Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Nonlinear dynamics of vortex pairing in transitional jets

Akhil Nekkanti\aff1,2    Tim Colonius\aff2       Oliver T. Schmidt\aff1 \corresp oschmidt@ucsd.edu \aff1Department of Mechanical and Aerospace Engineering, University of California San Diego, La Jolla, CA 92093, USA \aff2 Division of Engineering and Applied Science, California Institute of Technology, Pasadena, CA 91125, USA
Abstract

This study investigates the onset of linear instabilities and their later nonlinear interactions in the shear layer of an initially-laminar jet using a combination of stability analysis and data from high-fidelity flow simulations. We provide a complete picture of the vortex-pairing process. Hydrodynamic instabilities initiate the transition to turbulence, causing the shear layer to spread rapidly. In this process, the shear layer rolls up to form vortices, accompanied by the exponential growth of the fundamental frequency. As the fundamental frequency grows, it gains energy from the mean flow. Subsequently, as it saturates and begins to decay, the fundamental vortices start to pair. During this vortex pairing process, the subharmonic vortex acquires energy both linearly from the mean flow and nonlinearly through a reverse cascade from the fundamental. The process concludes when the subharmonic vortex eventually saturates. Similarly, two subharmonic vortices merge to form a second subharmonic vortex. Our results confirm Kelly (1967)’s hypothesis of a resonance mechanism between the fundamental and subharmonic, which supplies energy to the subharmonic. In this multi-tonal, convective-dominated flow, we clarify the ambiguity surrounding the fundamental frequency by demonstrating that the spatially most amplified frequency should be considered fundamental, rather than the structure associated with the spectral energy peak. For the initially-laminar jet considered here, the fundamental frequency corresponds to the fourth largest spectral peak, highlighting the important distinction between the energetically and dynamical significance of a tone. Despite its low energy, the fundamental frequency is dynamically dominant as it determines all other spectral peaks and supplies energy to the subharmonics through a reverse energy cascade.

keywords:

1 Introduction

The hydrodynamic near-field and the far-field acoustics of jets are significantly influenced by the inflow conditions and the state of the boundary layer at the nozzle’s exit. In particular, the flow field is sensitive to parameters such as momentum thickness, fluctuation level, and whether the boundary layer is laminar or turbulent. Several experimental (Hill Jr et al., 1976; Hussain & Zedan, 1978a, b; Husain & Hussain, 1979; Bridges & Hussain, 1987; Zaman, 1985, 2012; Fontaine et al., 2015) and numerical studies (Bogey & Bailly, 2005, 2010; Bogey et al., 2012; Kim & Choi, 2009; Brès et al., 2018) have explored the effect of these parameters. For a jet with a laminar boundary layer at laboratory-scale Reynolds number, the flow emerging from the nozzle mixes with the surrounding fluid and transitions to turbulence within the first few jet diameters. This rapid mixing results in roll-up and pairing of vortices, consequently leading to an increase in the radiated noise (Zaman, 1985; Bridges & Hussain, 1987; Bogey & Bailly, 2010). Zaman (1985) showed that the initially-laminar jet exhibits a 4 dB increase in the radiated noise compared to its turbulent counterpart. Bogey & Bailly (2010) demonstrated that initially-laminar jets with larger momentum thickness exhibit stronger vortex pairing that increases the sound pressure levels in the sideline direction.

Vortex pairing is a main characteristic of mixing layers (Brown & Roshko, 1974; Winant & Browand, 1974; Ho & Huang, 1982; Metcalfe et al., 1987; Moser & Rogers, 1993) and jet flows (Becker & Massaro, 1968; Zaman & Hussain, 1980; Hussain & Zaman, 1980; Meynart, 1983). It significantly contributes to turbulent mixing (Brown & Roshko, 1974), the production of Reynolds stresses (Zaman & Hussain, 1980), entrainment (Winant & Browand, 1974), and triggers the transition to turbulence (Ho & Huang, 1982; Moser & Rogers, 1993). Given its importance, vortex pairing has been the subject of numerous studies. In jets, it was first visualized by Becker & Massaro (1968). In the seminal study by (Crow & Champagne, 1971), pairing was found to be more regular at low Reynolds numbers and became increasingly chaotic at higher Reynolds numbers. Detailed experimental studies on forced jets conducted by Zaman & Hussain (1980) and Hussain & Zaman (1980) reveal that pairing occurs at two distinct frequencies: one around Stθ=fθ/U0.012subscriptSt𝜃𝑓𝜃𝑈0.012\mbox{\it St}_{\theta}=f\theta/U\approx 0.012St start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = italic_f italic_θ / italic_U ≈ 0.012, termed the shear layer mode, and the other around StD=fD/U0.85subscriptSt𝐷𝑓𝐷𝑈0.85\mbox{\it St}_{D}=fD/U\approx 0.85St start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT = italic_f italic_D / italic_U ≈ 0.85, referred to as the jet column mode. Here, f𝑓fitalic_f is the frequency, U𝑈Uitalic_U is the jet velocity, D𝐷Ditalic_D is the diameter, and θ𝜃\thetaitalic_θ is the momentum thickness. Kibens (1980) forced the jet at the shear-layer instability frequency Stθ𝑆subscript𝑡𝜃St_{\theta}italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT and found that this forcing results in three successive vortex pairings, producing the subharmonic frequencies Stθ/2𝑆subscript𝑡𝜃2St_{\theta}/2italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT / 2, Stθ/4𝑆subscript𝑡𝜃4St_{\theta}/4italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT / 4, and Stθ/8𝑆subscript𝑡𝜃8St_{\theta}/8italic_S italic_t start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT / 8. Similarly, in mixing layers, Ho & Huang (1982) observed multiple pairing when the flow was forced at the subharmonic frequency.

The nonlinear mechanisms behind the growth of the subharmonic during the vortex pairing process have been the focus of many studies (Kelly, 1967; Monkewitz, 1988; Paschereit et al., 1995; Husain & Hussain, 1995). Kelly (1967) proposed a resonance mechanism that provides energy to the subharmonics. Monkewitz (1988) used weakly nonlinear spatial theory to further support Kelly (1967)’s resonance mechanism. They demonstrated that the fundamental mode needs to reach a critical amplitude before both the fundamental and subharmonic can phase-lock, resulting in a energy transfer to the subharmonic. Experiments by Hajj et al. (1992) show that the subharmonic gains energy through this resonance mechanism. Works by Husain & Hussain (1995) and Cho et al. (1998) demonstrate that vortex pairing can be either enhanced or attenuated by controlling the phase difference between the fundamental and the subharmonic. Mankbadi (1985) employed an energy-integral method to show that the subharmonic gains energy both from the fundamental and the mean flow. Paschereit et al. (1995) estimated the energy transfer to the subharmonic wave based on the production terms. Their findings reveal that the subharmonic primarily gains its energy from the mean flow, with the fundamental wave acting as a catalyst.

Using classical linear theory and spectral modal analysis, this work answers the following questions: How does the growth of the shear layer differ between an initially-laminar and turbulent jet? What is the influence of vortex pairing on the development of the shear layer? What should be considered as the fundamental frequency in a multi-tonal flow? In particular, our study sheds light on the ambiguity of energetic vs dynamical significance in this context. We further demonstrate how the energy transfer during vortex pairing can be quantified.

Linear stability theory (LST) has been used in the past with great success, first by Michalke (1964, 1965) on a hyperbolic tangent profile. Here, we apply LST to the temporally averaged mean flow of the jet. This approach corresponds to applying the parallel flow assumption locally to the zero-frequency component, which has been utilized in cylinder wakes (Pier, 2002; Barkley, 2006) and jets (Suzuki & Colonius, 2006; Gudmundsson & Colonius, 2011; Schmidt et al., 2017). The mathematical framework of spectral property orthogonal decomposition (SPOD) dates back to the work of Lumley (1967, 1970). SPOD identifies the most energetic coherent structures at each time scale. Early applications of SPOD include the work of Glauser et al. (1987); Glauser & George (1992); Delville (1994). More recently, this method has attracted significant interest following its application to large flow databases by Schmidt et al. (2018) and the establishment of its relationship to other methods by Towne et al. (2018). Since then, SPOD has become a mainstay of physical exploration, in particular for identifying the different modal and non-modal instabilities (Schmidt et al., 2018; Nogueira et al., 2019; Pickering et al., 2020).

Vortex-pairing is a nonlinear process involving energy transfer between different frequencies or scales. This interscale energy transfer arises from the quadratic nonlinearity of the Navier-Stokes equations, leading to triadic interactions. Various approaches have been employed to analyze the energy transfer between different scales, including the Karman-Howarth equation (Von Karman & Howarth, 1938; Danaila et al., 1999; Hill, 2001) and bispectrum analysis (Lii et al., 1976; Kim & Powers, 1979; Herring, 1980). The former is based on structure functions, and the latter on third-order statistics. is the frequency domain representation of third-order moments. Recently, a modal decomposition based on bispectral analysis was developed by Schmidt (2020). Consistent with our data analysis approach, we quantify the production, dissipation, and nonlinear transfer between the leading SPOD modes associated with different harmonics and the mean flow, based on the spectral turbulent kinetic energy (TKE) equation. Previous studies (Mizuno, 2016; Cho et al., 2018; Lee & Moser, 2019; Gomé et al., 2023) have investigated interscale energy transfer in turbulent channel flows using the spectral TKE equation. Additionally, researchers have employed various bases, such as resolvent modes (Symon et al., 2021; Jin et al., 2021), Fourier modes (Nekkanti et al., 2023), dynamic mode decomposition (DMD) modes (Kinjangi & Foti, 2023), and optimal mode decomposition (OMD) modes (Biswas et al., 2022) to estimate the spectral energy budget.

The paper is organized as follows: In §2, the methodologies of LST, SPOD, and SPOD-based spectral analysis budget are discussed. Results focusing on shear layer instability, vortex pairing, and spectral energy transfer are presented in §3. The paper concludes with discussions and conclusions in §4.

2 Methodology

2.1 Linear stability theory

We employ the spatial form of the linear stability theory. This LST determines the spatial growth rates of the corresponding frequency. Here, we use it to identify the most unstable frequency. LST is also referred to as local linear theory, and throughout this paper, we will use the terms linear stability theory and local linear theory interchangeably. In LST, the frequency ω𝜔\omegaitalic_ω is assumed to be real, and the eigenvalue problem is solved for a complex α𝛼\alphaitalic_α. The real part of α𝛼\alphaitalic_α is the streamwise wavenumber, and the imaginary part is its amplification rate. We start off with the Reynolds decompostion, 𝐪(x,r,θ,t)=𝐪¯(x,r)+𝐪(x,r,θ,t)𝐪𝑥𝑟𝜃𝑡¯𝐪𝑥𝑟superscript𝐪𝑥𝑟𝜃𝑡\mathbf{q}\pqty{x,r,\theta,t}=\bar{\mathbf{q}}\pqty{x,r}+\mathbf{q}^{\prime}% \pqty{x,r,\theta,t}bold_q ( start_ARG italic_x , italic_r , italic_θ , italic_t end_ARG ) = over¯ start_ARG bold_q end_ARG ( start_ARG italic_x , italic_r end_ARG ) + bold_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( start_ARG italic_x , italic_r , italic_θ , italic_t end_ARG ), where 𝐪¯¯𝐪\bar{\mathbf{q}}over¯ start_ARG bold_q end_ARG is the mean flow and 𝐪superscript𝐪\mathbf{q}^{\prime}bold_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is the fluctuation. The basic assumption of the local linear theory that the flow is locally streamwise parallel, i.e., the flow is homogeneous in the streamwise direction. Using the normal mode ansatz, the fluctuation is expressed as

𝐪(x,r,θ,t)=𝐪~(r)ei(αx+mθωt),superscript𝐪𝑥𝑟𝜃𝑡~𝐪𝑟superscript𝑒𝑖𝛼𝑥𝑚𝜃𝜔𝑡\mathbf{q}^{\prime}(x,r,\theta,t)=\tilde{\mathbf{q}}(r)e^{i\pqty{\alpha x+m% \theta-\omega t}},bold_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x , italic_r , italic_θ , italic_t ) = over~ start_ARG bold_q end_ARG ( italic_r ) italic_e start_POSTSUPERSCRIPT italic_i ( start_ARG italic_α italic_x + italic_m italic_θ - italic_ω italic_t end_ARG ) end_POSTSUPERSCRIPT , (1)

where the streamwise wavenumber α𝛼\alphaitalic_α is complex, and 𝐪~~𝐪\tilde{\mathbf{q}}over~ start_ARG bold_q end_ARG is the radial profile. Linearizing the Navier-Stokes equation about the base flow yields the equation,

(iω𝐈+𝐋)𝐪~=0,𝑖𝜔𝐈𝐋~𝐪0\pqty{i\omega\mathbf{I}+\mathbf{L}}\tilde{\mathbf{q}}=0,( start_ARG italic_i italic_ω bold_I + bold_L end_ARG ) over~ start_ARG bold_q end_ARG = 0 , (2)

where 𝐈𝐈\mathbf{I}bold_I is the identity matrix and 𝐋𝐋\mathbf{L}bold_L is the linearized compressible Navier–Stokes operator. The domain is extended to the far field by mapping the original domain, r[0,6]𝑟06r\in[0,6]italic_r ∈ [ 0 , 6 ] to r[0,)𝑟0r\in[0,\infty)italic_r ∈ [ 0 , ∞ ) by the mapping function suggested by Lesshafft & Huerre (2007). Dirichlet boundary conditions are used for a far-field boundary and the radial direction is discretised using Chebyshev collocation points. Finally, the eigenvalue problem is solved using the using the methodology of Maia et al. (2021, 2022). Solving this eigenvalue problems yields monochromatic amplification rates, αisubscript𝛼𝑖\alpha_{i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT; For αi<0subscript𝛼𝑖0\alpha_{i}<0italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT < 0, the disturbances will grow exponentially downstream, whereas for αi>0subscript𝛼𝑖0\alpha_{i}>0italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > 0, they will decay downstream.

2.2 Spectral proper orthogonal decomposition

SPOD decomposes stationary flow data into monochromatic modes that represent the spatial flow structures optimized in terms of the flow’s energy. The eigenvalues corresponding to these modes represent their energy. We employ a SPOD algorithm based on Welch’s method (Welch, 1967). For the mathematical derivation and computational details, refer to Towne et al. (2018) and Schmidt & Colonius (2020). The SPOD modes and eigenvalues are obtained by solving the eigenvalue problem

Ω𝑺(𝐱,𝐱,f)𝐖(𝐱)𝚽(𝐱,f)d𝐱=𝚽(𝐱,f)𝝀(f),subscriptΩ𝑺𝐱superscript𝐱𝑓𝐖superscript𝐱𝚽superscript𝐱𝑓superscript𝐱𝚽𝐱𝑓𝝀𝑓\int_{\Omega}\bm{S}\pqty{\mathbf{x},\mathbf{x}^{\prime},f}\mathbf{W}\pqty{% \mathbf{x}^{\prime}}\bm{\Phi}\pqty{\mathbf{x}^{\prime},f}\differential\mathbf{% x}^{\prime}=\bm{\Phi}\pqty{\mathbf{x},f}\bm{\lambda}(f),∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT bold_italic_S ( start_ARG bold_x , bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_f end_ARG ) bold_W ( start_ARG bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG ) bold_Φ ( start_ARG bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_f end_ARG ) start_DIFFOP roman_d end_DIFFOP bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = bold_Φ ( start_ARG bold_x , italic_f end_ARG ) bold_italic_λ ( italic_f ) , (3)

where 𝑺𝑺\bm{S}bold_italic_S is the cross-spectral density matrix, 𝐖𝐖\mathbf{W}bold_W the positive definite matrix that accounts for component-wise and numerical quadrature weights, 𝚽𝚽\bm{\Phi}bold_Φ the SPOD modes and 𝝀𝝀\bm{\lambda}bold_italic_λ the eigenvalues. The modes ϕ(i)(𝐱,f)superscriptbold-italic-ϕ𝑖𝐱𝑓\bm{\phi}^{(i)}(\mathbf{x},f)bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) and eigenvalues λ(i)(f)superscript𝜆𝑖𝑓\lambda^{(i)}(f)italic_λ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_f ) are sorted by energy, where i𝑖iitalic_i is the mode number index. At each frequency, the SPOD modes are orthogonal in space,

Ωϕ(i)(𝐱,f)𝐖(𝐱)ϕ(j)(𝐱,f)d𝐱=δij,subscriptΩsuperscriptbold-italic-ϕ𝑖𝐱𝑓𝐖𝐱superscriptbold-italic-ϕ𝑗𝐱𝑓𝐱subscript𝛿𝑖𝑗\int_{\Omega}\bm{\phi}^{(i)}(\mathbf{x},f)\mathbf{W}\pqty{\mathbf{x}}\bm{\phi}% ^{(j)}(\mathbf{x},f)\differential\mathbf{x}=\delta_{ij},∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) bold_W ( start_ARG bold_x end_ARG ) bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) start_DIFFOP roman_d end_DIFFOP bold_x = italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT , (4)

where δijsubscript𝛿𝑖𝑗\delta_{ij}italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the Kronecker delta function. The Fourier modes of each flow realization are expanded as

𝐪^(x,f)=ia(i)(f)ϕ(i)(𝐱,f).𝐪^𝑥𝑓subscript𝑖superscript𝑎𝑖𝑓superscriptbold-italic-ϕ𝑖𝐱𝑓\mathbf{\hat{q}}(x,f)=\sum\limits_{i}a^{(i)}(f)\bm{\phi}^{(i)}(\mathbf{x},f).start_ID over^ start_ARG bold_q end_ARG end_ID ( italic_x , italic_f ) = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_f ) bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) . (5)

Recently, Nekkanti & Schmidt (2021) proposed a convolution approach that computes time-continuous expansion coefficients

a(i)(f,t)=ΔTΩ(ϕ(i)(𝐱,f))𝐖(𝐱)q(𝐱,t+τ)ei2πfτdxdτ,superscript𝑎𝑖𝑓𝑡subscriptΔ𝑇subscriptΩsuperscriptsuperscriptbold-italic-ϕ𝑖𝐱𝑓𝐖𝐱q𝐱𝑡𝜏superscript𝑒𝑖2𝜋𝑓𝜏𝑥𝜏a^{(i)}(f,t)=\int_{\Delta T}\int_{\Omega}\quantity({\bf\it\phi}^{(i)}(\mathbf{% x},f))^{*}\mathbf{W}(\mathbf{x})\textbf{q}(\mathbf{x},t+\tau)e^{-i2\pi f\tau}% \differential x\differential\tau,italic_a start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_f , italic_t ) = ∫ start_POSTSUBSCRIPT roman_Δ italic_T end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT ( start_ARG bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) end_ARG ) start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT bold_W ( bold_x ) q ( bold_x , italic_t + italic_τ ) italic_e start_POSTSUPERSCRIPT - italic_i 2 italic_π italic_f italic_τ end_POSTSUPERSCRIPT start_DIFFOP roman_d end_DIFFOP italic_x start_DIFFOP roman_d end_DIFFOP italic_τ , (6)

which facilitates obtaining the time-continuous Fourier modes

𝐪^(𝐱,f,t)ia(i)(f,t)ϕ(i)(𝐱,f).𝐪^𝐱𝑓𝑡subscript𝑖superscript𝑎𝑖𝑓𝑡superscriptbold-italic-ϕ𝑖𝐱𝑓\mathbf{\hat{q}}(\mathbf{x},f,t)\approx\sum\limits_{i}a^{(i)}(f,t){\bf\it\phi}% ^{(i)}(\mathbf{x},f).start_ID over^ start_ARG bold_q end_ARG end_ID ( bold_x , italic_f , italic_t ) ≈ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_f , italic_t ) bold_italic_ϕ start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( bold_x , italic_f ) . (7)

2.3 SPOD-based spectral energy budget

The optimality of the SPOD expansion, equation (7), can be leveraged to quantify the nonlinear interactions between the most salient flow features and their global energy budget. We specifically focus on production and nonlinear energy transfer. The starting point is the spectral TKE equation,

k^t=[u¯ik^xiu^juiujxi^¯𝒯nlu^ju^iu¯jxi¯𝒫2Res^ijs^ij¯𝒟xj(u^jp^¯)+2Rexi(u^js^ij¯)],^𝑘𝑡subscript¯𝑢𝑖^𝑘subscript𝑥𝑖subscript¯superscriptsubscript^𝑢𝑗^subscript𝑢𝑖subscript𝑢𝑗subscript𝑥𝑖subscript𝒯𝑛𝑙subscript¯superscriptsubscript^𝑢𝑗subscript^𝑢𝑖subscript¯𝑢𝑗subscript𝑥𝑖𝒫subscript2𝑅𝑒¯superscriptsubscript^𝑠𝑖𝑗subscript^𝑠𝑖𝑗𝒟subscript𝑥𝑗¯superscriptsubscript^𝑢𝑗^𝑝2𝑅𝑒subscript𝑥𝑖¯superscriptsubscript^𝑢𝑗subscript^𝑠𝑖𝑗\frac{\partial\hat{k}}{\partial t}=\mathcal{R}\bqty{-\bar{u}_{i}\frac{\partial% \hat{k}}{\partial x_{i}}\underbrace{\vphantom{\displaystyle\sum_{i=1}^{N}}-% \overline{\hat{u}_{j}^{*}\widehat{u_{i}\frac{\partial u_{j}}{\partial x_{i}}}}% }_{\mathcal{T}_{nl}}\underbrace{\vphantom{\displaystyle\sum_{i=1}^{N}}-% \overline{\hat{u}_{j}^{*}\hat{u}_{i}\frac{\partial\bar{u}_{j}}{\partial x_{i}}% }}_{\mathcal{P}}\underbrace{\vphantom{\displaystyle\sum_{i=1}^{N}}-\frac{2}{Re% }\overline{\hat{s}_{ij}^{*}\hat{s}_{ij}}}_{\mathcal{D}}-\frac{\partial}{% \partial x_{j}}\pqty{\overline{\hat{u}_{j}^{*}\hat{p}}}+\frac{2}{Re}\frac{% \partial}{\partial x_{i}}\pqty{\overline{\hat{u}_{j}^{*}\hat{s}_{ij}}}},divide start_ARG ∂ over^ start_ARG italic_k end_ARG end_ARG start_ARG ∂ italic_t end_ARG = caligraphic_R [ start_ARG - over¯ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ over^ start_ARG italic_k end_ARG end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG under⏟ start_ARG - over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG end_ARG end_ARG start_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_n italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT under⏟ start_ARG - over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ over¯ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG end_ARG start_POSTSUBSCRIPT caligraphic_P end_POSTSUBSCRIPT under⏟ start_ARG - divide start_ARG 2 end_ARG start_ARG italic_R italic_e end_ARG over¯ start_ARG over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG end_ARG start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT - divide start_ARG ∂ end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ( start_ARG over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_p end_ARG end_ARG end_ARG ) + divide start_ARG 2 end_ARG start_ARG italic_R italic_e end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG end_ARG ) end_ARG ] , (8)

where, s^ij=1/2(u^i/xj+u^j/xi)subscript^𝑠𝑖𝑗12subscript^𝑢𝑖subscript𝑥𝑗subscript^𝑢𝑗subscript𝑥𝑖\hat{s}_{ij}=1/2\pqty{\partial\hat{u}_{i}/\partial x_{j}+\partial\hat{u}_{j}/% \partial x_{i}}over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 1 / 2 ( start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ) is the spectral strain rate, \mathcal{R}caligraphic_R denotes the real part, and k^(f,m)=u^i(f,m)u^i(f,m)¯/2^𝑘𝑓𝑚¯superscriptsubscript^𝑢𝑖𝑓𝑚subscript^𝑢𝑖𝑓𝑚2\hat{k}(f,m)=\overline{\hat{u}_{i}^{*}(f,m)\hat{u}_{i}(f,m)}/2over^ start_ARG italic_k end_ARG ( italic_f , italic_m ) = over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_f , italic_m ) over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_f , italic_m ) end_ARG / 2. For brevity, we suppress the spatial dependence of all quantities. For statistically stationary flows, the left-hand side goes to zero. In this work, we focus on the scale-specific production term

𝒫(f,m)=[u^j(f,m)u^i(f,m)u¯jxi]¯,𝒫𝑓𝑚¯superscriptsubscript^𝑢𝑗𝑓𝑚subscript^𝑢𝑖𝑓𝑚subscript¯𝑢𝑗subscript𝑥𝑖\mathcal{P}\pqty{f,m}=-\mathcal{R}\overline{\bqty{\hat{u}_{j}^{*}\pqty{f,m}% \hat{u}_{i}\pqty{f,m}\frac{\partial\bar{u}_{j}}{\partial x_{i}}}},caligraphic_P ( start_ARG italic_f , italic_m end_ARG ) = - caligraphic_R over¯ start_ARG [ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f , italic_m end_ARG ) over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f , italic_m end_ARG ) divide start_ARG ∂ over¯ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG ] end_ARG , (9)

the scale-specific dissipation term

𝒟(f,m)=2Re[(u^ixj(f,m)+u^jxi(f,m))(u^ixj(f,m)+u^jxi(f,m))]¯,𝒟𝑓𝑚2𝑅𝑒¯superscriptsubscript^𝑢𝑖subscript𝑥𝑗𝑓𝑚superscriptsubscript^𝑢𝑗subscript𝑥𝑖𝑓𝑚subscript^𝑢𝑖subscript𝑥𝑗𝑓𝑚subscript^𝑢𝑗subscript𝑥𝑖𝑓𝑚\mathcal{D}\pqty{f,m}=-\frac{2}{Re}\mathcal{R}\overline{\bqty{\pqty{\frac{% \partial\hat{u}_{i}^{*}}{\partial x_{j}}\pqty{f,m}+\frac{\partial\hat{u}_{j}^{% *}}{\partial x_{i}}\pqty{f,m}}\pqty{\frac{\partial\hat{u}_{i}}{\partial x_{j}}% \pqty{f,m}+\frac{\partial\hat{u}_{j}}{\partial x_{i}}\pqty{f,m}}}},caligraphic_D ( start_ARG italic_f , italic_m end_ARG ) = - divide start_ARG 2 end_ARG start_ARG italic_R italic_e end_ARG caligraphic_R over¯ start_ARG [ start_ARG ( start_ARG divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f , italic_m end_ARG ) + divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f , italic_m end_ARG ) end_ARG ) ( start_ARG divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f , italic_m end_ARG ) + divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f , italic_m end_ARG ) end_ARG ) end_ARG ] end_ARG , (10)

and the scale-specific nonlinear transfer term

𝒯nl(f,m)=[u^j(f,m)uiujxi^(f,m)]¯.subscript𝒯nl𝑓𝑚¯superscriptsubscript^𝑢𝑗𝑓𝑚^subscript𝑢𝑖subscript𝑢𝑗subscript𝑥𝑖𝑓𝑚\mathcal{T}_{\text{nl}}\pqty{f,m}=-\mathcal{R}\overline{\bqty{\hat{u}_{j}^{*}% \pqty{f,m}\widehat{u_{i}\frac{\partial u_{j}}{\partial x_{i}}}\pqty{f,m}}}.caligraphic_T start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT ( start_ARG italic_f , italic_m end_ARG ) = - caligraphic_R over¯ start_ARG [ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f , italic_m end_ARG ) over^ start_ARG italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG ( start_ARG italic_f , italic_m end_ARG ) end_ARG ] end_ARG . (11)

The term uiujxi^^subscript𝑢𝑖subscript𝑢𝑗subscript𝑥𝑖\widehat{u_{i}\frac{\partial u_{j}}{\partial x_{i}}}over^ start_ARG italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG in (11) entails the contributions from all other frequencies and azimuthal wavenumbers to frequency f𝑓fitalic_f and azimuthal wavenumber m𝑚mitalic_m. Following Cho et al. (2018), we isolate the energy transfer of individual triads, i.e., frequency and wavenumber triplets related by the resonance conditions, m1±m2±m3=0plus-or-minussubscript𝑚1subscript𝑚2subscript𝑚30m_{1}\pm m_{2}\pm m_{3}=0italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ± italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ± italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0 and f1±f2±f3=0plus-or-minussubscript𝑓1subscript𝑓2subscript𝑓30f_{1}\pm f_{2}\pm f_{3}=0italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ± italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ± italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0 by splitting uiujxi^^subscript𝑢𝑖subscript𝑢𝑗subscript𝑥𝑖\widehat{u_{i}\frac{\partial u_{j}}{\partial x_{i}}}over^ start_ARG italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG ∂ italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG end_ARG using the discrete convolution to obtain

𝒯nl(f3,m3)=[u^j(f3,m3)f1+f2=f3m1+m2=m3u^i(f1,m1)u^jxi(f2,m2)¯].subscript𝒯nlsubscript𝑓3subscript𝑚3¯superscriptsubscript^𝑢𝑗subscript𝑓3subscript𝑚3subscriptsubscript𝑓1subscript𝑓2subscript𝑓3subscript𝑚1subscript𝑚2subscript𝑚3subscript^𝑢𝑖subscript𝑓1subscript𝑚1subscript^𝑢𝑗subscript𝑥𝑖subscript𝑓2subscript𝑚2\mathcal{T}_{\text{nl}}\pqty{f_{3},m_{3}}=-\mathcal{R}\bqty{\overline{\hat{u}_% {j}^{*}\pqty{f_{3},m_{3}}\sum\limits_{\begin{subarray}{c}f_{1}+f_{2}=f_{3}\\ m_{1}+m_{2}=m_{3}\end{subarray}}\hat{u}_{i}\pqty{f_{1},m_{1}}\frac{\partial% \hat{u}_{j}}{\partial x_{i}}\pqty{f_{2},m_{2}}}}.caligraphic_T start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) = - caligraphic_R [ start_ARG over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG end_ARG ] . (12)

To obtain a bispectral representation of the energy transfer, we further split the sum in (12) into individual frequency and wavenumber components that are triadically compatible

𝓉nl(f1,f2,m1,m2)=[u^j(f1+f2,m1+m2)u^i(f1,m1)u^jxi(f2,m2)¯],subscript𝓉nlsubscript𝑓1subscript𝑓2subscript𝑚1subscript𝑚2¯superscriptsubscript^𝑢𝑗subscript𝑓1subscript𝑓2subscript𝑚1subscript𝑚2subscript^𝑢𝑖subscript𝑓1subscript𝑚1subscript^𝑢𝑗subscript𝑥𝑖subscript𝑓2subscript𝑚2\mathscr{t}_{\text{nl}}\pqty{f_{1},f_{2},m_{1},m_{2}}=-\mathcal{R}\bqty{% \overline{\hat{u}_{j}^{*}\pqty{f_{1}+f_{2},m_{1}+m_{2}}\hat{u}_{i}\pqty{f_{1},% m_{1}}\frac{\partial\hat{u}_{j}}{\partial x_{i}}\pqty{f_{2},m_{2}}}},script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) = - caligraphic_R [ start_ARG over¯ start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG end_ARG ] , (13)

such that 𝒯nl(f3,m3)=f1+f2=f3m1+m2=m3𝓉nl(f1,f2,m1,m2)subscript𝒯nlsubscript𝑓3subscript𝑚3subscriptsubscript𝑓1subscript𝑓2subscript𝑓3subscript𝑚1subscript𝑚2subscript𝑚3subscript𝓉nlsubscript𝑓1subscript𝑓2subscript𝑚1subscript𝑚2\mathcal{T}_{\text{nl}}(f_{3},m_{3})=\sum_{\begin{subarray}{c}f_{1}+f_{2}=f_{3% }\\ m_{1}+m_{2}=m_{3}\end{subarray}}\mathscr{t}_{\text{nl}}\pqty{f_{1},f_{2},m_{1}% ,m_{2}}caligraphic_T start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ).

At each frequency, the dominant flow structure is represented by the leading SPOD mode. To characterize the production and nonlinear kinetic energy transfer of these structures, we use a rank-1 approximation of the velocity field, following equation (7). As the focus is on self-interactions of the axisymmetric component, we further suppress the azimuthal wavenumber and SPOD mode number dependence with the understanding that m1=m2=m3=0subscript𝑚1subscript𝑚2subscript𝑚30m_{1}=m_{2}=m_{3}=0italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_m start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0 and i=1𝑖1i=1italic_i = 1. We have confirmed that higher azimuthal wavenumber components with m>0𝑚0m>0italic_m > 0 do not play a significant role in the dynamics of interest. Substituting equation (7) into equations (9), (10), (12), and (13) we get

𝒫rank-1(f1)=a(f1)a(f1)¯ϕj(f1)ϕi(f1)uj¯xi=λ(f1)ϕj(f1)ϕi(f1)u¯jxi,superscript𝒫rank-1subscript𝑓1¯superscript𝑎subscript𝑓1𝑎subscript𝑓1superscriptsubscriptitalic-ϕ𝑗subscript𝑓1subscriptitalic-ϕ𝑖subscript𝑓1¯subscript𝑢𝑗subscript𝑥𝑖𝜆subscript𝑓1superscriptsubscriptitalic-ϕ𝑗subscript𝑓1subscriptitalic-ϕ𝑖subscript𝑓1subscript¯𝑢𝑗subscript𝑥𝑖\mathcal{P}^{\text{rank-1}}\pqty{f_{1}}=-\overline{a^{*}\pqty{f_{1}}{a}\pqty{f% _{1}}}\phi_{j}^{*}\pqty{f_{1}}\phi_{i}\pqty{f_{1}}\frac{\partial\bar{u_{j}}}{% \partial x_{i}}=-\lambda(f_{1})\phi_{j}^{*}\pqty{f_{1}}\phi_{i}\pqty{f_{1}}% \frac{\partial\bar{u}_{j}}{\partial x_{i}},caligraphic_P start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) = - over¯ start_ARG italic_a start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_a ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) end_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ over¯ start_ARG italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = - italic_λ ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ over¯ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG , (14)
𝒟rank-1(f1)=λ(f1)(ϕixj(f1)+ϕjxi(f1))(ϕixj(f1)+ϕjxi(f1)),superscript𝒟rank-1subscript𝑓1𝜆subscript𝑓1superscriptsubscriptitalic-ϕ𝑖subscript𝑥𝑗subscript𝑓1superscriptsubscriptitalic-ϕ𝑗subscript𝑥𝑖subscript𝑓1subscriptitalic-ϕ𝑖subscript𝑥𝑗subscript𝑓1subscriptitalic-ϕ𝑗subscript𝑥𝑖subscript𝑓1\mathcal{D}^{\text{rank-1}}\pqty{f_{1}}=-\lambda(f_{1})\pqty{\frac{\partial% \phi_{i}^{*}}{\partial x_{j}}\pqty{f_{1}}+\frac{\partial\phi_{j}^{*}}{\partial x% _{i}}\pqty{f_{1}}}\pqty{\frac{\partial\phi_{i}}{\partial x_{j}}\pqty{f_{1}}+% \frac{\partial\phi_{j}}{\partial x_{i}}\pqty{f_{1}}},caligraphic_D start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) = - italic_λ ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ( start_ARG divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) + divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) end_ARG ) ( start_ARG divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) + divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) end_ARG ) , (15)
𝓉nlrank-1(f1,f2)superscriptsubscript𝓉nlrank-1subscript𝑓1subscript𝑓2\displaystyle\mathscr{t}_{\text{nl}}^{\text{rank-1}}\pqty{f_{1},f_{2}}script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) =(a(f3)a(f1)a(f2)¯ϕj(f3)ϕi(f1)ϕjxi(f2)),absent¯superscript𝑎subscript𝑓3𝑎subscript𝑓1𝑎subscript𝑓2superscriptsubscriptitalic-ϕ𝑗subscript𝑓3subscriptitalic-ϕ𝑖subscript𝑓1subscriptitalic-ϕ𝑗subscript𝑥𝑖subscript𝑓2\displaystyle=-\mathcal{R}\pqty{\overline{a^{*}\pqty{f_{3}}a\pqty{f_{1}}a\pqty% {f_{2}}}\phi_{j}^{*}\pqty{f_{3}}\phi_{i}\pqty{f_{1}}\frac{\partial\phi_{j}}{% \partial x_{i}}\pqty{f_{2}}},= - caligraphic_R ( start_ARG over¯ start_ARG italic_a start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) italic_a ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_a ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG ) , (16)
𝒯nlrank-1(f3)superscriptsubscript𝒯nlrank-1subscript𝑓3\displaystyle\mathcal{T}_{\text{nl}}^{\text{rank-1}}(f_{3})caligraphic_T start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) =(f1+f2=f3a(f3)a(f1)a(f2)¯ϕj(f3)ϕi(f1)ϕjxi(f2)).absentsubscriptsubscript𝑓1subscript𝑓2subscript𝑓3¯superscript𝑎subscript𝑓3𝑎subscript𝑓1𝑎subscript𝑓2superscriptsubscriptitalic-ϕ𝑗subscript𝑓3subscriptitalic-ϕ𝑖subscript𝑓1subscriptitalic-ϕ𝑗subscript𝑥𝑖subscript𝑓2\displaystyle=-\mathcal{R}\pqty{\sum\limits_{f_{1}+f_{2}=f_{3}}\overline{a^{*}% \pqty{f_{3}}a\pqty{f_{1}}a\pqty{f_{2}}}\phi_{j}^{*}\pqty{f_{3}}\phi_{i}\pqty{f% _{1}}\frac{\partial\phi_{j}}{\partial x_{i}}\pqty{f_{2}}}.= - caligraphic_R ( start_ARG ∑ start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_POSTSUBSCRIPT over¯ start_ARG italic_a start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) italic_a ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_a ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG ) italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( start_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) divide start_ARG ∂ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( start_ARG italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) end_ARG ) . (17)

The term 𝒫rank-1(f1)superscript𝒫rank-1subscript𝑓1\mathcal{P}^{\text{rank-1}}(f_{1})caligraphic_P start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) signifies energy transfer from frequency f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT to the mean flow. 𝒟rank-1(f1)superscript𝒟rank-1subscript𝑓1\mathcal{D}^{\text{rank-1}}(f_{1})caligraphic_D start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) represents energy dissipation at f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, 𝓉nlrank-1(f1,f2)superscriptsubscript𝓉nlrank-1subscript𝑓1subscript𝑓2\mathscr{t}_{\text{nl}}^{\text{rank-1}}(f_{1},f_{2})script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) denotes nonlinear interactions between f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and 𝒯nlrank-1(f3)superscriptsubscript𝒯nlrank-1subscript𝑓3\mathcal{T}_{\text{nl}}^{\text{rank-1}}(f_{3})caligraphic_T start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) indicates net nonlinear energy transfer into f3subscript𝑓3f_{3}italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. In §3.4, we use these equations to shed light on the energy transfer during the vortex pairing process.

3 Results

This study investigates the nonlinear dynamics of the initially-laminar shear layer of jets at laboratory-scale Reynolds numbers. To validate against experiments and contrast to jets with different shear layers, two large-eddy simulations of subsonic jets are conducted at a Reynolds number of \Rey=UjD/ν=50000\Reysubscript𝑈𝑗𝐷𝜈50000\Rey=U_{j}D/\nu=50000= italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_D / italic_ν = 50000. Here, Ujsubscript𝑈𝑗U_{j}italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is jet exit velocity, D𝐷Ditalic_D is the diameter, and ν𝜈\nuitalic_ν is the kinematic viscosity. In both simulations, the jet plume is turbulent. The main difference between the two cases is the state of the boundary layer inside the nozzle and consequently, the initial free shear layer over the first few jet diameters downstream of the nozzle. In the first case, the boundary layer inside the nozzle is laminar but quickly transitions within the first jet diameter. In the second case, the boundary layer is tripped inside the nozzle and, hence, turbulent from the start. We refer to the first case as the initially-laminar jet and the second case as the turbulent jet. Independent of the boundary layer tripping, both jets exhibit a potential core length that extends over several jet diameters.

The large-eddy simulations are performed using the compressible flow solver “Charles” developed at Cadence, formerly Cascade Technologies (Brès et al., 2017, 2018) and the reader is referred to Brès et al. (2017, 2018) for further details on the numerical method and validation on jet flows. To ensure the accuracy of our simulations, we first validate them by comparing them against companion experiments conducted by Maia et al. (2022). This experimental nozzle geometry is meshed using the same strategy as by Brès et al. (2018), resulting in a total grid size of 16.6 million control volumes. The LES are performed at the experimental Reynolds number. The Mach number is artificially increased to Mj=0.4subscript𝑀𝑗0.4M_{j}=0.4italic_M start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0.4 to avoid the very small explicit time steps associated with the incompressible limit. It has been confirmed by comparison with the experimental data that the effects of compressibility at this relatively small Mach number are negligible for the purpose of this study (see figure 2).

3.1 Turbulent jets

Refer to caption
Figure 1: Comparison of the initially-laminar and turbulent jets: instantaneous fluctuating streamwise velocity field of (a𝑎aitalic_a) initially-laminar jet, (b𝑏bitalic_b) turbulent jet; RMS of streamwise velocity at (c𝑐citalic_c) x=0𝑥0x=0italic_x = 0; (d𝑑ditalic_d) x=1𝑥1x=1italic_x = 1; (e𝑒eitalic_e) x=2𝑥2x=2italic_x = 2; (f𝑓fitalic_f) x=5𝑥5x=5italic_x = 5; (g𝑔gitalic_g) x=15𝑥15x=15italic_x = 15. The potential core and the jet width are indicated as lines of constant uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT at 95% and 10% of the jet velocity Ujsubscript𝑈𝑗U_{j}italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, respectively.

We compare the initially laminar and turbulent jets using instantaneous visualizations and RMS of the fluctuating streamwise velocity in figure 1. The 95% (white solid line) and 5% (white dashed line) contour lines of the mean streamwise velocity outline the potential core and the jet plume, respectively. We refer to the initial shear layer as the free shear layer between the potential core and the ambient free stream, extending over about five jet diameters downstream of the nozzle exit. The instantaneous streamwise velocity fluctuating fields shown in figure 1(a𝑎aitalic_a,b𝑏bitalic_b) reveal that the initially-laminar jet exhibits a shorter potential core and larger jet width. The initial shear layer over the first two jet diameters is highlighted in the insets of figure 1(a𝑎aitalic_a) and (b𝑏bitalic_b). The hallmark of the untripped jet is the initially-laminar shear layer that is easily distinguished from the turbulent shear layer by its growth rate; the laminar shear layer exiting the nozzle has a significantly lower spreading rate untill x0.8𝑥0.8x\approx 0.8italic_x ≈ 0.8, where its growth rate rapidly increases after it transitions to turbulence. This distinction between the laminar and turbulent portions of the shear layer is quantitatively confirmed by the RMS profiles in panels (c𝑐citalic_c)-(g𝑔gitalic_g). At the nozzle exit in figure 1 (c𝑐citalic_c), the RMS profile of the turbulent jet peaks at 10%greater-than-or-equivalent-toabsentpercent10\gtrsim 10\%≳ 10 % of the freestream velocity close to the nozzle wall. On the other hand, expectedly, the laminar shear layer of the untripped jet has zero RMS. By x=1𝑥1x=1italic_x = 1, the RMS profile of the initially-laminar jet already resembles that of the turbulent jet. Further downstream, at x=5𝑥5x=5italic_x = 5, the RMS of the initially-laminar jet surpasses that of the turbulent jet, and eventually, in the region of self-similarity, the RMS profiles become nearly identical again. We will later show that the rapid growth of the shear layer is associated with exponential growth of the hydrodynamics instability modes supported by the initially laminar shear layer. We explore the nonlinear dynamics of the jet that lead this rapid growth and explain previous observations by (Zaman & Hussain, 1980; Kim & Choi, 2009; Bogey & Bailly, 2010).

Refer to caption
Figure 2: Experimental validation of turbulent jet LES and comparison of initially-laminar jet LES with literature (Zaman & Hussain, 1980; Bogey & Bailly, 2010): (a𝑎aitalic_a) mean and (b𝑏bitalic_b) RMS of the streamwise velocity on the centerline. The intersection of the black dashed line at ux¯/Uj=0.95¯subscript𝑢𝑥subscript𝑈𝑗0.95\overline{u_{x}}/U_{j}=0.95over¯ start_ARG italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG / italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0.95 with the mean streamwise velocity defines the length of the potential core.

Figure 2 shows a four-way comparison between initially laminar, turbulent jets, LES, and experiments. The mean and RMS streamwise velocities on the centerline are shown in figure 2 (a𝑎aitalic_a) and (b𝑏bitalic_b), respectively. In the absence of the initially-laminar jet for the considered nozzle geometry, the turbulent jet is used to validate the numerical setup. Good agreement is observed for the turbulent jet in terms of the mean and RMS streamwise velocities, but the LES of the initially laminar jet shows significant differences. Along the centerline, the mean flow of the initially laminar jet exhibits a dip at x2𝑥2x\approx 2italic_x ≈ 2 (see the inset of figure 2(a𝑎aitalic_a)) and decays rapidly beyond the end of the potential core. The mean flow velocity profile of Bogey & Bailly (2010) exhibits the same phenomenon. The RMS of the initially-laminar jet in figure 2(b𝑏bitalic_b) is notably higher than the fully turbulent case and exhibits a distinct hump at x2.8𝑥2.8x\approx 2.8italic_x ≈ 2.8. This hump, to a certain degree, was also observed by Zaman & Hussain (1980) and Bogey & Bailly (2010). This hump and elevated RMS were previously associated with vortex-pairing by Kim & Choi (2009); Bogey & Bailly (2010); Bogey et al. (2012). The authors also demonstrated that both these phenomena strongly depend on the initial shear-layer thickness, with the thicker shear layer growing faster and exhibiting enhanced vortex-pairing. In contrast, a thinner shear layer will grow earlier but slower, resulting in lower uxrmssuperscriptsubscript𝑢𝑥𝑟𝑚𝑠u_{x}^{rms}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_m italic_s end_POSTSUPERSCRIPT on the centerline (Bogey & Bailly, 2010). In the remainder of this paper, we go beyond this phenomenological description and analyze the underlying nonlinear mechanism in detail. To this end, we use spectral modal decomposition techniques and local linear theory.

3.2 Shear layer instability

Refer to caption
Figure 3: Premultiplied radially integrated PSD along x𝑥xitalic_x for initially laminar (a𝑎aitalic_a,b𝑏bitalic_b) and turbulent jets (c𝑐citalic_c,d𝑑ditalic_d): (a𝑎aitalic_a,c𝑐citalic_c) streamwise velocity, uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT; (b𝑏bitalic_b,d𝑑ditalic_d) radial velocity, ursubscript𝑢𝑟u_{r}italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. Dashed lines indicate the three tones of the initially-laminar jet, and the dotted line corresponds to the most energetic frequency of the turbulent jet.

Figure 3 shows the premultiplied PSD of the streamwise and radial velocities for the initially-laminar and turbulent jets, respectively. The PSD is computed at each spatial location and integrated radially for each streamwise location. For the initially-laminar jet, the PSDs of uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and ursubscript𝑢𝑟u_{r}italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT show the prominence of three frequency components, St=1.76,0.88St1.760.88\mbox{\it St}=1.76,0.88St = 1.76 , 0.88, and 0.44. The three peaks are located at (St𝑆𝑡Stitalic_S italic_t, x𝑥xitalic_x) = (1.76, 1.0), (0.88, 1.45), and (0.44, 2.4). The most upstream peak occurs at St=1.76𝑆𝑡1.76St=1.76italic_S italic_t = 1.76 as a result of the convective Kelvin-Helmholtz instability of the initial shear layer. Notably, it is an integer multiple of the second and third peaks. This observation suggests that St=1.76𝑆𝑡1.76St=1.76italic_S italic_t = 1.76 is the fundamental frequency. We will later confirm this using local stability theory and demonstrate in §3.3 that the second and third peaks arise from vortex-pairing, i.e., due to nonlinearity as opposed to hydrodynamic instability. The fundamental frequency translates to Stθ=fθ/Uj=0.0132subscriptSt𝜃𝑓𝜃subscript𝑈𝑗0.0132\mbox{\it St}_{\theta}=f\theta/U_{j}=0.0132St start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = italic_f italic_θ / italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 0.0132, which closely matches that of the shear layer mode observed in forced jet experiments Zaman & Hussain (1980). For all three frequencies, the PSD of the ursubscript𝑢𝑟u_{r}italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT component is greater than the uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT component. The PSDs for the turbulent jet in figure 3 (c𝑐citalic_c,d𝑑ditalic_d) are significantly lower than the initially-laminar jet. For the turbulent jet, the first tonal component is observed at x2.6𝑥2.6x\approx 2.6italic_x ≈ 2.6 and St=0.56St0.56\mbox{\it St}=0.56St = 0.56 (exact frequency determined from figure 6), denoted by the dotted line. No peaks are observed at St=1.76𝑆𝑡1.76St=1.76italic_S italic_t = 1.76 and 0.880.880.880.88, indicating that the dynamics associated with these frequencies are absent in the fully turbulent jet.

Refer to caption
Figure 4: Total RMS velocities and selected contributing frequency components along x𝑥xitalic_x: (a𝑎aitalic_a) uxrmssuperscriptsubscript𝑢𝑥𝑟𝑚𝑠u_{x}^{rms}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_m italic_s end_POSTSUPERSCRIPT on the centerline, r=0𝑟0r=0italic_r = 0; (b𝑏bitalic_b) uxrmssuperscriptsubscript𝑢𝑥𝑟𝑚𝑠u_{x}^{rms}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_m italic_s end_POSTSUPERSCRIPT on the lipline, r=0.5𝑟0.5r=0.5italic_r = 0.5; (c𝑐citalic_c) urrmssuperscriptsubscript𝑢𝑟𝑟𝑚𝑠u_{r}^{rms}italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_m italic_s end_POSTSUPERSCRIPT on the lipline. The total RMS (black curve) is plotted on the left ordinate, and the remaining curves are plotted on the right ordinate. The red curve is the sum of five frequencies St=1.76St1.76\mbox{\it St}=1.76St = 1.76, 0.88, 0.44, 0.22, and 0.11.

Figure 4 shows the contribution of tonal frequencies to the total RMS of the initially-laminar jet. This contribution is shown for the streamwise velocity on the centerline (r=0𝑟0r=0italic_r = 0) in (a𝑎aitalic_a), the streamwise velocity on the lipline (r=0.5𝑟0.5r=0.5italic_r = 0.5) in (b𝑏bitalic_b), and the radial velocity on the lipline in (c𝑐citalic_c). The total RMS is plotted on the left ordinate, whereas the RMS of five frequencies and their sum (denoted by a red line) are plotted on the right ordinate. In all cases, the frequency St=0.44𝑆𝑡0.44St=0.44italic_S italic_t = 0.44 exhibits the maximum RMS. The lower frequencies are more significant on the centerline, whereas the higher frequencies are dominant on the lipline. In figure 4(a𝑎aitalic_a), the curve representing the sum of the five frequencies, while of a lower magnitude, closely resembles the shape of the total RMS. Upon comparing the sum curve to the individual frequencies, it becomes evident that the frequencies St=0.44𝑆𝑡0.44St=0.44italic_S italic_t = 0.44 (magenta) and St=0.22𝑆𝑡0.22St=0.22italic_S italic_t = 0.22 (cyan) correspond to the accumulation of RMS at x2.8𝑥2.8x\approx 2.8italic_x ≈ 2.8 and x6𝑥6x\approx 6italic_x ≈ 6, respectively.

Along the lipline, the total RMS of streamwise and radial velocity peaks at x1.8𝑥1.8x\approx 1.8italic_x ≈ 1.8. The curve representing the sum of five frequencies matches the shape of the total curve only up to x1.8𝑥1.8x\lessapprox 1.8italic_x ⪅ 1.8. This is because the flow downstream exhibits a broadband-like nature, necessitating more frequencies to capture the fluctuating dynamics. The higher frequencies, St=1.76,0.88𝑆𝑡1.760.88St=1.76,0.88italic_S italic_t = 1.76 , 0.88, and 0.44, are more prominent on the lipline than the centerline. In figure 4(c𝑐citalic_c), these three frequency components peak successively as the preceding frequency starts to decline. This observation is consistent with the expectation that lower frequency components peak at more downstream locations where the shear layer is thicker. Our findings are in agreement with the research of Hajj et al. (1992), which demonstrates that the simultaneous decay of the fundamental frequency and growth of the subharmonic is a result of the energy transfer from the fundamental to the subharmonic through a resonance mechanism (Monkewitz, 1988). Later in §3.4, we will investigate this energy transfer using the spectral kinetic energy equation.

Refer to caption
Figure 5: Comparison of TKE and amplification rate predicted from linear stability theory with empirical data: TKE of the dominant frequencies for (a𝑎aitalic_a) initially-laminar jet and (b𝑏bitalic_b) turbulent jet. Amplification rate predicted from LST (c,d𝑐𝑑c,ditalic_c , italic_d) and empirically from data (e,f𝑒𝑓e,fitalic_e , italic_f), using equation (19), for the initially-laminar jet (c,e𝑐𝑒c,eitalic_c , italic_e) and turbulent jet (d,f𝑑𝑓d,fitalic_d , italic_f). The solid and dashed lines in (a,b𝑎𝑏a,bitalic_a , italic_b) represent the TKE computed from data and LST using equation (20), respectively. The green line in (c𝑐citalic_c) denotes the most unstable frequency at each streamwise location. The neutral stability curve is represented by the black line in (c𝑐citalic_c-f𝑓fitalic_f).

Next, we perform local linear stability analysis to understand the origin of the fundamental frequency. In particular, we seek a quantitative comparison between the empirical growth rates deducted from data and theory. We define the local amplitude of 𝐪^^𝐪\hat{\mathbf{q}}over^ start_ARG bold_q end_ARG as

A(x,m,ω)=r𝐪^𝐪^rdr=kLES,𝐴𝑥𝑚𝜔subscript𝑟superscript^𝐪^𝐪𝑟𝑟superscript𝑘LESA(x,m,\omega)=\sqrt{\int_{r}\hat{\mathbf{q}}^{*}\hat{\mathbf{q}}r\differential r% }=\sqrt{k^{\text{LES}}},italic_A ( italic_x , italic_m , italic_ω ) = square-root start_ARG ∫ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT over^ start_ARG bold_q end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG bold_q end_ARG italic_r start_DIFFOP roman_d end_DIFFOP italic_r end_ARG = square-root start_ARG italic_k start_POSTSUPERSCRIPT LES end_POSTSUPERSCRIPT end_ARG , (18)

where kLESsuperscript𝑘LESk^{\text{LES}}italic_k start_POSTSUPERSCRIPT LES end_POSTSUPERSCRIPT is the turbulent kinetic energy computed from the data. Using equation (18) and the normal mode ansatz, equation (1), allows us to compute the empirical growth rate as

αiemp=1AdAdx.superscriptsubscript𝛼𝑖emp1𝐴𝐴𝑥\alpha_{i}^{\rm{emp}}=-\frac{1}{A}\frac{\differential A}{\differential x}.italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_emp end_POSTSUPERSCRIPT = - divide start_ARG 1 end_ARG start_ARG italic_A end_ARG divide start_ARG start_DIFFOP roman_d end_DIFFOP italic_A end_ARG start_ARG start_DIFFOP roman_d end_DIFFOP italic_x end_ARG . (19)

The TKE can also be predicted from the local linear theory by integrating and squaring the theoretical amplification rate

kLST=(α(x)𝑑x)2.superscript𝑘LSTsuperscript𝛼𝑥differential-d𝑥2k^{\rm{LST}}=\pqty{\int\alpha(x)dx}^{2}.italic_k start_POSTSUPERSCRIPT roman_LST end_POSTSUPERSCRIPT = ( start_ARG ∫ italic_α ( italic_x ) italic_d italic_x end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (20)

In the absence of an amplitude in local linear theory, only a qualitative comparison of the TKE distribution can be made. Here, we choose to normalize kLSTsuperscript𝑘LSTk^{\rm{LST}}italic_k start_POSTSUPERSCRIPT roman_LST end_POSTSUPERSCRIPT by the maximum kLESsuperscript𝑘LESk^{\text{LES}}italic_k start_POSTSUPERSCRIPT LES end_POSTSUPERSCRIPT of each frequency.

Figure 5 shows the TKE (a,b𝑎𝑏a,bitalic_a , italic_b) and spatial growth rate (c𝑐citalic_c-f𝑓fitalic_f) for the axisymmetric component m=0𝑚0m=0italic_m = 0 in both jets. The spatial growth rate is computed from local linear theory in panels (c,d𝑐𝑑c,ditalic_c , italic_d) and empirically using equation (19) in panels (e,f𝑒𝑓e,fitalic_e , italic_f). The black line represents the neutral stability curve, while the green line denotes the frequency associated with the largest growth rate at each streamwise location. The TKE for St=1.76St1.76\mbox{\it St}=1.76St = 1.76, 0.88, and 0.44 in the initially laminar jet and for St=0.56St0.56\mbox{\it St}=0.56St = 0.56 in the turbulent jet, is shown in figure 5(a𝑎aitalic_a) and (b𝑏bitalic_b), respectively. The TKE estimated from the linear stability theory (kLSTsuperscriptkLST\text{k}^{\rm{LST}}k start_POSTSUPERSCRIPT roman_LST end_POSTSUPERSCRIPT), denoted by the dotted-dashed lines, is also shown, demonstrating a good fit with the data. It is observed that the TKE peak of each frequency is approximately located on the neutral stability curve. This observation is in agreement with the fact that the local growth rate at the maximum is zero. As a visual aid, dashed lines indicating the locations of the maximum TKE extend into figure 5(c𝑐citalic_c) and (d𝑑ditalic_d), which show the spatial growth rates of the initially laminar and turbulent jets, respectively. Figure 5(c𝑐citalic_c) reveals that St1.76St1.76\mbox{\it St}\approx 1.76St ≈ 1.76 is the most unstable frequency at the nozzle’s exit for the initially-laminar jet. The coalescence of the observations that St=1.76St1.76\mbox{\it St}=1.76St = 1.76 is the most unstable frequency and occurs most upstream at the nozzle’s exit confirms our previous interpretation of St1.76St1.76\mbox{\it St}\approx 1.76St ≈ 1.76 as the fundamental frequency. The absence of a peak at St=0.88St0.88\mbox{\it St}=0.88St = 0.88 and 0.44 indicates that these frequencies are not a result of hydrodynamic instability. Despite being spatially unstable in their own right, these tonal frequencies arise from the nonlinearity involving the fundamental frequency. On the other hand, for the turbulent jet (figure 5(d𝑑ditalic_d)), the growth rates are much lower in magnitude, and the most unstable frequency is St0.90St0.90\mbox{\it St}\approx 0.90St ≈ 0.90. The empirical spatial growth rates in (e,f)𝑒𝑓(e,f)( italic_e , italic_f ) are similar to those estimated by local linear theory. In particular, the most unstable frequencies St1.76𝑆𝑡1.76St\approx 1.76italic_S italic_t ≈ 1.76 and 0.9, for the initially laminar and turbulent jet, respectively, are well estimated. The notable similarity between the empirical and theoretical results reveals two key points: first, it confirms the validity of the assumption of local linear theory, and second, it supports the physical interpretation that the observed amplitude disturbances of individual frequency components represent hydrodynamic waves.

Refer to caption
Figure 6: SPOD eigenspectra with a focus on the shear layer untill the end of the potential core: (a𝑎aitalic_a) initially-laminar jet; (b𝑏bitalic_b) turbulent jet. The white-shaded area in the top row denotes the focus region of SPOD. Dashed lines indicate the three tones of the initially-laminar jet. Dotted lines in (a𝑎aitalic_a) correspond to the ultra harmonics St=0.66St0.66\mbox{\it St}=0.66St = 0.66 and 1.32, and the dotted line in (b𝑏bitalic_b) to the most energetic frequency of the turbulent jet.

As both jets are statistically stationary flows, we use SPOD to extract the spatiotemporal coherent structures. To emphasize the dynamics in the initial shear layer, SPOD is computed using a weighting function that assigns zero weights to the region we wish to exclude. Specifically, zero weights are assigned to the areas outside the shear layer and beyond the end of the potential core. These weighting functions are shown in the top row of figure 6, where the black regions represent zero weights. Figure 6 shows the SPOD eigenspectra of the initially-laminar and turbulent jets. Corresponding to the previously observed tones in figure 3, SPOD identifies the tonal peaks at St=1.76St1.76\mbox{\it St}=1.76St = 1.76, 0.88, and 0.44 for the initially-laminar jet in figure 6(a𝑎aitalic_a). Interestingly, a broader peak is observed at the fundamental frequency St=1.76St1.76\mbox{\it St}=1.76St = 1.76, which has lower energy compared to the two subharmonics. Additionally, the spectra reveal the presence of ultra and superharmonics at St2.7St2.7\mbox{\it St}\approx 2.7St ≈ 2.7 and 3.5, respectively. We have confirmed from the spatial linear theory that St=3.5St3.5\mbox{\it St}=3.5St = 3.5 is indeed a higher harmonic, as there was no associated peak in the amplification rate. No distinct peaks are observed at the ultraharmonic frequencies St=0.66St0.66\mbox{\it St}=0.66St = 0.66 and 1.321.321.321.32 (denoted by gray dotted lines). Later, in figure 12, we will demonstrate that despite not being energetically important, they play an active role in triadic interactions. The SPOD spectrum of the turbulent jet is relatively broadband and exhibits a low-rank behaviour for 0.3St10.3St10.3\leq\mbox{\it St}\leq 10.3 ≤ St ≤ 1 (Schmidt et al., 2018). Overall, figure 6 indicates that the presence of discrete tones leads to a much higher fluctuation level than the turbulent case.

Refer to caption
Figure 7: Leading SPOD modes of the fundamental and four of its subharmonic frequencies: (a𝑎aitalic_a,b𝑏bitalic_b) St=1.76St1.76\mbox{\it St}=1.76St = 1.76; (c𝑐citalic_c,d𝑑ditalic_d) St=0.88St0.88\mbox{\it St}=0.88St = 0.88; (e𝑒eitalic_e,f𝑓fitalic_f) St=0.44St0.44\mbox{\it St}=0.44St = 0.44; (g𝑔gitalic_g,hhitalic_h) St=0.22St0.22\mbox{\it St}=0.22St = 0.22; (i𝑖iitalic_i,j𝑗jitalic_j) St=0.11St0.11\mbox{\it St}=0.11St = 0.11. The left column represents the streamwise velocity component, uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT and the right column represents the radial velocity, ursubscript𝑢𝑟u_{r}italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. The potential core and the jet width are indicated as lines of constant uxsubscript𝑢𝑥u_{x}italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT at 95% and 10% of the jet velocity Ujsubscript𝑈𝑗U_{j}italic_U start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, respectively.

The leading SPOD modes associated with St=1.76,0.88,0.44,0.22St1.760.880.440.22\mbox{\it St}=1.76,0.88,0.44,0.22St = 1.76 , 0.88 , 0.44 , 0.22, and 0.11, are shown in figure 7. Due to the spreading of the shear layer, structures with lower frequencies are supported farther downstream than those associated with higher frequencies. At the fundamental frequency, the leading SPOD mode materializes as a Kelvin-Helmholtz wavepacket with compact support in the region 0.8x1.20.8𝑥1.20.8\leq x\leq 1.20.8 ≤ italic_x ≤ 1.2 around the lip line. At a lower frequency of St=0.44St0.44\mbox{\it St}=0.44St = 0.44, the leading SPOD mode exhibits larger spatial support within the range 1x61𝑥61\leq x\leq 61 ≤ italic_x ≤ 6 and 0r10𝑟10\leq r\leq 10 ≤ italic_r ≤ 1. Notably, it exhibits high amplitude in 2x32𝑥32\leq x\leq 32 ≤ italic_x ≤ 3, where the streamwise velocity is concentrated on the centerline and the radial velocity on the lipline. Furthermore, the presence of this Kelvin-Helmholtz-type wavepacket structure is responsible for the accumulation of the RMS streamwise velocity on the centerline at x3𝑥3x\approx 3italic_x ≈ 3. Similarly, the structures at other frequencies are also associated with the peaks of the RMS velocities in figure 7. For instance, the global maximum of the spatial structure of the radial velocity for St=0.88St0.88\mbox{\it St}=0.88St = 0.88 is at (x,r)(1.5,0.5)𝑥𝑟1.50.5(x,r)\approx(1.5,0.5)( italic_x , italic_r ) ≈ ( 1.5 , 0.5 ), the corresponding to the peak of the blue curve in figure 4 (c𝑐citalic_c).

3.3 Vortex pairing

Refer to caption
Figure 8: Time traces exemplifying two successive vortex pairing events are visualized in terms of the azimuthal vorticity for the m=0𝑚0m=0italic_m = 0 component. The green, blue, and magenta rectangles enclose the fundamental, subharmonic, and second subharmonic vortices, respectively.

The vortex pairing process is visualized in figure 8. Eight time instances of the vorticity ωθsubscript𝜔𝜃\omega_{\theta}italic_ω start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT for the axisymmetric component m=0𝑚0m=0italic_m = 0 are shown. These snapshots follow two successive vortex pairing events, where four fundamental vortices merge into two subharmonic vortices, eventually coalescing into a single vortex corresponding to the second subharmonic frequency. This process involves the formation of a larger vortex associated with half the frequency and half the wavenumber of the previous vortices. The highlighted vortical structures are unambiguously associated with their respective frequencies. This association is determined by considering their spatial support consistently with figures 4(c𝑐citalic_c), 5(a𝑎aitalic_a), and 7(a𝑎aitalic_a-f𝑓fitalic_f). The first snapshot highlights two developing and two developed vortices enclosed in green rectangles, corresponding to a frequency of St=1.76St1.76\mbox{\it St}=1.76St = 1.76. These vortices are formed due to the roll-up of the shear layer. The following three snapshots show the pairing of these vortices in the region 1.2x2.1less-than-or-similar-to1.2𝑥less-than-or-similar-to2.11.2\lesssim x\lesssim 2.11.2 ≲ italic_x ≲ 2.1, denoted by the blue rectangle in panels 8(b𝑏bitalic_b-d𝑑ditalic_d). This vortex pairing results in the formation of the St=0.88St0.88\mbox{\it St}=0.88St = 0.88 vortex. Another instance of vortex pairing is evident in panels 8(c𝑐citalic_c-e𝑒eitalic_e). Next, the two St=0.88St0.88\mbox{\it St}=0.88St = 0.88 vortices, denoted by blue rectangles in (e𝑒eitalic_e), undergo pairing to form the St=0.44St0.44\mbox{\it St}=0.44St = 0.44 vortex. Panels 8(e𝑒eitalic_e-f𝑓fitalic_f) demonstrate this process. In this process, the vortex at a more upstream location accelerates and catches up with the decelerating downstream vortex. Eventually, they wrap around each other and form a single vortex at x3.5𝑥3.5x\approx 3.5italic_x ≈ 3.5. These consecutive vortex pairing events are qualitative evidence of an inverse cascade, transferring energy from smaller to larger structures.

Refer to caption
Figure 9: x𝑥xitalic_x-t𝑡titalic_t plots along the lipline showing the vorticity fluctuations, ωθsubscriptsuperscript𝜔𝜃\omega^{\prime}_{\theta}italic_ω start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT: (a𝑎aitalic_a) representative time interval; (b𝑏bitalic_b) conditional average of the SPOD-band filtered data about the spatial location x3.5𝑥3.5x\approx 3.5italic_x ≈ 3.5. The successive vortex-pairing events shown in figure 8 are enclosed by the yellow box in (a𝑎aitalic_a). The green, blue, and magenta lines correspond to the fundamental, subharmonic, and second subharmonic vortices, respectively.

A different perspective on vortex pairing is presented in terms of the x𝑥xitalic_x-t𝑡titalic_t plots in figure 9. Figure 9(a𝑎aitalic_a) shows the vorticity fluctuations, ωθsuperscriptsubscript𝜔𝜃\omega_{\theta}^{\prime}italic_ω start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, of the m=0𝑚0m=0italic_m = 0 component along the lipline in the time interval, 35t6535𝑡6535\leq t\leq 6535 ≤ italic_t ≤ 65. Merging lines with slopes greater than zero indicate the pairing of vortices. For instance, the vortex pairing process in figure 8 is highlighted in yellow dashed lines. However, note that this represents a single event of vortex pairing chosen for its clarity in figure 8. In order to confirm that this vortex-pairing sequence is indeed a prevailing flow feature, we deploy a statistical perspective. To this end, we use conditional averaging and SPOD-band pass filtering developed by Nekkanti & Schmidt (2021). To isolate the two successive vortex-pairing events, we band-pass filter the data using SPOD, retaining only the fundamental frequency and its first two subharmonics. Next, we select the location x=3.5,r=0.5formulae-sequence𝑥3.5𝑟0.5x=3.5,r=0.5italic_x = 3.5 , italic_r = 0.5 and extract all local peaks that exceed 25% of their global maximum. We obtain the i𝑖iitalic_i-th realization of the vortex-pairing sequence by collecting all the snapshots in the interval [t0(i)20,t0(i)+10]superscriptsubscript𝑡0𝑖20superscriptsubscript𝑡0𝑖10[t_{0}^{(i)}-20,t_{0}^{(i)}+10][ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT - 20 , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT + 10 ], where t0(i)superscriptsubscript𝑡0𝑖t_{0}^{(i)}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT is the time instant of the i𝑖iitalic_i-th local peak. Finally, the statistical representation of the vortex-pairing sequence is obtained by averaging over all realizations. The x𝑥xitalic_x-t𝑡titalic_t plot along the lipline obtained from this conditional averaging is shown in figure 9(b𝑏bitalic_b). Evident here are the first and second vortex-pairing events at x1.5𝑥1.5x\approx 1.5italic_x ≈ 1.5 and x3𝑥3x\approx 3italic_x ≈ 3, respectively. For illustrative purposes only, we use a polynomial curve fitting to demonstrate this process. The merging of two lines indicates the acceleration and deceleration of the vortices in the vortex pairing event. We observe that the phase speed outside of the vortex pairing events is constant. This is expected, as these result from the Kelvin-Helmholtz-type instability waves governed by the dispersion relation ω/k=cph𝜔𝑘subscript𝑐𝑝\omega/k=c_{ph}italic_ω / italic_k = italic_c start_POSTSUBSCRIPT italic_p italic_h end_POSTSUBSCRIPT. Here, for the second subharmonic frequency, cph=0.55subscript𝑐𝑝0.55c_{ph}=0.55italic_c start_POSTSUBSCRIPT italic_p italic_h end_POSTSUBSCRIPT = 0.55.

3.4 SPOD-based spectral energy transfer

We now investigate the spectral energy balance of turbulent kinetic energy as outlined in §2.3. In particular, we use it to identify the net production and nonlinear transfer of TKE associated with the frequencies involved in the vortex pairing process. The nonlinear energy transfer term could alternatively be estimated using bispectral mode decomposition (BMD, Schmidt, 2020). However, to maintain consistency with the modes employed in this analysis, we estimate the nonlinear energy transfer using the leading SPOD modes. Moreover, employing the SPOD modal basis ensures direct comparability among different terms such as production, nonlinear transfer, and dissipation. In appendix A, we tailor BMD to estimate nonlinear energy transfer and compare it with those estimated from SPOD. We find that both methods yield qualitatively similar results.

Figure 10 shows the production and nonlinear energy transfer at frequencies St=1.76,0.88St1.760.88\mbox{\it St}=1.76,0.88St = 1.76 , 0.88, and 0.44. These terms are radially integrated and plotted as a function of the streamwise location. The production term is computed using equation (14). Positive production indicates the energy gain from the mean flow, and negative production represents energy loss to the mean flow. Figure 10(a𝑎aitalic_a) shows that St=1.76St1.76\mbox{\it St}=1.76St = 1.76 is the earliest to gain energy from the mean flow, and as it saturates, St=0.88St0.88\mbox{\it St}=0.88St = 0.88 begins its growth. Subsequently, St=0.88St0.88\mbox{\it St}=0.88St = 0.88 attains its global maximum at x=1.25𝑥1.25x=1.25italic_x = 1.25, which also corresponds to the location of the global minimum of St=1.76St1.76\mbox{\it St}=1.76St = 1.76. Similarly, the global maximum of St=0.44St0.44\mbox{\it St}=0.44St = 0.44 and the global minimum of St=0.88St0.88\mbox{\it St}=0.88St = 0.88 are in close proximity. The saturation of the subharmonic frequency has been linked to the onset of vortex-pairing by Ho & Huang (1982) and was confirmed by Hajj et al. (1992, 1993). Our findings are in agreement with these studies. The production curves, in combination with figure 8, demonstrate this, where the location corresponding to the peak production of the subharmonic also marks the beginning of the merging process. Furthermore, the production of St=1.76,0.88St1.760.88\mbox{\it St}=1.76,0.88St = 1.76 , 0.88, and 0.44 becomes negative at x=1.07𝑥1.07x=1.07italic_x = 1.07, 1.64, and 2.64 respectively. These locations closely correspond to the location of the neutral stability points as predicted by linear stability theory (denoted by dashed lines, also see figure 5). The correspondence is expected as the positive production is associated with the amplification rate, and the negative production is associated with the decay rate of each frequency.

Refer to caption
Figure 10: Production and nonlinear energy transfer terms for St=1.76,0.88St1.760.88\mbox{\it St}=1.76,0.88St = 1.76 , 0.88, and 0.44 integrated in r𝑟ritalic_r and as a function of streamwise location. Dashed lines indicate neutral stability points, predicted by the linear stability theory, of the corresponding frequency. The black dotted line indicates the onset of nonlinear interactions.

The net nonlinear energy transfer, computed using equation (17), is shown in figure 10(b𝑏bitalic_b). The net nonlinear energy transfer is always negative for St=1.76St1.76\mbox{\it St}=1.76St = 1.76. On the other hand, for St=0.88St0.88\mbox{\it St}=0.88St = 0.88, and 0.44, the energy is initially transferred into these frequencies and as the flow evolves downstream, energy is extracted from these frequencies. The blue and magenta curves in the region 1.3x2.31.3𝑥2.31.3\leq x\leq 2.31.3 ≤ italic_x ≤ 2.3 exhibit a similar shape but are opposite in sign, indicating that nonlinear interactions result in energy loss for St=0.88St0.88\mbox{\it St}=0.88St = 0.88 and energy gain for St=0.44St0.44\mbox{\it St}=0.44St = 0.44. This phenomenon is expected for a vortex pairing process, as the energy is transferred from St=0.88St0.88\mbox{\it St}=0.88St = 0.88 to its subharmonic, St=0.44St0.44\mbox{\it St}=0.44St = 0.44. The black dotted line denotes the location for the onset of nonlinearity at x=0.72𝑥0.72x=0.72italic_x = 0.72. This also corresponds to the location where St=0.88St0.88\mbox{\it St}=0.88St = 0.88 starts to grow (see figure 10(a𝑎aitalic_a)), implying that the nonlinear interactions are triggered due to the growth of the subharmonic. Overall these observations suggest that the spatially unstable fundamental grows linearly, gaining energy from the mean flow, followed by successive nonlinear growth and decay of its subharmonics. As pointed out earlier, this successive growth and decay is characteristic of the vortex-pairing process.

Refer to caption
Figure 11: Spatial fields of production (a𝑎aitalic_a,d𝑑ditalic_d,g𝑔gitalic_g), dissipation (b𝑏bitalic_b,e𝑒eitalic_e,hhitalic_h), and net nonlinear energy transfer (c𝑐citalic_c,f𝑓fitalic_f,i𝑖iitalic_i).

Figure 11 shows the spatial fields of production, dissipation, and net nonlinear transfer at St=1.76St1.76\mbox{\it St}=1.76St = 1.76, 0.88, and 0.44. These fields are computed using equations (14), (15), and (17), respectively. As in figure 10, the production fields in panels (a𝑎aitalic_a,d𝑑ditalic_d,g𝑔gitalic_g) indicate that the fundamental frequency and its subharmonics initially gain energy from the mean flow and subsequently transfer energy back to the mean flow. These fields also reveal that this energy transfer from the mean flow is localized in the region about the lip line. Figure 11 (b𝑏bitalic_b,e𝑒eitalic_e,hhitalic_h) shows that the spatial dissipation fields are significantly lower in magnitude and, hence, have minimal impact on the energy budget of the vortex pairing process. The net nonlinear energy transfer fields, shown in panels (c𝑐citalic_c,f𝑓fitalic_f,i𝑖iitalic_i), exhibit multilobe structures. These fields elucidate the spatial dependence of the nonlinear kinetic energy transfer. Figure 11 highlights that individual terms in the spectral TKE budget can exhibit high local values, yet their overall contributions may remain small when integrated across the entire spatial domain.

Refer to caption
Figure 12: Nonlinear energy transfer using SPOD: transfer term bispectrum for (a𝑎aitalic_a) entire domain and (b𝑏bitalic_b) shear layer subdomain, ΩrsubscriptΩ𝑟\Omega_{r}roman_Ω start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT; spatial fields for the triads (c𝑐citalic_c) (1.76,-0.88,0.88) and (d𝑑ditalic_d) (0.88,-0.44,0.44) are compared to the radially integrated TKE of St=1.76,0.88St1.760.88\mbox{\it St}=1.76,0.88St = 1.76 , 0.88, and 0.44.

Figure 12(a𝑎aitalic_a) and (b𝑏bitalic_b) show the triadic energy transfer obtained by integrating equation (16) over the entire spatial domain and a domain ΩrsubscriptΩ𝑟\mathrm{\Omega}_{r}roman_Ω start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT that focuses on the initial shear layer within the first two jet diameters in x,r[0,2]×[0,6]𝑥𝑟0206x,r\in[0,2]\times[0,6]italic_x , italic_r ∈ [ 0 , 2 ] × [ 0 , 6 ], respectively. In panels (a𝑎aitalic_a) and (b𝑏bitalic_b), positive values (red color) denote energy transfer to frequency St3subscriptSt3\mbox{\it St}_{3}St start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT from frequencies St1subscriptSt1\mbox{\it St}_{1}St start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and St2subscriptSt2\mbox{\it St}_{2}St start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, while negative values (blue color) indicate the extraction of energy from St3subscriptSt3\mbox{\it St}_{3}St start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT by St1subscriptSt1\mbox{\it St}_{1}St start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and St2subscriptSt2\mbox{\it St}_{2}St start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The two triads with the highest positive and negative intensity are (0.88,0.44,0.44)0.880.440.44(0.88,-0.44,0.44)( 0.88 , - 0.44 , 0.44 ) and (0.44,0.44,0.88)0.440.440.88(0.44,0.44,0.88)( 0.44 , 0.44 , 0.88 ). These triads exhibit positive and negative energy transfer, respectively. However, both triads convey the same information. The triad (0.88,0.44,0.44)0.880.440.44(0.88,-0.44,0.44)( 0.88 , - 0.44 , 0.44 ) has a positive value, signifying the transfer of energy from St=0.88St0.88\mbox{\it St}=0.88St = 0.88 and -0.44 to St=0.44St0.44\mbox{\it St}=0.44St = 0.44. On the other hand, the triad (0.44,0.44,0.88)0.440.440.88(0.44,0.44,0.88)( 0.44 , 0.44 , 0.88 ) has a negative value, indicating that St=0.44St0.44\mbox{\it St}=0.44St = 0.44 extracts energy from St=0.88St0.88\mbox{\it St}=0.88St = 0.88. These triads clearly suggest that the energy is transferred from the subharmonic to the second subharmonic frequency. Other significant triads include (0.44, 0.88, 1.32), (1.32, -0.44, 0.88), and (1.32, -0.88, 0.44), highlighting the presence of the ultraharmonic frequency St=1.32St1.32\mbox{\it St}=1.32St = 1.32. The occurrence of St=1.32St1.32\mbox{\it St}=1.32St = 1.32 is interesting in itself as its relevance was not apparent from the SPOD analysis, i.e., no distinct peak was found at St=1.32St1.32\mbox{\it St}=1.32St = 1.32. The nonlinear energy transfer analysis reveals that the ultraharmonic St=1.32St1.32\mbox{\it St}=1.32St = 1.32 is created from the sum interaction of 0.44 and 0.88.

In figure 12(a𝑎aitalic_a), the triads (1.76, -0.88, 0.88) and (0.88, 0.88, 1.76) exhibit lower amplitudes. This is because the fundamental frequency St=1.76St1.76\mbox{\it St}=1.76St = 1.76 is dynamically significant but not energetically dominant. Therefore, to shed light on the nonlinear interactions of the fundamental frequency, we narrow our focus to the first two jet diameters. Figure 12(b𝑏bitalic_b) now illuminates the triads (1.76,-0.88,0.88) and (0.88,0.88,1.76). These triads show a similar energy transfer behavior characteristic of the vortex-pairing process. These findings, in combination with observations from figures 5, 8, and 9, suggest that during the vortex pairing process, the nonlinear interactions cause the energy to be initially transferred from the fundamental (St=1.76St1.76\mbox{\it St}=1.76St = 1.76) to its first subharmonic (St=0.88St0.88\mbox{\it St}=0.88St = 0.88), and then from the first subharmonic to the second subharmonic (St=0.44St0.44\mbox{\it St}=0.44St = 0.44). Our results hence support the hypothesis of a parametric resonance mechanism proposed by Monkewitz (1988).

The spatial fields of 𝓉nlrank-1superscriptsubscript𝓉nlrank-1\mathscr{t}_{\text{nl}}^{\text{rank-1}}script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT for the dynamically dominant and energetically significant triad are shown in figure 12(c𝑐citalic_c) and (d𝑑ditalic_d), respectively. The TKE of the fundamental, subharmonic, and second subharmonic frequencies are overlaid on the contours of the 𝓉nlrank-1superscriptsubscript𝓉nlrank-1\mathscr{t}_{\text{nl}}^{\text{rank-1}}script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT field. For the (1.76, -0.88, 0.88) triad, the 𝓉nlrank-1superscriptsubscript𝓉nlrank-1\mathscr{t}_{\text{nl}}^{\text{rank-1}}script_t start_POSTSUBSCRIPT nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT rank-1 end_POSTSUPERSCRIPT field is concentrated in the region 0.95x1.4less-than-or-similar-to0.95𝑥less-than-or-similar-to1.40.95\lesssim x\lesssim 1.40.95 ≲ italic_x ≲ 1.4, corresponding to where the fundamental frequency decays and the subharmonic frequency grows. A similar trend is evident in panel (d𝑑ditalic_d), demonstrating that the nonlinear energy transfer of the triad (0.88, -0.44, 0.44) is localized to the region of subharmonic decay and second subharmonic growth. In summary, SPOD-based transfer analysis allowed us to systematically catalog the triadic energy transfer among the SPOD modes, thus establishing a direct link between energy flow analysis and the phenomenon of vortex pairing.

4 Discussion and conclusions

Many studies have investigated the vortex pairing process in mixing layers, free shear layers, and natural and forced jets. It is a nonlinear process that involves the merging of two smaller vortices into a larger vortex with half the frequency. Previous studies have focussed on the hydrodynamics instabilities (Michalke, 1965; Kelly, 1967), phase locking of vortices (Monkewitz, 1988; Husain & Hussain, 1995), and specific aspects of energy transfer (Mankbadi, 1985; Paschereit et al., 1995) during the vortex pairing process. This study reveals the complete physical picture of the vortex-pairing process, establishes a clear concept for identifying the fundamental frequency, and provides a framework to characterize the energy transfer between the most energetic coherent structures.

The analysis was conducted for two LES of initially laminar and turbulent jets at \Rey=50000\Rey50000\Rey=50000= 50000. PSDs and local linear theory, confirm the expectation that the hydrodynamic instabilities are more pronounced in the initially-laminar jet. These instabilities cause the initially-laminar jet to transition quickly into turbulence within the first two jet diameters. A comparison of the two jets shows that the initially-laminar jet starts to develop from a more downstream location but grows more rapidly. This delayed but faster growth is triggered by the hydrodynamic instabilities. The boundary layer of the turbulent jet is tripped inside the nozzle, and the shear layer is already fully turbulent at the nozzle’s exit. Therefore, its shear layer’s growth rate is more gradual compared to that of the initially-laminar jet. Local linear theory identifies the fundamental frequency, i.e., the frequency with the largest spatial growth rate, as St=1.76St1.76\mbox{\it St}=1.76St = 1.76 for the initially-laminar jet. At this frequency, the shear layer rolls up into vortices. The tones at St=0.88St0.88\mbox{\it St}=0.88St = 0.88 and 0.44 are not distinguished as distinct peaks in the local linear theory and are therefore a consequence of the nonlinear interactions of the fundamental frequency. A remarkable agreement was found between the spatial growth rates predicted from theory and empirically from data. This validates the parallel flow assumption and strongly suggests that the purely empirical approach can be utilized to estimate spatial growth rates in turbulent flows. Conversely, linear stability theory can provide an accurate prediction of growth rates of coherent structures despite the presence of nonlinear dynamics. The growth rate spectrum estimated empirically is also broadband, which re-emphasizes that the peaks in PSD at first and second subharmonics are due to nonlinear effects.

Different visualization techniques give a clear phenomenological understanding of the vortex pairing process, wherein the accelerating upstream vortex and decelerating downstream vortex merge to form a larger vortex with twice the wavelength of the preceding ones. The process starts upstream with two vortices associated with the fundamental frequency merging to form a vortex associated with the subharmonic frequency. Subsequently, the two subharmonic vortices merge to result in a second subharmonic vortex. The second subharmonic frequency is energetically the most significant, while the fundamental frequency, despite its low energy, is dynamically the most important as it dictates the entire nonlinear dynamics. This dynamical significance can be clearly inferred from LST. While LST predicts the fundamental frequency to have the highest amplification rate, it does not indicate the presence of any other peaks. As the flow is convectively dominated, the fundamental frequency primarily influences the dynamics, confining them mainly to its subharmonic frequencies, which result from nonlinear interactions. This is a manifestation of the importance of distinguishing between dynamical relevance and energetic significance, as highlighted, among others, by Schmid (2010).

SPOD-based spectral TKE analysis was performed to characterize the energy transfer during the vortex pairing process. The focus was on the interactions between the mean flow, fundamental frequency, and subharmonic frequencies. The production and nonlinear transfer terms are major contributors to energy transfer, while the effect of dissipation is negligible. The fundamental frequency gains energy from the mean flow, while its subharmonics gain energy from both the mean flow and their harmonic. Quantitatively, the energy gained from the mean flow is greater than from its harmonic. In a spatial sense, as two fundamental vortices merge, there is a backscatter of energy from the fundamental to its subharmonic. This process repeats itself for higher subharmonics. Few studies (Hajj et al., 1992, 1993) argue that the dominant interaction during vortex pairing is between the subharmonic and the fundamental frequency, while other studies (Mankbadi, 1985; Paschereit et al., 1995) argue that the subharmonic gains most of its energy from the mean flow and the fundamental-subharmonic interaction only acts as a catalyst. Our findings show that both views are correct in their own right. In an energetic sense, the energy extracted by the subharmonic from the mean flow is more significant. However, from a dynamical perspective, the fundamental-subharmonic interaction is more significant.

The entire process can be summarized in the following steps:

  1. 1.

    The hydrodynamic instabilities initiate the transition into turbulence, causing the shear layer to grow rapidly.

  2. 2.

    Through exponential growth, the fundamental frequency attains significant amplitude and triggers the roll-up of the shear layer. As the fundamental frequency grows, it extracts energy from the mean flow.

  3. 3.

    The saturation and subsequent decay of the fundamental frequency mark the onset of vortex pairing.

  4. 4.

    As the vortex pairing continues, the subharmonic frequency acquires energy linearly from the mean flow and nonlinearly through backscatter from the fundamental frequency. The eventual saturation of the subharmonic frequency signals the completion of the vortex pairing process.

  5. 5.

    Processes analogous to steps (ii) and (iii) then repeat to create higher subharmonics.

Acknowledgments

OTS and AN are grateful for support from the Office of Naval Research under grant N00014-20-1-2311. T.C. acknowledges support from ONR grants, N00014-20-1-2311 and N00014-23-12650. The main LES calculations were performed at SDSC through UC@HPC.

Appendix A Nonlinear energy transfer using SPOD and BMD

Here, we qualitatively compare the nonlinear energy transfer estimated from SPOD and BMD. BMD is modal decomposition technique that can be understood as an extension of classical bispectral analysis to multidimensional and multivariate data. It identifies the spatially coherent structures associated with the triadic interactions by maximizing the integrated point-wise bispectrum

b(fk,fl)=E[Ω𝐪^1(𝐱,fk)𝐪^2(𝐱,fl)𝐪^3(𝐱,fk+fl)𝑑𝐱].𝑏subscript𝑓𝑘subscript𝑓𝑙𝐸subscriptΩsuperscriptsubscript^𝐪1𝐱subscript𝑓𝑘superscriptsubscript^𝐪2𝐱subscript𝑓𝑙subscript^𝐪3𝐱subscript𝑓𝑘subscript𝑓𝑙differential-d𝐱b(f_{k},f_{l})=E\bqty{\int_{\Omega}\hat{\mathbf{q}}_{1}^{*}(\mathbf{x},f_{k})% \circ\hat{\mathbf{q}}_{2}^{*}(\mathbf{x},f_{l})\circ\hat{\mathbf{q}}_{3}(% \mathbf{x},f_{k}+f_{l})d\mathbf{x}}.italic_b ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = italic_E [ start_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT over^ start_ARG bold_q end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ∘ over^ start_ARG bold_q end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ∘ over^ start_ARG bold_q end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) italic_d bold_x end_ARG ] . (21)

For further details on computing the mode bispectrum, the reader is referred to Schmidt (2020). We tailor BMD to estimate the nonlinear energy transfer by maximizing the following point-wise integral

bnlBMD(fk,fl)=E[Ω𝐮^i(𝐱,fk)𝐮^j𝐱i(𝐱,fl)𝐮^j(𝐱,fk+fl)𝑑𝐱].superscriptsubscript𝑏nlBMDsubscript𝑓𝑘subscript𝑓𝑙𝐸subscriptΩsuperscriptsubscript^𝐮𝑖𝐱subscript𝑓𝑘superscriptsubscript^𝐮𝑗subscript𝐱𝑖𝐱subscript𝑓𝑙subscript^𝐮𝑗𝐱subscript𝑓𝑘subscript𝑓𝑙differential-d𝐱b_{\rm{nl}}^{\rm{BMD}}(f_{k},f_{l})=E\bqty{\int_{\Omega}\hat{\mathbf{u}}_{i}^{% *}(\mathbf{x},f_{k})\circ\frac{\partial\hat{\mathbf{u}}_{j}}{\partial\mathbf{x% }_{i}}^{*}(\mathbf{x},f_{l})\circ\hat{\mathbf{u}}_{j}(\mathbf{x},f_{k}+f_{l})d% \mathbf{x}}.italic_b start_POSTSUBSCRIPT roman_nl end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_BMD end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = italic_E [ start_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT over^ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ∘ divide start_ARG ∂ over^ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ bold_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) ∘ over^ start_ARG bold_u end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x , italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) italic_d bold_x end_ARG ] . (22)
Refer to caption
Figure 13: Comparison of nonlinear energy transfer: (a𝑎aitalic_a) SPOD; (b𝑏bitalic_b) BMD.

Figure 13 shows the nonlinear energy transfer estimated using SPOD (equation 16) and BMD (equation 22). Qualitatively similar trends are observed, particularly, the direction of energy transfer is same for both methods, i.e., all the significant triads exhibit the same sign. Additionally, the triad with highest intensity for both methods is (0.88,-0.44,0.44), which is representative of energy transfer from subharmonic to second subharmonic. Note that, the BMD-based nonlinear energy transfer exhibits more noise in comparison to that of the SPOD-based 𝓉nlsubscript𝓉nl\mathscr{t}_{\rm{nl}}script_t start_POSTSUBSCRIPT roman_nl end_POSTSUBSCRIPT, this is because the BMD in comparison to SPOD requires more data for convergence.

References

  • Barkley (2006) Barkley, D. 2006 Linear analysis of the cylinder wake mean flow. Europhys. Lett. 75 (5), 750.
  • Becker & Massaro (1968) Becker, H.A. & Massaro, T.A. 1968 Vortex evolution in a round jet. J. Fluid Mech. 31 (3), 435–448.
  • Biswas et al. (2022) Biswas, N., Cicolin, M.M. & Buxton, O.R.H. 2022 Energy exchanges in the flow past a cylinder with a leeward control rod. J. Fluid Mech. 941, A36.
  • Bogey & Bailly (2005) Bogey, C. & Bailly, C. 2005 Effects of inflow conditions and forcing on subsonic jet flows and noise. AIAA J. 43 (5), 1000–1007.
  • Bogey & Bailly (2010) Bogey, C. & Bailly, C. 2010 Influence of nozzle-exit boundary-layer conditions on the flow and acoustic fields of initially laminar jets. J. Fluid Mech. 663, 507–538.
  • Bogey et al. (2012) Bogey, C., Marsden, O. & Bailly, C. 2012 Influence of initial turbulence level on the flow and sound fields of a subsonic jet at a diameter-based reynolds number of 105. J. Fluid Mech. 701, 352–385.
  • Brès et al. (2017) Brès, G.A., Ham, F.E., Nichols, J.W. & Lele, S.K. 2017 Unstructured large-eddy simulations of supersonic jets. AIAA J. pp. 1164–1184.
  • Brès et al. (2018) Brès, G.A., Jordan, P., Jaunet, V., Le Rallic, M., Cavalieri, A.V.G., Towne, A., Lele, S.K., Colonius, T. & Schmidt, O.T. 2018 Importance of the nozzle-exit boundary-layer state in subsonic turbulent jets. J. Fluid Mech. 851, 83–124.
  • Bridges & Hussain (1987) Bridges, J.E. & Hussain, A.K.M.F. 1987 Roles of initial condition and vortex pairing in jet noise. J. Sound Vib. 117 (2), 289–311.
  • Brown & Roshko (1974) Brown, G.L. & Roshko, A. 1974 On density effects and large structure in turbulent mixing layers. J. Fluid Mech. 64 (4), 775–816.
  • Cho et al. (2018) Cho, M., Hwang, Y. & Choi, H. 2018 Scale interactions and spectral energy transfer in turbulent channel flow. J. Fluid Mech. 854, 474–504.
  • Cho et al. (1998) Cho, S.K., Yoo, J.Y. & Choi, H. 1998 Vortex pairing in an axisymmetric jet using two-frequency acoustic forcing at low to moderate strouhal numbers. Experiments in Fluids 25 (4), 305–315.
  • Crow & Champagne (1971) Crow, S.C. & Champagne, F.H. 1971 Orderly structure in jet turbulence. J. Fluid Mech. 48 (3), 547–591.
  • Danaila et al. (1999) Danaila, L., Anselmet, F., Zhou, T. & Antonia, R.A. 1999 A generalization of yaglom’s equation which accounts for the large-scale forcing in heated decaying turbulence. J. Fluid Mech. 391, 359–372.
  • Delville (1994) Delville, J. 1994 Characterization of the organization in shear layers via the proper orthogonal decomposition. Appl. Sci. Res. 53 (3), 263–281.
  • Fontaine et al. (2015) Fontaine, R.A., Elliott, G.S., Austin, J.M. & Freund, J.B. 2015 Very near-nozzle shear-layer turbulence and jet noise. J. Fluid Mech. 770, 27–51.
  • Glauser & George (1992) Glauser, M.N. & George, W.K. 1992 Application of multipoint measurements for flow characterization. Exp. Therm. and Fluid Sci. 5 (5), 617–632.
  • Glauser et al. (1987) Glauser, M.N., Leib, S.J. & George, W.K. 1987 Coherent structures in the axisymmetric turbulent jet mixing layer. In Turbulent Shear Flows 5, pp. 134–145. Springer.
  • Gomé et al. (2023) Gomé, S., Tuckerman, L.S. & Barkley, D. 2023 Patterns in transitional shear turbulence. part 1. energy transfer and mean-flow interaction. J. Fluid Mech. 964, A16.
  • Gudmundsson & Colonius (2011) Gudmundsson, K. & Colonius, T. 2011 Instability wave models for the near-field fluctuations of turbulent jets. J. Fluid Mech. 689, 97–128.
  • Hajj et al. (1992) Hajj, M.R., Miksad, R.W. & Powers, E.J. 1992 Subharmonic growth by parametric resonance. J. Fluid Mech. 236, 385–413.
  • Hajj et al. (1993) Hajj, M.R., Miksad, R.W. & Powers, E.J. 1993 Fundamental–subharmonic interaction: effect of phase relation. J. Fluid Mech. 256, 403–426.
  • Herring (1980) Herring, J.R. 1980 Theoretical calculations of turbulent bispectra. J. Fluid Mech. 97 (1), 193–204.
  • Hill (2001) Hill, R.J. 2001 Equations relating structure functions of all orders. J. Fluid Mech. 434, 379–388.
  • Hill Jr et al. (1976) Hill Jr, W.G., Jenkins, R.C. & Gilbert, B.L. 1976 Effects of the initial boundary-layer state on turbulent jet mixing. AIAA J. 14 (11), 1513–1514.
  • Ho & Huang (1982) Ho, C.M. & Huang, L.S. 1982 Subharmonics and vortex merging in mixing layers. J. Fluid Mech. 119, 443–473.
  • Husain & Hussain (1995) Husain, H.S. & Hussain, F. 1995 Experiments on subharmonic resonance in a shear layer. J. Fluid Mech. 304, 343–372.
  • Husain & Hussain (1979) Husain, Z.D. & Hussain, A.K.M.F. 1979 Axisymmetric mixing layer: influence of the initial and boundary conditions. AIAA J. 17 (1), 48–55.
  • Hussain & Zaman (1980) Hussain, A.K.M.F. & Zaman, K.B.M.Q. 1980 Vortex pairing in a circular jet under controlled excitation. part 2. coherent structure dynamics. J. Fluid Mech. 101 (3), 493–544.
  • Hussain & Zedan (1978a) Hussain, A.K.M.F. & Zedan, M.F. 1978a Effects of the initial condition on the axisymmetric free shear layer: Effect of the initial fluctuation level. Phys. Fluids 21 (9), 1475–1481.
  • Hussain & Zedan (1978b) Hussain, A.K.M.F. & Zedan, M.F. 1978b Effects of the initial condition on the axisymmetric free shear layer: Effects of the initial momentum thickness. Phys. Fluids 21 (7), 1100–1112.
  • Jin et al. (2021) Jin, B., Symon, S. & Illingworth, S.J. 2021 Energy transfer mechanisms and resolvent analysis in the cylinder wake. Phys. Rev. Fluids 6 (2), 024702.
  • Kelly (1967) Kelly, R.E. 1967 On the stability of an inviscid shear layer which is periodic in space and time. J. Fluid Mech. 27 (4), 657–689.
  • Kibens (1980) Kibens, V. 1980 Discrete noise spectrum generated by acoustically excited jet. AIAA Journal 18 (4), 434–441.
  • Kim & Choi (2009) Kim, J. & Choi, H. 2009 Large eddy simulation of a circular jet: effect of inflow conditions on the near field. J. Fluid Mech. 620, 383–411.
  • Kim & Powers (1979) Kim, Y.C. & Powers, E.J. 1979 Digital bispectral analysis and its applications to nonlinear wave interactions. IEEE Trans. Plasma Sci. 7 (2), 120–131.
  • Kinjangi & Foti (2023) Kinjangi, D.K. & Foti, D. 2023 Characterization of energy transfer and triadic interactions of coherent structures in turbulent wakes. J. Fluid Mech. 971, A7.
  • Lee & Moser (2019) Lee, M. & Moser, R.D. 2019 Spectral analysis of the budget equation in turbulent channel flows at high reynolds number. J. Fluid Mech. 860, 886–938.
  • Lesshafft & Huerre (2007) Lesshafft, L. & Huerre, P. 2007 Linear impulse response in hot round jets. Phys. Fluids 19 (2).
  • Lii et al. (1976) Lii, K.S., Rosenblatt, M. & Van Atta, C. 1976 Bispectral measurements in turbulence. J. Fluid Mech. 77 (1), 45–62.
  • Lumley (1967) Lumley, J.L. 1967 The structure of inhomogeneous turbulent flows. Atmospheric Turbulence and Radio Wave Propagation pp. 166–178.
  • Lumley (1970) Lumley, J.L. 1970 Stochastic Tools in Turbulence. Academic Press.
  • Maia et al. (2022) Maia, I.A., Jordan, P. & Cavalieri, A.V.G. 2022 Wave cancellation in jets with laminar and turbulent boundary layers: The effect of nonlinearity. Phys. Rev. Fluids 7 (3), 033903.
  • Maia et al. (2021) Maia, I.A., Jordan, P., Cavalieri, A.V.G., Martini, E., Sasaki, K. & Silvestre, F.J. 2021 Real-time reactive control of stochastic disturbances in forced turbulent jets. Phys. Rev. Fluids 6 (12), 123901.
  • Mankbadi (1985) Mankbadi, R.R. 1985 On the interaction between fundamental and subharmonic instability waves in a turbulent round jet. J. Fluid Mech. 160, 385–419.
  • Metcalfe et al. (1987) Metcalfe, R.W., Orszag, S.A., Brachet, M.E., Menon, S. & Riley, J.J. 1987 Secondary instability of a temporally growing mixing layer. J. Fluid Mech. 184, 207–243.
  • Meynart (1983) Meynart, R. 1983 Speckle velocimetry study of vortex pairing in a low-re unexcited jet. Phys. Fluids 26 (8), 2074–2079.
  • Michalke (1964) Michalke, A. 1964 On the inviscid instability of the hyperbolictangent velocity profile. J. Fluid Mech. 19 (4), 543–556.
  • Michalke (1965) Michalke, A. 1965 On spatially growing disturbances in an inviscid shear layer. J. Fluid Mech. 23 (3), 521–544.
  • Mizuno (2016) Mizuno, Y. 2016 Spectra of energy transport in turbulent channel flows for moderate reynolds numbers. J. Fluid Mech. 805, 171–187.
  • Monkewitz (1988) Monkewitz, P.A. 1988 Subharmonic resonance, pairing and shredding in the mixing layer. J. Fluid Mech. 188, 223–252.
  • Moser & Rogers (1993) Moser, R.D. & Rogers, M.M. 1993 The three-dimensional evolution of a plane mixing layer: pairing and transition to turbulence. J. Fluid Mech. 247, 275–320.
  • Nekkanti et al. (2023) Nekkanti, A., Nidhan, S., Schmidt, O.T. & Sarkar, S. 2023 Large-scale streaks in a turbulent bluff body wake. J. Fluid Mech. 974, A47.
  • Nekkanti & Schmidt (2021) Nekkanti, A. & Schmidt, O.T. 2021 Frequency–time analysis, low-rank reconstruction and denoising of turbulent flows using spod. J. Fluid Mech. 926, A26.
  • Nogueira et al. (2019) Nogueira, P.A.S., Cavalieri, A.V.G., Jordan, P. & Jaunet, V. 2019 Large-scale streaky structures in turbulent jets. J. Fluid Mech. 873, 211–237.
  • Paschereit et al. (1995) Paschereit, C.O., Wygnanski, I. & Fiedler, H.E. 1995 Experimental investigation of subharmonic resonance in an axisymmetric jet. J. Fluid Mech. 283, 365–407.
  • Pickering et al. (2020) Pickering, E., Rigas, G., Nogueira, P.A.S., Cavalieri, A.V.G., Schmidt, O.T. & Colonius, T. 2020 Lift-up, kelvin–helmholtz and orr mechanisms in turbulent jets. J. Fluid Mech. 896, A2.
  • Pier (2002) Pier, B. 2002 On the frequency selection of finite-amplitude vortex shedding in the cylinder wake. J. Fluid Mech. 458, 407–417.
  • Schmid (2010) Schmid, P.J. 2010 Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech. 656, 5–28.
  • Schmidt (2020) Schmidt, O.T. 2020 Bispectral mode decomposition of nonlinear flows. Nonlinear Dyn. 102 (4), 2479–2501.
  • Schmidt & Colonius (2020) Schmidt, O.T. & Colonius, T. 2020 Guide to spectral proper orthogonal decomposition. AIAA J. 58 (3), 1023–1033.
  • Schmidt et al. (2017) Schmidt, O.T., Towne, A., Colonius, T., Cavalieri, A.V.G., Jordan, P. & Brès, G.A. 2017 Wavepackets and trapped acoustic modes in a turbulent jet: coherent structure eduction and global stability. J. Fluid Mech. 825, 1153–1181.
  • Schmidt et al. (2018) Schmidt, O.T., Towne, A., Rigas, G., Colonius, T. & Brès, G.A. 2018 Spectral analysis of jet turbulence. J. Fluid Mech. 855, 953–982.
  • Suzuki & Colonius (2006) Suzuki, T. & Colonius, T. 2006 Instability waves in a subsonic round jet detected using a near-field phased microphone array. J. Fluid Mech. 565, 197–226.
  • Symon et al. (2021) Symon, S., Illingworth, S.J. & Marusic, I. 2021 Energy transfer in turbulent channel flows and implications for resolvent modelling. J. Fluid Mech. 911, A3.
  • Towne et al. (2018) 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, 821–867.
  • Von Karman & Howarth (1938) Von Karman, T. & Howarth, L. 1938 On the statistical theory of isotropic turbulence. Proceedings of the Royal Society of London. Series A-Mathematical and Physical Sciences 164 (917), 192–215.
  • Welch (1967) Welch, P. 1967 The use of fast fourier transform for the estimation of power spectra: a method based on time averaging over short, modified periodograms. IEEE Trans. Audio Electroacoust. 15 (2), 70–73.
  • Winant & Browand (1974) Winant, C.D. & Browand, F.K. 1974 Vortex pairing: the mechanism of turbulent mixing-layer growth at moderate reynolds number. J. Fluid Mech. 63 (2), 237–255.
  • Zaman (1985) Zaman, K.B.M.Q. 1985 Effect of initial condition on subsonic jet noise. AIAA J. 23 (9), 1370–1373.
  • Zaman (2012) Zaman, K.B.M.Q. 2012 Effect of initial boundary-layer state on subsonic jet noise. AIAA J. 50 (8), 1784–1795.
  • Zaman & Hussain (1980) Zaman, K.B.M.Q. & Hussain, A.K.M.F. 1980 Vortex pairing in a circular jet under controlled excitation. part 1. general jet response. J. Fluid Mech. 101 (3), 449–491.