This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Disk wakes in nonlinear stratification

Divyanshu Gola \aff1    S. Nidhan\aff1    J. L. Ortiz-Tarin\aff1    Hieu T. Pham\aff1    S. Sarkar\aff1\corresp [email protected] \aff1Department of Mechanical and Aerospace Engineering, University of California San Diego, CA 92093, USA
Abstract

Nonlinearity of density stratification modulates buoyancy effects. We report results from a body-inclusive large eddy simulation (LES) of a wake in nonlinear stratification, specifically for a circular disk at diameter-based Reynolds number (\Rey\Rey) of 50005000. Five density profiles are considered; the benchmark has linear stratification and the other four have hyperbolic tangent profiles to model a pycnocline. The disk moves inside the central core of the pycnocline in two of those four cases and, in the other two cases with a shifted density profile, the disk moves partially/completely outside the pycnocline. The maximum buoyancy frequency (NmaxN_{max}) for all the profiles is the same. The first part of the study investigates the centered cases. Nonuniform stratification results in increasing wake turbulence relative to the benchmark owing to reduced suppression of turbulence production as well as the wave trapping in the pycnocline. Steady lee waves are also quantified to understand limitations of linear theory. The second part pays attention to the effect of a relative shift between the pycnocline and the disk. The wake defect velocity decays faster in the cases with a shift. The effect of disk location on the Kelvin wake waves (a family of steady waves within the pycnocline) and its modal form is obtained and explained by solving the Taylor-Goldstein equation. The family of unsteady internal gravity waves that are generated by the wake is also studied and the effect of disk shift is quantified.

1 Introduction

A pycnocline is a layer of fluid whose density changes rapidly with depth and so does the stratification described by the buoyancy frequency (NN), where N2=(g/ρ0)ρ/zN^{2}=-(g/\rho_{0})\partial\rho/\partial z. Such layers in the ocean or atmosphere affect environmental turbulence and waves, the transport of tracers (nutrients, pollutants, etc.) and the interaction of the environment with engineered structures. The upper-ocean pycnocline is flanked by a mixed layer on the top and a deeper layer at the bottom. The mixed layer and the deep layer can also be stratified but their stratification levels are typically lower than that in the pycnocline. Since much of our knowledge of turbulence in stratified wakes is derived from canonical wakes in uniform stratification, it is useful to study the wake of a canonical body inside and near a pycnocline where the change in stratification is rapid and nonlinear.

Most of the past work on the flow features of bodies moving through a pycnocline has involved the internal wave structure. Robey (1997) studied internal waves generated by a sphere moving below a pycnocline using experimental and numerical techniques for a wide range of body-based \Rey\Rey and Fr. Nicolaou et al. (1995) experimentally studied the waves of an accelerating sphere in a thermocline. The phase configuration for two-dimensional trapped internal waves was theoretically and experimentally studied by Stevenson et al. (1986) for a cylinder moving in a thermocline generated using salt brine. Often, these experimental observations of wave patterns are compared to the theoretical results of earlier studies, e.g., Barber (1993) and Keller & Munk (1970), where analytical results involving the dispersion relation of internal waves and their modal wave forms for nonlinear stratification profiles are provided.

A hyperbolic tangent profile bridging two regions with different values of NN has been often used to model nonlinear density variation within and density jump across a pycnocline. Ermanyuk & Gavrilov (2002, 2003) used a hyperbolic tangent profile to study forces on an oscillating cylinder and sphere, while Grisouard et al. (2011) used a hyperbolic tangent profile with a constant bottom NN to study internal solitary waves in a pycnocline using direct numerical simulation. A hyperbolic profile of NN has been used for studying a weakly stratified shear layer adjacent to a uniformly stratified region (Pham et al., 2009) as well as an asymmetrically stratified jet (Pham & Sarkar, 2010).

There have been numerous studies on evolution of turbulent wakes under stratification. An early study of Froude number O(101000)O(10-1000) wakes by Lin & Pao (1979) showed that stratification starts to affect the wake at buoyancy time scale (NtNt) of O(1)O(1) resulting in a non-axisymmetric evolution. The multistage wake decay was later characterized by Spedding (1997) into three separate regimes, namely three-dimensional (3-D), non-equilibrium (NEQ), and quasi-two-dimensional (Q2D) and quantitative turbulence measurements were obtained in this and following laboratory studies.

Temporal simulations have enabled longer simulations in terms of evolved time (TT) or streamwise distance (xx) using the transformation T=x/UT=x/U_{\infty}, where UU_{\infty} is the free-stream velocity. Gourlay et al. (2001) found the appearance of pancake vortices in the Q2D regime in their temporal DNS of a wake at \Rey=104\Rey=10^{4} and Fr=10\mbox{{Fr}}=10. Brucker & Sarkar (2010) in their DNS of towed and self-propelled wakes at \Rey=104\Rey=10^{4} and 5×1045\times 10^{4} quantified the wake turbulence statistics and found that the higher defect velocity in stratified wakes is a result of the buoyancy induced reduction of the Reynolds shear stress and ultimately the turbulent production term. This results in a longer lifetime of the stratified wake, also verified later by Redford et al. (2015) in their DNS of a weakly stratified turbulent wake. They also found that the horizontal nature of the wake in Q2D regime resulted in the dominance of the lateral Reynolds shear stress so that the decay of the mean wake velocity was faster (compared to NEQ regime) and was accompanied by an increase in the turbulent kinetic energy. Other studies that use temporal wake models include Dommermuth et al. (2002), Diamessis et al. (2005), de Stadler & Sarkar (2012) and Abdilghanie & Diamessis (2013).

Temporal simulations have been shown to be sensitive to the choice of initial conditions (Redford et al., 2012). Besides, inclusion of vortex shedding at the body and lee wave generation presents a challenge to temporal models. In recent work on stratified wakes, these limitations have been overcome by body-inclusive simulations. Orr et al. (2015) conducted a numerical study of a sphere wake at \Rey=200\Rey=200 and 10001000 where they identified vortex shedding and lee waves. Pal et al. (2016) simulated the wake of a sphere at \Rey=3700\Rey=3700 using DNS and found that there is a resurgence of turbulent fluctuations below a critical Fr instead of a monotone suppression. This observation was further examined by Chongsiripinyo et al. (2017) who found that, despite the low Fr, vortex stretching was dominant in the near wake, resulting in small-scale, three dimensional turbulence. The decay of a disk wake at \Rey=5×104\Rey=5\times 10^{4} was examined by Chongsiripinyo & Sarkar (2020) who decomposed its evolution into three regimes based on buoyancy time scale and horizontal Froude number of the fluctuations (instead of the mean wake): weakly stratified, intermediately stratified and strongly stratified turbulence. Turbulence scales were also used to characterize wake transitions by Zhou & Diamessis (2019). Ortiz-Tarin et al. (2019) performed LES of flow past a prolate spheroid with laminar boundary layer separation into the near and intermediate wake and found that at FrO(1)\mbox{{Fr}}\sim O(1), buoyancy effects are stronger for slender bodies as compared to bluff bodies.

Interaction of the turbulent wake with the stratified background leads to the generation of unsteady internal waves. Abdilghanie & Diamessis (2013) found that the NEQ regime is prolonged at high \Rey\Rey, resulting in internal wave radiation that persists up to Nt100Nt\approx 100. Brucker & Sarkar (2010) showed that internal wave flux dominates the turbulent dissipation during 20<Nt<7520<Nt<75 for a wake at \Rey=5×104\Rey=5\times 10^{4} and Fr=4\mbox{{Fr}}=4. In contrast, Redford et al. (2015) showed that for very high Fr, the internal wave activity, although still more pronounced at the start of the NEQ regime, makes a negligible contribution to the turbulent energy budget. Rowe et al. (2020) characterized the angles for the strongest internal waves and found that after Nt=10Nt=10, internal wave radiation is an important sink for wake kinetic energy. Nidhan et al. (2022) used spectral proper orthogonal decomposition and connected the leading wake eigenmodes to the wave field. Besides numerical techniques, many experimental (Gilreath & Brandt, 1985; Hopfinger et al., 1991; Chomaz et al., 1993; Bonneton et al., 1993; Brandt & Rottier, 2015; Meunier et al., 2018) as well as theoretical (Sturova, 1974; Voisin, 1991, 1994, 2003, 2007) studies have been done to analyze the internal wave field in a stratified flow.

Study of the wake in the vicinity of a pycnocline with a characterization of turbulence and its interaction with trapped internal waves is limited. Pham & Sarkar (2010) studied the interaction between internal waves from a shear layer and an adjacent stratified jet, and classified waves that are trapped/reflected by the jet and waves that transmit through the jet. Sutherland & Yewchuk (2004) and Sutherland (2016) calculated analytical expressions for internal wave transmission through density staircases for a stationary, two-dimensional Boussinesq fluid. Voropayeva & Chernykh (1997) simulated a temporally evolving wake in a nonlinear stratification using a Reynolds-averaged turbulence model. To the best of the authors’ knowledge, the present study is the first body-inclusive wake LES for a nonlinearly stratified fluid. We aim to assess the effect of nonlinear stratification on wake turbulence, its interaction with trapped waves as well as characterize the far-field lee waves. The disk is chosen as a canonical generator of a bluff-body wake to avoid the computational expense incurred in resolving the boundary layer that develops on a long body.

Section 2 describes the numerical setup and the simulation parameters. The results are divided into two sections. The basic differences between linear and nonlinear stratification when the disk is centered in the pycnocline layer are studied in section 3. The effect of relative shift between the density profile and the disk, so that the disk is partially/completely outside the pycnocline layer, is discussed in section 4. We summarise and conclude the study in section 5.

2 Methodology

2.1 Governing equations and numerical scheme

The wake of a disk is simulated by solving the three dimensional, incompressible, unsteady form of the conservation equations for mass, momentum and density. A high-resolution large eddy simulation (LES) with the Boussinesq approximation for density effects is used. The disk is immersed perpendicular to a flow with velocity UU_{\infty}. The equations are numerically solved in cylindrical coordinates but both Cartesian (x,y,z)(x,y,z) and cylindrical (r,θ,x)(r,\theta,x) coordinates are appropriately used in the discussion. Here, xx is streamwise, yy is spanwise and positive along θ=0\theta=0^{\circ}, and zz is vertical and positive along θ=90\theta=90^{\circ} (Fig. 1). The density field (ρ(𝐱,t))(\rho(\mathbf{x},t)) is split into a constant reference density (ρ0)(\rho_{0}), the variation of the background (Δρb(z))(\Delta\rho_{b}(z)), and the flow induced deviation, (ρd(𝐱,t))(\rho_{d}(\mathbf{x},t)) so that ρ(𝐱,t)=ρ0+Δρb(z)+ρd(𝐱,t)\rho(\mathbf{x},t)=\rho_{0}+\Delta\rho_{b}(z)+\rho_{d}(\mathbf{x},t). The Reynolds number of the flow, defined as \Rey=UD/ν\Rey=U_{\infty}D/\nu with DD being the diameter of the disk, is 50005000. Since the stratification is non-uniform, the Froude number defined by Fr(z)=U/N(z)DFr(z)=U_{\infty}/N(z)D is variable and its case-dependent behavior is described in section 2.2. The minimum value of the Froude number (FrminFr_{min}) is set to 11 for all the cases.

The filtered nondimensional equations are as follows

Refer to caption

Figure 1: Schematic showing the simulation setup and domain.
uixi=0,\frac{\partial u_{i}}{\partial x_{i}}=0, (1)
uit+ujuixj=pxi+1\Reyxj[(1+νsgsν)uixj]ρdFrmin2δi3,\frac{\partial u_{i}}{\partial t}+u_{j}\frac{\partial u_{i}}{\partial x_{j}}=-\frac{\partial p}{\partial x_{i}}+\frac{1}{\Rey}\frac{\partial}{\partial x_{j}}\Big{[}(1+\frac{\nu_{sgs}}{\nu})\frac{\partial u_{i}}{\partial x_{j}}\Big{]}-\frac{\rho_{d}}{Fr_{min}^{2}}\delta_{i3}, (2)
ρt+ujρxj=1\ReyPrxj[(1+κsgsκ)ρxj],\frac{\partial\rho}{\partial t}+u_{j}\frac{\partial\rho}{\partial x_{j}}=\frac{1}{\Rey Pr}\frac{\partial}{\partial x_{j}}\Big{[}(1+\frac{\kappa_{sgs}}{\kappa})\frac{\partial\rho}{\partial x_{j}}\Big{]}, (3)

where uiu_{i} refers to the filtered velocities in xx, yy, and zz directions for i=1,2i=1,2, and 33 respectively. νsgs\nu_{sgs} and ν\nu in Eq. (2) are the subgrid-scale kinematic viscosity and the kinematic viscosity, respectively, while κsgs\kappa_{sgs} and κ\kappa in Eq. (3) are the subgrid scale diffusion coefficient and the diffusion coefficient, respectively. The Prandtl number defined by Pr=ν/κPr=\nu/\kappa is set to one. No major qualitative influence of PrPr was found in the wake study by de Stadler et al. (2010) who varied PrPr between 0.2 and 7. The dynamic eddy viscosity model (Germano et al., 1991) is employed to obtain νsgs\nu_{sgs}, following the implementation of Chongsiripinyo & Sarkar (2020), and the subgrid Prandtl number is also set to unity to obtain κsgs\kappa_{sgs}. The parameters used for nondimensionalization are as follows: free-stream velocity (UU_{\infty}) for velocity, disk diameter (DD) for length, advection time (D/UD/U_{\infty}) for time, dynamic pressure (ρU2\rho U_{\infty}^{2}) for pressure, and characteristic change in background density (D(Δρb/z)|max-D(\partial\Delta\rho_{b}/\partial z)|_{max}) for the flow induced density deviation.

The periodicity in the azimuthal direction is leveraged in solving the discretised pressure Poisson equation in the predictor step, reducing it to a pentadiagonal system of linear equations, which is then solved using a direct solver (Rossi & Toivanen, 1999). The disk is represented by the immersed body method of Balaras (2004) and Yang & Balaras (2006). At the inlet boundary, a uniform stream of velocity (UU_{\infty}) is imposed while an Orlanski-type convective boundary condition is used for the outflow (Orlanski, 1976). Neumann boundary condition is imposed at the radial boundary of the domain for density as well as the three velocity components. To prevent spurious reflection of waves back into the domain, sponge layers are used at the inlet (streamwise length 10D10D), outlet (streamwise length 10D10D) and the cylindrical walls (radial length 15D15D) of the domain to gradually relax the velocities and the density to their respective background values at the boundaries.

2.2 Density profiles

The profiles chosen for the variation of background density in the five cases are shown in Fig. 2. Except for the benchmark 111 case, the profiles have nonuniform N(z)N(z) and Fr(z)=U/N(z)DFr(z)=U/N(z)D. Profile 111 with linear stratification has a constant Fr(z)Fr(z) which is the conventional body-based Froude number of Fr=U/ND=1\mbox{{Fr}}=U_{\infty}/ND=1. The maximum value of the density gradient is the same for all four nonlinear profiles and corresponds to Frmin=1Fr_{min}=1. In Profile 614 (Fig. 2b), a hyperbolic tangent function is used to bridge two linearly stratified regions (6 and 4 represent the value, rounded to an integer, of the local FrFr in the linearly stratified regions). Profile-I1I (I standing for “Infinity”) is a complete hyperbolic tangent profile between two constant-density regions, therefore having infinite Froude number far above and below the body.

For the I1I and 614 profiles, the disk center coincides with the center of the density profile and, furthermore, the disk lies entirely within the nonlinearly stratified region. Profiles 614614- and 614+614+ are obtained after shifting profile 614 (and not the disk) vertically by 2.5D-2.5D and 3D3D, respectively. The disk center is located at z=0z=0 for all five profiles and the negative and positive signs in the names 614614- and 614+614+ indicate whether the profile is shifted downwards or upwards, respectively. The vertical shift in 614614- (Fig. 2d) is chosen so that the upper half of the disk lies in the linearly stratified region with Fr=6\mbox{{Fr}}=6 and the lower half in the pycnocline. The vertical shift in 614+614+ (Fig. 2e) is chosen so that the disk lies entirely in the linearly stratified region with Fr=4\mbox{{Fr}}=4 while being very close to the pycnocline, specifically, its upper edge is 0.5D0.5D below the pycnocline.

The profile I1I is given by the following function, which was also used by Ermanyuk & Gavrilov (2002, 2003) and Nicolaou et al. (1995) in their studies:

Δρb(z)=ρ0ε2tanh2zδ.\Delta\rho_{b}(z)=-\frac{\rho_{0}\varepsilon}{2}\tanh{\frac{2z}{\delta}}. (4)

The central nonlinear region of the 614 profile is obtained from the I1I profile. The linearly stratified regions of profile 614 are added by matching the slopes at z=1.25δz=1.25\delta and z=δz=-\delta so that the density profile is continuously differentiable in zz. This results in Froude numbers of 6.136.13 and 3.743.74 above and below the pycnocline layer respectively, for profile 614.

Refer to caption

Figure 2: Variation of background density for (a) 111, (b) 614, (c) I1I, (d) 614614-, (e) 614+614+, along with (f) respective buoyancy frequency profiles. Solid black line at the center of (a-e) represents the disk (not to scale). Profiles are summarized in table 1 and also discussed in text. For 614614-, the upper half of the disk is in a constant-NN region and the lower is in the pycnocline. For 614+614+, the disk is in the bottom constant-NN region with its upper edge 0.5D0.5D below the pycnocline.

The profiles 614614- and 614+614+ are shifted with respect to 614 and their central nonlinear region is as follows:

Δρb(z)=ρ0ε2tanh2(zs)δ,\Delta\rho_{b}(z)=-\frac{\rho_{0}\varepsilon}{2}\tanh{\frac{2(z-s)}{\delta}}, (5)

where ss denotes the shift and takes values 2.5D-2.5D and 3D3D for 614614- and 614+614+, respectively.

The local Froude number corresponding to this pycnocline region can be calculated using the local background buoyancy frequency:

Fr(z)=UεgDδD|cosh2(zs)δ|.Fr(z)=\frac{U_{\infty}}{\sqrt{\varepsilon gD}}\sqrt{\frac{\delta}{D}}\;\Big{|}\cosh{\frac{2(z-s)}{\delta}}\Big{|}. (6)

The minimum value of Fr(z)Fr(z) occurs at z=sz=s. Thus, the minimum Froude number for profiles 614 and I1I is

Frmin=Fr(z=s=0)=UεgDδD.Fr_{min}=Fr(z=s=0)=\frac{U_{\infty}}{\sqrt{\varepsilon gD}}\sqrt{\frac{\delta}{D}}. (7)

Eq. 7 highlights two important non-dimensional parameters for a body moving through a pycnocline: U/εgDU_{\infty}/\sqrt{\varepsilon gD}, which is the conventional Froude number defined using reduced gravity εg\varepsilon g, and δ/D\delta/D, which is the nondimensional thickness of the pycnocline. In the simulations, U/εgD=1/2U_{\infty}/\sqrt{\varepsilon gD}=1/\sqrt{2} and δ/D=2\delta/D=2 for profiles 614 and I1I to obtain Frmin=1Fr_{min}=1. Note that for 111, which has constant linear stratification throughout, Frmin=1Fr_{min}=1 still holds. Parameters for each case are given in Table 1.

2.3 Domain and Grid

The domain extends from x/D=Lx=30x/D=-L_{x}^{-}=-30 to x/D=Lx+=102x/D=L_{x}^{+}=102 in the streamwise direction and from r/D=0r/D=0 to r/D=Lr=60r/D=L_{r}=60 in the radial direction. The number of grid points employed for discretizing the domain is Nx=2176N_{x}=2176, Nθ=128N_{\theta}=128, and Nr=479N_{r}=479 in the streamwise, azimuthal and radial directions respectively, resulting in approximately 130130 million grid points. The LES grid is non uniform in the streamwise and radial direction and is designed to have high resolution. In terms of the Kolmogorov length (η=(ν3/ϵkT)1/4\eta=(\nu^{3}/\epsilon_{k}^{T})^{1/4}), the maximum value of Δx/η\Delta x/\eta is 4.394.39 and that of Δr/η\Delta r/\eta is 5.035.03. Also, the turbulent dissipation rate used for the calculation of η\eta includes the resolved scale dissipation (ϵ=2νsijsij\epsilon=2\nu\langle s_{ij}^{\prime}s_{ij}^{\prime}\rangle) as well as the subgrid scale dissipation (ϵsgs=τsgsijsij\epsilon_{sgs}=-\langle\tau_{sgs_{ij}}^{\prime}s_{ij}^{\prime}\rangle with τsgsij=2νsgsSij\tau_{sgs_{ij}}=-2\nu_{sgs}S_{ij}), where sijs_{ij}^{\prime} and SijS_{ij} are the fluctuating and mean strain rates respectively. Most of the turbulent dissipation resides in the resolved scales.

Case δD\frac{\delta}{D} Frmin\mbox{{Fr}}_{min} Frz+\mbox{{Fr}}_{z\rightarrow+\infty} Frz\mbox{{Fr}}_{z\rightarrow-\infty} sD\frac{s}{D} Comment
111111 N/A 11 11 11 0 N(z)=N(z)= constant (Fig. 1a)
I11I 22 11 \infty \infty 0 N(z±)=0N(z\rightarrow\pm\infty)=0 (Fig. 1c)
614614 22 11 6.136.13 3.743.74 0 N(z+)=0.163N(z\rightarrow+\infty)=0.163, N(z)=0.267N(z\rightarrow-\infty)=0.267 (Fig. 1b)
614614- 22 11 6.136.13 3.743.74 2.5-2.5 N(z)N(z) of 614614 shifted down (Fig. 1d)
614+614+ 22 11 6.136.13 3.743.74 33 N(z)N(z) of 614614 shifted up (Fig. 1e)
Table 1: Physical parameters of the simulated cases. For each case, \Rey=5000\Rey=5000 and Pr=1Pr=1. The domain with Lr=60L_{r}=60, Lθ=2πL_{\theta}=2\pi, Lx=30L_{x}^{-}=30, and Lx+=102L_{x+}=102 is discretised using Nr=479N_{r}=479, Nθ=128N_{\theta}=128, and Nx=2176N_{x}=2176 points.

3 Linear vs. nonlinear stratification

3.1 Steady lee waves

Disturbances in stratified environments generate internal waves. These can be steady lee waves generated by the body or unsteady internal waves generated by turbulence in the wake of body. In this section, the characteristics of the body generated lee waves on the vertical center plane for 111 and 614 will be analyzed using linear asymptotic theory. Note that since I1I is essentially unstratified away from the pycnocline layer, it does not show any lee waves.

Instantaneous contours of vertical velocity (ww) for 111 (top half) and 614 (top and bottom halves) on a vertical plane passing through the centerline are plotted in Fig. 3(a-c). The amplitude of steady lee waves decays moving away from the body. The lee waves observed in 111 are symmetric with respect to z=0z=0 while 614 has two different sets of lee waves (top and bottom) owing to the two different local values of Fr(z)Fr(z) (6.136.13 above and 3.743.74 below) in the linearly stratified regions above and below the pycnocline layer. The waves in the 614 case have a much smaller amplitude than in the 111 case and a larger wavelength.

Linear theory, given by Voisin (1991, 2003, 2007) is used to analytically predict the wavelength and amplitude of the lee waves. The theory computes a wave function, χ\chi from a linearized set of inviscid equations involving a source term qq on the right hand side of the continuity equation to model the moving body using a Green’s function approach for large times (Nt>>1Nt>>1), which in our case translates to x/D>>Frx/D>>Fr. Once χ\chi is known, ww is calculated. The expression for ww for the case of a Rankine ovoid as derived by Ortiz-Tarin et al. (2019) and employed for a 4:1 spheroid wake can be reduced to give the wave pattern on the vertical centerline plane as:

w(x,y=0,z)mNπUrxzcosψsin(NaUcosψ)sin(NUrxz),w(x,y=0,z)\sim-\frac{mN}{\pi U_{\infty}r_{xz}}\cos{\psi}\sin{\Big{(}\frac{Na}{U_{\infty}}\cos{\psi}\Big{)}}\sin{\Big{(}\frac{N}{U_{\infty}}r_{xz}\Big{)}}, (8)

where rxz=x2+z2r_{xz}=\sqrt{x^{2}+z^{2}} and ψ=arctan(z/x)\psi=\arctan(z/x). Also, mm and aa are calculated from the potential flow stagnation point solution for a Rankine ovoid of length LroL_{ro} and cross-sectional diameter DroD_{ro}:

(Lro24a2)2=(8maπU)Lro,Dro2=(8maπU)14a2+Dro2(L_{ro}^{2}-4a^{2})^{2}=\Big{(}\frac{8ma}{\pi U_{\infty}}\Big{)}L_{ro},\qquad D_{ro}^{2}=\Big{(}\frac{8ma}{\pi U_{\infty}}\Big{)}\frac{1}{\sqrt{4a^{2}+D_{ro}^{2}}} (9)

Note that Ortiz-Tarin et al. (2019) calculated mm and aa using the potential flow solution of a Rankine oval, which is a 2-dimensional solution. Eq. (9) is based on the solution for a Rankine ovoid (which is 3-dimensional), and serves as a correction to Ortiz-Tarin et al. (2019). Note that, as per Eq. 8, the wave amplitude decreases with decreasing NN or increasing Fr.

The justification for using the expression for a Rankine ovoid instead of a disk is as follows. To apply linear theory to a bluff body such as a disk, the separation bubble created by the disk should also be included as a part of the extended body involved in wave generation. The complete extended body for 111, I1I and 614 cases resembles an ovoid with Lro2.5DL_{ro}\approx 2.5D and Dro1.5DD_{ro}\approx 1.5D. It should be noted that the exact body shape is insignificant as far as the wavelength of the waves is considered, but the wave amplitude depends on the body shape.

Refer to caption

Figure 3: Vertical velocity contours showing lee waves and their amplitudes along the dashed arrow labeled by d/Dd/D. (a,d) - 111 (9090^{\circ} plane); (b,e) - 614 (9090^{\circ} plane); (c,f) - 614 (270270^{\circ} plane).

Fig. 3(d) contrasts the wave amplitude between simulation and the theoretical estimate of Eq. 8. The comparison is along the dashed line with an arrow in Fig. 3(a), which is at an angle of 4545^{\circ} to the xx-axis. The theory, when used with the Rankine ovoid approximation for the disk and the separated flow behind it, is able to accurately capture the amplitude variation specially moving away from the body for the 111 case which has linear stratification with constant NN.

A similar analysis is performed for the lee waves of 614 after taking the Froude number in the analysis to be that of the linearly stratified region, i.e. Fr=6.13Fr=6.13 for Fig. 3(e) and Fr=3.74Fr=3.74 for Fig. 3(f). This procedure results in the correct prediction of the wavelength of the propagating lee waves. Theory is able to capture the order of magnitude of the significantly reduced (relative to 111) wave amplitude but the theoretical estimate of the wave amplitude is an under-prediction. The disk moves through a local stratification corresponding to Fr=1\mbox{{Fr}}=1 but the lee waves form and propagate in the linearly stratified regions where the Froude number is larger (Fr=6.13\mbox{{Fr}}=6.13 and Fr=3.74\mbox{{Fr}}=3.74). Therefore, the true wave amplitude corresponds to Fr somewhat lower than that used in the theory and, thus, larger than the theoretical estimate. Since the wavelength is comparable to the variability scale of N(z)N(z), WKB theory cannot be used and we do not proceed further with linear analysis.

3.2 Mean defect velocity

All cases show a strong effect of buoyancy on the wake since Fr=1\mbox{{Fr}}=1 at the disk is common to them but there are differences among the cases as elaborated below. Fig. 4(a) shows the mean centerline defect velocity for the three cases plotted against the streamwise distance. The 111 case shows strong oscillatory modulation of the wake and its defect velocity, similar to the sphere wake (Pal et al., 2017). In contrast to 111, the modulation in 614 and I1I, although present, is small. The wavelength of modulation for 111 is 2πDFr2\pi DFr as in the linear-theory result (Eq. 3.1), and as noted in previous work.

All cases show a similar decay law of U0x0.18U_{0}\propto x^{-0.18} in the NEQ regime, which was also observed by Chongsiripinyo & Sarkar (2020) for a disk at a higher Reynolds number of 5×1045\times 10^{4}. The decay law transitions to x0.45x^{-0.45} around x/D=50(Nt=50)x/D=50(Nt=50) signaling the beginning of the Q2D regime. A notable difference in the defect velocity profiles occurs at x/D4x/D\sim 4, where the initial dip in the profile for 111 is larger in magnitude and occurs earlier than that of 614 and I1I, owing to the stronger lee wave in 111 as compared to 614 and I1I. The mean velocity contours on a vertical streamwise cut (Fig. 4b), show that the lee wave modulation in the wake for 111 is stronger and its separation bubble is smaller, which is consistent with the earlier dip in the profile of defect velocity (Fig. 4a).

Refer to caption

Figure 4: (a) Mean defect velocity at centerline. (b) Mean streamwise velocity contours on vertical plane passing through centerline. Dashed yellow line represents the separation bubble (U=0U=0).

3.3 Turbulent kinetic energy

The nonlinearly stratified cases exhibit stronger levels of turbulent kinetic energy (TKE). This difference is illustrated by the contours of TKE at x/D=20x/D=20 in Fig. 5(a), plotted for the three stratification profiles. The pycnocline cases have a higher TKE (by 50 - 100 %) relative to the 111 case over the entire wake length as shown by the area-averaged values of TKE plotted in Fig. 5(b). The area averaging at any streamwise location is done in the region covered by a circle of radius 3D3D, therefore containing most of the turbulent zone. The TKE is substantially larger in I1I and 614 relative to 111, especially in the NEQ regime. For example, I1I has almost twice the TKE of 111 at x/D=45x/D=45. It is worth noting the slight increase in TKE for the case 111 at x/D50x/D\approx 50 which is where the decay law for the mean defect velocity changes (Section 3.2), again pointing towards the transition to Q2D regime.

To explain the relatively high turbulence, we inspect the terms in the TKE budget, especially the lateral production (the production by vertical Reynolds shear stress is small) and the wave flux. Although the hyperbolic tangent profile is close to a linear profile near the centerline, the weaker stratification regions above and below are not as effective in suppressing near-wake turbulence stresses, which results in higher TKE production further downstream. Also, the weaker far-field stratification in 614 and I1I relative to 111 does not allow the full frequency range of waves generated in the Fr=1Fr=1 central region to escape into the far field. The reduction of wave energy flux traps fluctuations in the wake and results in a TKE increase. The two quantities, lateral TKE production and wave flux, are diagnosed in the following subsections.

=

Refer to caption

Figure 5: (a) Turbulent kinetic energy contours at x/D=20x/D=20. (b) Area averaged values of TKE as a function of streamwise distance.

3.3.1 TKE production

Contours of lateral production PxyP_{xy}, which is the dominant component of TKE production for all three cases, are plotted in Fig. 6(a) at x/D=20x/D=20. The increased PxyP_{xy} in 614 and I1I is partly a result of the higher lateral Reynolds shear stress uv\langle-u^{\prime}v^{\prime}\rangle.

Refer to caption

Figure 6: (a) Lateral production contours at xD=20\frac{x}{D}=20. (b) Area-averaged values of lateral production as a function of streamwise distance.

Area averaged-values of the lateral production over a circle of radius 3D3D are shown in Fig. 6(b). The lateral production is consistently larger in the 614 and I1I cases over 10<x/D<5010<x/D<50 and is as much as twice at some locations.

3.3.2 Energy flux of wake generated unsteady waves

The dispersion relation for internal gravity waves in a stratified non-rotating medium,

Ω=NcosΘ,\Omega=N\cos{\Theta}\,, (10)

where Θ\Theta is the angle of the phase line with the vertical, limits the maximum allowable frequency to NN.

Refer to caption

Figure 7: (a) Radial waveflux contours at xD=20\frac{x}{D}=20. (b) Total waveflux integrated over rectangular perimeter as a function of streamwise distance.

Since weaker stratification supports a narrower frequency band of wave propagation, 614 and I1I can be expected to allow a smaller wave energy flux to radiate through the pycnocline layer as compared to 111. This is indeed the case as can be seen by the contours of the radial wave flux pur\langle p^{\prime}u_{r}^{\prime}\rangle in Fig. 7(a), where it is evident that the wave flux is restricted within the pycnocline layer (2<z/D<2-2<z/D<2) instead of being radiating away as in 111. Note that the mean of the radial velocity and pressure are subtracted when computing the flux so as to discard the component from the steady lee waves.

A quantitative comparison among the cases is performed by computing the line integral of the internal wave flux over a rectangular perimeter. Specifically, the integral

C1ρ0𝐮p.𝐧dl\oint_{C}\frac{1}{\rho_{0}}\langle{\bf u}^{\prime}p^{\prime}\rangle.{\bf n}\,dl (11)

which is a function of x/Dx/D, is calculated using a rectangle CC with vertical sides at y/D=±4y/D=\pm 4 and top/bottom sides at z/D=±2.5z/D=\pm 2.5 (thereby containing the pycnocline layer). Fig.7(b) shows that the flux of wave energy radiated away from the wake of 111 is much larger than 614 and I1I in the early part of the NEQ regime. However, at x/D=40x/D=40 in the late NEQ stage, the fluxes for 614 and I1I overtake the flux of 111. Both of these trends can be explained by the schematic shown in Fig. 8(a). The wake generated internal waves that are trapped inside undergo complex wave-wave interaction after getting vertically restricted, resulting in sideward escape of the wave flux. The sideward escape can be verified by using an appropriate scaling argument for the vertical as well as lateral wave flux at the top and side boundaries of the rectangle, respectively.

pwprmswrms,pvprmsvrms\langle p^{\prime}w^{\prime}\rangle\sim p_{rms}w_{rms},\langle p^{\prime}v^{\prime}\rangle\sim p_{rms}v_{rms} (12)

The values of wrmsw_{rms} at the point PtopP_{top}, vrmsv_{rms} at the point PsideP_{side} and prmsp_{rms} at PtopP_{top} and PsideP_{side} are shown in Fig. 8(b-e), respectively. Comparison of Fig. 8(d) and 8(e) shows that, although pressure fluctuations at the top boundary are highest initially for 111 among all cases and boundaries, the pressure fluctuations for 614 and I1I increase with time at the side boundaries because of the wave trapping and the resulting sideward escape. The velocity fluctuations at the boundaries also show similar trends. The sideward escape of the wave energy takes some time to manifest which, in the simulation frame, translates to streamwise distance, hence causing the slight delay in the increase of integrated wave flux for 614 and I1I. Fig. 8(b-e) are consistent with the following result: most of the wave flux for 614 and I1I that escapes the wake core does so in the form of lateral wave flux pv\langle p^{\prime}v^{\prime}\rangle unlike 111 where the vertical wave flux pw\langle p^{\prime}w^{\prime}\rangle is the major contributor.

Refer to caption

Figure 8: (a) Schematic showing trapping of internal waves for 614 and I1I compared to 111 (b) wrmsw_{rms} at PtopP_{top} (c) vrmsv_{rms} at PsideP_{side} (d) prmsp_{rms} at PtopP_{top} (e) prmsp_{rms} at PsideP_{side}.

4 Centered vs. shifted nonlinear stratifications

4.1 Taylor-Goldstein Equation and Kelvin wake waves

In cases involving pycnoclines, there is a family of steady waves which resemble a Kelvin ship wave pattern, on horizontal planes inside the pycnocline layer. Intuitively, a stratification profile with large central value of buoyancy frequency (FrO(1)\mbox{{Fr}}\leq O(1)) that changes rapidly in the vertical (over δ/DO(1)\delta/D\leq O(1)) to a small value at the edges can be thought of as a smoothed density jump and thus one can expect a pattern resembling that generated on the air-water interface by a moving ship. Both centered and shifted stratification profiles show Kelvin wake wave patterns. However, as demonstrated here, the wake wave structure differs qualitatively from the centered 614 wake when the stratification profile is shifted with respect to the disk center as in 614614- and 614+614+ .

The dispersion relation of air-water surface gravity waves leads to the waveforms of air-water Kelvin ship waves. To obtain an appropriate dispersion relation for the nonlinear continuously stratified profiles, the Taylor-Goldstein Equation is numerically solved:

d2ϕ(z)dz2+k2(N2(z)Ω21)ϕ(z)=0.\frac{d^{2}\phi(z)}{dz^{2}}+k^{2}\Big{(}\frac{N^{2}(z)}{\Omega^{2}}-1\Big{)}\phi(z)=0. (13)

Here, ϕ(z)\phi(z) is the vertical displacement eigenfunction, and kk and Ω(k)\Omega(k) are the wavenumber and angular frequency corresponding to the dispersion relation. The TGE is numerically solved as an eigenvalue problem for ϕ(z)\phi(z) by treating Ω\Omega as an eigenvalue and kk as a parameter, eg. by Robey (1997). The boundary conditions used are

ϕ(z=Lr)=ϕ(z=Lr)=0.\phi(z=-L_{r})=\phi(z=L_{r})=0. (14)

Also, based on Sturm-Liouville theory, the eigenfunctions are normalized using:

z=Lrz=LrϕmN2(z)ϕn𝑑z=δmn.\int_{z=-L_{r}}^{z=L_{r}}\phi_{m}N^{2}(z)\phi_{n}\,dz=\delta_{mn}. (15)

Fig. 9(a) and 9(b) respectively show the mode 1 and mode 2 eigenfunctions for different values of kk. The numerically calculated dispersion relation is also compared in Fig. 9(c) to the approximation given by Barber (1993):

Ω(k)=cp0k1+cp0kNmax.\Omega(k)=\frac{c_{p_{0}}k}{1+\frac{c_{p_{0}}k}{N_{max}}}. (16)

Here, cp0c_{p_{0}} is the limiting long-wave phase speed of a given mode at k=0k=0, which is obtained by extrapolation, i.e., cp0c_{p_{0}} is the slope at the origin of the mode-specific curve in Fig. 9(c). For the chosen profiles, cp0/Uc_{p_{0}}/U_{\infty} for mode 1 and mode 2 are 2.232.23 and 0.680.68, respectively. Since Eq. 16 provides a very good approximation to the dispersion relation, it will be used for the remaining analysis of this section.

Refer to caption

Figure 9: (a) Mode-1 and (b) Mode-2 eigenfunctions for different values of wavenumber in Eq. 13. (c) Numerically calculated dispersion relation compared with approximation given by Barber (1993).(d) Mode-1 and (e) Mode-2 waveforms as calculated using Eq. 17.

Using the dispersion relation, phase and group velocities associated with each mode can be calculated using cp=Ω/kc_{p}=\Omega/k and cg=dΩ/dkc_{g}=d\Omega/dk, and then substituted in the expressions given by Keller & Munk (1970) to obtain the modal wave patterns corresponding to each eigenfunction:

x=φU(1cpcgU2)k(cpcg),y=φcg(1cp2U2)12k(cpcg),x=\frac{\varphi U_{\infty}\Big{(}1-\frac{c_{p}c_{g}}{U_{\infty}^{2}}\Big{)}}{k(c_{p}-c_{g})},y=\frac{\varphi c_{g}\Big{(}1-\frac{c_{p}^{2}}{U_{\infty}^{2}}\Big{)}^{\frac{1}{2}}}{k(c_{p}-c_{g})}, (17)

where (x,y)(x,y) corresponds to a locus of points on the horizontal plane parametrically given in terms of kk for a constant value of phase, φ\varphi. Fig. 9(d) and 9(e) respectively show mode 1 and mode 2 wave patterns, respectively, plotted using Eq. 17.

Fig. 10 shows the instantaneous radial velocity contours for 111, 614, 614614- and 614+614+ plotted on a half-horizontal plane passing through the respective pycnocline center. (For 111, the plot is on θ=0\theta=0^{\circ} plane). The waves in 614 have constant-phase lines which resemble mode 2 (shown earlier in Fig. 9e) but not mode 1. In contrast, the phase lines for 614614- and 614+614+ are more complex. In the region 0<z<100<z<10 they resemble the mode 1 pattern of Fig. 9(d) but, for z>10z>10, they resemble the mode 2 pattern. The manifestation of any mode by a disturbance in the pycnocline layer depends on the location of the disturbance with respect to the pycnocline layer. In 614, the disk center travels along the center of the pycnocline layer to displace fluid symmetrically in the upward and downward direction, corresponding to the antisymmetric mode 2 eigenfunction for displacement and therefore generates mode 2 waves. Any vertical offset from the pycnocline center modifies the antisymmetric pattern of the displacement so as to also involve some contribution from the mode 1 waveform. The presence of a mode 2 wave pattern in the central region of the lateral plane and a mode 1 pattern otherwise is in agreement with the visualizations in the experiment of Robey (1997). For profile 614+614+, the contours look similar to that of 614614- but with slightly lower intensity because of the disk being further away from the center of the pycnocline. (Note that the wavepattern for I1I is the same as that of 614 because they have the same hyperbolic tangent function modeling the nonlinearity.)

Refer to caption

Figure 10: Instantaneous contours of radial velocity on a half-pycnocline-center plane: (a) 111, (b) 614614-, (c) 614, and (d) 614+614+. Dashed lines correspond to mode 2 waves which were shown in Fig. 9(e).

4.2 Distinction between lee waves and Kelvin wake waves

Refer to caption


Figure 11: Instantaneous contours of radial velocity for the centered cases on vertically offset horizontal planes for 111 (left column), 614 (middle) and I1I (right). The vertical offset of each plane increases from the top to bottom row.

Kelvin wake waves are steady in the frame of reference of the moving body like the lee waves but constitute a distinct family of waves and appear only when the stratification is nonuniform. Fig. 11 shows instantaneous contours of radial velocity on horizontal planes for the centered profiles 111, 614, and I1I at different vertical locations. At z/D=2.5z/D=-2.5 (Fig. 11 (a-c)), the contours have superficial resemblance but they represent fundamentally different waves. For 111, since there is no nonlinear gradient, Kelvin wake waves are not formed on the horizontal plane passing through the domain centerline at z/D=0z/D=0, (e.g. the previously shown Fig. 10a), and so Fig. 11(a) has the horizontal imprint of only lee waves. However, for 614 and I1I, the contours at z/D=2.5z/D=-2.5 in Fig. 11(b,c) are mostly the imprints of the Kelvin wake waves in Fig. 10(c). For 614, Kelvin wake waves disappear farther away from the region of nonlinear stratification and the contours transition to those of lee waves corresponding to Fr=3.74\mbox{{Fr}}=3.74 (Fig. 11(e,h,k)). For I1I, once the Kelvin wake waves disappear, lee waves are not observed since the profile is unstratified outside the pycnocline region (Fig. 11(f,i,l)). Turning back to case 111, we see the lee waves weakening as z/Dz/D is increased (Fig. 11(d,g,j)) but their pattern is still clear at z/D=30z/D=-30.

4.3 Wake generated internal gravity waves

Refer to caption

Figure 12: w/z\partial w^{\prime}/\partial z contours showing wake generated internal gravity waves on the θ=90\theta=90^{\circ} plane for 111. Black lines are inclined at an angle of 3939^{\circ} with the vertical.

The internal gravity waves generated by the turbulent wake are unsteady. These waves are visualized in Fig. 12 for case 111 by contours of w/z\partial w/\partial z on the θ=90\theta=90^{\circ} vertical plane. The phase lines of the radiated waves in the far field are seen to cluster around a a characteristic inclination angle which suggests a narrow frequency band in the far field according to the dispersion relationship, Eq. 10, for the intrinsic (in a frame where the background has zero velocity) wave frequency. Fig. 12 also shows solid black lines plotted at an angle of 3939^{\circ} from the vertical, which also seems to be the angle for the waves (39±239^{\circ}\pm 2^{\circ}). This gives Ω/N=0.78±0.03\Omega/N=0.78\pm 0.03.

Refer to caption

Figure 13: w/z\partial w^{\prime}/\partial z contours showing wake generated internal gravity waves on the θ=90\theta=90^{\circ} and 180180^{\circ}plane for 614+614+.

For profile 614+614+, the w/z\partial w^{\prime}/\partial z contours on the entire vertical plane (θ=90\theta=90^{\circ} and 270270^{\circ}) are plotted in Fig. 13. Phase lines with a dominant inclination angle are seen in the negative z/Dz/D region where the waves with downward group velocity travel in an entirely linear stratification (Fr=3.74\mbox{{Fr}}=3.74). The inclination angles in this region of z/D<0z/D<0 cluster around Θ=44\Theta=44^{\circ} which gives Ω/N0.72\Omega/N\approx 0.72 (Note that NN here corresponds to the local Fr=3.74\mbox{{Fr}}=3.74). The upward moving waves generated by the disk enter a region of nonlinear stratification where there is significant small-scale variability. As the waves exit the pycnocline layer from the top, they have near-zero inclination with respect to the vertical, i.e. Θ0\Theta\approx 0^{\circ}, which implies that only the high-frequency near-NN (NN here again corresponds to the local Fr=6.13\mbox{{Fr}}=6.13) part of the wake generated waves escapes into the upper region of weak stratification. Thus, a significant portion of the wake generated waves is trapped within the pycnocline leading to the small-scale variability in that region.

4.4 Wake characteristics

Refer to caption

Figure 14: Evolution of centerline mean defect velocity for the five simulated cases.

Wake characteristics, namely, the mean defect velocity and turbulent kinetic energy, show qualitatively different behavior for the shifted pycnocline cases 614614- and 614+614+ relative to the centered 614 profile. Fig. 14 compares the streamwise evolution of centerline mean defect velocity (U0/UU_{0}/U_{\infty}) among all five simulated cases. Recall that for 614+614+, the disk is entirely in the Fr=3.74\mbox{{Fr}}=3.74 region with its upper edge in the pycnocline and, for 614614-, half of the disk is in the Fr=6.13\mbox{{Fr}}=6.13 region and half in the pycnocline. Since, for both 614614- and 614+614+, the disk wake evolves in relatively weaker stratification relative to the Fr=1Fr=1 stratification seen by the disk in 614, there is a faster rate of decay of U0U_{0} (x0.35\propto x^{-0.35}) in these cases (Fig. 14). 614614- has a weaker effective stratification relative to 614+614+ and, therefore, has a somewhat lower U0U_{0}.

Refer to caption

Figure 15: Turbulent kinetic energy contours at x/D=20x/D=20 for all five cases.

The TKE contours plotted in Fig. 15 also show that the buoyancy effect is weaker for the shifted profiles. 614614- and 614+614+ show far higher TKE than the other three cases that have Fr=1\mbox{{Fr}}=1 at the disk center. Since the bottom half of the disk of 614614- is inside the pycnocline layer, the wake core is not symmetric about the horizontal axis and appears to be somewhat compressed from the bottom where the effective stratification is higher. The TKE profile at x/D=20x/D=20 is more symmetric for 614+614+.

5 Summary and Conclusions

We present results from a body-inclusive LES of turbulent wakes in a background stratification which is nonlinear instead of the linear stratification that is typically considered. Four density profiles with a hyperbolic tangent pycnocline are selected along with a standard constant linear stratification profile. The nonlinear profiles are 614 (weak upper stratification with Fr6Fr\approx 6 and weak lower stratification with Fr4Fr\approx 4 bounding the pycnocline), I1I (no stratification surrounding pycnocline), 614614- (614 vertically shifted by 2.5D-2.5D), and 614+614+ (614 vertically shifted by 3D3D). A disk of diameter DD is chosen as the canonical bluff body, the relative velocity is UU_{\infty} and the Reynolds number is 50005000. The Froude number Fr(z)=U/N(z)DFr(z)=U_{\infty}/N(z)D, which is based on local value of buoyancy frequency N(z)N(z), varies and takes a minimum value of 11 for the cases with nonuniform N(z)N(z) and is set to 1 in the benchmark constant-NN case. The inverse of Fr(z)Fr(z) can be viewed as the local buoyancy frequency, nondimensionalized with the flow scales, so that nondimensional NmaxN_{max} is the same among all cases. The pycnocline thickness normalized by DD is also held constant.

Our main conclusion is that the non-constant NN profile studied here substantially alters both the turbulent wake and the internal wave field. In the first part of the study, results with the nonlinear profiles 614 and I1I are compared with the linear profile 111. The mean defect velocity (U0/UU_{0}/U_{\infty}) in the near wake (x/D<8x/D<8) is quite different for 111 which shows an initial oscillation with stronger amplitude than 614 and I1I. This initial oscillation of the defect velocity is the oscillatory modulation of the wake, which has been shown to be an imprint of the steady lee wave field in linear stratification (Pal et al., 2017). The I1I case, which is unstratified outside the pycnocline, does not support a steady lee wave in the far field. Nevertheless, its near wake exhibits oscillatory modulation (wavelength is slightly larger than that for 111) because of an evanescent steady wave. In the NEQ regime, which follows the initial oscillation, U0/UU_{0}/U_{\infty} in all three cases was found to decay as x0.18x^{-0.18}. Chongsiripinyo & Sarkar (2020) found the same power-law exponent of 0.18-0.18 in their disk wake, which had an order of magnitude larger \Rey=5×104\Rey=5\times 10^{4}. The decay rate transitions to x0.45x^{-0.45} around x/D=45x/D=45, suggesting the end of the NEQ regime and beginning of the regime with Q2D power law. The Q2D regime was not found by Chongsiripinyo & Sarkar (2020) at their lowest Froude number, Fr=2\mbox{{Fr}}=2.

Wakes formed within a pycnocline layer are found here to be more turbulent than in linear stratification given the same value of N(z)N(z) at the disk center. The TKE for cases 614 and I1I exceeds that for 111 even though N(z)N(z) is very similar in the wake core among the three cases. The difference is specially large in the NEQ regime where I1I has about twice the TKE of 111. The reasons are as follows. First, the most dominant production term, PxyP_{xy}, is higher for 614 and I1I. The turbulent momentum flux is sensitive to the weak stratification, even though the weakening occurs away from and not in the wake core and, therefore, the suppression of the turbulent flux by stabilizing buoyancy is weaker in 614 and I1I. Second, the internal wave energy flux out of the wake, pur\langle p^{\prime}u_{r}^{\prime}\rangle is smaller for 614 and I1I because there is a range of internal waves that is generated but cannot propagate out vertically since their frequency exceeds the smaller NN outside the pycnocline. This wave energy is trapped inside the pycnocline layer. Both of these effects dominate at the beginning of the NEQ regime, resulting in a more turbulent wake. It is worth noting that, a side lobe of wave radiation forms in 614 and I1I (center and left panels of Fig. 7a), but the overall integrated wave flux in the NEQ regime is still smaller than 111.

Based on the theoretical asymptotic analysis for steady lee waves by Voisin (2007), the potential flow solution for an oval was used by Ortiz-Tarin et al. (2019) to develop an expression for the lee wave field of a spheroid in linear stratification. We find that the disk, along with its separation bubble, can be approximated as a Rankine ovoid and have used the ovoid shape for the potential flow solution instead of an oval. There is excellent agreement between the 111 simulation and the analytical expression with regards to both wave length and wave amplitude. The 614 simulation has a lee wave field whose wavelength is well predicted by constant-NN linear theory upon using Fr=6.13\mbox{{Fr}}=6.13 above and Fr=3.74\mbox{{Fr}}=3.74 below the pycnocline to approximate the steady lee wave in those regions. However, the simulation amplitude for 614 is higher than that given by this approximation since the wave-generation region, which includes the pycnocline, has an effective NN that is larger then the constant far-field NN and, therefore, a stronger wave field.

In the second part of the study, the influence of shifting the pycnocline layer relative to the disk center is found to be strong. In 614-, half of the disk lies in the upper region of weak (Fr=6.13\mbox{{Fr}}=6.13) linear stratification and, in 614+, the disk lies entirely in the lower linearly stratified region (Fr=3.74\mbox{{Fr}}=3.74) but still very close to the nonuniform-NN region. Overall, the relative shift between pycnocline and disk weakens the buoyancy effect on the mean wake, since the wake feels a weaker effective stratification. Thus, different from the 614 centered case, the defect velocity in the shifted cases do not show an oscillation in x/D<8x/D<8 and also decays more quickly than 614 in the far wake. The wakes are also more turbulent in 614614- and 614+614+ as shown by the TKE. The asymmetric placement (with respect to the profile) of the disk in 614614- is also clearly manifested in its TKE contour where the lower half, which is inside the pycnocline layer is vertically compressed relative to the upper half.

With regards to the wave field, in addition to the usual steady far-field lee waves and unsteady wake generated waves, the pycnocline supports a third family of waves, the steady Kelvin wake waves. The Kelvin wake waves can be analytically described by solving the Taylor-Goldstein equation as an eigenvalue problem and they take a fundamentally different form than the lee waves as visualized by Fig. 11. Furthermore, shifting the disk alters the modal wave form. The dominant waveform in 614614 corresponds to the mode-2 eigenfunction while 614614- and 614+614+ exhibit a mix of two waveforms, corresponding to mode-1 and mode-2 eigenfunction. In 614, the disk moves right at the center of the pycnocline layer leading to the symmetry property of the mode-2 wave. Any vertical offset from the center leads to the appearance of mode-1 waves. These modal waveforms are well corroborated by the experiments of Robey (1997) as well as Nicolaou et al. (1995).

Phase lines of unsteady internal gravity waves in the linearly stratified far field are found to cluster around a characteristic inclination angle which takes the value of θ39\theta\approx 39^{\circ} in 111. This result of phase-angle clustering is consistent with several previous studies of wave radiation from a turbulent flow (Sutherland & Linden, 1998; Dohan & Sutherland, 2003; Taylor & Sarkar, 2007; Pham et al., 2009) but it is also worth noting that as \Rey\Rey increases, Abdilghanie & Diamessis (2013) find a broader range of wave phase angles. In case 614+614+, the wake generated waves that propagate downward do so in a linear stratification and cluster around θ44\theta\approx 44^{\circ}. The upward propagating waves have to go through the pycnocline layer and only the near-NN waves are able to propagate in the weak linear stratification above the pycnocline. Some of the wave energy that reflects off the pycnocline boundaries is trapped and leads to small-scale variability while a fraction escapes through the sides of the wake.

We have explored a limited portion of the parameter space applicable to turbulent wakes in nonlinear background stratification. Depending on the application and the environmental setting, there can be wide variability in the nondimensional pycnocline thickness (δ/D\delta/D), the shape of the nonuniform stratification, and the variability of N(z)N(z). Body characteristics such as its nominal Froude number, Reynolds number and its shape can vary. Future examination of this wide parameter space in laboratory experiments as well as simulations (body-inclusive hybrid spatial or temporal as well as the cheaper body-exclusive temporal with a good choice for initial conditions) will be useful.

\backsection

[Funding]The authors gratefully acknowledge the support of Office of Naval Research Grant N00014-20-1-2253.

\backsection

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

References

  • Abdilghanie & Diamessis (2013) Abdilghanie, Ammar M & Diamessis, Peter J 2013 The internal gravity wave field emitted by a stably stratified turbulent wake. Journal of Fluid Mechanics 720, 104–139.
  • Balaras (2004) Balaras, E. 2004 Modeling complex boundaries using an external force field on fixed Cartesian grids in large-eddy simulations. Comput. Fluids 33 (3), 375–404.
  • Barber (1993) Barber, BC 1993 On the dispersion relation for trapped internal waves. Journal of Fluid Mechanics 252, 31–49.
  • Bonneton et al. (1993) Bonneton, P, Chomaz, JM & Hopfinger, EJ 1993 Internal waves produced by the turbulent wake of a sphere moving horizontally in a stratified fluid. Journal of Fluid Mechanics 254, 23–40.
  • Brandt & Rottier (2015) Brandt, A & Rottier, JR 2015 The internal wavefield generated by a towed sphere at low froude number. Journal of Fluid Mechanics 769, 103–129.
  • Brucker & Sarkar (2010) Brucker, K. A. & Sarkar, S. 2010 A comparative study of self-propelled and towed wakes in a stratified fluid. J. Fluid Mech. 652, 373–404.
  • Chomaz et al. (1993) Chomaz, J. M., Bonneton, P. & Hopfinger, E. J. 1993 The structure of the near wake of a sphere moving horizontally in a stratified fluid. J. Fluid Mech. 254, 1–21.
  • Chongsiripinyo et al. (2017) Chongsiripinyo, Karu, Pal, Anikesh & Sarkar, Sutanu 2017 On the vortex dynamics of flow past a sphere at re= 3700 in a uniformly stratified fluid. Physics of Fluids 29 (2), 020704.
  • Chongsiripinyo & Sarkar (2020) Chongsiripinyo, K. & Sarkar, S. 2020 Decay of turbulent wakes behind a disk in homogeneous and stratified fluids. J. Fluid Mech. 885, A31.
  • Diamessis et al. (2005) Diamessis, Peter J, Domaradzki, JA & Hesthaven, Jan S 2005 A spectral multidomain penalty method model for the simulation of high reynolds number localized incompressible stratified turbulence. Journal of Computational Physics 202 (1), 298–322.
  • Dohan & Sutherland (2003) Dohan, K & Sutherland, BR 2003 Internal waves generated from a turbulent mixed region. Physics of Fluids 15 (2), 488–498.
  • Dommermuth et al. (2002) Dommermuth, Douglas G, Rottman, James W, Innis, George E & Novikov, Evgeny A 2002 Numerical simulation of the wake of a towed sphere in a weakly stratified fluid. Journal of Fluid Mechanics 473, 83–101.
  • Ermanyuk & Gavrilov (2002) Ermanyuk, Eugeny V & Gavrilov, Nikolai V 2002 Force on a body in a continuously stratified fluid. part 1. circular cylinder. Journal of Fluid Mechanics 451, 421.
  • Ermanyuk & Gavrilov (2003) Ermanyuk, Eugeny V & Gavrilov, Nikolai V 2003 Force on a body in a continuously stratified fluid. part 2. sphere. Journal of Fluid Mechanics 494, 33.
  • Germano et al. (1991) Germano, M., Piomelli, U., Moin, P. & Cabot, W. H. 1991 A dynamic subgrid‐scale eddy viscosity model. Phys. Fluids 3 (7), 1760–1765.
  • Gilreath & Brandt (1985) Gilreath, HE & Brandt, A 1985 Experiments on the generation of internal waves in a stratified fluid. AIAA journal 23 (5), 693–700.
  • Gourlay et al. (2001) Gourlay, M. J., Arendt, S. C., Fritts, D. C. & Werne, J. 2001 Numerical modeling of initially turbulent wakes with net momentum. Phys. Fluids 13 (12), 3783–3802.
  • Grisouard et al. (2011) Grisouard, Nicolas, Staquet, Chantal & Gerkema, Theo 2011 Generation of internal solitary waves in a pycnocline by an internal wave beam: a numerical study. Journal of Fluid Mechanics 676, 491–513.
  • Hopfinger et al. (1991) Hopfinger, Emil J, Flor, J-B, Chomaz, J-M & Bonneton, P 1991 Internal waves generated by a moving sphere and its wake in a stratified fluid. Experiments in Fluids 11 (4), 255–261.
  • Keller & Munk (1970) Keller, Joseph B & Munk, Walter H 1970 Internal wave wakes of a body moving in a stratified fluid. The Physics of Fluids 13 (6), 1425–1431.
  • Lin & Pao (1979) Lin, Jung-Tai & Pao, Yih-Ho 1979 Wakes in stratified fluids. Annual Review of Fluid Mechanics 11 (1), 317–338.
  • Meunier et al. (2018) Meunier, P, Le Dizès, Stéphane, Redekopp, L & Spedding, GR 2018 Internal waves generated by a stratified wake: experiment and theory. Journal of Fluid Mechanics 846, 752–788.
  • Nicolaou et al. (1995) Nicolaou, D, Garman, JFR & Stevenson, TN 1995 Internal waves from a body accelerating in a thermocline. Applied scientific research 55 (2), 171–186.
  • Nidhan et al. (2022) Nidhan, Sheel, Schmidt, Oliver T & Sarkar, Sutanu 2022 Analysis of coherence in turbulent stratified wakes using spectral proper orthogonal decomposition. Journal of Fluid Mechanics 934.
  • Orlanski (1976) Orlanski, I. 1976 A simple boundary condition for unbounded hyperbolic flows. J. Comput. Phys. 21 (3), 251–269.
  • Orr et al. (2015) Orr, Trevor S, Domaradzki, J Andrzej, Spedding, Geoffrey R & Constantinescu, George S 2015 Numerical simulations of the near wake of a sphere moving in a steady, horizontal motion through a linearly stratified fluid at re= 1000. Physics of Fluids 27 (3), 035113.
  • Ortiz-Tarin et al. (2019) Ortiz-Tarin, Jose L, Chongsiripinyo, KC & Sarkar, S 2019 Stratified flow past a prolate spheroid. Physical Review Fluids 4 (9), 094803.
  • Pal et al. (2016) Pal, Anikesh, Sarkar, Sutanu, Posa, Antonio & Balaras, Elias 2016 Regeneration of turbulent fluctuations in low-froude-number flow over a sphere at a reynolds number of 3700. Journal of Fluid Mechanics 804.
  • Pal et al. (2017) Pal, Anikesh, Sarkar, Sutanu, Posa, Antonio & Balaras, Elias 2017 Direct numerical simulation of stratified flow past a sphere at a subcritical reynolds number of 3700 and moderate froude number. Journal of Fluid Mechanics 826, 5.
  • Pham & Sarkar (2010) Pham, Hieu T & Sarkar, Sutanu 2010 Internal waves and turbulence in a stable stratified jet. Journal of fluid mechanics 648, 297–324.
  • Pham et al. (2009) Pham, Hieu T, Sarkar, Sutanu & Brucker, Kyle A 2009 Dynamics of a stratified shear layer above a region of uniform stratification. Journal of fluid mechanics 630, 191–223.
  • Redford et al. (2015) Redford, JA, Lund, TS & Coleman, GN 2015 A numerical study of a weakly stratified turbulent wake. Journal of Fluid Mechanics 776, 568.
  • Redford et al. (2012) Redford, J. A., Castro, I. P. & Coleman, G. N. 2012 On the universality of turbulent axisymmetric wakes. J. Fluid Mech. 710, 419–452.
  • Robey (1997) Robey, Harry F 1997 The generation of internal waves by a towed sphere and its wake in a thermocline. Physics of Fluids 9 (11), 3353–3367.
  • Rossi & Toivanen (1999) Rossi, T. & Toivanen, J. 1999 A Parallel Fast Direct Solver for Block Tridiagonal Systems with Separable Matrices of Arbitrary Dimension. SIAM J. Sci. Comput. 20 (5), 1778–1793.
  • Rowe et al. (2020) Rowe, KL, Diamessis, PJ & Zhou, Q 2020 Internal gravity wave radiation from a stratified turbulent wake. Journal of Fluid Mechanics 888.
  • Spedding (1997) Spedding, GR 1997 The evolution of initially turbulent bluff-body wakes at high internal froude number. Journal of Fluid Mechanics 337, 283–301.
  • de Stadler & Sarkar (2012) de Stadler, Matthew B & Sarkar, Sutanu 2012 Simulation of a propelled wake with moderate excess momentum in a stratified fluid. Journal of fluid mechanics 692, 28–52.
  • de Stadler et al. (2010) de Stadler, Matthew B, Sarkar, Sutanu & Brucker, Kyle A 2010 Effect of the prandtl number on a stratified turbulent wake. Physics of Fluids 22 (9), 095102.
  • Stevenson et al. (1986) Stevenson, TN, Kanellopulos, D & Constantinides, M 1986 The phase configuration of trapped internal waves from a body moving in a thermocline. Applied scientific research 43 (2), 91–105.
  • Sturova (1974) Sturova, IV 1974 Wave motions produced in a stratified liquid from flow past a submerged body. Journal of Applied Mechanics and Technical Physics 15 (6), 796–805.
  • Sutherland (2016) Sutherland, Bruce R 2016 Internal wave transmission through a thermohaline staircase. Physical Review Fluids 1 (1), 013701.
  • Sutherland & Linden (1998) Sutherland, Bruce R & Linden, Paul F 1998 Internal wave excitation from stratified flow over a thin barrier. Journal of Fluid Mechanics 377, 223–252.
  • Sutherland & Yewchuk (2004) Sutherland, Bruce R & Yewchuk, Kerianne 2004 Internal wave tunnelling. Journal of Fluid Mechanics 511, 125–134.
  • Taylor & Sarkar (2007) Taylor, John R & Sarkar, Sutanu 2007 Internal gravity waves generated by a turbulent bottom ekman layer. Journal of Fluid Mechanics 590, 331–354.
  • Voisin (1991) Voisin, Bruno 1991 Internal wave generation in uniformly stratified fluids. part 1. green’s function and point sources. Journal of Fluid Mechanics 231, 439–480.
  • Voisin (1994) Voisin, Bruno 1994 Internal wave generation in uniformly stratified fluids. part 2. moving point sources. Journal of Fluid Mechanics 261, 333–374.
  • Voisin (2003) Voisin, Bruno 2003 Limit states of internal wave beams. Journal of Fluid Mechanics 496, 243–293.
  • Voisin (2007) Voisin, Bruno 2007 Lee waves from a sphere in a stratified flow. Journal of Fluid Mechanics 574, 273–315.
  • Voropayeva & Chernykh (1997) Voropayeva, OF & Chernykh, GG 1997 Numerical model of the dynamics of a momentumless turbulent wake in a pycnocline. Journal of applied mechanics and technical physics 38 (3), 391–406.
  • Yang & Balaras (2006) 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), 12–40.
  • Zhou & Diamessis (2019) Zhou, Qi & Diamessis, Peter J 2019 Large-scale characteristics of stratified wake turbulence at varying reynolds number. Physical Review Fluids 4 (8), 084802.