B\}$ scale in panel a relative to panel b). Interestingly, the R30F1.9 wake exhibits a smaller amplitude of $\{B\}$ than R30F6, despite having a higher $N^2$ . The R30F1.9 near wake is quasi-laminar and lacking small-scale structures (see figure 4c of NJOS25), it experiences smaller density and velocity fluctuations (analogous to its low level of TKE and $\{B\}$ ) until $x/D\approx 12$ . Consequently, the buoyancy flux is small in this case. Figure 18(d) shows the evolution of the ratio $\{B\}/\{\varepsilon \}$ in the R30 series. For both stratified cases, $\{B\}/\{\varepsilon \}$ is smaller than for their higher ${\textit{Re}}$ counterparts and shows similar buoyancy-induced oscillations. The R30F1.9 wake shows slightly higher $\{B\}/\{\varepsilon \}$ towards the end of domain, but is less than $0.5$ for the most part.

The contrast in the TKE balance between ${\textit{Fr}}=\infty$ and $6$ can be summed up at both ${\textit{Re}}$ as follows. Production for ${\textit{Fr}}=6$ is lower than ${\textit{Fr}}=\infty$ , bringing in less TKE into the stratified case budget. Dissipation for ${\textit{Fr}}=6$ is also lower than ${\textit{Fr}}=\infty$ , taking less TKE out of the stratified budget. Buoyancy flux for ${\textit{Fr}}=6$ is not sufficient to directly influence the stratified budget. Also, just after the flow separates from the spheroid, TKE is similar for both wakes. Taking the aforementioned features into account, it follows that the reduced production at ${\textit{Fr}}=6$ is the key reason for faster TKE decay compared with ${\textit{Fr}}=\infty$ . Although the buoyancy flux is small, it plays a significant role in indirectly affecting the turbulence statistics, as will be discussed in the next subsection.

5.3. Components of turbulent production

The production term in (5.2) can be further expanded into

(5.5) \begin{equation} P = -\big\langle u^{\prime}_{x}u^{\prime}_{j}\big\rangle \frac {\partial U_x}{\partial x_j}-\big\langle u^{\prime}_{y}u^{\prime}_{j}\big\rangle \frac {\partial U_y}{\partial x_j}-\big\langle u^{\prime}_{z}u^{\prime}_{j}\big\rangle \frac {\partial U_z}{\partial x_j}. \end{equation}

Figure 19. (a) Area-integrated values of components $\{P_{xy}\}$ , $\{P_{\textit{xz}}\}$ and total production $\{P\}$ , and (b) the fractional contributions of the dominant components, $\{P_{\textit{ij}}\}/\{P\}$ , for the R90F $\infty$ A0 wake.

The previous section established that the area-integrated production, $\{P\}$ , exhibits a faster decay rate at ${\textit{Fr}}=6$ compared with ${\textit{Fr}}=\infty$ at both Reynolds numbers and leads to lower TKE. To delve deeper into the role of buoyancy in this behaviour, the contribution of all dominant production component terms, denoted as $P_{\textit{ij}}=-\langle u^{\prime}_{i}u^{\prime}_{j}\rangle \partial _jU_i$ , are quantified. A spheroid wake that develops in the $x$ direction is a laterally thin shear flow implying that $U_y, U_z \ll U_x$ and $|\partial U_x/\partial x|\ll |\partial U_x/\partial y|,|\partial U_x/\partial z|$ . Consequently, $P$ is dominated by transverse terms, specifically $P_{xy}=-\langle u^{\prime}_{x}u^{\prime}_{y}\rangle \partial _yU_x$ and $P_{\textit{xz}}=-\langle u^{\prime}_{x}u^{\prime}_{z}\rangle \partial _zU_x$ . In the present simulations of an unstratified spheroid wake at zero angle of attack (R90F $\infty$ A0), it is confirmed that $P_{xy}$ and $P_{\textit{xz}}$ indeed constitute the dominant components of production, as illustrated in figure 19(a). Similar to $\{P\}$ , both $P_{xy}$ and $P_{\textit{xz}}$ exhibit a monotonic decline after reaching a peak at $x/D\approx 5$ . Each of $P_{xy}$ and $P_{\textit{xz}}$ contributes approximately $50\,\%$ of the total $P$ . Note that the total contribution of these two terms near body at $x/D=4$ is more than 100 % of $P$ . This is because of the significant negative contribution by $P_{xx}=-\langle u^{\prime}_{x}u^{\prime}_{x}\rangle \partial _xU_x$ in the near wake caused by rapid acceleration of the mean flow near the separation region.

The situation is different for a pitched $6:1$ spheroid, despite a moderate $\alpha =10^\circ$ . Owing to the pitch and flow separation into a counter-rotating vortex pair, significant vertically downward velocity is imparted to the fluid, as was illustrated in § 4 (figure 11). Therefore, in addition to $P_{xy}$ and $P_{\textit{xz}}$ , terms related to mean vertical velocity, such as $P_{zy}=-\langle u^{\prime}_{z}u^{\prime}_{y}\rangle \partial _yU_z$ and $P_{zz}=-\langle u^{\prime}_{z}u^{\prime}_{z}\rangle \partial _zU_z$ , also contribute crucially to the total production. Since the relative behaviour of production components for ${\textit{Fr}}=\infty$ and ${\textit{Fr}}=6$ is qualitatively similar for both ${\textit{Re}}$ cases, except during an initial transition period, the following discussion is limited to the higher ${\textit{Re}}=9\times 10^4$ cases for brevity.

Figure 20. Area-integrated production components $\{P_{xy}\}$ , $\{P_{\textit{xz}}\}$ , $\{P_{zy}\}$ , $\{P_{zz}\}$ and total production $\{P\}$ for (a) R90F $\infty$ and (b) R90F6 wakes. Ratio of production components and total production, $\{P_{\textit{ij}}\}/\{P\}$ , for (c) R90F $\infty$ and (d) R90F6.

Figure 20(a) illustrates the evolution of the dominant production components alongside the total production for R90F $\infty$ . Similar to the total production, all the production components also reach a peak near $x/D\approx 3$ followed by a monotonic decline. Notably, $P_{zy}$ and $P_{zz}$ consistently demonstrate higher levels compared with $P_{xy}$ and $P_{\textit{xz}}$ for $x/D\gtrsim 5$ . As illustrated by figure 20(c), beyond $x/D\approx 5$ , $P_{zy}$ contributes approximately 35 % and $P_{zz}$ contributes approximately 30 % to the total $P$ . In contrast, $P_{xy}$ and $P_{\textit{xz}}$ both contribute only approximately 20 % each. These results highlight the importance of spanwise and vertical gradients of $U_z$ to TKE production. This is in stark contrast to the zero pitch case R90F $\infty$ A0 discussed previously (figure 19) where only the gradients of streamwise velocity ( $U_x$ ) are important.

The $P_{xy}$ for R90F6 exhibit nearly the same levels as those for R90F $\infty$ (figure 20 b), while $P_{\textit{xz}}$ shows a slight reduction. However, $P_{zy}$ and $P_{zz}$ show a much faster decay after $x/D\approx 4$ compared with their unstratified wake counterparts. Eventually, for the stratified wake, $P_{xy}$ and $P_{\textit{xz}}$ components become the dominant ones after $x/D\approx 7$ and remain so until the end of the domain. The ratio $\{P_{\textit{ij}}\}/\{P\}$ (figure 20 d) for both $P_{xy}$ and $P_{\textit{xz}}$ consistently increases for $x/D\gtrsim 5$ , with $P_{xy}$ contributing approximately 55 % and $P_{\textit{xz}}$ contributing approximately 40 % of the total $P$ towards the end of the domain. In contrast, $P_{zy}$ and $P_{zz}$ show reduced contributions, each contributing only approximately 15 % of the total $P$ towards the end of the domain. Since $P_{xy}$ is similar at both ${\textit{Fr}}=6$ and ${\textit{Fr}}=\infty$ , it follows that the reduction in $P_{\textit{xz}}$ , $P_{zy}$ and $P_{zz}$ contributions is responsible for the faster decay of the total production in the stratified wake.

Similar to R90F $\infty$ A0, the sum of all dominant production terms in both R90F $\infty$ and R90F6 wakes is slightly greater than 100 % of the total $P$ . This is due to the small negative contribution of certain production terms like $P_{xx}, P_{yy}$ and $P_{yz}$ .

5.4. Indirect effects of buoyancy flux on the turbulent statistics

Quantification of the Reynolds stresses ( $R_{\textit{ij}}=\langle u^{\prime}_{i}u^{\prime}_{j} \rangle$ ) involved in the dominant $P_{\textit{ij}}$ terms leads to further insights into the differential decay of production components under stratification. Figure 21 illustrates the streamwise evolution of the area-integrated absolute Reynolds stress components: $\{|R_{xy}|\}, \{|R_{yz}|\}, \{|R_{\textit{xz}}|\}, \{|R_{zz}|\}$ for R90F $\infty$ and R90F $6$ wakes. Since $R_{\textit{ij}}$ can be either positive or negative, we computed the area integral of its magnitude to capture the local impact of the Reynolds stress. It is evident that all $\{|R_{\textit{ij}}|\}$ components in both R90F $\infty$ and R90F6 cases commence with remarkably similar values after the flow separates from the body at approximately $x/D\approx 3$ . The $\{|R_{xy}|\}$ for both cases exhibit a close resemblance until $x/D\approx 20$ , beyond which a slightly faster decay is observed for the stratified case (figure 21 a). Since $\{|R_{xy}|\}$ undergoes minimal changes with stratification, it can be inferred that buoyancy has a negligible influence on altering $P_{xy}$ .

Figure 21. Area-integrated absolute Reynolds stress (a) $\{|R_{xy}|\}$ , (b) $\{|R_{yz}|\}$ , (c) $\{|R_{\textit{xz}}|\}$ and (d) $\{|R_{zz}|\}$ for R90F $\infty$ and R90F6.

The other Reynolds stresses, $\{|R_{yz}|\}, \{|R_{\textit{xz}}|\}$ and $\{|R_{zz}|\}$ , experience a faster decay when subjected to stratification of ${\textit{Fr}}=6$ compared with the unstratified case. For instance, $\{|R_{\textit{xz}}|\}$ for R90F6 begins to decay faster than R90F $\infty$ at $x/D\approx 4$ , with a significant drop observed after $x/D\approx 20$ (figure 21 c). This aligns with the observation of $P_{\textit{xz}} \lt P_{xy}$ in the ${\textit{Fr}}=6$ wake throughout the domain, unlike the ${\textit{Fr}}=\infty$ wake where $P_{\textit{xz}} \approx P_{xy}$ was observed (figure 20). Similarly, $\{|R_{yz}|\}$ and $\{|R_{zz}|\}$ also decay much faster for ${\textit{Fr}}=6$ than ${\textit{Fr}}=\infty$ for $x\gtrsim 4$ (figure 21 b, d). These results lead to a reduced contribution of $P_{yz}$ and $P_{zz}$ to the total production, as evident in figure 20.

One crucial point to note is that only the Reynolds stresses with $u^{\prime}_z$ , such as $R_{iz}=\langle u^{\prime}_i u^{\prime}_z \rangle$ , experience a reduction in the stratified case. In contrast, stratification appears to have a negligible impact on other Reynolds stresses. This phenomenon can be attributed to the role of buoyancy in the transport equations for $R_{\textit{ij}}$ . The $R_{\textit{ij}}$ transport equation for an incompressible flow without buoyancy, e.g. Pope (Reference Pope2000), needs to be supplemented by the following buoyancy term on its right-hand side,

(5.6) \begin{equation} B_{\textit{ij}} = -\frac {g}{\rho _0}\big(\big\langle \rho ^{\prime}u^{\prime}_{i}\big\rangle \delta _{jz} + \big\langle \rho ^{\prime}u^{\prime}_{j}\big\rangle \delta _{iz}\big). \end{equation}

The $B_{\textit{ij}}$ term specifically affects the Reynolds stress terms with $u^{\prime}_z$ , leading to a reduction in $\{|R_{\textit{xz}}|\}, \{|R_{yz}|\} \text{ and } \{|R_{zz}|\}$ . Consequently, this reduction also impacts the associated production components $P_{\textit{xz}}, P_{yz} \text{ and }P_{zz}$ . Note that $B_{zz} = 2B$ , where $B$ is the buoyancy term in the TKE equation.

It was shown in figure 18(a) that $\{B_{zz}\}$ ( $=2\{B\}$ ) remains negative throughout the evolution and, since it appears on the right-hand side of the $R_{zz}$ transport equation, would tend to decrease $R_{zz}$ . Moreover, $\{|R_{zz}|\}$ begins to decay rapidly at the location where $\{B_{zz}\}$ peaks, approximately $x/D\approx 4$ . This marks the threshold beyond which buoyancy starts influencing wake turbulence. The $B_{\textit{xz}}$ and $B_{yz}$ terms are not presented here for brevity.

To summarise, the results of this section show that, although buoyancy is not a significant sink of TKE, it plays an indirect role in significantly altering it. Buoyancy reduces turbulent production by suppressing its specific components that depend on the three Reynolds stress components that involve vertical velocity fluctuations ( $u^{\prime}_z$ ).

6. Summary and conclusions

The present work examines stratified wakes of a $6:1$ prolate spheroid at a moderate angle of attack $\alpha =10^\circ$ using high-fidelity large eddy simulation (LES). NJOS25 (Nidhan et al. Reference Nidhan, Jain, Ortiz-Tarin and Sarkar2025) discussed buoyancy effects on the behaviour of the mean wake – deficit velocity, topology and body forces – for a length-based ${\textit{Re}}=3\times 10^4$ and four stratification levels ( ${\textit{Fr}}=\infty ,6,1.9\text{ and }1$ ). In contrast, the current study focusses on wake turbulence – its energetics and also its modal analysis to identify unsteady coherent structures. Two values of Reynolds number ( ${\textit{Re}}=3\times 10^4\text{ and }9\times 10^4$ ) are considered for the same four stratification levels. The lower ${\textit{Re}}=3\times 10^4$ case of NJOS25 did not exhibit the distinctive unsteady coherent structures at the higher ${\textit{Re}}$ simulated here and the turbulence, especially for the stratified cases, was much weaker at ${\textit{Re}}=3\times 10^4$ . The mean wake also exhibits notable changes between the two cases.

The instantaneous wakes of R90F1.9 and R90F6 (figure 2) exhibit spanwise flapping just after flow separation. This flapping stays laterally symmetric at all times and has a frequency $St\sim 1.2$ as evidenced by point spectra (figure 4) in the shear layers formed after flow separation and SPOD eigenspectra (figures 5 and 6). The $St \sim 1.2$ frequency is higher than that of vortex shedding from a body and is typically associated with shear-layer instability. The leading SPOD eigenspectra also show a peak at $St \sim 1.2$ and, furthermore, the corresponding eigenmode of its spanwise velocity shows two laterally asymmetric patches. Thus, the flapping shear layer mode is the most energetic coherent structure in the near wake. Since both the wake structures stay symmetric at all times, no net spanwise unsteady force is imparted on the wake generator. The flapping mode does not persist far downstream and no significant peaks are found in the spectra at $x/D=10$ .

The near wake of R90F $\infty$ exhibits not only a flapping shear-layer mode ( $St\approx 1.28$ in figures 4 and 7) but also an oscillating asymmetric mode with a very low frequency, $St\approx 0.04$ . The $St\approx 0.04$ mode is responsible for the periodically alternating polarity of the wake asymmetry. This behaviour is unlike R30F $\infty$ , where the asymmetry was found to be non-oscillating. A similar transition from steady asymmetry to oscillating asymmetry was previously discussed by Tezuka & Suzuki (Reference Tezuka and Suzuki2006) for spheroid wakes at moderate pitch. The leading SPOD modes capture the coherent structures. The radial-velocity eigenmodes corresponding to the dominant peaks of $St\approx 1.28$ and $St\approx 0.04$ exhibit spanwise symmetry properties that are consistent with the temporal variability seen in wake visualisations. A significant cyclic spanwise force with an amplitude $\approx 10\,\%$ of drag and lift (figure 8) is found. This cyclic force has a frequency $St\approx 0.04$ , identical to that of the oscillating asymmetric mode. Similar to R90F6 and R90F1.9, both these dominant modes are limited to the near wake and are absent at $x/D = 10$ located after the wake transitions to broadband turbulence.

At ${\textit{Re}}=9\times 10^4$ , the mean wake evolution is found to be significantly different compared with that at ${\textit{Re}}=3\times 10^4$ (figure 9). All wakes at ${\textit{Re}}=9\times 10^4$ exhibit symmetry of the mean defect velocity in the spanwise direction. In contrast, the Re30F $\infty$ and Re30F6 wake are asymmetric in the mean. Although both values of ${\textit{Re}}$ exhibit laminar boundary-layer separation, there is evidently a qualitative difference in the details of flow separation. Interestingly, even though the instantaneous wake of R90F $\infty$ is found to be asymmetric (see the discussion of coherent structures in the preceding paragraph), the mean profile stays symmetric throughout the domain. Furthermore, all cases at ${\textit{Re}}=9\times 10^4$ exhibit higher rates of momentum mixing. Therefore, buoyancy-induced layering of the defect velocity at ${\textit{Fr}}=1.9$ is substantially reduced at ${\textit{Re}}=9\times 10^4$ relative to ${\textit{Re}}=3\times 10^4$ . The ${\textit{Fr}}=1$ wake exhibits a similar topology for both ${\textit{Re}}$ cases throughout the domain. Strong buoyancy effects on flow separation and on downstream mixing lead to a secondary wake above a primary wake (figure 10) at ${\textit{Re}}=9\times 10^4$ similar to the secondary wake found at the lower ${\textit{Re}}$ by NJOS25.

Inclined spheroid wakes are characterised by a streamwise vortex pair (figure 12) that advects downwards and induces considerable vertical velocity (figure 11). When compared with its zero-pitch counterpart, the vertical velocity in the $\alpha =10^\circ$ wake remains approximately 10 times stronger for the entire domain. This becomes quite important for the overall turbulence statistics as discussed with the TKE budget.

Stratification significantly alters the vortex pair dynamics through the generation of baroclinic torque with non-dimensional mean value denoted by $\omega _{BT}= - (1/\textit{Fr})^2\partial \langle \rho \rangle /\partial y$ . The magnitude of $\omega _{BT}$ in the near wake not only depends on stratification level, but also on the geometry of the wake generator. Simplified analysis for a $6:1$ spheroid placed at a pitch angle $\alpha$ (figure 14) shows that $\omega _{BT} \sim \pm 12 \sin (\alpha )/\textit{Fr}^2$ , which is established to be quite accurate when compared with simulation results. Buoyancy generated $\omega _{BT}$ is responsible for extinguishing the primary vortex pair and generating two secondary vortex pairs, directionally opposite in circulation to the primary pair (figure 13). These secondary vortex pairs advect vertically upwards. Thus, the baroclinic torque leads to a down-up motion of streamwise vorticity that is not seen when the fluid is unstratified.

Turbulence statistics are heavily influenced by buoyancy. A faster decay of turbulent kinetic energy (TKE) is seen in stratified cases when compared with the unstratified counterpart at both ${\textit{Re}}=3\times 10^4\text{ and }9\times 10^4$ (figure 15). There are some differences in the buoyancy effect at low ${\textit{Re}}$ . For example, the R30F1.9 wake has almost 2 orders of magnitude lower TKE initially than its R90F1.9 counterpart and remains transitional. The $\textit{Fr} =1$ wake is quasi-laminar with low values of TKE at both values of ${\textit{Re}}$ .

To understand the differential TKE decay between stratified and unstratified wakes, budget terms of the TKE transport equation are analysed. Both production and dissipation are reduced by buoyancy (figure 16) and over the entire domain. Furthermore, the buoyancy flux is found to be significantly smaller than other budget terms, especially at ${\textit{Fr}}=6$ (figure 18), and does not directly influence the TKE budget. Since dissipation, a sink of TKE, takes out less energy from the stratified wake budget as compared with the unstratified counterpart, it can be concluded that the faster TKE decay in stratified wakes can only be due to its reduced production.

Components of turbulent production were examined to gain insights into the buoyancy-related reduction of TKE discussed previously. For spheroid wakes at zero pitch angle, only streamwise velocity ( $U_x$ ) related terms $P_{xy}=-\langle u^{\prime}_{x}u^{\prime}_y\rangle \partial _yU_x$ and $P_{\textit{xz}}=-\langle u^{\prime}_{x}u^{\prime}_{z}\rangle \partial _zU_x$ are found to be significant (figure 19). However, even for moderate pitch, the vertical velocity ( $U_z$ ) is found to be important ( $U_z\sim U_x$ ) due to the presence of a counter-rotating vortex pair in the wake (figure 11). Therefore, along with $P_{xy}$ and $P_{\textit{xz}}$ , production components related to $U_z$ , namely, $P_{zy}=-\langle u^{\prime}_{z}u^{\prime}_y\rangle \partial _yU_z$ and $P_{zz}=-\langle u^{\prime}_{z}u^{\prime}_{z}\rangle \partial _zU_z$ , also become significant (figure 20). Indeed, in the unstratified wake, $P_{zy}$ and $P_{zz}$ dominate with values that substantially exceed those of $P_{xy}$ and $P_{\textit{xz}}$ . Thus, the presence of pitch angle fundamentally changes the TKE budget.

In the presence of density stratification and buoyancy, some of the production components are selectively altered. Out of all the dominant terms, a reduction in $P_{\textit{xz}}$ , $P_{zy}$ and $P_{zz}$ is observed even at the weak stratification of ${\textit{Fr}}=6$ , while $P_{xy}$ maintains a level similar to ${\textit{Fr}}=\infty$ (figure 20). Buoyancy-related suppression of vertical velocity fluctuations ( $u^{\prime}_z$ ) leads to reduction in the Reynolds stresses (figure 21) that appear in the production terms. Anisotropic suppression of $u^{\prime}_z$ , and the decrease of $\langle u^{\prime}_{z}u^{\prime}_y\rangle$ and $\langle u^{\prime}_{z}u^{\prime}_{x}\rangle$ have been previously documented for wakes (Meunier et al. Reference Meunier, Diamessis and Spedding2006; Brucker & Sarkar Reference Brucker and Sarkar2010; Chongsiripinyo & Sarkar Reference Chongsiripinyo and Sarkar2020) of bodies without pitch. The dominant production term $P_{zz}$ in the case with pitch involves $\langle u^{\prime}_{z}u^{\prime}_{z}\rangle$ , which is also suppressed by buoyancy. The buoyancy effects responsible for selective reduction of $P_{\textit{xz}}$ , $P_{zy}$ and $P_{zz}$ reduce the total production for stratified cases. It can be said that, although buoyancy is not strong enough to influence the TKE budget directly through the buoyancy flux, it indirectly influences the budget through reduction of the production terms involving vertical velocity.

The present results suggest that wake turbulence and dominant coherent structures are strongly influenced by the characteristics of flow separation. Given the ${\textit{Re}}$ effects found here in two cases, both with laminar boundary layer separation, a parametric study of higher- ${\textit{Re}}$ cases would help in understanding better the role of ${\textit{Re}}$ on the wake and body forces in laminar as well as turbulent boundary layer separation scenarios. Furthermore, larger pitch angles need to be studied. Finally, a detailed study of the vortex dynamics and the role of baroclinic torque in the high- ${\textit{Re}}$ stratified wake are worthy of exploration.

Supplementary movies

Supplemementary movies are available at https://doi.org/10.1017/jfm.2025.10923.

Funding

We gratefully acknowledge the support of ONR grant N00014-20-1-2253. Computational resources were provided by the Department of Defense High Performance Computing Modernization Program.

Declaration of interests

The authors report no conflict of interest.

References

Balaras, E. 2004 Modeling complex boundaries using an external force field on fixed Cartesian grids in large-eddy simulations. Comput. Fluids 33 (3), 375404.10.1016/S0045-7930(03)00058-6CrossRefGoogle Scholar
Berger, E., Scholz, D. & Schumm, M. 1990 Coherent vortex structures in thewake of a sphere and a circular disk at rest and under forced vibrations. J. Fluid. Struct. 4 (3), 231257.10.1016/S0889-9746(05)80014-3CrossRefGoogle Scholar
Bridges, D. 2006 The asymmetric vortex wake problem - asking the right question. In 36th AIAA Fluid Dynamics Conference and Exhibit. American Institute of Aeronautics and Astronautics.10.2514/6.2006-3553CrossRefGoogle Scholar
Brucker, K.A. & Sarkar, S. 2010 A comparative study of self-propelled and towed wakes in a stratified fluid. J. Fluid Mech. 652, 373404.10.1017/S0022112010000236CrossRefGoogle Scholar
Bucci, M.A., Puckert, D.K., Andriano, C., Loiseau, J.-C., Cherubini, S., Robinet, J.-C. & RistU. 2018 Roughness-induced transition by quasi-resonance of a varicose global mode. J. Fluid Mech. 836, 167191.10.1017/jfm.2017.791CrossRefGoogle Scholar
Chesnakas, C.J. & Simpson, R.L. 1997 Detailed investigation of the three-dimensional separation about a 6 : 1 prolate spheroid. AIAA J. 35 (6), 990999.10.2514/2.208CrossRefGoogle Scholar
Chevray, R. 1968 The turbulent wake of a body of revolution. J. Basic Eng-Trans. ASME 90 (2), 275284.10.1115/1.3605089CrossRefGoogle Scholar
Chongsiripinyo, K., Pal, A. & Sarkar, S. 2017 On the vortex dynamics of flow past a sphere at Re = 3700 in a uniformly stratified fluid. Phys. Fluids 29 (2), 020704.10.1063/1.4974503CrossRefGoogle Scholar
Chongsiripinyo, K., Pal, A. & Sarkar, S. 2019 Scaling laws in the axisymmetric wake of a sphere. In Direct and Large-Eddy Simulation XI, pp. 439444. Springer.10.1007/978-3-030-04915-7_58CrossRefGoogle Scholar
Chongsiripinyo, K. & Sarkar, S. 2020 Decay of turbulent wakes behind a disk in homogeneous and stratified fluids. J. Fluid Mech. 885, A31.10.1017/jfm.2019.1013CrossRefGoogle Scholar
Dairay, T., Obligado, M. & Vassilicos, J.C. 2015 Non-equilibrium scaling laws in axisymmetric turbulent wakes. J. Fluid Mech. 781, 166195.10.1017/jfm.2015.493CrossRefGoogle Scholar
Dommermuth, D.G., Rottman, J.W., Innis, G.E. & Novikov, E.A. 2002 Numerical simulation of the wake of a towed sphere in a weakly stratified fluid. J. Fluid Mech. 473, 83101.10.1017/S0022112002002276CrossRefGoogle Scholar
El Khoury, G.K., Andersson, H.I. & Pettersen, B. 2012 Wakes behind a prolate spheroid in crossflow. J. Fluid Mech. 701, 98136.10.1017/jfm.2012.135CrossRefGoogle Scholar
Fu, T.C., Shekarriz, A., Katz, J. & Huang, T.T. 1994 The flow structure in the lee of an inclined 6 : 1 prolate spheroid. J. Fluid Mech. 269, 79106.10.1017/S0022112094001497CrossRefGoogle Scholar
Germano, M., Piomelli, U., Moin, P. & Cabot, W.H. 1991 A dynamic subgrid-scale eddy viscosity model. Phys. Fluids 3 (7), 17601765.10.1063/1.857955CrossRefGoogle Scholar
Guo, P., Kaiser, F. & Rival, D.E. 2023 Vortex-wake formation and evolution on a prolate spheroid at subcritical reynolds numbers. Exp. Fluids 64 (10), 167.10.1007/s00348-023-03702-yCrossRefGoogle Scholar
Han, T. & Patel, V.C. 1979 Flow separation on a spheroid at incidence. J. Fluid Mech. 92 (4), 643657.10.1017/S002211207900080XCrossRefGoogle Scholar
Jacobitz, F.G. & Sarkar, S. 1998 The effect of nonvertical shear on turbulence in a stably stratified medium. Phys. Fluids 10, 11581168.10.1063/1.869640CrossRefGoogle Scholar
Jacobitz, F.G., Sarkar, S. & Van Atta, C.W. 1997 Direct numerical simulations of the turbulence evolution in a uniformly sheared and stably stratified flow. J. Fluid Mech. 342, 231261.10.1017/S0022112097005478CrossRefGoogle Scholar
Jiang, F., Andersson, H.I., Gallardo, J.P. & Okulov, V.L. 2016 On the peculiar structure of a helical wake vortex behind an inclined prolate spheroid. J. Fluid Mech. 801, 112.10.1017/jfm.2016.428CrossRefGoogle Scholar
Jiang, F., Gallardo, J.P., Andersson, H.I. & Zhang, Z. 2015 The transitional wake behind an inclined prolate spheroid. Phys. Fluids 27 (9), 093602.10.1063/1.4929764CrossRefGoogle Scholar
Jimenez, J.M., Hultmark, M. & Smits, A.J. 2010 The intermediate wake of a body of revolution at high Reynolds numbers. J. Fluid Mech. 659, 516539.10.1017/S0022112010002715CrossRefGoogle Scholar
Kumar, P. & Mahesh, K. 2018 Large-eddy simulation of flow over an axisymmetric body of revolution. J. Fluid Mech. 853, 537563.10.1017/jfm.2018.585CrossRefGoogle Scholar
Li, J.J.L., Yang, X.I.A. & Kunz, R.F. 2024 Direct numerical simulation of temporally evolving stratified wakes with ensemble average. J. Fluid Mech. 980, A3.10.1017/jfm.2023.1099CrossRefGoogle Scholar
Lin, Q., Lindberg, W.R., Boyer, D.L. & Fernando, H.J.S. 1992 Stratified flow past a sphere. J. Fluid Mech. 240, 315354.10.1017/S0022112092000119CrossRefGoogle Scholar
Lumley, J. 1970 Stochastic tools in turbulence. In Applied Mathematics and Mechanics (tech. rep.), vol. 12. Pennsylvania State University, Department of Aerospace Engineering.Google Scholar
Madison, T.J., Xiang, X. & Spedding, G.R. 2022 Laboratory and numerical experiments on the near wake of a sphere in a stably stratified ambient. J. Fluid Mech. 933, A12.10.1017/jfm.2021.1037CrossRefGoogle Scholar
Meunier, P., Diamessis, P.J. & Spedding, G.R. 2006 Self-preservation in stratified momentum wakes. Phys. Fluids 18 (10), 106601.10.1063/1.2361294CrossRefGoogle Scholar
Meunier, P. & Spedding, G.R. 2006 Stratified propelled wakes. J. Fluid Mech. 552, 229256.10.1017/S0022112006008676CrossRefGoogle Scholar
Nedić, J., Vassilicos, J.C. & Ganapathisubramani, B. 2013 Axisymmetric turbulent wakes with new nonequilibrium similarity scalings. Phys. Rev. Lett. 111 (14), 15.10.1103/PhysRevLett.111.144503CrossRefGoogle ScholarPubMed
Nelson, R.C., Corke, T.C. & Matsuno, T. 2006 Visualization and control of fore-body vortices. In Proceedings of the 12th ISFV (306).Google Scholar
Nelson, R.C. & Pelletier, A. 2003 The unsteady aerodynamics of slender wings and aircraft undergoing large amplitude maneuvers. Prog. Aerosp. Sci. 39 (2–3), 185248.10.1016/S0376-0421(02)00088-XCrossRefGoogle Scholar
Nidhan, S., Chongsiripinyo, K., Schmidt, O.T. & Sarkar, S. 2020 Spectral proper orthogonal decomposition analysis of the turbulent wake of a disk at Re = 50 000. Phys. Rev. Fluids 5 (12), 124606.10.1103/PhysRevFluids.5.124606CrossRefGoogle Scholar
Nidhan, S., Jain, S., Ortiz-Tarin, J.L. & Sarkar, S. 2025 Stratified wake of a 6 : 1 prolate spheroid at a moderate pitch angle. J. Fluid Mech. 1009, A58.10.1017/jfm.2025.204CrossRefGoogle Scholar
Nidhan, S., Schmidt, O.T. & Sarkar, S. 2022 Analysis of coherence in turbulent stratified wakes using spectral proper orthogonal decomposition. J. Fluid Mech. 934, A12.10.1017/jfm.2021.1096CrossRefGoogle Scholar
Ohh, C. & Spedding, G.R. 2024 The effects of stratification on the near wake of 6 : 1 prolate spheroid. J. Fluid Mech. 997, A43.10.1017/jfm.2024.829CrossRefGoogle Scholar
Orlanski, I. 1976 A simple boundary condition for unbounded hyperbolic flows. J. Comput. Phys. 21 (3), 251269.10.1016/0021-9991(76)90023-1CrossRefGoogle Scholar
Ortiz-Tarin, J.L., Chongsiripinyo, K.C. & Sarkar, S. 2019 Stratified flow past a prolate spheroid. Phys. Rev. Fluids 4 (9), 094803.10.1103/PhysRevFluids.4.094803CrossRefGoogle Scholar
Ortiz-Tarin, J.L., Nidhan, S. & Sarkar, S. 2021 High-reynolds-number wake of a slender body. J. Fluid Mech. 918, A30.10.1017/jfm.2021.347CrossRefGoogle Scholar
Ortiz-Tarin, J.L., Nidhan, S. & Sarkar, S. 2023 The high-reynolds-number stratified wake of a slender body and its comparison with a bluff-body wake. J. Fluid Mech. 957, A7.10.1017/jfm.2023.24CrossRefGoogle Scholar
Pal, A., Sarkar, S., Posa, A. & Balaras, E. 2017 Direct numerical simulation of stratified flow past a sphere at a subcritical Reynolds number of 3700 and moderate Froude number. J. Fluid Mech. 826, 531.10.1017/jfm.2017.398CrossRefGoogle Scholar
Patel, V. & Kim, S. 1994 Topology of laminar flow on a spheroid at incidence. Comput. Fluids 23 (7), 939953.10.1016/0045-7930(94)90062-0CrossRefGoogle Scholar
Pope, S.B. 2000 Turbulent Flows. Cambridge University Press.Google Scholar
Posa, A. & Balaras, E. 2016 A numerical investigation of the wake of an axisymmetric body with appendages. J. Fluid Mech. 792, 470498.10.1017/jfm.2016.47CrossRefGoogle Scholar
Rigas, G., Oxlade, A.R., Morgans, A.S. & Morrison, J.F. 2014 Low-dimensional dynamics of a turbulent axisymmetric wake. J. Fluid Mech. 755, R5.10.1017/jfm.2014.449CrossRefGoogle Scholar
Rowe, K.L., Diamessis, P.J. & Zhou, Q. 2020 Internal gravity wave radiation from a stratified turbulent wake. J. Fluid Mech. 888, A25.10.1017/jfm.2020.40CrossRefGoogle Scholar
Schmidt, O.T. & Colonius, T. 2020 Guide to spectral proper orthogonal decomposition. AIAA J. 58 (3), 10231033.10.2514/1.J058809CrossRefGoogle Scholar
Sirovich, L. 1987 Turbulence and the dynamics of coherent structures. i. coherent structures. Q. Appl. Math. 45 (3), 561571.10.1090/qam/910462CrossRefGoogle Scholar
Spedding, G.R. 1997 The evolution of initially turbulent bluff-body wakes at high internal froude number. J. Fluid Mech. 337, 283301.10.1017/S0022112096004557CrossRefGoogle Scholar
Spedding, G.R., Browand, F.K. & Fincham, A.M. 1996 Turbulence, similarity scaling and vortex geometry in the wake of a towed sphere in a stably stratified fluid. J. Fluid Mech. 314, 53103.10.1017/S0022112096000237CrossRefGoogle Scholar
de Stadler, M.B. & Sarkar, S. 2012 Simulation of a propelled wake with moderate excess momentum in a stratified fluid. J. Fluid Mech. 692, 2852.10.1017/jfm.2011.489CrossRefGoogle Scholar
Stanley, S. & Sarkar, S. 1997 Simulations of spatially developing two-dimensional shear layers and jets. Theor. Comput. Fluid Dyn. 9, 121147.10.1007/s001620050036CrossRefGoogle Scholar
Tezuka, A. & Suzuki, K. 2006 Three-dimensional global linear stability analysis of flow around a spheroid. AIAA J. 44 (8), 16971708.10.2514/1.16632CrossRefGoogle Scholar
Towne, A., Schmidt, O.T. & Colonius, T. 2018 Spectral proper orthogonal decomposition and its relationship to dynamic mode decomposition and resolvent analysis. J. Fluid Mech. 847, 821867.10.1017/jfm.2018.283CrossRefGoogle Scholar
Yang, J. & Balaras, E. 2006 An embedded-boundary formulation for large-eddy simulation of turbulent flows interacting with moving boundaries. J. Comput. Phys. 215 (1), 1240.10.1016/j.jcp.2005.10.035CrossRefGoogle Scholar
Zhou, Q. & Diamessis, P.J. 2019 Large-scale characteristics of stratified wake turbulence at varying reynolds number. Phys. Rev. Fluids 4 (8), 084802.10.1103/PhysRevFluids.4.084802CrossRefGoogle Scholar
Zhu, T. & Morrison, J.F. 2021 Simulation of the turbulent axisymmetric bluff body wake with pulsed jet forcing. Phys. Rev. Fluids 6 (12), 124604.10.1103/PhysRevFluids.6.124604CrossRefGoogle Scholar
Figure 0

Figure 1. Schematic of the flow configuration in a cylindrical computational domain (not to scale). $L_x^{-}$, $L_x^{+}$ and $L_r$ refer to the upstream, downstream and radial domain distance, respectively. $\alpha$ is the pitch angle. The centre of the spheroid is at the origin of the coordinate system.

Figure 1

Table 1. Simulation parameters. $N_r, N_\theta , N_x$ correspond to the number of grid points in radial, azimuthal and streamwise directions, respectively. Here, ${\textit{Re}}_L$ and $\textit{Fr}$ are the major-axis Reynolds number and minor-axis-based Froude number, respectively. Case labels given as RxxFyy refers to a case with ${\textit{Re}}=$ xx $\times 10^3$ and ${\textit{Fr}}=$ yy. The label with A0 at the end refers to a zero angle of attack case

Figure 2

Figure 2. Contours of instantaneous defect velocity ($u_d=U_\infty -u_x$) at $x/D = 3$ for (a–c) R90F1.9, (d–f) R90F6 and (g–i) R90F$\infty$ at three distinct timestamps for each case. Radial extent for each panel is $r/D=0.75$.

Figure 3

Figure 3. Instantaneous vertical vorticity ($\omega _z$) contours plotted on the centre horizontal plane ($z=0$) for R90F1.9 at three distinct timestamps similar to figure 2(ac). The separated shear layers exhibit Kelvin–Helmholtz billows.

Figure 4

Figure 4. Point spectra of spanwise velocity fluctuation ($u^{\prime}_y$) for (a) R90F1.9, (b) R90F6 and (c,d) R90F$\infty$ at different locations in the wake, where panel (d) is computed for a longer time series. Location of points A, B and C for each $\textit{Fr}$ case is marked atop the $u_d$ contours shown for $x/D=3$. At $x/D=10$, spectra shown in dashed red represent a point at the edge of the wake and dashed green represent a point on the vertical centre-plane.

Figure 5

Table 2. SPOD parameters for ${\textit{Re}}=9\times 10^4$ wakes

Figure 6

Figure 5. (a) SPOD eigenspectra, $\lambda ^{(1)}, \lambda ^{(2)},\ldots , \lambda ^{(n)}$, for all eigenmodes of R90F1.9 at a streamwise location of $x/D=3$. (b) Real part of the leading SPOD mode for the spanwise velocity, ${\varPhi}_{v}^{(1)}(y,z,St;x/D)$, corresponding to the peak in the eigenspectrum of $\lambda ^{(1)}$.

Figure 7

Figure 6. Same as figure 5 for R90F6.

Figure 8

Figure 7. Same as figure 5 for R90F$\infty$. Here, panel (b) corresponds to the leading eigenmode at $St\approx 1.28$ peak and panel (c) corresponds to eigenmode of low-frequency mode at $St\approx 0.04$.

Figure 9

Figure 8. (a) Time evolution of total spanwise force on the body, (b) Spectra of total force in the spanwise, vertical and streamwise directions for R90F$\infty$.

Figure 10

Figure 9. Contours of mean defect velocity ($U_d$) for two streamwise locations (a,b,e, f,i,j) $x/D=3$ and (c,d,g,h,k,l) $x/D=10$ are shown for (a–d) ${\textit{Fr}}=\infty$, (e–h) ${\textit{Fr}}=6$ and (i–l) ${\textit{Fr}}=1.9$ at (a,c,e,g,i,k) ${\textit{Re}}=9\times 10^4$ and (b,d, f,h,j,l) ${\textit{Re}}=3\times 10^4$ (adapted from NJOS25). Isopycnals are overlaid on the plots (red) for stratified cases. Radial extent ($r/D$) of domain is $r/D=1$ unless explicitly mentioned.

Figure 11

Figure 10. Contours of mean defect velocity ($U_d$) for R90F1 at four streamwise locations: (a) $x/D=3$; (b) $x/D=5$; (c) $x/D=10$ and (d) $x/D=20$. Isopycnals are overlaid on the plots (red). Radial extent of domain is $r/D=1$ for each panel.

Figure 12

Figure 11. Contours of mean vertical velocity ($U_z$) at three streamwise locations $x/D=3$, $x/D=10$ and $x/D=20$. Panels (a)–(c) corresponds to the $\alpha =10^\circ$ case R90F$\infty$ and panels (d)–( f) to the $\alpha =0^\circ$ case R90F$\infty$A0. Radial extent ($r/D$) of the domain is shown on each panel.

Figure 13

Figure 12. Contours of mean streamwise vorticity ($\langle \omega _x\rangle$) for R90F$\infty$ at three streamwise locations: (a) $x/D=5$; (b) $x/D=10$ and (c) $x/D=20$. Radial extent ($r/D$) of domain is shown on each panel.

Figure 14

Figure 13. Contours of (a–d) mean streamwise vorticity ($\langle \omega _x\rangle$) and (e–h) baroclinic torque ($\omega _{BT}$) for R90F6 at four streamwise locations $x/D=5$, $x/D=10$, $x/D=20$ and $x/D=30$. Isopycnals are overlaid on the $\omega _{BT}$ contours (black). Radial extent ($r/D$) of the domain is shown on each panel.

Figure 15

Figure 14. Schematic for estimating the streamwise component of baroclinic torque ($\omega _{BT}=-(1/{Fr})^2\partial \langle \rho \rangle /\partial y$).

Figure 16

Figure 15. Area-integrated TKE evolution in the streamwise direction at (a) ${\textit{Re}}=9\times 10^4$ and (b) ${\textit{Re}}=3\times 10^4$ for $\textit{Fr} = \infty , 6$ and $1.9$. All cases are for $\alpha =10^\circ$. Dotted lines denote empirical curve fit.

Figure 17

Figure 16. Area-integrated turbulent (a,c) production and (b,d) dissipation evolution in streamwise direction at (a,c) ${\textit{Re}}=9\times 10^4$ and (b,d) ${\textit{Re}}=3\times 10^4$ for $\textit{Fr} = \infty , 6$ and $1.9$. All cases are for $\alpha =10^\circ$. Dotted lines denote empirical curve fit.

Figure 18

Figure 17. Ratio of area-integrated production and dissipation with respect to streamwise coordinate at (a) ${\textit{Re}}=9\times 10^4$ and (b) ${\textit{Re}}=3\times 10^4$. All cases are for $\alpha =10^\circ$.

Figure 19

Figure 18. (a,b) Area-integrated turbulent buoyancy flux $\{B\}$ and (c,d) ratio of area-integrated buoyancy and dissipation for ${\textit{Re}}=9\times 10^4$ and ${\textit{Re}}=3\times 10^4$, respectively. All cases are for $\alpha =10^\circ$. Note that the $\{ B \}$-scale in panel (b) is an order of magnitude smaller than that in panel (a).

Figure 20

Figure 19. (a) Area-integrated values of components $\{P_{xy}\}$, $\{P_{\textit{xz}}\}$ and total production $\{P\}$, and (b) the fractional contributions of the dominant components, $\{P_{\textit{ij}}\}/\{P\}$, for the R90F$\infty$A0 wake.

Figure 21

Figure 20. Area-integrated production components $\{P_{xy}\}$, $\{P_{\textit{xz}}\}$, $\{P_{zy}\}$, $\{P_{zz}\}$ and total production $\{P\}$ for (a) R90F$\infty$ and (b) R90F6 wakes. Ratio of production components and total production, $\{P_{\textit{ij}}\}/\{P\}$, for (c) R90F$\infty$ and (d) R90F6.

Figure 22

Figure 21. Area-integrated absolute Reynolds stress (a) $\{|R_{xy}|\}$, (b) $\{|R_{yz}|\}$, (c) $\{|R_{\textit{xz}}|\}$ and (d) $\{|R_{zz}|\}$ for R90F$\infty$ and R90F6.

Supplementary material: File

Jain et al. supplementary movie 1

Time evolution of instantaneous defect velocity (ud = U∞ − ux) contours at x/D = 3 for R90F1.9 (Re = 9×104 and Fr = 1.9) wake. Three frames of this movie at distinct timestamps are shown in figure 2 (a-c).
Download Jain et al. supplementary movie 1(File)
File 4.6 MB
Supplementary material: File

Jain et al. supplementary movie 2

Same as movie 1 for R90F6 (Re = 9 × 104 and Fr = 6) wake. Three frames of this movie at distinct timestamps are shown in figure 2 (d-f).
Download Jain et al. supplementary movie 2(File)
File 4.5 MB
Supplementary material: File

Jain et al. supplementary movie 3

Same as movie 1 for R90F∞ (Re = 9×104 and Fr = ∞) wake. Three frames of this movie at distinct timestamps are shown in figure 2 (g-i).
Download Jain et al. supplementary movie 3(File)
File 8.9 MB
Supplementary material: File

Jain et al. supplementary movie 4

Time evolution of instantaneous vertical vorticity (ωz) contours at horizontal plane (z = 0) for R90F1.9 (Re = 9 × 104 and Fr = 1.9) wake. Three frames of this movie at distinct timestamps are shown in figure 3 (a-c).
Download Jain et al. supplementary movie 4(File)
File 9.2 MB