1. Introduction
The flow over a step is of interest for a number of engineering applications, such as wind turbines, heat exchangers, combustion chambers, as well as transport vehicles. Its behaviour is dictated by the physics of separation, as it is characterized by two recirculation zones or bubbles, one at the base of the step and one on top of the step (as illustrated in figure 1). The dynamics of separated flows involve a wide range of frequencies, which are generally divided into three ranges: low frequencies, generally associated with ‘flapping’ or ‘breathing’ of the recirculation; medium frequencies, associated with large-scale vortex shedding downstream of the recirculation; and high frequencies, associated with instabilities in the developing shear layer and small-scale turbulence (Kiya & Sasaki Reference Kiya and Sasaki1983; Weiss, Mohammed-Taifour & Schwaab Reference Weiss, Mohammed-Taifour and Schwaab2015; Tenaud et al. Reference Tenaud, Podvin, Fraigneau and Daru2016).
Although the flow can be assumed to be statistically two dimensional if the width-to-height ratio of the step is large enough ($W/H > 6\unicode{x2013}10$ Martinuzzi & Tropea Reference Martinuzzi and Tropea1993; Sherry, Jacono & Sheridan Reference Sherry, Jacono and Sheridan2010), its fluctuations are inherently three dimensional (Largeau & Moriniere Reference Largeau and Moriniere2007). Hammond & Redekopp (Reference Hammond and Redekopp1998) showed that globally unstable dynamics can be expected in a recirculation bubble if the backflow velocity reaches 30 % of the incoming velocity. Lanzerstorfer & Kuhlmann (Reference Lanzerstorfer and Kuhlmann2012) carried out linear stability analysis of channel flows with various constriction ratios and identified a three-dimensional (3-D) global critical mode in the recirculation zone on top of the step, consisting of stationary streaks with a wavelength equal to three times the step height. The sensitivity of the flow to upstream conditions has been documented in experimental studies (Stuer, Gyr & Kinzelbach Reference Stuer, Gyr and Kinzelbach1999), as well as in numerical studies (Wilhelm, Hartel & Kleiser Reference Wilhelm, Hartel and Kleiser2003; Marino & Luchini Reference Marino and Luchini2009; Lanzerstorfer & Kuhlmann Reference Lanzerstorfer and Kuhlmann2012). A major finding of these studies is that convective instabilities developing upstream can influence the formation of instabilities at the step face.
The upstream region of the step has therefore been the focus of specific investigation (Pearson, Goulart & Ganapathisubramani Reference Pearson, Goulart and Ganapathisubramani2013; Graziani et al. Reference Graziani, Kerhervé, Martinuzzi and Keirsbulck2018). The existence of streaky structures fed by spanwise motions along the step face was supported by the experimental observations of Martinuzzi & Tropea (Reference Martinuzzi and Tropea1993), Stuer et al. (Reference Stuer, Gyr and Kinzelbach1999), Largeau & Moriniere (Reference Largeau and Moriniere2007) and Pearson et al. (Reference Pearson, Goulart and Ganapathisubramani2013). In the case of an immerged step, where the step height $H$ is smaller than the boundary layer thickness $\delta (\delta /H=1.47)$, Pearson et al. (Reference Pearson, Goulart and Ganapathisubramani2013) were able to relate ejections over the step with the convection of low-velocity regions originating in the upstream boundary layer, thereby establishing a connection between the upstream and downstream region. Graziani et al. (Reference Graziani, Kerhervé, Martinuzzi and Keirsbulck2018) examined the case of a protruding step ($\delta /H=0.49$) and found evidence that the downstream recirculation was at least partly driven by the low-frequency oscillations of the upstream recirculation, in that the enlargement of the upstream separation was associated with a contraction of the downstream separation. Evidence of an anti-correlation was also found by Fang et al. (Reference Fang, Tachie, Bergstrom, Yang and Wang2021) in the case of an immerged step ($\delta /H=2$), while a previous study by the same group (Fang & Tachie Reference Fang and Tachie2020) found a positive correlation when the step was placed in a rough turbulent boundary layer such that $\delta /H=6.5$.
These results confirm that the spatio-temporal characteristics of the recirculation zones depend on various parameters of the incoming boundary layer, such as the ratio $\delta /H$, the turbulence level and the Reynolds number based on the external velocity $U$ and the boundary layer thickness $\delta$. A discussion of the influence of these parameters can be found in Sherry et al. (Reference Sherry, Jacono and Sheridan2010). When the step is immersed in the boundary layer ($\delta /H > 1$) (Pearson et al. Reference Pearson, Goulart and Ganapathisubramani2013; Fang et al. Reference Fang, Tachie, Bergstrom, Yang and Wang2021), the flow on top of the step is then subject to strong mixing, which tends to reduce the size of the separation zone, but is also dependent on the velocity gradients and, therefore, the Reynolds number. As $\delta /H$ increases, the size of the downstream recirculation decreases (Largeau & Moriniere Reference Largeau and Moriniere2007). Moss & Baker (Reference Moss and Baker1980), Kiya & Sasaki (Reference Kiya and Sasaki1985) have shown that a higher turbulence intensity results in a higher growth rate of the roll-up vortices developing in the shear layer between the separation point and the reattachment zone. The higher entrainment rate associated with larger vortices leads to a reduction of the separation zone. Sherry et al. (Reference Sherry, Jacono and Sheridan2010) showed that for several different ratios $\delta /H$, two regimes could be identified, with a cutoff value of $Re_h=8500$ that appears to be independent of $\delta /H$: a low-Reynolds-number regime, where the recirculation length increases linearly with the Reynolds number, and a high-Reynolds-number regime, where it seems to be essentially independent from the Reynolds number, and which Sherry et al. (Reference Sherry, Jacono and Sheridan2010) attribute to the shear layer becoming turbulent at separation. In addition, the influence of the wall on the development of the shear layer is still an open question. McGuiness (Reference McGuiness1978) studied the internal separated flow at the inlet to a parallel pipe and found that the shear layer development was very similar to that of a mixing layer without the presence of the wall over most of the recirculation length, while wall effects apperated close to reattachment.
Different statistical tools have been used to understand the physics of the flow over the step, including cross-correlations (Largeau & Moriniere Reference Largeau and Moriniere2007), as well as quadrant analysis of the turbulent motion (Hattori & Nagano Reference Hattori and Nagano2010). Proper orthogonal decomposition (POD), a statistical technique (Lumley Reference Lumley1967) that extracts characteristic patterns from turbulent flows in an energetically optimal manner, has been applied to experimental measurements (Pearson et al. Reference Pearson, Goulart and Ganapathisubramani2013; Graziani et al. Reference Graziani, Kerhervé, Martinuzzi and Keirsbulck2018; Fang & Tachie Reference Fang and Tachie2020) as well as numerical data (Fang & Tachie Reference Fang and Tachie2020; Fang et al. Reference Fang, Tachie, Bergstrom, Yang and Wang2021). As far as we know, in experimental as well as numerical studies, POD has been only limited to plane sections in the case of the forward-facing step, despite the 3-D nature of the flow. However, application of POD to volumes has been shown to be well suited for the analysis and low-order model reduction of complex turbulent flows (see, for instance, Aubry et al. Reference Aubry, Holmes, Lumley and Stone1988; Podvin et al. Reference Podvin, Pellerin, Fraigneau, Bonnavion and Cadot2021).
In the present paper, a fully 3-D POD analysis of the flow is carried out for the flow over a step protruding from a turbulent boundary layer $\delta /H < 1$. Owing to spatial homogeneity of the geometry, POD modes in the spanwise direction are by construction Fourier modes, so that the decomposition is applied independently for each Fourier spanwise mode. In this framework, the approach is to construct a low-dimensional representation for the flow in order to better understand its underlying physics and to model its dynamics. The first step is therefore to identify the features of the most energetic patterns upstream and downstream of the step, which constitute the building blocks of the representation. The next step is to use this representation to investigate the connection between the upstream and the downstream regions. The goal is then to construct a reduced-order model that can capture the main flow dynamics. The paper is therefore organized as follows. After a description of the numerical set-up in § 2, we present a brief review of POD in § 3. The flow in the upstream region of the step is discussed in § 4. Section 5 provides a description of the dominant dynamics over the step. The connection between the two recirculation zones is examined in § 6. A model to predict the amplitudes of the most energetic fluctuations is presented in § 7. A conclusion is given in § 8.
2. Configuration
2.1. Set-up
The configuration, shown in figure 1, consists of a two-dimensional (2-D) forward-facing step of height $H$, which will be the reference length throughout the paper. The Reynolds number $Re$ based on the incoming velocity $U$ and step height is 8300. The streamwise, vertical and spanwise directions will be respectively referred to as $x,y,z$ and the velocity components as $u,v,w$ or, equivalently, $u_1,u_2,u_3$. The bottom end of the domain consists of an impermeable wall. A symmetry condition is imposed at the top end. No corrections for blockage effects are applied. Periodic boundary conditions are used in the spanwise direction.
The dimensions of the domain are $(L_x,L_y,L_z)=[37, 22, {\rm \pi}]H$. The step is located at the origin of the domain at a distance of $17H$ from the upstream end. A convective boundary condition is used at the outlet of the domain with a correction procedure to ensure the mass flow rate conservation. The domain length is long enough to prevent small perturbations generated by the outlet conditions from spreading into the region of interest for the study. Turbulent inlet boundary conditions were implemented using the synthetic eddy method of Pamies et al. (Reference Pamies, Weiss, Garnier, Deck and Sagaut2009) and the mean velocity profile was given by the numerical data of Spalart (Reference Spalart1988) (AGARD database) for a turbulent boundary layer with a zero pressure gradient at a Reynolds number based on the momentum thickness $Re_\theta = 300$. Validation of the synthetic procedure was carried out for a turbulent boundary layer in the absence of a step. It also made it possible to define the boundary layer thickness to step height ratio $\delta /H$, where $\delta$ was estimated as the flat plate boundary layer thickness at the same streamwise position as the step. It was found that $\delta /H \sim 0.8$, which means that the step is slightly protruding from the boundary layer. In all that follows, unless specified otherwise, all quantities will be expressed in terms of the step height $H$ and incoming velocity $U$.
2.2. Numerical method and implementation
The incompressible Navier–Stokes equations were solved with the in-house code SUNFLUIDH (Fraigneau Reference Fraigneau2024), which is based on a second-order finite volume formulation. The spatial discretisation of convective and viscous terms is carried out with a second-order centred scheme on a staggered Cartesian grid following a MAC approach (Harlow & Welch Reference Harlow and Welch1965). A second-order backward formulation is used for the time derivative, with a semi-implicit treatment of the viscous term in order to strengthen the numerical stability in regard to the time step. An incremental projection method is used to compute the pressure field (Goda Reference Goda1979; Guermond, Minev & Shen Reference Guermond, Minev and Shen2006). For each time step, a predicted velocity field is obtained from integration of the Navier–Stokes equations by means of the alternating-direction implicit method (Peaceman & Rachford Reference Peaceman and Rachford1955; Beam & Warming Reference Beam and Warming1976). The pressure field is then computed by solving Poisson's equation with a relaxed successive over-relaxation method coupled with a geometric multigrid method to improve the convergence of the solution. The solution of Poisson's equation allows us to update both the pressure field and the velocity field so that the latter has a divergence-free property. A complete description of the algorithm is presented in Faugaret et al. (Reference Faugaret, Duguet, Fraigneau and Martin-Witkowski2022). This code has already been used for a number of studies, including bluff-body flows (Podvin et al. Reference Podvin, Pellerin, Fraigneau, Bonnavion and Cadot2021).
The domain was divided into 200 subdomains ($5 \times 4 \times 4$ upstream of the step and $10 \times 3 \times 4$ downstream of the step), each of which had a grid size $(128 \times 192 \times 96)$. The cell sizes ranged from 0.01 to 0.027 in the streamwise direction $x$, 0.00013 to 0.075 in the wall-normal direction in order to better resolve the flow near the surfaces, near the step edge and the shear layer growth capping the recirculation bubble downstream of the step. The spanwise resolution was uniform and equal to 0.0082. Table 1 sums up the spatial resolution of the grid in the main regions of interest. The resolution is expressed in wall units based on conditions at the inlet ($Re_{\theta }=300$). The time step was constant and taken equal to $4.5 10^{-4}$ units, which yielded a Courant–Friedrichs–Lewy value of less than 0.4 over the range of integration. The simulation was run for 730 time units, which required about 900 000 CPU hours on the HPE SGI 8600 cluster at IDRIS (Institut du Développement et des Ressources en Informatique Scientifique). Data acquisition of 3-D flow fields was carried out once statistical convergence was established. The DNS results used in this study are part of the database https://datasetmeca.lisn.upsaclay.fr/.
2.3. Flow statistics
Sufficiently far away from the step, the flow corresponds to that of a flat plate boundary layer. A good agreement of the profiles with recent experimental results has been established in Larose et al. (Reference Larose, Kerhervé, Fraigneau, Podvin, Morton and Martinuzzi2024). The mean velocity profile and turbulent intensities at $x=-6.6$ are represented in figure 2 and agree well with the results of Spalart (Reference Spalart1988). The shape factor, defined as the ratio between the displacement and the momentum thicknesses $\delta _1/\theta$, is higher by 6 % (1.64 for a Reynolds number based on the momentum thickness of $Re_\theta =570$) than in the flat plate case (1.55, Erm & Joubert Reference Erm and Joubert1991), which suggests that the influence of the step begins to be felt at this distance.
The streamwise component of the time-averaged velocity $U= \langle u \rangle$, where $\langle {\cdot } \rangle$ represents a time average, is shown in figure 3(a). The flow experiences an adverse pressure gradient as it approaches the step. Two recirculation zones, one upstream and one downstream, are delimited by a black line corresponding to the contour $U=0$. The centre of the upstream recirculation zone is located at $x=-0.4$, with a separation point located at $x=-1.6$ and extends up to $y=0.55$. The recirculation length of the downstream zone is about $x_R= 4.3$, which is in good agreement with results from the literature (Moss & Baker Reference Moss and Baker1980; Largeau & Moriniere Reference Largeau and Moriniere2007). The height of the downstream recirculation is about $0.25$. The variance of each component of the fluctuations is shown in figure 3(c,d). The fluctuations are maximal in the shear layer developing at the edge of the step above the separation bubble. The streamwise component is dominant and the spanwise fluctuations are more intense than the wall-normal fluctuations.
The pressure coefficient on the step is shown in figure 4. Comparison with data from Farabee & Casarella (Reference Farabee and Casarella1986) and Hahn (Reference Hahn2008) corresponding to slightly immersed cases $\delta /H=2.4$ and $\delta /H=1.5$ shows a very good agreement upstream of the step, with similar maximum values in the downstream area. We observe however a variation in the size of the adverse pressure gradient region downstream of the step, which could be due to the difference between the experimental (immersed) set-ups and our numerical (slightly protruding) configuration. We note that better agreement is obtained for the ratio of 1.5 (Hahn Reference Hahn2008) that is closer to our case.
3. Proper orthogonal decomposition
Our analysis of velocity fluctuations will be based on POD. Proper orthogonal decomposition is a statistical technique (Lumley Reference Lumley1967) that decomposes any spatio-temporal field $\underline {q}$ defined over a domain $\varOmega$ into a sum of spatial modes $\underline {\varPhi }$, the amplitudes of which vary in time. The spatial modes $\underline {\varPhi }$ are the principal directions of the autocorrelation tensor, which is built from a collection of fields or snapshots. The modes are optimal in the sense that the reconstruction error between the original snapshots and any linear reconstruction based on $N$ modes is minimal for the POD modes.
Due to the statistical homogeneity in the spanwise direction ($z$), POD modes are Fourier modes in that direction (Holmes, Lumley & Berkooz Reference Holmes, Lumley and Berkooz1996). The decomposition is therefore applied in Fourier space in the spanwise direction and any field $\underline {q}$ can be written as
where ${\rm c.c.}$ represents the complex conjugate. Here $\underline {q}$ will be taken equal to the fluctuating velocity $\underline {q} = \underline {u} - \underline {U}$. The positive quantity $\sqrt {\lambda _k^n}$ represents the contribution of the mode $\underline {\varPhi }_k^n$ to the variance of the fluctuations. The spatial modes $\underline {\varPhi }_k^n$ and temporal amplitudes $a_k^n$ are complex for $k \neq 0$ and purely real for $k=0$. By construction, at each spanwise Fourier wavenumber, the spatial modes are orthogonal and the temporal amplitudes uncorrelated. They are also normalized so that $\int _{\varOmega } \underline {\varPhi }_k^n {\cdot } \underline {\varPhi }_k^{m*}\, {{\rm d}{\kern0.9pt}x}\, {{\rm d} y} = \delta _{nm}$ and $\langle a_k^n a_k^{m*} \rangle = \delta _{nm}$, where $*$ represents the complex conjugate. For each spanwise wavenumber $k$, the amplitudes $a_k^n(t)$ and modes $\underline {\varPhi }_k^n$ are obtained by applying the method of snapshots (Sirovich Reference Sirovich1987) to the temporal correlation matrix obtained from $N_s$ snapshots, i.e.
and solving the eigenproblem
where the $n$th column of the matrix $A$ is such that $A_k^{pn}= a_k^n(t_p)$, and a summation convention is used for the $p$ index. Here $\lambda _k^n$ represents the contribution of the $n$th POD mode to the $k$th Fourier mode of the correlation. Since $\lambda _{-k}^n= \lambda _k^n$ for $k > 0$, only the positive part of the spectrum will be represented. However, to account for the contribution of both negative and positive wavenumbers to the variance, we used a rescaling factor of 2 when representing the eigenvalues at non-zero wavenumbers.
The modes $\underline {\varPhi }_k^n(x,y)$ can be obtained from
Equation (3.3) shows that each vector of amplitudes $A_k^n$ and, therefore, each spatial mode $\underline {\varPhi }_{k}^n (x,y)$ is defined within an arbitrary phase.
We denote $\phi _k^{n,i}$ the $i$th velocity component of the $n$th POD mode associated with the $k$th Fourier wavenumber. The number of snapshots was taken to be 940 fields, separated by 0.5 convective time units. No significant change in the dominant modes was observed when only 600 fields were considered. Since the fluctuations are dominant downstream of the step, POD was first limited to the upstream region close to the step in order to extract local energetic motions there. We then applied POD to the velocity in a volume comprising the upstream region and a region downstream of the step. The labels $up$ and $ds$ will be used to refer respectively to the upstream and the downstream region to remove possible ambiguity, but will be omitted otherwise.
4. Upstream dynamics
4.1. Dominant motions
A first description of the flow is given by the spatial distribution of the lowest and most energetic spanwise Fourier modes for each velocity component, which is represented in figure 5 for the region directly upstream of the step. The vertical velocity is most important for the lower wavenumbers $k=1$ and $k=2$ with a maximum at the step mid-height. The streamwise component is most important for modes $k=3$ and $k=4$ with a maximum close to the step corner at $x=-0.3$ and $y=0.7$. It is also associated with strong spanwise components along the step face close to the edge.
Proper orthogonal decomposition was applied to the fluctuating velocity field in the upstream region $\varOmega _x^{up} \times \varOmega _y^{up} \times \varOmega _z^{up} = [-6,0] \times [0, 1.5] \times [0, {\rm \pi}]$. The POD eigenvalues $\lambda _k^{n,up}$ are shown in figure 6 for the first spanwise wavenumbers. As expected from the energy spectra in figure 5, the largest eigenvalues $\lambda _k^{1,up}$ correspond to the wavenumbers $k=2, 3, 4$, and are about equal. For these wavenumbers, the first eigenvalue $\lambda _k^n$ is more than twice as large as the next one, which indicates the importance of 3-D coherent motions associated with the first mode. The eigenvalues $\lambda ^n$ decay as $n^{-1}$ for higher-order modes.
The eigenvalue of the largest fluctuating spanwise invariant mode $\lambda _0^{1,up}$ is about an order of magnitude less than the dominant ones. The streamwise and the normal velocity components $\phi ^{1,1(up)}_0$ and $\phi ^{1,2(up)}_0$ are represented in figure 7 on which we also reported the streamline $\langle U\rangle =0$ delimiting the recirculation zone. The step height is delimited with a horizontal dashed line. The mode is characterized by a strong horizontal motion close to the edge of the recirculation in the region $y < 0.5$, which coincides with a vertical motion of nearly equivalent intensity along the step face above the recirculation in the region $x > -0.5$. A weaker horizontal (respectively vertical) motion of the opposite sign is observed in the region $y > 0.5$ (respectively $x < -0.5$). The mode $\underline {\varPhi }^{1(up)}_0$ therefore represents momentum transfer from the streamwise to the wall-normal direction: it constitutes a spanwise invariant vortex that oscillates non-harmonically back and forth.
The most energetic mode corresponds to $k=3$ or $\lambda _z \sim 1.03$ and is shown in figure 8. Other dominant wavenumbers $k=2$ and $k=4$ were found to have generally similar features. The main wavelength $\lambda _z$ is in good agreement with Fang et al.'s (Reference Fang, Tachie, Bergstrom, Yang and Wang2021) results, who found a dominant wavelength of 1.32. The streamwise component of the mode $\underline {\varPhi }_3^{1(up)}$ captures 70 % of its energy and is of constant sign over a region connecting the upstream boundary layer with the recirculation over the step. The mode is characterized by much weaker vertical fluctuations (5 % of the energy), which are negatively correlated with the streamwise component over the step, so that the corresponding pattern is consistent with ejections of low-speed fluid. Although the spanwise motion represents only 25 % of the total mode energy, it is characterized by intense fluctuations (twice as large as in the streamwise direction) in a very thin layer, which is less than $0.1$ thick and extends from the top of the recirculation to over the top of the step. A weaker fluctuation of the opposite sign is also noticeable upstream of the step. The presence of both positive and negative fluctuations could be consistent with vortical motion aligned with the vertical axis along the step face.
The spectra of the amplitudes $a_k^{1,up}$ are represented in figure 9. Since the amplitudes are complex for non-zero Fourier modes, their Fourier transform is generally not symmetric in frequency space $\hat {a}_{k}^n(-f) \neq \hat {a}_{k}^{*n}(f)$. The presence of a significant asymmetry in the spectrum can be associated with a persistent drift in physical space. In the present case the spectrum appears to be nearly symmetric, which does not make it possible to identify specific convective motions. The dominant part of the spectra for the different wavenumbers is located in the same low-frequency range $[0,0.05]$, which suggests that their dynamics are connected.
4.2. Characteristic structure
The label $up$ will be implicit throughout the paragraph. It would be helpful for physical interpretation to educe a characteristic spatial or coherent structure associated with the dominant POD modes, that could be compared with experimental observations such as particle tracking velocimetry visualizations (Stuer Reference Stuer1999) or numerical approaches such as stochastic estimation (Fang et al. Reference Fang, Tachie, Bergstrom, Yang and Wang2021). We note that reconstructing a physical structure from Fourier POD modes is different from applying POD directly in the physical space, since the temporal amplitude of each mode $a_k^n$ also depends on the wavenumber. This means that each term $\sqrt {\lambda _k^n} a_k^n(t) \underline {\varPhi }_k^n(x,y)$, which we rewrite as a product $a_k^n(t) \underline {\tilde {\varPhi }}_k^n(x,y)$, where $\underline {\tilde {\varPhi }}_k^n=\sqrt {\lambda _k^n} \underline {\varPhi }_k^n$ corresponds to a convolution in physical space. As a result, the expansion
If the average energy of the $k$th Fourier component of $a^1$ is taken to be $1$, it follows that the $k$th Fourier component of $\underline {\varPhi }^{1,C}$ is given by the POD mode $\lambda _k^n|\underline {\varPhi }_k^1(x,y)|^2$. However, this is not enough to reconstruct $\underline {\varPhi }^{1,C}$ in physical space as the phases of the Fourier components are arbitrary. Additional assumptions are therefore required in order to determine $\underline {\varPhi }^{n,C}$ in physical space. A discussion of this issue can be found in Holmes et al. (Reference Holmes, Lumley and Berkooz1996), who proposed a solution based on the determination of the bi-spectrum. Moin & Moser (Reference Moin and Moser1989) applied this method to the case of a turbulent channel flow in order to educe a ’characteristic eddy’ from 2-D Fourier mode components, but also explored alternative approaches based on eigenfunction continuity with respect to the wavenumber as well as spatial compactness.
Following their lead, we choose to guarantee spatial compactness of the structure, which is easy to do in one-dimensional Fourier space. As shown by Moin & Moser (Reference Moin and Moser1989), this means minimizing the phase difference between the modes – the Fourier coefficients should all be real and positive for the sum of the modes to reach its maximum. This will define a structure in physical space but will leave its absolute position in the spanwise direction arbitrary since due to homogeneity only the phase differences of the modes can be determined. We therefore computed for each mode $\underline {\varPhi }_k^1$ an average phase $\alpha _k^1$ over the domain and over each component, using
where $\phi _k^{1,i}$ is the $i$th component of the velocity POD mode and $Arg$ is its argument. We recall that $\iint _{\varOmega _x^{up}\times \varOmega _y^{up}} \phi _k^{1,i}(x,y) \phi _k^{1,i*}(x,y)\, {{\rm d}{\kern0.9pt}x}\, {{\rm d} y} = 1.$ A characteristic structure was then reconstructed from the dominant wavenumbers $k \le 4$ as
where the contribution $\underline {\varPhi }_k^1$ of each wavenumber $k$ to the sum was weighted by the characteristic intensity $\sqrt {\lambda _k^1}$. This determines the amplitude of the structure.
The streamlines of the field corresponding to the superposition of the mean flow and the reconstructed structure $\underline {\varPhi }^{1,C}$ are represented in figure 10 for three different sets of original locations (both forward and backward time integration are performed). Each set of locations consists of a horizontal line along the step face: one close to the inward corner $(x=-0.05, y=0.1)$, one just above the mean recirculation $(x=-0.05, y=0.6)$ and one just above the step edge $(x=-0.05, y=0.9)$. Each type of streamline is associated with the same well-delimited spanwise regions of fluid upstream, illustrating the compactness of the representation. Vizualization of the streaky pattern created by the structure is provided by representing the contour lines of the streamwise velocity in the vertical plane at the step corner $x=0$ in each subplot. Different flow patterns associated with the characteristic structure $\underline {\varPhi }^{1,C}$ can thus be identified.
Near the bottom wall (figure 10a), fluid from the incoming boundary layer gets trapped within the recirculation zone, but then leaves it after a few spins and is ejected over the step. A good agreement is found with experimental (Stuer et al. Reference Stuer, Gyr and Kinzelbach1999) and numerical observations (Fang et al. Reference Fang, Tachie, Bergstrom, Yang and Wang2021, figure 14a). The twisting and ejecting motions from within the recirculation zone are associated with open recirculation streamlines, which cannot be captured by a 2-D representation. Comparison of the incoming streamlines with the lifted-up streamwise velocity contours in the vertical plane at the step corner indicate that the ejected fluid is slower than the mean flow.
At mid-height (figure 10b), flow separation takes place, due to the strong spanwise motions close to the wall. Above the recirculation zone, the flow is collected into different low-speed streaks, which are then ejected over the step. Note that the location of these ejections generally coincides with the fluid leaving the recirculation. Just underneath the step at $y=0.9$ (figure 10c), most of the flow streamlines spread out from upstream locations further away from the wall and are less likely to get trapped in the recirculation zone. As a result, they reach the wall with a higher velocity (as evidenced by the streamwise contour lines in the vertical plane) and without being shifted in the spanwise direction.
4.3. Centrifugal instability
We now investigate the origin of this structure. Due to the presence of the step, the mean flow is characterized by a strong curvature (see figure 3), so that a possible formation mechanism could be a centrifugal instability. A necessary condition for the existence of this instability is given by an inviscid criterion (Drazin & Reid Reference Drazin and Reid1982) based on the Rayleigh discriminant $\varPsi$ that, following Beaudoin et al.'s (Reference Beaudoin, Cadot, Aider and Wesfreid2004) approach for a backward-facing step, can be defined as
where $U_m^2=U^2+V^2$, $\varOmega _z=\partial V/\partial x - \partial U /\partial y$ and $R$ is the local radius of curvature obtained from Sipp & Jacquin (Reference Sipp and Jacquin2000) as
where $a_x$ and $a_y$ correspond to the components of the convective acceleration $\underline{U}.\boldsymbol {\nabla } U$. When the Rayleigh discriminant is negative, the flow is potentially unstable to a centrifugal instability. The region where the Rayleigh discriminant is negative with a large modulus is represented in figure 11. It is characterized by two local minima corresponding to the respective local maxima of the streamwise and spanwise velocity fluctuations (see figure 5).
However, for instability to occur, curvature effects need to overcome the effect of viscosity, a relative measure of which is given by the Görtler number
where in the classical laminar Görtler problem (Saric Reference Saric1994), $\delta$ corresponds to the boundary layer thickness and $U_{fs}$ is the free-stream velocity (here $U_{fs}=1$). Note that although we used a description based on a Görtler number, we did not find any clear evidence of Görtler vortices in the flow. In the present case, the flow is (weakly) turbulent and, as proposed by Tani (Reference Tani1962), the stability characteristics can be determined in a turbulent case if the viscosity is replaced by the turbulent eddy viscosity $\nu _t$, for which we can use Clauser's law $\nu _t \sim 0.018 U \delta _1$. Following Dagaut et al. (Reference Dagaut, Negretti, Balarac and Brun2021), we computed a turbulent Görtler number
where the displacement thickness $\delta _1$ was estimated at the position of the step in the flat plate turbulent boundary layer.
Figure 11 shows that the upper region close to the step edge is characterized by both a high Görtler number and strongly negative values of the Rayleigh discriminant. Comparison of figure 11 with figure 5 suggests that the region around $x=-0.2$, $y=0.8$, corresponding to the maximum streamwise fluctuation intensity, could be a good candidate for the origin of the instability. An interesting parallel can be drawn with Brès & Colonius's (Reference Brès and Colonius2008) study of the flow over a cavity. Using 3-D linear stability analysis and direct numerical simulation (DNS), they provided evidence of a centrifugal instability mechanism with a spanwise wavelength of the order of the cavity depth for shallow cavities, which is comparable with the dominant wavelengths found upstream of the step.
5. Full domain
Proper orthogonal decomposition was then applied to the domain consisting of both the upstream domain $\varOmega _x^{up} \times \varOmega _y^{up} \times \varOmega _z^{up}= [-6, 0] \times [0, 1.5] \times [0, {\rm \pi}]$ and the downstream region $\varOmega _x^{ds} \times \varOmega _y^{ds} \times \varOmega _z^{ds}= [0, 7] \times [1, 7] \times [0, {\rm \pi}]$. Figure 12 represents the dominant POD eigenvalues $\lambda _k^n$ for the velocity for the first spanwise wavenumbers. As mentioned in § 3, the contribution of the negative wavenumbers was lumped with that of the positive ones, so that a factor $\alpha _k$ of 2 was used in the representation of the non-zero spanwise wavenumbers. Unlike the upstream case, the most energetic wavenumbers correspond to the lowest ones, i.e. $k=0$ and $k=1$, and several eigenmodes make nearly equally important contributions to the energy at each spanwise wavenumber, which indicates a relatively complex organization of the flow. The discussion in this section will be focused on the wavenumbers $k=0$ and $k=1$.
5.1. Zero spanwise modes
Figure 12 shows that the fluctuating velocity eigenvalues for the spanwise invariant modes $k=0$ are organized in pairs, of the form $2n-1, 2n$ for $n \ge 1$, particularly for the first three couples of modes. The corresponding velocity eigenmodes are represented in figure 13. One can see that they correspond to 2-D, roll-like structures shed from the downstream recirculation bubble. The streamwise component of the modes is important at the location of reattachment, which suggests that the modes could also be associated with the breathing or flapping motion of the separation bubble.
The spectra of the corresponding amplitudes are represented in figure 14. One can see that the range of frequencies covered by the spanwise-averaged velocity is the same for each pair $n$. As we will see below, the organization of the modes into pairs is consistent with convective motion. For the first pair of eigenmodes, two frequencies, one around 0.10 and one around 0.13, can be identified. These values are consistent with the characteristic value of $f x_R/U=0.6$ associated with vortex shedding (Kiya & Sasaki Reference Kiya and Sasaki1983) since we have $x_R=4.3$. For the higher-order pairs, the frequency at 0.10 can still be identified, but the dominant lower-frequency part of the spectrum is characterized by low frequencies while the high-frequency part shifts towards higher frequencies. A peak is observed around $0.03$ in the low-frequency range, which is in good agreement with the non-dimensional flapping frequency of $f x_R/U =0.12$ identified by Kiya & Sasaki (Reference Kiya and Sasaki1983).
5.2. Dominant spanwise modes
The dominant wavenumber $k=1$ corresponding to a wavelength $\lambda _z$ of ${\rm \pi}$ is in good agreement with the critical mode of $3$ identified by Lanzerstorfer & Kuhlmann (Reference Lanzerstorfer and Kuhlmann2012) and Wilhelm et al. (Reference Wilhelm, Hartel and Kleiser2003). The first six modes corresponding to this wavenumber are represented in figure 15. The spatial features of the modes are markedly different from those of the 2-D (spanwise invariant) modes: the signature of the shear layer extending above the separation bubble is clear, the modes extend within the recirculation zone and are characterized by a strong spanwise component, which is larger than the vertical component. The modes have different spatial scales: the first three modes are relatively large scale, while modes 4 and 5 are characterized by smaller scales, and mode 6 is characterized by a mixture of large and small scales.
Figure 16 represents the temporal spectrum of the corresponding amplitudes $a_1^n$. Unlike the upstream region, the modes are characterized by a strong dissymmetry, which is consistent with convective motion, as will be seen in more detail below. The first two modes are characterized by frequencies in the range 0.05–0.15 (in absolute value) with a peak slightly below 0.1, which is also observed for mode 3. However, mode 3 also contains energy in the range $[0.1, 0.25]$, which is also important for mode 4. Mode 5 is characterized by a narrow range of frequencies around 0.22, while mode 6 contains a wide range of frequencies in the range $[0.05, 0.3]$ with a low-frequency peak at 0.05. The overlap in frequency space supports the idea that the modes are physically connected with each other.
5.3. Convection velocity
Comparison of figures 15 and 16 suggests that large scales tend to be associated with low frequencies. However, as shown by Buxton, de Kat & Ganapathisubramani (Reference Buxton, de Kat and Ganapathisubramani2013) in the case of a mixing layer, different scales can be associated with different convection velocities. Convection velocities are typically determined from correlation-based methods (Wills Reference Wills1964). Here, for the mode $k=1$, we use the spatio-temporal splitting induced by POD to infer a convection velocity $C_n$ for each POD mode $n$ as follows.
(i) We first extract for each amplitude $a_1^n$ a characteristic frequency $f_{1}^n$ such that
(5.1)\begin{equation} f_{1}^n= \frac{\displaystyle \sum_{f} |\hat{a}_{1}^n(f)|^2 f }{\displaystyle \sum_f |\hat{a}_{1}^n(f)|^2 }, \end{equation}where $f$ spans the space of positive and negative discrete frequencies.(ii) We compute the streamwise Fourier components $\hat {\underline {\varPhi }}_{k}^n(l,y)$ of the spatial mode $\underline {\varPhi }_1^n(x,y)$. It is defined as $\hat {\underline {\varPhi }}_{k}^n(l,y)= \int _{\varOmega _x^{ds}} \underline {\varPhi }_k^n(x,y) \exp (-2 {\rm i} {\rm \pi}l x/L_{\varOmega _x}^{ds})\, {{\rm d}{\kern0.9pt}x}$, where $l$ is the streamwise wavenumber and $\varOmega _x^{ds}=[0,L_{\varOmega _x}^{ds}]$ is the streamwise domain over which the transform is taken. We then extract a characteristic wavenumber $l_m^n$ from
(5.2)\begin{equation} l_{1}^n= \frac{\displaystyle \sum_{l} \int_1^{y_f} |\hat{\underline{\varPhi}}_{1}^n(l,y)|^2 l\, {{\rm d} y}}{\displaystyle \sum_l \int_1^{y_f} |\hat{\underline{\varPhi}}_{1}^n(l,y)|^2\, {{\rm d} y}}, \end{equation}where $l$ spans the space of streamwise wavenumbers. The streamwise Fourier transform was applied to the downstream region $\varOmega _x^{ds}=[0,7]$ over a vertical region $\varOmega _y^{ds}=[1,7]$. End effects due to the non-periodicity of the domain were not found to affect significantly the value of $l_{1}^n$.(iii) The convection velocity for each POD mode is then computed as
(5.3)\begin{equation} C_1^n = \frac{f_1^n}{l_{1}^n}. \end{equation}
The method described above can be adapted to extract a convection velocity for each pair of spanwise invariant modes, by defining complex modes of the form $\underline {\varPsi }_0^n= \underline {\varPhi }_0^{2n-1} + i \underline {\varPhi }_0^{2n}$, associated with complex amplitudes $z_0^n=a_0^{2n-1} - i a_0^{2n}$, so that
For the $n$th pair of POD modes $(a_0^{2n-1}, a_0^{2n})$, one then defines a characteristic frequency $f_0^n$ as
and a characteristic streamwise wavenumber $l_0^n$ as
The convection velocity for the $n$th pair is then given by $C_0^n=f_0^n/l_0^n$.
Figure 17 shows the convection velocity for the most energetic POD modes for $k=0$ (a) and $k=1$ (b). For almost all modes, the convection velocity is between 0.5 and 0.7, which is in excellent agreement with results from Largeau & Moriniere (Reference Largeau and Moriniere2007) (0.5–0.62), Camussi et al. (Reference Camussi, Felli, Pereira, Aloisio and Marco2008) (0.4–0.7), Hoarau et al. (Reference Hoarau, Borée, Laumonier and Gervais2006) (0.7), Kiya & Sasaki (Reference Kiya and Sasaki1983) (0.5–0.7). The exception is mode $n=5$ at $k=1$, for which the convection velocity is closer to 1, which suggests that the convective motion there is driven by the external flow. This is consistent with the higher frequencies observed in figure 16, as well as the spatial localization of the mode, which appears to take large values in the external part of the shear layer, further away from the recirculation and from the wall, as evidenced by $\phi _1^{5,1}$ in figure 15.
6. Connection between the recirculation zones
We now determine how the dominant spatio-temporal features of the regions upstream and downstream of the step are related to their respective recirculations. We then examine whether the two recirculations can be connected with each other.
Following Pearson et al. (Reference Pearson, Goulart and Ganapathisubramani2013), Fang et al. (Reference Fang, Tachie, Bergstrom, Yang and Wang2021), we define reverse flow regions using the indicator
For each spanwise position $z$ and instant $t$, we define upstream and downstream areas of backflow as
and
The limits of the integration domains were chosen large enough to enclose the recirculation zones. The dynamics of the upstream and downstream recirculations can thus be monitored through the spatio-temporal variations of the corresponding backflow regions. The volume of these regions $V^{up}$ and $V^{ds}$ can be obtained by integrating the areas over the spanwise dimension:
and
The time evolution of the backflow volumes $V^{up}$ and $V^{ds}$ is represented in figure 18(a). The maximum cross-correlation between the upstream and downstream volumes, shown in figure 18b), is of the order of 0.15 for a time delay of 4 units, which does not suggest a clear connection. We note that generally due to experimental constraints, discussions of the areas of reverse flow in the literature (Pearson et al. Reference Pearson, Goulart and Ganapathisubramani2013; Graziani et al. Reference Graziani, Kerhervé, Martinuzzi and Keirsbulck2018; Fang et al. Reference Fang, Tachie, Bergstrom, Yang and Wang2021) are generally focused on the 2-D areas $A$ corresponding to the same longitudinal plane.
The spatio-temporal characteristics of the areas of reverse flow $A^{up}$ and $A^{ds}$ are shown in spectral space in figure 19. The dominant wavelengths are found to be respectively about $1$ and $3$ for the upstream and downstream regions and the dominant frequencies were in the range $[0.01, 0.1]$. This is in good agreement with the results of Fang et al. (Reference Fang, Tachie, Bergstrom, Yang and Wang2021), who computed similar spectra for an embedded step ($\delta /H > 1$) and identified characteristic wavelengths of $1.32$ and $2.68$ for respectively the upstream and downstream regions. They also found frequencies in the range $[0.03-0.1]$ for both recirculations.
6.1. Global dynamics
We first determine how the spatio-temporal modulations of the backflow regions are connected with the dominant POD modes identified in each region. For the upstream region, figure 20 shows the cross-correlation between the time variations of the volume $V^{up}$ and of the dominant spanwise-averaged fluctuating mode $a_0^{1,up}$. The correlation coefficient between the two curves is $-0.7$ and is maximal in absolute value for a zero time delay. The negative correlation indicates that the recirculation zone grows when the amplitude of the mode decreases, as discussed in § 4. This high correlation is consistent with the relative importance at a zero spanwise wavenumber of the dominant eigenvalue $\lambda _0^{1,up}$ (see figure 6), and shows that the variations of the backflow volume reflect those of the most energetic 2-D fluctuation.
As was also seen in § 4, a large part of the fluctuation energy is associated with non-zero spanwise wavenumbers. Figure 20 compares for each spanwise wavenumber $k=1$ to $5$, the temporal cross-correlation of the Fourier components of the area of reverse flow $|\hat {A}_{k}^{up}|$ with those of the dominant fluctuations $|a_k^{1,up}|^2$. The maximum correlation is relatively high (in the range 0.5–0.6) for an optimal time delay that is negative and less than one time unit for $k>1$. The physical significance of this time delay is not clear, especially as correlation values remain high at zero time delay. The high correlation indicates that the spatial variations of the upstream recirculation are strongly connected with the most energetic fluctuations in the upstream zone, which confirms the relevance of the characteristic structure identified in § 4.
In the downstream region, as more eigenmodes contribute to the motion at the dominant wavenumbers, the influence of the velocity fluctuations was examined using aggregated contributions $q_k=\sum _{n=1}^{N} E_k^n$, with $E_k^n= \lambda _k^n |a_k^n|^2$ and $N=10$ (which represents about 50 % of the kinetic energy at this wavenumber). The cross-correlation between these aggregated contributions and the downstream backflow volume $V^{ds}$ is shown in figure 21. No significant correlation was observed for the spanwise invariant modes, which can be expected in view of their low contribution to the fluctuations in the downstream recirculation. In contrast, a significant negative correlation ($-0.49$) was obtained for the dominant first wavenumber. The maximal anti-correlation reached $-0.7$ when higher spanwise wavenumbers were included (up to $k=9$), which confirms the importance of the lowest mode $k=1$. The optimal time delays were small, i.e. of the order of 1–2 time units, and negative. Although it may be difficult to interpret these time delays, the fact that they are negative is consistent with the idea that the downstream separation zone is influenced by the development of the shear layer above it as well as the dominant POD modes at $k=1$ that characterize it.
We also computed various cross-correlations between the upstream fluctuations and the downstream backflow area (not shown). No correlation larger than 0.25 in absolute value was obtained. This points to a significant reorganization of the flow past the edge of the step, which we attempt to track using local indicators in the next section.
6.2. Local dynamics at the edge
We now discuss the connection between the upstream velocity fluctuations and the downstream recirculation with a focus on the dynamics at the edge of the step. Figure 22 shows, for two different instants $t_{min}$ and $t_{max}$, a horizontal and a vertical section of the streamwise velocity field respectively located upstream of the step at a height of $y=0.9$ and directly downstream of the edge at $x=0.02$ (the sections are represented in figure 23a). The two instants correspond to the times $t_{min}$ and $t_{max}$ for which the downstream backflow volume $V^{ds}$ is respectively minimal and maximal. For both times, the streamwise velocity is organized into a similar streaky pattern that persists from the horizontal to the vertical plane. However, the streaks are more intense when the volume of reverse flow is maximal: the plane-averaged backflow in the horizontal plane at $t_{max}$ is larger by 70 %, and its maximum value is more than twice as high than at $t_{min}$, while the vertical extent of the most intense streaks in the vertical plane at $t_{max}$ is higher, and can reach above the recirculation height (0.25). This suggests that the downstream separation zone is influenced by the incoming flow and, in particular, the streaks associated with ejections of low-speed fluid at the step edge.
A quantity of interest to characterize instantaneous upstream dynamics is therefore the spanwise-averaged flux of streamwise velocity that crosses the plane $y=1$ at the edge, a measure of which is given by the momentum deficit
A local characteristic time scale $\tau ^{up}$ for the ejection process can be built from the velocity scale $U_e=M_e^{1/2}$ and a characteristic distance $d^{up}$, which should be a fraction of the step height and can be tentatively estimated as $d^{up} \approx 0.2-0.4$, based on the location of the fluctuation peaks. This yields
Downstream of the step, the development of perturbations in the shear layer should occur on a time scale $\tau ^{ds}$ that varies like $1/\delta _{\omega }$, where $\delta _{\omega }$ is a characteristic length scale. From Ho & Huerre (Reference Ho and Huerre1984), the initial maximal spatial growth rate of the shear layer instability is of the order of $0.12 / \delta _{S}$ where $\delta _{S}$ is the shear layer thickness, so that $\delta _{\omega } \sim C^{ds} \delta _{S}$ with $C^{ds}=1/0.12 \approx 8$. A measure of the shear layer thickness $\delta _{S}$ can be obtained from the inverse of the largest positive eigenvalue $\lambda _S$ of the mean strain rate $\frac {1}{2}(\boldsymbol {\nabla } \underline {U} + \boldsymbol {\nabla } \underline {U}^T)$, which was used by Fang et al. (Reference Fang, Tachie, Bergstrom, Yang and Wang2021) to track the position of the shear layer. In a similar fashion, we define an instantaneous shear layer thickness based on the maximal value of the largest positive eigenvalue of the spanwise-averaged strain rate in the vertical plane directly downstream of the edge:
The time-averaged location of the maximum corresponds to a vertical distance of $y=0.02$ above the edge. The local time scale at the edge of the shear layer $\tau ^{ds}$ can therefore be estimated as
Figure 23 compares the time evolution of the characteristic velocity for the ejection process $U_e=M_e^{1/2}$ with the shear layer time scale $\lambda _S^{max}$. Both signals are almost perfectly correlated with a correlation coefficient of $0.95$. We point out that we computed several other correlation coefficients based on the spanwise-averaged velocity, such as that between the wall shear and $M_e$, as well as that between $\lambda _S^{max}$ and the velocity modulus. All coefficients were significantly inferior to the 0.95 value, which means that the match does not simply reflect the fact that the two signals $M_e$ and $\lambda _S^{max}$ are based on the spanwise-averaged velocity measured at two nearby locations. In addition, both signals have nearly the same root mean square to mean value, which suggests that they are directly proportional to each other, with a ratio $\langle \lambda _S^{max}\rangle /\langle M_e^{1/2}\rangle$ of about 60. This leads to a ratio for $\tau ^{up}/\tau ^{ds}$ of about $60 d^{up}/C^{ds} \approx 2.25$, which is sufficiently close to 1 for $\tau ^{up}$ and $\tau ^{ds}$ to be considered of the same order, given the crudeness of our estimates. This suggests that the shear layer dynamics instantaneously adapt to the upstream ejections, so that a decrease in ejections will immediately trigger a faster development of the shear layer perturbations. A direct connection is therefore established between the momentum deficit associated with the ejections and the strength of the shear layer downstream of the step.
To get a better understanding of this connection, we define an extended correlation coefficient to measure the cross-correlation between the complex Fourier amplitudes of the momentum deficit $-\hat {M}_{e,k}$ and those of the backflow areas $\hat {A}_{k}^{\alpha }$ with $\alpha \in \{up, ds\}$. Since $M_e$ is a small-scale, highly intermittent quantity, the signals were time filtered using a moving average of $5$ convective time units corresponding to less than one characteristic shedding time. The extended correlation coefficient $\bar {C}$ is defined as
where $\bar {q}$ represents the time-filtered version of $q$, $\langle q\rangle$ its time average and $|\bar {q}|^{rms}$ the standard deviation of its filtered modulus. Results are presented in figure 24 for the upstream recirculation (a) and the downstream recirculation (b). For Fourier modes $1 \le k \le 5$, the correlation between $-\hat {M}_{e,k}$ and the upstream recirculation is strong – larger than $0.6$ – with optimal time delays inferior to one time unit (a slightly larger time delay of 3 units is observed for the less energetic mode $k=1$). This shows that the modulations of the recirculation at the bottom of the step are directly connected with the ejections at the edge, and that both match the variations of the characteristic structure identified in § 4.
It is interesting to contrast these results with those obtained for the downstream recirculation. Only the spanwise-averaged area (or volume) of reverse flow $V^{ds}$ is well correlated – with a correlation coefficient larger than $0.6$ – with the average momentum deficit $-\hat {M}_{e,0}$, indicating that an increase in momentum deficit at the edge of the step results in a larger recirculation downstream. The corresponding optimal time delay was positive and of the order of $2$ convective time units. Interestingly, no such correlation was detected for higher modes, which confirms a fast reorganization of the flow downstream of the step (and was also substantiated by examination of instantaneous fields).
We therefore propose the following description, summarized in figure 25: low-speed fluid is entrained into the upstream recirculation bubble and ejected with a strong spanwise motion into low-velocity streaks over the edge of the step. The global momentum deficit due to these local ejections impacts the development of the shear layer and the downstream recirculation, although the fluctuations are quickly reorganized. A higher ejection rate (low $M_e$) leads to a weaker development of the shear layer and a larger recirculation zone. The flow in the shear layer is characterized by a low spanwise modulation, but becomes more two dimensional downstream of the reattachment.
7. A model for the dominant POD velocity amplitudes
7.1. The POD-based models
In this section we propose a model to reconstruct the dynamics of the most energetic POD modes, which are characterized by a range of low frequencies. The model is based on a simplified version of the POD-based models obtained by projection of the Navier–Stokes equations onto the basis of the spatial POD modes. The construction of dynamical systems to predict the amplitudes of the POD modes was first proposed by Aubry et al. (Reference Aubry, Holmes, Lumley and Stone1988) and a thorough discussion can be found in Holmes et al. (Reference Holmes, Lumley and Berkooz1996).
Its general form for the set of normalized amplitudes $a_k^n$ with $k \in \{-K, \ldots K\}$ and $n \in \{1, \ldots N_T\}$ can be given as
where Einstein notation has been used.
The linear component $A_k^{nm}$ can be further decomposed as
where $D_k^{nm}$ corresponds to the diffusion term
and $L_k^{nm}$ represents the interaction with the mean mode
We denote by $Q_{kk'}^{nmp}$ the interaction coefficient due to the nonlinear term $\underline {q} . \boldsymbol {\nabla } \underline {q}$ in the Navier–Stokes equations ($\underline {q} . \boldsymbol {\nabla } \underline {q}$ represents the velocity fluctuation). It can also include the effect of the pressure gradient if it is not neglected (see Noack, Papas & Monkewitz Reference Noack, Papas and Monkewitz2005).
We denote by $T_k^n$ the effect of the modes that are excluded from the truncation
where $\hat {\tau }_k^i$ is the Fourier transform of $q_j ({\partial q_i}/{\partial x_j})$.
Here $T_k^{n}$ is similar to a subgrid-scale stress term that needs to be modelled with additional assumptions in order to close the equations. The most general form is to assume that
where $\alpha _k^n$ is a function that is equivalent to a turbulent eddy viscosity $\nu _t/\nu$. In early derivations of the model, $\alpha _k^n$ was assumed to be constant (Aubry et al. Reference Aubry, Holmes, Lumley and Stone1988) but in more recent implementations (see, for instance, Podvin et al. Reference Podvin, Pellerin, Fraigneau, Bonnavion and Cadot2021), $\alpha _k$ can be a nonlinear function of the resolved modes $a_k^n$. Following Osth et al. (Reference Osth, Noack, Krajnoci, Borée and Barros2014), the turbulent eddy viscosity is assumed to scale with the characteristic velocity of the resolved modes $\sqrt {\sum _{m=1}^{N_T} |a_k^m|^2}$, which can be expanded into a sum of a constant term and a quadratic term $1/2 \sum _{m=1}^{N_T} (\lambda _k^m + |a_k^m|^2)$ (Podvin & Sergent Reference Podvin and Sergent2017). The proportionality constant is negative so that the effect of $T_k^n$ is to prevent the unphysical growth of fluctuations in the model.
7.2. Linear model
We derive a linearized version of the model for the most energetic POD modes $a_k^n$, with $k=0,1$ and $n \le 10$. The linear model is based on the assumption that the driving mechanism for the fluctuations is their direct interaction with the mean flow – through convection and deformation by the mean shear. Some justification for this can be found on the large separation between the intensity of the mean velocity and that of the fluctuations (the largest POD eigenvalue is less than 1 %). We therefore assume that the quadratic interaction terms $Q$, the effect of which is to redistribute energy among the resolved modes, can be neglected.
A second assumption is that the effect of both molecular ($D$) and turbulent ($T$) viscosity is also small and is limited to preventing non-physical unbounded growth of the fluctuations without modifying their characteristic frequencies. Figure 26 (top row) shows the spectrum of the matrix $L_0$ and $L_1$ for a truncation of 10 modes. Each eigenvalue $\mu _k^n$ has a small positive real part, which is much smaller than its imaginary part $\mu _k^n = \mathrm {Re}[\mu _k^n]+ {\rm i}\, \mathrm {Im}[\mu _k^n]$ with $\mathrm {Re}[\mu _k^n] \ll |\mathrm {Im}[\mu _k^n]|$ (the fact that the real part is positive is physically consistent with the fact that the fluctuations are extracting energy from the mean flow). Figure 26 (bottom row) shows the values of the matrix $D_k$ for the wavenumber $k=1$ (a similar behaviour was observed for $k=0$). We can see that $D_1$ is essentially diagonal. Its values $D_1^{nn}$ are mostly real and negative, and their relatively small magnitudes increase only slightly with $n$. Therefore, $D_k$ can be considered roughly proportional to the identity matrix as a first approximation,
This means that $L_k+D_k$ can be assumed to diagonalize in the same basis $B$ as $L$ with eigenvalues $\mu _k^n - \eta _k = \mathrm {Re}[\mu _k^n] - \eta _k + {\rm i}\, \mathrm {Im}[\mu _k^n]$.
Since the closure for the unresolved stresses $T_k$ is based on the properties of the dissipation matrix $D_k$ (7.6), we can further assume that the additional effect of the unresolved stresses is to merely modify the real part of the eigenvalues, with $| {\rm Re}[\mu _k^n] - \eta _k ( 1 + \mathrm {Sup}_p \alpha _k^p) | \ll {\rm Im}[\mu _k^n]$. Equation (7.1) can then be approximated as
where $\tilde {L}_k$ diagonalizes in the same basis $B$ as $L_k$ with purely imaginary eigenvalues ${\rm i}\, \mathrm {Im}[\mu _k^n]$:
The $n$th column of the matrix $B$ contains the eigenvector $b_k^n$ associated with the eigenvalue ${\rm i}\, {\rm Im}[\mu _k^n]$ and expressed in the basis $a_k^n$. Conversely, each POD amplitude $a_k^n$ can be represented as a sum of eigenmodes $b_k^p: a_k^n = \sum _{p=1}^{N_{T}} \beta _k^{np} b_k^p$. Predicted amplitudes for the model can then be reconstructed using
Figure 27 compares the spectra of the POD amplitudes extracted from the DNS with the frequencies obtained by considering $N_T=10$ and $N_T=6$ systems for $k=0$ and $k=1$. The predicted spectra shown for each mode $n$ are arbitrarily rescaled by $\mathrm {Max}_p \beta _{np}/\mathrm {Max}_f |\hat {a}_k^n(f)|$ for comparison purposes. We see that in all cases, the prediction provides a good identification of the dominant frequencies of the amplitude spectra, which supports the assumption of the fundamentally linear nature of the dynamics. We emphasize that the model frequencies are solely determined from the spatial modes extracted from the sampled autocorrelation tensor, independently of the sampling rate used to compute the tensor.
7.3. Time-varying linear model
Unlike those of the DNS, the spectra shown in figure 27 are discrete and limited to $N_T$ frequencies, since the linear model has constant coefficients. To make the model more physical, we modify the model so that the strength of the mean mode is allowed to vary in time. As a first approximation, we assume that this variation is solely driven by the local time fluctuations of the mean shear at the edge, which is determined by the maximum stretching eigenvalue $\lambda _S^{max}$ and, therefore, the ejections over the edge (as evidenced in figure 23). The modified model will be called the forced model.
If $\epsilon$ is the time-dependent modulation of the shear layer $\epsilon = \lambda _S^{max}/ \langle \lambda _S^{max}\rangle - 1$, the forced model is defined as
For all wavenumbers, the model was integrated from the initial condition $a_k^n=1$. Using arbitrary initial phases between the modes did not affect the general characteristics reported below. Figures 28(a) and 28(b) show the spectra of the forced model for the modes $k=0$ and $k=1$, which can be respectively compared with figures 14 and 16. Although complete agreement should not be expected given the crudeness of our assumption, it can be seen that the model roughly reproduces the widths and heights of the spectral bands of $a_k^n$ observed in the DNS. Figures 28(c) and 28(d) compare the probability distribution functions of the amplitude moduli $|a_k^n|$ for $k=0$ and $k=1$ in the DNS, the linear and the forced models. Although some discrepancies are occasionally observed in the distribution tails, both models capture equally well the shape of the distributions, which are characterized by frequent low values for the spanwise invariant modes and a non-zero maximum around 0.5–1 for the spanwise modes.
8. Conclusion
The dynamics of the flow over a forward-facing step have been investigated in a 3-D POD framework using a Fourier representation in the spanwise direction. An energy-rich characteristic structure was extracted from the fluctuations in the upstream region. The spanwise size of the structure scales with the step height $H$ and its formation appears to be associated with a centrifugal instability, the origin of which is located close to the edge of the step. Through the flow modulation created by the structure, incoming fluid close to the bottom wall can get trapped in or around the upstream recirculation then be ejected over the step into longitudinal packets, along with low-speed fluid collected along the step face through strong spanwise motions.
The following scenario connecting the dynamics of the upstream and the downstream recirculations is proposed: the strong ejections associated with the upstream flow deformation create an instantaneous momentum deficit at the edge of the step, which directly modifies the maximal shear rate in the downstream shear layer, thus impacting its development and the size of the downstream recirculation. A higher volume of ejections reduces the shear rate and increases the downstream recirculation. The flow immediately behind the step is dominated by a spanwise modulation of characteristic wavelength $\lambda _z={\rm \pi} H$, while spanwise rolls are observed downstream of reattachment. Since the wavelength $\lambda _z$ corresponds to the span of the periodic domain, more investigations would be needed to establish whether this result can be generalized to wider domains. Both 3-D fluctuations in the shear layer and 2-D motions behind the recirculation are characterized by convection velocities of about 0.6 $U$, where $U$ is the incoming velocity.
A linear model based on the direct interaction of the mean flow with the POD spatial modes at these dominant wavenumbers was able to predict the characteristic frequencies of the mode amplitudes. The modulations in the frequency spectra appear to be consistent with the variations of the instantaneous maximal shear rate at the edge of the step, and therefore, with those of the ejections. This suggests that linear mechanisms play a key role in the global flow dynamics and that these dynamics could be monitored through local indicators at the step edge, which opens up interesting possibilities for control. However, more investigations are needed to confirm this potential. In particular, linear stability analysis as well as resolvent-based approaches would be helpful to better understand the formation and characteristics of the upstream structure.
Acknowledgements
The DNS results used in this study are part of the database https://datasetmeca.lisn.upsaclay.fr/. We are grateful to E. Larose, R. Martinuzzi and F. Kerhervé for several helpful discussions. We express our thanks to the anonymous referees for their insightful suggestions.
Funding
This work was performed using HPC resources from GENCI–IDRIS (grant 2023-AD012A02062R2).
Declaration of interest
The authors report no conflict of interest.