1. Introduction
Adverse pressure gradients (APGs) arise in many industrial applications, such as the flow over an airplane wing or within a diffuser. To isolate the effect of the APG from the surface curvature, many studies focus on the flat plate turbulent boundary layers (TBL) with an applied APG. In an experiment, the APG can be created by a converging/diverging ceiling (Volino Reference Volino2020), an obstacle in the free stream (Vishwanathan et al. Reference Vishwanathan, Fritsch, Lowe and Devenport2021) or wall-normal transpiration (Clauser Reference Clauser1954; Marusic & Perry Reference Marusic and Perry1995) to decelerate the free stream velocity while in simulations, the free stream velocity is prescribed through a boundary condition (Kitsios et al. Reference Kitsios, Atkinson, Sillero, Borrell, Gungor, Jiménez and Soria2016; Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Lee Reference Lee2017). Whereas the statistics of the canonical zero pressure gradient (ZPG) TBL are parameterized solely by its local Reynolds number, $Re$, the APG TBL statistics are parameterized by the local $Re$, a measure of the local pressure gradient strength, and a measure of the upstream pressure gradient conditions or history effect (Schlatter & Örlü Reference Schlatter and Örlü2010; Monty, Harun & Marusic Reference Monty, Harun and Marusic2011; Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Vinuesa et al. Reference Vinuesa, Örlü, Sanmiguel Vila, Ianiro, Discetti and Schlatter2017). A characterization of the history effect on the turbulent statistics remains difficult because measurements and analysis of the upstream, streamwise variation of the flow field is required.
The friction Reynolds number, $Re_{\tau }$, the local ratio of the outer boundary layer thickness, $\delta _{99}$, to the viscous length scale, $\ell _{\nu }$, is often used as the Reynolds number to parameterize wall-bounded turbulent flows, like a channel or a TBL. There have been various scaling relationships presented to collapse different regions of the mean turbulent flow for different $Re_{\tau }$ including the near-wall, log-region and wake regions (Coles Reference Coles1956; Pope Reference Pope2000; Wei et al. Reference Wei, Fife, Klewicki and McMurtry2005; Marusic et al. Reference Marusic, Monty, Hultmark and Smits2013). While both the outer-scaled wake structures and intermediate large-scale structures in the log region, such as the attached eddies or very large-scale motions, become increasingly energized with $Re_{\tau }$, the effect is most evident in the latter as a secondary peak in the log region emerges at sufficiently high $Re_{\tau }$ (Smits, McKeon & Marusic Reference Smits, McKeon and Marusic2011; Lee & Moser Reference Lee and Moser2015; Deshpande et al. Reference Deshpande, van den Bogaard, Vinuesa, Lindić and Marusic2023). This intensifies the Reynolds stresses in viscous units throughout the boundary layer, primarily in the outer region of the flow, for both ZPG and APG TBLs (Aubertine & Eaton Reference Aubertine and Eaton2005; Hutchins et al. Reference Hutchins, Nickels, Marusic and Chong2009).
One measure of the APG strength in attached APG TBLs is the Clauser parameter (Clauser Reference Clauser1954), $\beta = ({\delta ^{*}}/{\tau _{W}})({{\rm d}P}/{{\rm d}\kern0.7pt x})$, where $\delta ^{*}$ is the displacement thickness, $\tau _{W}$ is the wall-shear stress, $P$ is the free stream pressure and $x$ is the streamwise coordinate. This study will focus on mild to moderate APG with $\beta \le 5$ such that the flow remains attached. An increase in $\beta$ affects the APG TBL by increasing the wake in the mean velocity profile (Aubertine & Eaton Reference Aubertine and Eaton2005; Monty et al. Reference Monty, Harun and Marusic2011), energizing the large-scale structures that scale with $\delta _{99}$ in the wake region (Harun et al. Reference Harun, Monty, Mathis and Marusic2013; Lee Reference Lee2017; Sanmiguel Vila et al. Reference Sanmiguel Vila, Vinuesa, Discetti, Ianiro, Schlatter and Örlü2020; Pozuelo et al. Reference Pozuelo, Li, Schlatter and Vinuesa2022) and a subsequent intensification of the Reynolds stresses in viscous units in both the inner and outer regions of the flow (Sanmiguel Vila et al. Reference Sanmiguel Vila, Vinuesa, Discetti, Ianiro, Schlatter and Örlü2020). The energization of large-scales in the outer region creates secondary peaks in the turbulent fluctuations, even at low $Re_{\tau }$ (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Lee Reference Lee2017), though these energized motions are primarily the outer-scaled structures in the wake rather than log region structures as evidenced in high $Re_{\tau }$ studies (Sanmiguel Vila et al. Reference Sanmiguel Vila, Vinuesa, Discetti, Ianiro, Schlatter and Örlü2020; Deshpande et al. Reference Deshpande, van den Bogaard, Vinuesa, Lindić and Marusic2023). The differences in the APG TBL structure have inspired a variety of scaling relationships that collapse the turbulent statistics (Gungor et al. Reference Gungor, Maciel, Simens and Soria2016; Sekimoto et al. Reference Sekimoto, Kitsios, Atkinson and Soria2019; Romero et al. Reference Romero, Zimmerman, Philip and Klewicki2022; Wei & Knopp Reference Wei and Knopp2023). While there are differences between APG and ZPG TBLs, primarily in the outer region of the TBL, the near-wall structure remains mostly unchanged (Harun et al. Reference Harun, Monty, Mathis and Marusic2013; Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Pozuelo et al. Reference Pozuelo, Li, Schlatter and Vinuesa2022). The premultiplied streamwise kinetic energy spectra in experiments and simulations identify a self-similar near-wall peak at $y = 15\ell _{\nu }$, $\lambda _{x} = 1000\ell _{\nu }$ and $\lambda _{z} = 100\ell _{\nu }$ for canonical wall-bounded turbulent flows (Hoyas & Jiménez Reference Hoyas and Jiménez2006; Marusic, Mathis & Hutchins Reference Marusic, Mathis and Hutchins2010; Smits et al. Reference Smits, McKeon and Marusic2011; Eitel-Amor, Örlü & Schlatter Reference Eitel-Amor, Örlü and Schlatter2014; Lee & Moser Reference Lee and Moser2015; Baidya et al. Reference Baidya, Philip, Hutchins, Monty and Marusic2017) that are reflected in APG TBLs with similar length scales (Harun et al. Reference Harun, Monty, Mathis and Marusic2013; Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Lee Reference Lee2017; Pozuelo et al. Reference Pozuelo, Li, Schlatter and Vinuesa2022; Deshpande et al. Reference Deshpande, van den Bogaard, Vinuesa, Lindić and Marusic2023).
Due to the history effect, which we will explore here in terms of the upstream variations in $\beta (x)$, the local parameters $Re_{\tau }$ and $\beta$ are not able to fully parameterize the local turbulent statistics since the boundary layer adjusts from the upstream conditions over a finite distance. In order to characterize the history effect and its influence on the turbulent statistics, the analysis needs to account for the upstream variation in the flow. As the history effect influences the mean velocity field, it also affects quantities of engineering interest like the skin-friction. Vinuesa et al. (Reference Vinuesa, Örlü, Sanmiguel Vila, Ianiro, Discetti and Schlatter2017) proposed an empirical predictive skin-friction curve for APG TBLs based on ZPG TBL data by accounting for the history effect through an upstream average of $\beta$. Here, an upstream average of $\beta$ will also be used as a measure for the history effect to quantify the accumulated $\beta$ in the APG TBL at a given streamwise location. It has been shown that for two APG TBLs with locally matched $Re_{\tau }$ and $\beta$, the APG TBL with the larger accumulated $\beta$ exhibits intensified streamwise fluctuations in the outer region and a stronger wake (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Tanarro, Vinuesa & Schlatter Reference Tanarro, Vinuesa and Schlatter2020).
Through the resolvent analysis framework of McKeon & Sharma (Reference McKeon and Sharma2010), the Fourier transformed velocity fluctuations are shown to be equal to the resolvent operator acting on the nonlinear terms in the Navier–Stokes equations (NSE). The resolvent operator is the inverse of the linearized NSE (LNSE) and captures the scale-dependent linear amplification from the mean flow field. In resolvent analysis, the resolvent operator is decomposed into an orthonormal forcing basis, $\boldsymbol {\phi }_{i}$, an orthonormal response basis, $\boldsymbol {\psi }_{i}$, and linear gains $\sigma _{i}$. Due to the non-normality of the LNSE, the leading modes take advantage of the increased pseudoresonant amplification such that the leading forcing and response modes exploit sources of non-normal amplification in the TBL (Trefethen Reference Trefethen1999; Schmid & Henningson Reference Schmid and Henningson2002; Symon et al. Reference Symon, Rosenberg, Dawson and McKeon2018). For wall-bounded turbulent flows, the main sources of non-normality come from the componentwise non-normality from the mean shear responsible for the lift-up mechanism and Orr-like tilting (Jovanović & Bamieh Reference Jovanović and Bamieh2005; Hwang & Cossu Reference Hwang and Cossu2010). Linear analyses that take advantage of these non-normal mechanisms have been shown to identify two distinct spectral peaks when the LNSE is augmented with an eddy viscosity (Del Alamo & Jimenez Reference Del Alamo and Jimenez2006; Cossu, Pujals & Depardon Reference Cossu, Pujals and Depardon2009). For streamwise developing flow fields, the convective non-normality is also present. It has been shown that modes that take advantage of the lift-up mechanism from the componentwise non-normality lead to low-rank behaviour in the resolvent operator. For these modes, $\boldsymbol {\psi }_{1}$ agrees well with the data-driven modes (Abreu et al. Reference Abreu, Cavalieri, Schlatter, Vinuesa and Henningson2020, Reference Abreu, Tanarro, Cavalieri, Schlatter, Vinuesa, Hanifi and Henningson2021) and the $\sigma _{i}$ identify length scales akin to the near-wall cycle (Moarref et al. Reference Moarref, Sharma, Tropp and McKeon2013). The convective non-normality manifests itself as upstream forcing that leads to downstream responses (Chomaz Reference Chomaz2005; Sipp & Marquet Reference Sipp and Marquet2013; Schmidt et al. Reference Schmidt, Towne, Rigas, Colonius and Brès2018; Symon et al. Reference Symon, Rosenberg, Dawson and McKeon2018; Pickering et al. Reference Pickering, Rigas, Schmidt, Sipp and Colonius2021). The convective and componentwise non-normalities are present in the biglobal resolvent approach, which has been applied to, for example, ZPG boundary layers (Sipp & Marquet Reference Sipp and Marquet2013; Abreu et al. Reference Abreu, Tanarro, Cavalieri, Schlatter, Vinuesa, Hanifi and Henningson2021), supersonic boundary layers (Kamal et al. Reference Kamal, Rigas, Lakebrink and Colonius2020) and spatially developing jets (Jeun, Nichols & Jovanović Reference Jeun, Nichols and Jovanović2016; Schmidt et al. Reference Schmidt, Towne, Rigas, Colonius and Brès2018). The linear biglobal studies of Ran et al. (Reference Ran, Zare, Hack and Jovanović2019) on stochastic receptivity in a laminar boundary layer and Davis, Uzun & Alvi (Reference Davis, Uzun and Alvi2019) on streamwise energy growth in a ZPG TBL also take advantage of the non-normal mechanisms discussed herein.
While the resolvent framework gives accurate predictions for systems with uncorrelated white-noise forcing, turbulent flows have correlated and structured nonlinear forcing that can limit comparisons between the resolvent modes and turbulent statistics (Towne, Schmidt & Colonius Reference Towne, Schmidt and Colonius2018; Morra et al. Reference Morra, Nogueira, Cavalieri and Henningson2021). This study focuses solely on the linear amplification mechanisms in APG TBLs that amplify the nonlinear terms in the NSE while drawing similarities to the turbulent statistics. Here, biglobal resolvent analysis is used to include the non-parallel effects of the APG TBL encoded within the mean flow field. This includes the history effect, which will be shown to affect the linear amplification and can be characterized through a scaling analysis of the LNSE.
This paper is organized as follows. Section 2 will describe the resolvent analysis framework, details about the discretization and a description of the datasets used in this study. Section 3 focuses on the local $\beta$ and $Re_{\tau }$ effects on the optimal response, $\sigma _{1}\boldsymbol {\psi }_{1}$, on both the small and large scales in low-moderate $Re_{\tau }$ TBLs. Section 4 demonstrates that the resolvent amplification, $\sigma _{i}$, of large-scale structures grows monotonically with a measure of the history effect. This linear monotonic growth with the history effect is then used to augment the hybrid velocity scale which improves the collapse of the Reynolds stresses (Sekimoto et al. Reference Sekimoto, Kitsios, Atkinson and Soria2019; Romero et al. Reference Romero, Zimmerman, Philip and Klewicki2022). An explanation for the linear growth is provided in § 5 through a simplified Orr–Sommerfeld-Squire (OSS) operator. A discussion of results is provided in § 6. Conclusions are presented in § 7.
2. Methodology
In this section, the biglobal resolvent approach is discussed for an incompressible TBL. This includes a description of the LNSE, the resolvent analysis framework, details of the numerical analysis and a description of the mean flow fields used herein.
2.1. Linearized Navier–Stokes equations
A statistically stationary incompressible TBL develops over a flat plate with a specified free stream velocity, $U_{\infty }(x)$. The streamwise, wall-normal and spanwise coordinates are $x$, $y$ and $z$ with $y=0$ denoting the location of the wall and $z$ assumed to be statistically homogeneous. The streamwise, wall-normal and spanwise velocities are $u$, $v$ and $w$ and the pressure divided by density is $p$. The domain is taken as a subset of the flat plate such that $x \in [x_{i},x_{o}]$, where the leading edge and virtual origin of the TBL are both outside of the domain and $y \in [0,y_{max}]$, where $y_{max}$ is in the free stream. Within the domain, reference locations, $x_{r}$, are used to define the length and velocity scales that non-dimensionalize the equations. Variables without superscripts are non-dimensionalized with the reference free stream velocity, $U_{r} = U_{\infty }(x_{r})$, and the reference length scale, $\delta _{99,r} = \delta _{99}(x_{r})$, where $\delta _{99}$ denotes the boundary layer thickness. Variables with $+$ superscripts are non-dimensionalized with the reference friction velocity $u_{\tau,r} = u_{\tau }(x_{r})$ and reference viscous length scale $\ell _{\nu } = \nu /u_{\tau,r}$, where $u_{\tau } = \sqrt {\tau _{W}/\rho }$. These reference length and velocity scales are then used to define the local outer Reynolds number, $Re = U_{r}\delta _{99,r}/\nu$ and local friction Reynolds number, $Re_{\tau } = u_{\tau,r}\delta _{99,r}/\nu$.
The instantaneous flow field, $\breve {\boldsymbol {q}}(\boldsymbol {x},t) = [\breve {\boldsymbol {u}},\breve {p}]^{T}$, can be written as the sum of a known mean flow field, $\bar {\boldsymbol {Q}}(x,y) = [\bar {\boldsymbol {U}}(x,y),\bar {P}(x,y)]^{T}$, and a fluctuation, $\boldsymbol {q}(\boldsymbol {x},t) = [\boldsymbol {u}(\boldsymbol {x},t),p(\boldsymbol {x},t)]^{T}$. For the boundary layer, $\bar {\boldsymbol {U}} = \bar {U}(x,y)\boldsymbol {e}_{x} + \bar {V}(x,y)\boldsymbol {e}_{y}$. Due to the spanwise homogeneity and statistical stationarity of the flow, the NSE may be Fourier transformed in $z$ and $t$. The Fourier transformed variables are denoted by $\hat {\cdot }$, such that
where $k_{z} = 2{\rm \pi} /\lambda _{z}$ and $\omega = 2{\rm \pi} /\lambda _{t}$ are the spanwise wavenumber and temporal frequency. $\lambda _{z}$ and $\lambda _{t}$ then denote the spanwise and temporal wavelengths. The equations governing the Fourier modes with $(\omega,k_{z}) \ne (0,0)$ are then
where $\boldsymbol {\nabla } = [\partial _{x},\partial _{y},\mathrm {i} k_{z}]$, $\nabla ^{2} = \partial _{xx} + \partial _{yy} - k_{z}^{2}$ and $\epsilon _{s}$ is the sponge used for artificial boundary conditions. Following McKeon & Sharma (Reference McKeon and Sharma2010), the nonlinear terms in (2.2) are treated as a nonlinear forcing with spanwise wavenumber $k_{z}$ and temporal frequency $\omega$, $\hat {\boldsymbol {f}} = [\hat {\boldsymbol {f}}_{\boldsymbol {u}},0]^{T}$. To focus on what insight can be obtained from the linear resolvent operator, the forcing is assumed to be uncorrelated from the turbulent fluctuations. While eddy viscosity can be used to incorporate turbulent interactions into the LNSE (Del Alamo & Jimenez Reference Del Alamo and Jimenez2006; Cossu et al. Reference Cossu, Pujals and Depardon2009; Hwang & Cossu Reference Hwang and Cossu2010), their specific forms are data-driven or ad hoc and can obfuscate the true linear amplification in the NSE and the interpretability of the nonlinearities (Morra et al. Reference Morra, Nogueira, Cavalieri and Henningson2021; Pickering et al. Reference Pickering, Rigas, Schmidt, Sipp and Colonius2021). In this approach, $\hat {\boldsymbol {f}}$ can be interpreted as an externally applied body force.
No-slip and no-penetration boundary conditions are applied such that $\hat {\boldsymbol {u}}(x,y=0) = \boldsymbol {0}$ is enforced. In the free stream, Neumann boundary conditions are applied such that $\partial _{y}\hat {\boldsymbol {u}}(x,y=y_{max}) = \boldsymbol {0}$. Since the streamwise domain is a subset of the flat plate with $x \in [x_{i},x_{o}]$, artificial boundary conditions are applied in the streamwise direction such that $\hat {\boldsymbol {q}}$ remains compact within the domain. This is enforced through the sponge, $\epsilon _{s}(x)$, in (2.2), which is $0$ for $90\,\%$ of the domain and ramps up quadratically from $0$ to $\epsilon _{0}$ at the endpoints of the domain. The explicit form of $\epsilon _s(x)$ is
where $L = x_{o} - x_{i}$ and $x_{c}$ is the centre of the domain. $\epsilon _{s}$ causes $\hat {\boldsymbol {q}}$ to decay to $\boldsymbol {0}$ in the vicinity of the boundaries. Similar strategies have been applied for linear analyses of incompressible flows in Ran et al. (Reference Ran, Zare, Hack and Jovanović2019) and Abreu et al. (Reference Abreu, Tanarro, Cavalieri, Schlatter, Vinuesa, Hanifi and Henningson2021). The inlet and outlet of the domains are then treated with Dirichlet boundary conditions such that $\hat {\boldsymbol {q}}(x = x_{i},y) = \hat {\boldsymbol {q}}(x = x_{o},y) = \boldsymbol {0}$ to enforce the compact nature of the resolvent modes. The effect of the sponge is qualitatively similar to fringe zones that are commonly used in TBL simulations to dampen the outflow and allow for the treatment of periodic boundary conditions (Chevalier et al. Reference Chevalier, Schlatter, Lundbladh and Henningson2007). In the calculations presented herein, $\epsilon _{s} = 30$. Varying this value was found to have a negligible effect on the linear amplification, provided that $\epsilon _{s}$ was large enough to dampen $\hat {\boldsymbol {q}}$ at the boundaries.
2.2. Resolvent analysis
Equation (2.2) can be written compactly as
where $\mathcal {B}$ projects away the $p$ component as $\mathcal {B}\hat {\boldsymbol {q}} = [\hat {\boldsymbol {u}},0]^{T}$. The kinetic energy of the disturbances, integrated over the domain, is a norm induced by the inner product,
By using this inner product, the adjoint of (2.2) can also be written compactly as
where the daggers denote adjoint operators and tildes denote adjoint variables. An explicit form of the adjoint operator can be found in Gomez (Reference Gomez2024). It is assumed that $\omega$ is not an eigenvalue of $\mathcal {L}$ so that the resolvent, $\mathcal {H} = (-\mathrm {i}\omega \mathcal {B} + \mathcal {L})^{-1}\mathcal {B}$, and adjoint resolvent, $\mathcal {H}^{{\dagger} } = (\mathrm {i}\omega \mathcal {B} + \mathcal {L}^{{\dagger} })^{-1}\mathcal {B}$, exist.
In resolvent analysis, one is interested in finding the forcing inputs of $\mathcal {H}$ that lead to the largest amplification. Here, the forcing inputs and response outputs of $\mathcal {H}$ are measured with the induced norms $\| \cdot \|_{f}$ and $\| \cdot \|_{r}$ from the inner products $\langle \boldsymbol {a},\boldsymbol {b} \rangle _{f} = \langle \boldsymbol {a},\mathcal {W}_{f}\boldsymbol {b} \rangle$ and $\langle \boldsymbol {a},\boldsymbol {b} \rangle _{r} = \langle \boldsymbol {a},\mathcal {W}_{r}\boldsymbol {b} \rangle$, respectively. Here, $\mathcal {W}_{r}$ and $\mathcal {W}_{f}$ are positive definite operators that dictate the components and spatial domains of $\hat {\boldsymbol {q}}$ and $\hat {\boldsymbol {f}}$ that are used to measure the response and the forcing. The exact form of $\mathcal {W}_{r}$ and $\mathcal {W}_{f}$ will be specified for the studies considered. An example of the regions included in the norms $\| \cdot \|_{f}$ and $\| \cdot \|_{r}$ are depicted in figure 1.
The linear amplification, $\sigma$, of $\mathcal {H}$ is here defined as the Rayleigh quotient,
To find the largest amplification, one seeks the forcing, $\boldsymbol {\phi }$, that produces the largest $\sigma$ such that
It can be shown that (2.8) can be solved via the eigenvalue problem,
Since $\mathcal {H}\mathcal {W}_{r}\mathcal {H}^{{\dagger} }$ and $\mathcal {W}_{f}$ are positive definite operators, the eigenvalues, $\sigma _{j}^{2}$, are positive and the eigenvectors, $\boldsymbol {\phi }_{j}$, are orthonormal with respect to the inner product $\langle {\cdot },{\cdot } \rangle _{f}$ such that $\langle \boldsymbol {\phi }_{j},\boldsymbol {\phi }_{k} \rangle _{f} = \delta _{jk}$, where $\delta _{jk}$ is the Kronecker delta. The $\sigma _{j}$ are ordered such that $\sigma _{1} \ge \sigma _{2} \ge \dots \ge 0$. $\boldsymbol {\phi }_{j}$ are here identified as the forcing modes with corresponding linear gain, $\sigma _{j}$. The response modes, $\boldsymbol {\psi }_{j}$, are defined as $\boldsymbol {\psi }_{j} = \sigma _{j}^{-1}\mathcal {H}\boldsymbol {\phi }_{j}$. It can be shown that the response modes are orthonormal with respect to the inner product $\langle {\cdot },{\cdot } \rangle _{r}$ such that $\langle \boldsymbol {\psi }_{j},\boldsymbol {\psi }_{k} \rangle _{r} = \delta _{jk}$.
In this study, two sets of norms are considered. The first set is tailored to study the effect of local $Re_{\tau }$ and $\beta$ on the linear amplification in § 3. This requires that the response modes are supported at $x_{r_{1}}$ where $\beta (x_{r_{1}})$ and $Re_{\tau }(x_{r_{1}})$ are defined. If there is no spatial masking applied to the inner product such that $\langle {\cdot },{\cdot } \rangle _{r} = \langle {\cdot },{\cdot } \rangle _{f} = \langle {\cdot },{\cdot } \rangle$, then the response and forcing modes can be supported anywhere along the streamwise domain. In order to capture the modes at $x_{r_{1}}$, spatial masking is applied to the response modes so that
where $L_{d}^+ = 2150$, while $\langle {\cdot },{\cdot } \rangle _{f} = \langle {\cdot },{\cdot } \rangle$. The forcing modes can extend across the entire domain allowing forcing from upstream the spatial mask to influence the linear response within the spatial mask. This helps capture non-local amplification mechanisms like those from the convective non-normality. Here $L_{d}$ is held constant in viscous units since the small-scale modes are expected to scale in inner units. For a mode to be optimal under the new inner products via the definition in (2.7), the optimal response modes must be supported in the streamwise region within $L_{d}^+/2$ of $x_{r_{1}}$. The effect of $L_{d}^+$ has been shown in the appendix of Gomez (Reference Gomez2024), where it is shown that increasing $L_{d}^+$ reduces the response of the near-wall small-scale modes at $x_{r_{1}}$. The second set of inner products is used to study the history effects in § 4. In order to only consider the effect of the pressure gradient upstream of $x_{r_{2}}$, the inner products $\langle {\cdot },{\cdot } \rangle _{r}$ and $\langle {\cdot },{\cdot } \rangle _{f}$ are masked such that
where $x_{i}$ is 12 $\delta _{99}(x_{r_{2}})$ upstream of $x_{r_{2}}$. This makes the resolvent modes optimal in the sense of (2.8) if they are supported upstream of $x_{r_{2}}$ and also prevents the amplification mechanisms present downstream of $x_{r_{2}}$ from contributing to the linear amplification.
It is important to note that $\boldsymbol {\psi }_{j}$, $\boldsymbol {\phi }_{j}$ and $\sigma _{j}$ are dependent on the $(\lambda _{z},\lambda _{t})$ pair, $\bar {\boldsymbol {U}}$, and choice of norms. Here $\boldsymbol {\psi }_{j}$ can be used to reconstruct the flow field as $\hat {\boldsymbol {q}} = \sum _{j}\xi _{j}\sigma _{j}\boldsymbol {\psi }_{j}$, where $\xi _{j}$ are coefficients that are found through fitting to data or appealing to the full NSE. The studies herein will primarily consider the optimal response, $\sigma _{1}\boldsymbol {\psi }_{1}$, as a rank-one approximation for the Fourier modes without any nonlinear closure. By taking an integral over $\omega$, one can also define the premultiplied amplification,
Here $\hat {E}$ can be thought of as the premultiplied streamwise kinetic energy spectrum for a velocity signal under the rank-one representation, $\hat {\boldsymbol {u}} = \sigma _{1}\boldsymbol {\psi }_{\boldsymbol {u},1}$. Here, $\hat {E}(y^+;k_{z}^+)$ is a measure of the linear amplification across all scales.
2.3. Numerical discretization
The domain is discretized with $N_{x}$ equispaced points in the streamwise direction and $N_{y}$ points in the wall-normal direction. The wall-normal grid points are stretched such that half the points are below $y_{min}$. Each wall-normal grid point, $y_{k}$, is mapped from an equispaced grid point, $y'_{k}$ using $y_{k} = a y_{k}'/(b-y_{k}')$ where $a = y_{max}y_{min}/(y_{max}-2y_{min})$ and $b = 1 + a/y_{max}$. This is similar to the grids used in the supersonic linear analyses of Kamal et al. (Reference Kamal, Rigas, Lakebrink and Colonius2020) and Malik (Reference Malik1990), except here, the grid stretching is used to increase the resolution in the near-wall region of the TBLs studied. Both streamwise and wall-normal directions use a fourth-order summation by parts scheme for the differentiation operators (Mattsson & Nordström Reference Mattsson and Nordström2004). Once discretized, the operators $\mathcal {B}$, $\mathcal {L}$, $\mathcal {L}^{{\dagger} }$, $\mathcal {H}$ and $\mathcal {H}^{{\dagger} }$ become the $4N_{x}N_{y} \times 4N_{x}N_{y}$ matrices ${\boldsymbol{\mathsf{B}}}$, ${\boldsymbol{\mathsf{L}}}$, ${\boldsymbol{\mathsf{L}}}^{{\dagger} }$, ${\boldsymbol{\mathsf{H}}}$ and ${\boldsymbol{\mathsf{H}}}^{{\dagger} }$.
In order to perform the integration in (2.5), the streamwise and wall-normal integration is performed with a trapezoidal scheme. The positive definite matrix, ${\boldsymbol{\mathsf{W}}}_{T}$, performs the integration in (2.5) such that
In a similar fashion, the operators $\mathcal {W}_{r}$ and $\mathcal {W}_{f}$ are discretized with the diagonal matrices ${\boldsymbol{\mathsf{M}}}_{r}$ and ${\boldsymbol{\mathsf{M}}}_{f}$. The diagonal entries of ${\boldsymbol{\mathsf{M}}}_{r}$ and ${\boldsymbol{\mathsf{M}}}_{f}$ are $0$ for the entries omitted and $1$ for the entries included in the inner products. As a result, the inner products for the discretized space are defined as
and
In the discretized space, the eigenvalue problem in (2.9) becomes
while $\boldsymbol {\psi }_{j} = \sigma _{j}^{-1}{\boldsymbol{\mathsf{H}}}{\boldsymbol {\phi }_{j}}$. Equation (2.16) is solved using the Arnoldi algorithm (Saad Reference Saad2011). The matrices ${\boldsymbol{\mathsf{B}}}$, ${\boldsymbol{\mathsf{L}}}$, ${\boldsymbol{\mathsf{L}}}^{{\dagger} }$, ${\boldsymbol{\mathsf{W}}}_{r}$ and ${\boldsymbol{\mathsf{W}}}_{f}$ are sparse, while the resolvent matrices, ${\boldsymbol{\mathsf{H}}}$ and ${\boldsymbol{\mathsf{H}}}^{{\dagger} }$, are dense matrices that are costly to compute. The Arnoldi approach requires repeated calculations of $\boldsymbol {v} = {\boldsymbol{\mathsf{H}}}\boldsymbol {b}$, or equivalently solving $(-\mathrm {i}\omega {\boldsymbol{\mathsf{B}}} + {\boldsymbol{\mathsf{L}}})\boldsymbol {v} = {\boldsymbol{\mathsf{B}}}\boldsymbol {b}$, and similarly for the adjoint. Following the approach described in Sipp & Marquet (Reference Sipp and Marquet2013) and Schmidt et al. (Reference Schmidt, Towne, Rigas, Colonius and Brès2018), LU decompositions of $(-\mathrm {i}\omega {\boldsymbol{\mathsf{B}}} + {\boldsymbol{\mathsf{L}}})$ and $(\mathrm {i}\omega {\boldsymbol{\mathsf{B}}} + {\boldsymbol{\mathsf{L}}}^{{\dagger} })$ are computed once. The LU factors can then efficiently solve $(-\mathrm {i}\omega {\boldsymbol{\mathsf{B}}} + {\boldsymbol{\mathsf{L}}})\boldsymbol {v} = {\boldsymbol{\mathsf{B}}}\boldsymbol {b}$ through Gaussian elimination. Thus, the overall cost of the Arnoldi algorithm is the cost of the LU decompositions (Jeun et al. Reference Jeun, Nichols and Jovanović2016). The LU decomposition and Gaussian elimination is computed using the Intel oneAPI Math Kernel Library PARDISO (Schenk & Gärtner Reference Schenk and Gärtner2004), which is optimized for the sparse form of the matrices and uses parallelized subroutines. See the appendix of Gomez (Reference Gomez2024) for an algorithm explaining this strategy in more detail.
2.4. Description of datasets
This study focuses on mild to moderate APG TBL where $\beta \in [0,5]$. Since the large-scale modes require large streamwise domains to both be resolved and achieve significant amplification, only studies with large, well-resolved streamwise domains, relative to $\delta _{99}(x_{r})$, are considered. The mean flow fields of interest are the large eddy simulation flat plate TBL data of Bobke et al. (Reference Bobke, Vinuesa, Örlü and Schlatter2017), Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022) and Eitel-Amor et al. (Reference Eitel-Amor, Örlü and Schlatter2014) which each have domains upstream of $x_{r_{2}}$ larger than $20\delta _{99}(x_{r_{2}})$. The datasets all use a constant free stream velocity, $U_{\infty }(x)$, followed by an APG imposed by the boundary condition $U_{\infty }(x) = (x-x_{0})^{m}$, where $x_{0}$ and $m$ are specified to create varying degrees of APG strengths. Under this regime, the APG TBLs are in the ‘near-equilibrium’ regime (Mellor & Gibson Reference Mellor and Gibson1966; Townsend Reference Townsend1976; Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017).
A brief description of the parameters of interest in the datasets are described in table 1. In this study, the Clauser parameter is defined as
as in Bobke et al. (Reference Bobke, Vinuesa, Örlü and Schlatter2017) and Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022). Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022) notes that the free stream boundary conditions in the simulations cause $\bar {U}$ to decrease in the wall-normal location from its peak at the boundary layer edge such that ${\rm d}P/{{\rm d}\kern0.7pt x}$ at $y = \delta _{99}$ is different from the imposed pressure gradient (PG) through the free stream boundary condition. Due to this, the parameter
will also be used to account for the effect of the free stream boundary condition. Here, $\beta _{\infty }$ is used to parameterize the effect of history by accounting for the streamwise variation of the imposed APG boundary condition. In this study, the history effect is quantified by
where $x_{i,2}$ is an upstream location. Here, $x_{i,2} = x_{r_{2}} - 12\delta _{99}(x_{r_{2}})$ corresponds to the inlet location of the domains used to study the history effect in § 4. Here $\langle \beta _{\infty } \rangle$ is similar to the streamwise average that Vinuesa et al. (Reference Vinuesa, Örlü, Sanmiguel Vila, Ianiro, Discetti and Schlatter2017) considered, though they averaged $\beta$ over the momentum thickness-based Reynolds number. The domains used in Vinuesa et al. (Reference Vinuesa, Örlü, Sanmiguel Vila, Ianiro, Discetti and Schlatter2017) for the averaging were $O(10\delta _{99})$ large, which is consistent with the averaging domain used here. Furthermore, (2.19) offers a measure of the APG history in a manner analogous to the streamwise averaging induced by the inner product in (2.11).
Figure 2(a) shows the variation of $\beta$ with $Re_{\tau }$ for the datasets used, along with the values at the reference locations used in this study. Figure 2(b–d) show $\bar {U}^+$, $\overline {uu}^+$ and $\bar {V}^+$ at fixed $x_{r_2}$. Consistent with the observations from previous studies described above, these mean flow fields demonstrate monotonic growth in the wake velocity and decreasing $\bar {U}^+$ in the log-layer region with increasing $\beta$. Monotonic growth in the secondary peak of $\overline {uu}^+$ is observed with $\langle \beta _{\infty } \rangle$, with the near-wall peak of $\overline {uu}^+$ also affected. The non-monotonic behaviour in $\bar {V}^+$ with $\langle \beta \rangle _{\infty }$ highlight the non-equilibrium nature of the datasets, with different levels of deceleration in the flow at $x_{r_{2}}$
3. Effect of local parameters on linear amplification
In this section, the local effects of $\beta (x_{r})$ and $Re_{\tau }(x_{r})$ on the linear amplification are examined. Here, the reference viscous and outer length and velocity scales are defined at $x_{r_{1}}$. Table 2 describes quantities of interest used for the seven case studies in this section. In order to support the large-scale modes, the domains are set with $L_{x} = 43.44$, which corresponds to the entire domain of the b2 dataset, except for case 4 whose entire domain is $L_{x} = 36.6$ long. To resolve small-scale structures near the wall, $y_{min} = 0.24 \sqrt {537/Re_{\tau }}$ in cases 5–7 so that $y_{min}$ is proportional to the start of the log-layer (Wei et al. Reference Wei, Fife, Klewicki and McMurtry2005). Since $\delta _{99}$ does not grow as rapidly for larger $Re_{\tau }$ TBL, cases 5–7 reduce the $y_{max}$ to $3$ to further increase the resolution within the boundary layer while maintaining $y_{max}> \delta _{99}(x)$.
3.1. Effect of masking on the response and forcing modes
Here, the goal is to study the linear response of large- and small-scale modes subject to different local parameters. This requires that the response modes are supported at $x_{r_{1}}$ where $\beta (x_{r_{1}})$ and $Re_{\tau }(x_{r_{1}})$ are defined. Without a spatial mask, the response and forcing modes can be supported anywhere along the streamwise domain. It was shown that the small-scale modes, characterized by $O({\lambda _{z}^+}) \le O({100})$, were amplified via the componentwise non-normality from the mean shear near the wall while the large-scale modes, characterized by $O({\lambda _{z}}) \ge O({1})$, extended into the outer region of the flow and were amplified via the convective non-normality (Gomez Reference Gomez2024). As a result, the unmasked small-scales are expected to be concentrated in the upstream region where they experience the largest mean shear within the domain. The unmasked large-scale modes can extend across the entire domain, with the largest forcing amplitude concentrated upstream and the largest response amplitude concentrated downstream (Sipp & Marquet Reference Sipp and Marquet2013; Symon et al. Reference Symon, Rosenberg, Dawson and McKeon2018).
As an example for how the leading resolvent modes behave without masking, $\boldsymbol {\psi }_{1}$ and $\boldsymbol {\phi }_{1}$ are computed for a representative small-scale mode with $(\lambda _{z}^+,\lambda _{t}^+) = (50,50)$ and a representative large-scale mode with $(\lambda _{z},\lambda _{t}) = (1,1.4)$ using case 3. In figure 3(a,b), $\phi _{v,1}$ and $\sigma _{1}\psi _{u,1}$ are plotted for the representative small-scale and in figure 3(c,d) for the representative large-scale. As alluded to earlier, without masking, the small-scale modes are concentrated upstream of $x_{r_{1}}$ while the large-scale mode is supported at $x_{r_{1}}$. For these unmasked modes, the amplitude of $\psi _{1}(y,x_{r_{1}})$ is negligible for the small-scale mode and serves as a poor measure of the linear response at $x_{r_{1}}$.
In figure 3(e,f), the modes from figure 3(a,b) are recomputed using the spatial masking described in (2.10). Both the representative small-scale $\phi _{v,1}$ and $\psi _{u,1}$ are supported at $x_{r_{1}}$ with the spatial mask. Due to the larger $Re_{\tau }$ at $x_{r_{1}}$ compared with the inlet location and fixed $\lambda _{z}$ and $\lambda _{t}$ across the domain in the calculation, the modes plotted in figure 3(e,f) have significantly smaller wall-normal extents and streamwise wavelengths than those in figure 3(a,b). This is because the small-scale modes are amplified by the large mean shear concentrated in the viscous subregion, whose wall-normal extent scales with $\ell _{\nu }$. As $Re_{\tau }$ increases downstream, $\ell _{\nu }$ decreases and constrains the length scales of the modes. Now the representative large-scale modes from figure 3(c,d) are recomputed using the same spatial masking and plotted in figure 3(g,h). Since $\langle {\cdot },{\cdot } \rangle _{f}$ is left unchanged, the modes can still be amplified via the convective non-normality. As a result, the large-scale forcing can still be supported upstream of $x_{r_{1}}$ while its response is supported downstream of $x_{r_{1}}$. However, because of the spatial masking, the optimal forcing is primarily concentrated upstream of $x_{r_{1}}$ since any forcing downstream of $x_{r_{1}}+L_{d}/2$ would lead to a response outside of the spatial mask without contributing to $\sigma _{1}$. This also highlights the non-local nature in the amplification of large-scale structures and their susceptibility to upstream history effects that is explored in more detail in § 4.
For the rest of this section, the spatial masking described in (2.10) is applied to identify optimal small and large-scale responses at $x_{r_{1}}$.
3.2. Effect of $\beta$ on linear amplification at low $Re_{\tau }$
To investigate the effect of local $\beta (x_{r_{1}})$, cases 1–4 are examined since they have a similar $Re_{\tau }(x_{r_{1}})$ and different values of $\beta (x_{r_{1}})$. Furthermore, the slow variation in $\beta (x)$ near $x_{r_{1}}$ minimizes the influence of ${{\rm d}\beta }/{{\rm d}\kern0.7pt x}$ or history on the linear amplification between cases $1$–$4$. Here, the effect of $\beta (x_{r_{1}})$ on the resolvent amplification is compared for a wide range of scales.
First, the streamwise component of the linear response of the ZPG TBL of case 1 and the APG TBL of case 3 are compared for a large-scale mode ($\lambda _{t}^+ = 200, \lambda _{z} = 1$) and small-scale mode ($\lambda _{t}^+ = 50, \lambda _{z}^+ = 50$) in figure 4. Plotted in figure 4(a–c), the effect of the APG on the large-scale response at $x_{r_{1}}$ is an increase in the magnitude and peak amplitude location farther from the wall than the ZPG counterparts. Since these large-scale modes have support, i.e. non-negligible amplitude, in the outer region of the TBL, it is expected that they are most susceptible to the APG effects. On the other hand, the near-wall small-scale modes plotted in figure 4(d–f) are only slightly affected by the change in the mean flow field. This is because for this mild APG, the near-wall is still dominated by viscous effects and the APG hardly affects the inner scaling of $\bar {U}^+$. These near-wall resolvent modes have been shown to be self-similar in the region where the mean flow field demonstrates self-similarity (Moarref et al. Reference Moarref, Sharma, Tropp and McKeon2013; Gomez Reference Gomez2024).
While it is illustrative to compare the effect of $\beta$ on individual modes, we now consider the effect of increasing $\beta$ over a range of scales. The leading masked response modes are computed over a range of $31$ logarithmically spaced $k_{z}^+$ between $2{\rm \pi} /9.5$ and $2{\rm \pi} /11\,000$ and $33$ logarithmically spaced $\omega ^+$ between $2{\rm \pi} /1.25$ and $2{\rm \pi} /20\,000$. This range of $k_{z}^+$ and $\omega ^+$ is wide enough to capture the small-scale and large-scale modes and encompasses the characteristic length and time scales of the expected energy-containing structures in the actual TBLs. The premultiplied amplification spectra, $\hat {E}(y^+,k_{z}^+)$, (see (2.12)) are then computed for each $\beta$.
In figure 5, $\hat {E}^+$ is computed for cases 1–4. For the ZPG in case 1, $\hat {E}^+$ identifies only a near-wall peak with a local maximum near $y^+ = 20$ and $\lambda _{z}^+ = 77$. The lack of a secondary peak in the premultiplied amplification at this low $Re_{\tau }$ is similar to what is observed in the true premultiplied kinetic energy spectra for low $Re_{\tau }$ ZPG TBL and channels, which are dominated by the near-wall cycle. For the APG TBLs in cases 2-4, a near-wall peak is also observed. For case 2, this local maximum occurs at the same location as case 1, while for cases 3 and 4, the local maxima is at $y^+ = 20$ and $\lambda _{z}^+= 60$.
Here $\hat {E}^+$ predicts higher wall-normal locations and smaller wavelengths for the near-wall peak than observed in the turbulent data of ZPG and mild to moderate APG TBLs. Besides the assumption of uncorrelated (uncoloured) forcing, one reason for this discrepancy is that in cases 1–4, $\delta _{99}$ varies substantially while $\lambda _{z}$ remains fixed across the domain. As an example, in case 3, the $\lambda _{z}^+ = 100$ mode has a $\lambda _{z} = 0.19\delta _{99}(x_{r_{1}}) = 1.20\delta _{99}(x_{i})$. That is, near the inlet, the $\lambda _{z}^+ = 100$ mode is larger than the $\delta _{99}$ at the inlet. This in turn causes the $\lambda _{z}^+=100$ mode at $x_{r_{1}}$ to behave like a large-scale mode and experiences amplification via the convective non-normality, rather than the componentwise non-normality characteristic of the small-scale modes. As a result, the $\lambda _{z}^+$ modes are absent from the near-wall peak for these low $Re_{\tau }$ TBLs. The changing nature of the mode behaviour with streamwise distance is a challenge unique to the TBL relative to the canonical internal flows. It was shown in Gomez (Reference Gomez2024) that by also applying masking to the forcing such that $\langle {\cdot },{\cdot } \rangle _{f}= \langle {\cdot },{\cdot } \rangle _{r}$, a local maximum could be found in $\hat {E}^+$ with $\lambda _{z}^+ \approx 100$ for both ZPG and APG TBLs.
Despite the simplified forcing assumption, $\hat {E}^+$ also demonstrates a secondary peak for large-scale modes farther from the wall for the APG TBLs in cases 2–4. The secondary peak in $\hat {E}^+$ is amplified monotonically with increasing $\beta$. Similarly, a monotonic increase in the secondary peak of the spanwise kinetic energy spectra was observed in the DNS data of Lee (Reference Lee2017) at similar $Re_{\tau }$ to the mean data used here. For larger $Re_{\tau }$, the streamwise energy spectra in the experiment also demonstrate large-scale secondary peaks that are energized with increasing $\beta$ (Sanmiguel Vila et al. Reference Sanmiguel Vila, Vinuesa, Discetti, Ianiro, Schlatter and Örlü2020). Despite the similar trends seen with $\beta$, there is again a discrepancy in the location of the secondary peaks in $\hat {E}^+$ and those of similar low $Re_{\tau }$ spanwise energy spectra (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Lee Reference Lee2017). The secondary peaks in figure 5(b–d) occur at $\lambda _{z}^+ \approx 0.5Re_{\tau }$ whereas the turbulent data reports $\lambda _z^+ \approx 0.8Re_{\tau }-0.9Re_{\tau }$. The discrepancy is likely attributable to the lack of nonlinear closure, but may again be in part related to the change in $\delta _{99}$ across the domain, as a mode of the order of $\delta _{99}(x_{r_{1}})$ at $x_{r_{1}}$ would be around $6\delta _{99}(x_{i})$ for case 3. Since the large-scale modes are forced from the upstream location, the smaller upstream boundary layer can affect the behaviour of the amplification at $x_{r_{1}}$ by geometrically constraining the modes that can be amplified downstream. Nonetheless, the masked biglobal resolvent approach is able to capture the presence of both a small-scale, near-wall peak and a large-scale peak that is amplified monotonically with increasing $\beta$.
3.3. Effect of $Re_{\tau }$ on linear amplification
Linear amplification is now considered using the ZPG TBL dataset of Eitel-Amor et al. (Reference Eitel-Amor, Örlü and Schlatter2014) for the low $Re_{\tau }$ case 1 and larger $Re_{\tau }$ case 6. The large $Re_{\tau }$ APG TBL dataset of Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022) (case 7) is also used to study the $\beta$ effects in the linear amplification at higher $Re_{\tau }$ by comparing with case 5 at the matched $Re_{\tau } = 1500$. The sweeps are computed using the same $k_{z}$ and $\omega$ range as the previous results while the wall-normal stretching has been adjusted to resolve the smaller inner region in the larger $Re_{\tau }$ datasets by shrinking $y_{min}$.
Figure 6(a,b) show the variation of $\hat {E}^+$ for cases 1 and 6. Along with the near-wall peaks, $\hat {E}^+$ also demonstrate increased amplification for the large-scale modes with increasing $Re_{\tau }$ with case 6 showing evidence of a secondary peak. This secondary peak occurs at a wall-normal location slightly closer to the wall than the secondary peaks of the APG results in figure 5, potentially pointing to a difference between log and APG-amplified wake structures. However, the $Re_{\tau }$ is too small to adequately distinguish between the two. Figure 6(d) provides direct comparison of isocontours of $\hat {E}^+$ for cases 1 and 6. These plots demonstrate that $\hat {E}^+$ is self-similar in the near-wall region due to the inner-scaled mask in $\langle {\cdot },{\cdot } \rangle _{r}$ and self-similar mean flows, plotted in figure 6(c) for reference. The near-wall self-similarity of the resolvent amplification in individual modes had been studied for channel flows in Moarref et al. (Reference Moarref, Sharma, Tropp and McKeon2013) and spatially developing TBLs in Gomez (Reference Gomez2024) where the response and forcing modes are spatially constrained via small, inner-scaled, streamwise domains. Figure 6(d) demonstrates that the self-similarity in the resolvent modes persists in $\hat {E}^+$ even when the forcing modes are not spatially constrained with a spatial mask or limited domain length. As the large-scale outer-scaled resolvent modes become more amplified farther from the wall, the near-wall self-similarity of $\hat {E}^+$ breaks because the inner-scaled self-similarity of $\bar {\boldsymbol {U}}$ no longer holds.
Now, the effect of $\beta$ is considered for the large $Re_{\tau }$ APG TBL from case 7. These results are directly compared with those of case 5 which is at the same $Re_{\tau }$, with their mean velocity profiles compared in figure 7(a). At this larger $Re_{\tau }$ and mild $\beta$, the inner region of the mean flow is hardly affected by the APG. As a result, $\hat {E}^+$ exhibits near-wall self-similarity in figure 7(b), like case 6. For $\hat {E}^+$, the self-similarity breaks at larger $\lambda _{z}^+$, indicating that these large-scale modes are more sensitive to the changes in the mean flow field than the near-wall small-scale modes. Just as was seen in figure 5, the large-scale modes are more amplified for the APG TBL than the ZPG TBL in the higher $Re_{\tau }$ case, similar to the energization of the outer-scaled spectral peak in the APG TBL observed in Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022).
The masked biglobal resolvent analysis approach is able to capture $Re_{\tau }$ trends seen in the spanwise energy spectra. A self-similar near-wall peak and increased amplification in the large-scale structures with $Re_{\tau }$ are observed. At matched $Re_{\tau }$, the linear amplification increases with $\beta$. For sufficiently large $Re_{\tau }$, a secondary peak is also observed. Although this analysis is is restricted to a rank-one approximation, it nonetheless suggests that changes in the linear amplification from a change in the mean flow field due to $Re_{\tau }$ and $\beta$ can explain features seen in experimental and simulation data without any nonlinear closure in the form of a scale-dependent coefficient. It can thus be surmised that these are phenomena associated with the dominant behaviour of the linear resolvent whose predictions can be improved by more sophisticated modelling of the details of the nonlinear forcing.
4. History effect on the linear amplification of large-scale modes
It was noted in several studies, and plotted in figure 2(c), that an increase in the measure of the accumulated upstream pressure gradient strength results in an increase in the streamwise turbulent fluctuations (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Tanarro et al. Reference Tanarro, Vinuesa and Schlatter2020; Pozuelo et al. Reference Pozuelo, Li, Schlatter and Vinuesa2022). In this section, the effect of the upstream pressure gradient, or history effect, on the linear amplification is studied by using the biglobal resolvent analysis. Although the PG does not enter the LNSE, the history effect is manifested in the LNSE through the pressure gradient's effect on $\bar {\boldsymbol {U}}$. The calculations in this section will use the datasets from table 1 with the numerical details described in table 3. From figure 2(a), it can be seen that all the datasets begin with different values of $\beta (x_{i})$ while at $x_{r_{2}}$ all the domains have matched $Re_{\tau }(x_{r_{2}})$ and some have similar $\beta (x_{r_{2}})$, like m16, b2 and b14. The rich $\beta$ values allow for the study of history effects and allows for the comparison of cases where the local $\beta$ and $Re_{\tau }$ are matched. Including the m18 dataset also allows for a study over a relatively large range of accumulated APG strength with $\beta _{\infty } \in [0.9,3.5]$ and the ZPG dataset with $\langle \beta _{\infty } \rangle = 0$. Here, the modes are computed using the masking described in (2.11) and the parameters described in table 3. Additionally, $L_{x} = 15.89$ to avoid the low $Re_{\tau }$ region in the database where $\beta$ ramps up quickly as seen in figure 2(a), particularly for the m18 dataset. Because the APG effects are present mostly in the outer region of the TBL, this section will focus on the history effect on the large-scale modes, like those plotted in figure 4(a–c), since they extend into the outer region of the flow.
In § 5, quantities are normalized with the length and velocity scales at $x_{r_{2}}$ and the inner products in (2.11) are used. Biglobal resolvent analysis is computed for a representative large-scale, $\lambda _{z} = 2,\lambda _t = 4.48$. Figure 8(a) demonstrates a monotonic increase in $| \sigma _{1}^+\psi _{u,1}^+(x_{r_{2}},y^+) |$ with $\langle \beta _{\infty } \rangle$, similar to the monotonic growth in $\overline {uu}^+$ with $\langle \beta _{\infty } \rangle$ in figure 2(c). Here, cases b14 and m16 are specifically highlighted since their $\beta (x_{r_{2}}) \approx 1.65$ and $Re_{\tau }(x_{r_{2}}) = 777$ while $\langle \beta _{\infty } \rangle$ differ by approximately $48\,\%$. The variation in their linear responses, $| \sigma _{1}^+\psi _{u,1}^+(x_{r_{2}},y^+) |$, evidences non-negligible sensitivity to the upstream flow conditions. This is unsurprising given the convective non-normality that amplifies these large-scale modes through upstream forcing present in this biglobal approach. It is also worth noting that the ZPG mean flow field produces a mode that peaks closer to the wall than the modes computed with the APG TBL mean flow fields. This may point to an amplification mechanism for the log-layer structures in the ZPG TBL while the APG TBL amplifies modes in the wake. However, studies with sufficiently high $Re_{\tau }$ with adequate scale separation would be needed to adequately disambiguate between the log and wake region structures. Nonetheless, the differences in the ZPG and APG response modes are likely due to the significantly different $\bar {U}$ in the ZPG TBL that produces a wake that does not follow the outer scaling for APG TBL studied in Wei & Knopp (Reference Wei and Knopp2023) as well as the negligible wall-normal convection from $\bar {V}$ in the ZPG TBL.
To show that the increase in $| \sigma _{1}^+\psi _{u,1}^+ |$ is due to the change in $\sigma$, figure 8(b) plots $| \psi _{u,1} |$, demonstrating that for the APG TBL modes, the mode shape amplitude is approximately $\langle \beta _{\infty } \rangle$ independent. The modes with increased $\langle \beta _{\infty } \rangle$ are farther from the wall due to the increased wall-normal advection throughout the domain. Figure 8(c) plots $\sigma _{j}$ for $j = 1,2,3,4,5$. Here $\sigma _{j}$ grows approximately linearly with $\langle \beta \rangle$ for $\langle \beta _{\infty } \rangle >0$, as evidenced by the lines of best fit. The $\sigma _{j}$ for $\langle \beta _{\infty } \rangle = 0$ do not fall on the line, likely because of the difference in the outer region of $\bar {\boldsymbol {U}}$ that also causes $\psi _{u,1}$ to peak closer to the wall. The effects of the wall-normal advection and non-parallel shear terms absent in the ZPG TBL will be expanded upon later in § 5.
In figure 9, $\sigma _{j}$ for $j = 1,2,3,4$ is plotted against $\langle \beta _{\infty } \rangle$ for a variety of $\lambda _{z}$ and $\lambda _{t}$, characteristic of large-scale structures. Just like in figure 8(b), the $\sigma _{j}$ grows approximately linearly with $\langle \beta _{\infty } \rangle$ such that for $\langle \beta _{\infty } \rangle >0$,
where $m_{j,0}$ and $m_{j,1}$ are $\lambda _{z}$ and $\lambda _{t}$ dependent positive constants for each gain, $\sigma _{j}$. Once again, the $\sigma _{j}$ computed with the ZPG TBL $\bar {\boldsymbol {U}}$ do not fall on this linear fit. Equation (4.1) encapsulates the influence of the history effect on the linear amplification through a streamwise averaged measure of $\beta$. As mentioned before, similar streamwise averaged measures of $\beta$ have been proposed before to account for the history effect in turbulent statistics (Vinuesa et al. Reference Vinuesa, Örlü, Sanmiguel Vila, Ianiro, Discetti and Schlatter2017) and the effect of an increased accumulated pressure gradient has been shown to intensify the streamwise fluctuations (Tanarro et al. Reference Tanarro, Vinuesa and Schlatter2020). Here, the history effect is shown to produce a monotonic increase in the linear amplification.
Now we study if the monotonic growth in $\overline {uu}^+$ with $\langle \beta _{\infty } \rangle$ is reflected in the linear growth of $\sigma _{1}$ with $\langle \beta _{\infty } \rangle$. The hybrid velocity scale, $u_{h}$, studied in Romero et al. (Reference Romero, Zimmerman, Philip and Klewicki2022) and Sekimoto et al. (Reference Sekimoto, Kitsios, Atkinson and Soria2019) showed that by rescaling the Reynolds stresses with $u_{h}^{2}$, the APG induced outer peaks are mitigated. Here $u_{h}$ augments $u_{\tau }$ with a wall-normally varying pressure velocity, $u_{p}^{2}(y) = y {{\rm d}p}/{{\rm d}\kern0.7pt x} = {\beta u_{\tau }^{2} y/\delta ^{*}}$, as
Notably, $u_{h}^{2}/u_{\tau }^{2}$ demonstrates linear growth with $\beta$, if the effect of $\beta$ on $\delta ^{*}$ is neglected. However, $u_{h}$ is not explicitly parameterized by any terms related to the history effect as it constructed from local mean flow quantities. Through the observation of linear growth of $\sigma _{1}$ with $\langle \beta _{\infty } \rangle$, an augmented hybrid velocity scale is proposed as
which simply replaces $\beta$ in (4.2) with $\langle \beta _{\infty } \rangle$. This form of $u_{hyb}$ is inspired through two assumptions in mild to moderate APG TBLs. The first is that the characteristic velocity scale close to the wall is $u_{\tau }$, regardless of $\beta$ or $\langle \beta _{\infty } \rangle$. The second is that $u_{hyb}^{2}$ has an explicit linear dependence on $\langle \beta _{\infty } \rangle$ like $\sigma _{j}$ in (4.1) because of the increased linear amplification of large-scale turbulent structures present in the outer region of the TBL. The velocity scale accounts for history effects by increasing the role of $u_{p}$ if $\langle \beta _{\infty } \rangle >\beta$ and decreasing its influence otherwise.
In figure 10(b), $\overline {uu}$ is normalized with $u_{h}^{2}$ as in Romero et al. (Reference Romero, Zimmerman, Philip and Klewicki2022) and Sekimoto et al. (Reference Sekimoto, Kitsios, Atkinson and Soria2019) demonstrating that it can remove the secondary peaks in $\overline {uu}$. However, because $u_{h}$ is a local measure, it cannot account for the monotonic growth in $\overline {uu}^+$ with $\langle \beta _{\infty } \rangle$. This is reflected with the b14 (orange) falling below the curves of b2 (blue) and m16 (purple) in the outer region despite their similar values of $\beta (x_{r_{2}})$. Furthermore, m18 (green) falls above the other curves because of its larger upstream $\beta$ values. Figure 10(c) demonstrates improved collapse across the entire boundary layer compared with $\overline {uu}/u_{h}^{2}$ by accounting for history effects. In a similar sense, figure 10(d–i) demonstrate that while rescaling $\overline {uv}$ and $\overline {vv}$ with $u_{h}^{2}$ mitigates the $\langle \beta _{\infty } \rangle$ intensified outer region turbulent fluctuations, using $u_{hyb}^{2}$ improves the collapse of the Reynolds stresses by accounting for the larger $\beta _{\infty }$ effects upstream. This serves to show that a linear dependence in $\langle \beta _{\infty } \rangle$ is reflected in the APG turbulent statistics of Bobke et al. (Reference Bobke, Vinuesa, Örlü and Schlatter2017) and Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022), not just the linear amplification of their mean flow fields. Furthermore, it suggests that a characterization of the turbulent statistics requires knowledge of the pressure gradient history.
Extending this velocity scale to other datasets requires a measurement of $\beta (x)$ over a long upstream region of the flow to compute the streamwise averages, which is often not reported. For the calculation of $\langle \beta _{\infty } \rangle$ used in $u_{hyb}$, this study averaged $\beta _{\infty }$ over a streamwise domain of $12\delta _{99}$ upstream of $x_{r_{2}}$, which coincides with the domain of integration used in the inner product in (2.11) that quantifies the linear amplification. Since, $u_{hyb}$ is based on an analogy with (4.1), which is obtained by the linear amplification in ‘near-equilibrium’ APG TBLs (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017), we caution that the scaling of the Reynolds stresses is not a general result for APG TBLs. As explained in more detail in the following section, the linear scaling on $\sigma _{j}$ is not guaranteed to hold for general non-equilibrium APG TBLs.
5. Explanation of the linear $\langle \beta _{\infty } \rangle$ dependence on the amplification
Here, it will be shown that the linear relationship between $\sigma _{j}$ and $\langle \beta _{\infty } \rangle$ can be predicted from the LNSE. In order to do so, several simplifying assumptions are made. First, it is assumed that the modes computed using the APG TBLs are independent of the mean flow field based on the observation from figure 8(b). Because of this, only the APG TBL modes will be considered since the ZPG TBL modes have significantly different behaviour since the outer region of the ZPG TBL has a different structure from the APG TBL (Wei & Knopp Reference Wei and Knopp2023). Second, the dominant terms in the LNSE are determined by comparing the advection ($\bar {U}\partial _{x}$, $\bar {V}\partial _{y}$) and mean shear ($\bar {U}_{y}$, $\bar {U}_{x}$, $\bar {V}_{x}$) terms. Specifically, the terms are compared in the outer region by using the norm
where $y_{o}^+ = 40$ since the modes of interest are large-scale modes supported in the outer region of the flow as plotted in figure 8(a). This also enables the assumption that $\| \partial _{x}\psi _{i,j} \|_{o} \sim O({\delta _{99}^{-1}\| \psi _{i,j} \|_{o}})$, $\| \partial _{y}\psi _{i,j} \|_{o} \sim O({\delta _{99}^{-1}\| \psi _{i,j} \|_{o}})$ and $k_{z} \sim O({\delta _{99}^{-1}})$ since the length scales of the large-scale modes scale with the outer length scale. Note that due to the non-dimensionalization with the outer units, $\delta _{99} \sim O({1})$. As a result, the measure of the advection terms is approximated as $\| \bar {U}\partial _{x} \| \sim O({\| \bar {U} \|_{o}})$ and $\| \bar {V}\partial _{y} \| \sim O({\| \bar {V} \|_{o}})$.
In figure 11, the shear and advection terms are compared using the norm in (5.1). Due to the increasing APG, the non-parallel terms increase in magnitude with $\langle \beta _{\infty } \rangle$. Here $\| \bar {U} \|_{o}$ increases monotonically with $\langle \beta _{\infty } \rangle$ because the norm is a measure of $\bar {U}$ upstream of $x_{r_{2}}$. Due to the streamwise deceleration in APG flows, the stronger APG have larger upstream $\bar {U}$ relative to $x_{r_{2}}$. Because of the increased boundary layer growth in APG flows, $\delta _{99}$ is significantly smaller upstream relative to $\delta _{99}(x_{r_{2}})$. These two effects create stronger upstream $\bar {U}_{y}$ with increasing $\langle \beta _{\infty } \rangle$. For the datasets studied here, $\| \bar {U}_{y} \|_{o}$ grows linearly with $\langle \beta _{\infty } \rangle$, as shown in figure 11(d). As one may expect from the typical boundary layer assumptions (Pope Reference Pope2000), the streamwise advection is an order of magnitude larger than the wall-normal advection while $\bar {U}_{y}$ is at least an order of magnitude larger than $\bar {U}_{x}$ and $\bar {V}_{x}$ for the APG TBL datasets used herein.
To see if the non-parallel terms $\bar {V}$, $\bar {U}_{x}$ and $\bar {V}_{x}$ are responsible for the linear growth of $\sigma _{j}$ with $\langle \beta _{\infty } \rangle$, the non-parallel terms are artificially set to $0$ in the LNSE. Although the streamwise derivatives are not explicitly included in the LNSE, $\bar {U}$ and $\bar {U}_{y}$ are allowed to vary in $x$ for this analysis. In figure 11(f), this assumption is tested for the same modes plotted in figure 8(a) by computing $\tilde {\sigma }_{1}$, the leading singular values of the biglobal resolvent operator with the non-parallel terms set to $0$. By neglecting the non-parallel terms, the linear growth with $\langle \beta _{\infty } \rangle$ is still observed in $\tilde {\sigma }_{1}$. The non-parallel terms are a non-negligible source of amplification for the linear operator as the difference between $\sigma _{1}$ and $\tilde {\sigma }_{1}$ increases with $\langle \beta _{\infty } \rangle$. However, their omission in the analysis can still be used to explain the growth with $\langle \beta _{\infty } \rangle$. As a consequence, an additional assumption is that the non-parallel terms can be neglected in the explanation of the linear growth of $\sigma _{1}$ with $\langle \beta _{\infty } \rangle$.
In this section, the linear growth of $\sigma _{1}$ with $\langle \beta _{\infty } \rangle$ is first investigated using the assumption that the non-parallel terms can be neglected. Then the effects of the non-parallel terms on the linear amplification will be discussed. Finally, the section concludes with implications of this analysis for more general APG TBLs.
5.1. Neglecting the non-parallel terms to explain the linear growth
By neglecting $\bar {V}$, $\bar {V}_{x}$ and $\bar {U}_{x}$, the LNSE in (2.2) can be approximated using the simplified forced biglobal OSS system,
where $\mathcal {L}_{SQ} = \!-\!\mathrm {i}\omega + \bar {U}\partial _{x} - Re^{-1}\nabla ^{2}$, $\mathcal {L}_{OS} = (-\mathrm {i}\omega + \bar {U}\partial _{x} - ({1}/{Re})\nabla ^{2}) -2\nabla ^{-2}\partial _{y}(\bar {U}_{y}\partial _{x})$, $\hat {f}_{\omega } = \mathrm {i} k_{z}\hat {f}_{x} - \partial _{x}\hat {f}_{z}$ and $\hat {\omega }_{2} = \mathrm {i} k_{z}\hat {u} - \partial _{x}\hat {w}$. See the Appendix for more details. The resolvent gains of this system will then be denoted as $\tilde {\sigma }_{j}$. The forced biglobal OSS system has been studied by Ran et al. (Reference Ran, Zare, Hack and Jovanović2019) to identify the stochastic receptivity of a laminar boundary layer using the LNSE. In that case, terms involving $\bar {U}_{xx}$ and $\bar {V}_{x}$ were neglected, consistent with the Blasius similarity solution. Equations (5.2) and (5.3) make a stronger assumption by neglecting the non-parallel shear terms and wall-normal advection all together.
The next approximation comes from the observation that the most linearly amplified modes have small wall-normal components relative to the spanwise and streamwise components. This is plotted in figure 12 where $\| \psi _{v,1} \|_{r}$ contributes less than $5\,\%$ of $\| \boldsymbol {\psi }_{1} \|_{r}$. As a result, the contribution from (5.2) is negligible in the numerator of (2.7) and $\| \boldsymbol {\psi }_{1} \|_{r}^{2} \approx \| \boldsymbol {\psi }_{u,1} \|_{r}^{2} + \| \boldsymbol {\psi }_{w,1} \|_{r}^{2} = \| (\partial _{xx} - k_{z}^{2})\boldsymbol {\psi }_{\omega,1} \|_{r}^{2}$, where $\psi _{\omega,1} = \mathrm {i} k_{z}\psi _{u,1} - \partial _{x}\psi _{w,1}$ (see the Appendix). Because it was assumed that $k_{z}$ and the streamwise derivatives are $O(1)$, it can be further assumed that $\| \boldsymbol {\psi }_{1} \|_{r}^{2} \sim \| \boldsymbol {\psi }_{\omega,1} \|_{r}^{2}$. By using the approximations that $\psi _{v,1}$ is negligible and that only outer region forcing is optimal, the linear amplification can be approximated as
Through application of operator norm inequalities and the triangle inequality using $\| \hat {\omega }_{2} \|_{o}$ and (5.3), the approximation $\tilde {\sigma }_{1}$ is bounded above as
Equation (5.5) gives an upper bound to the approximation of $\tilde {\sigma }_{1}$ based on linear growth with $\bar {U}_{y}$. Similarly, if the forcing is constrained to only wall-normal forcing, it must be the case that
In other words, the largest possible gain that stems from solely wall-normal forcing is less than the optimal gain. Equation (5.6) can be written explicitly by setting $\hat {f}_{\omega } = 0$ in (5.3) to establish the lower bound as
Together, (5.5) and (5.7) suggest that $\tilde {\sigma }_{1}$ is bounded below by a quantity that scales with $\| \bar {U}_{y} \|_{o}$ and bounded above by a quantity that grows linearly with $\| \bar {U}_{y} \|_{o}$.
It must be noted that the $\bar {U}_{y}$ dependence in both (5.6) and (5.7) also depends on the amplification from $\mathcal {L}_{OS}^{-1}$ and $\mathcal {L}_{SQ}^{-1}$, measured in the outer region of the flow. To show that the amplification from $\mathcal {L}^{-1}_{SQ}$ will not produce a significant additional $\bar {\boldsymbol {U}}$ dependence on the upper and lower bounds, its resolvent gains, $\sigma _{j,SQ}$, are plotted in figure 13(a) for the representative mode from figure 8. Recall that $\sigma _{1,SQ} = \| \mathcal {L}_{SQ}^{-1} \| \approx \| \mathcal {L}_{SQ}^{-1} \|_{o}$. Here $\sigma _{j,SQ}$ has a slight $\langle \beta _{\infty } \rangle$ dependence for the optimal modes, where $\sigma _{1,SQ}$ grows by $18\,\%$ between $\langle \beta _{\infty } \rangle = 0$ to $\langle \beta _{\infty } \rangle = 3.47$. For reference, $\sigma _{1}$ grows by $187\,\%$ over the same $\langle \beta _{\infty } \rangle$ range. For the higher-order modes, the $\sigma _{j,SQ}$ variation with $\langle \beta _{\infty } \rangle$ is negligible compared with the variation in $\sigma _{5}$. Furthermore, $\mathcal {L}^{-1}_{SQ}$ is not low-rank as the first five $\sigma _{j,SQ}$ are all within the same order of magnitude. This means that it is more than likely that $(\,\hat {f}_{\omega } + \mathrm {i} k_{z}\bar {U}_{y}\mathcal {L}_{OS}^{-1}\;\hat {f}_{v})$ in (5.3) is composed of higher-order forcing modes of $\mathcal {L}^{-1}_{SQ}$ that have weak $\bar {\boldsymbol {U}}$ dependence. As a result, it can be concluded that the upper bound in (5.5) is weakly dependent on the $\bar {\boldsymbol {U}}$-dependence on $\| \mathcal {L}_{SQ}^{-1} \|_{o}$.
To estimate the amplification from $\mathcal {L}^{-1}_{OS}$, componentwise masking is applied to consider the amplification from $\hat {f}_{v}$ to $\hat {v}$ so as to only consider the $\mathcal {L}_{OS}^{-1}\;\hat {f}_{v}$ term in (5.2). These resolvent gains are denoted as $\sigma _{j,OS}$ and are plotted in figure 13(b). Here $\sigma _{j,OS}$ demonstrates a slight decrease with $\langle \beta _{\infty } \rangle$ with a decrease of approximately $20\,\%$ between the $\langle \beta _{\infty } \rangle$ range. The $\sigma _{j,OS}$ also demonstrate a weak $\bar {\boldsymbol {U}}$ dependence on its leading gains. Although the amplification has contributions from $\mathcal {L}_{SQ}^{-1}$ and $\mathcal {L}_{OS}^{-1}$, it is expected that the $\bar {\boldsymbol {U}}$-dependence on the amplification only enters the $\bar {U}_{y}$ term.
To show that the $\bar {\boldsymbol {U}}$ dependence on $\mathcal {L}_{SQ}^{-1}$ and $\mathcal {L}_{OS}^{-1}$ does not affect the lower bound, figure 13(c) plots $\sigma _{1,v\rightarrow (u,w)}$, the resolvent gains when componentwise masking is applied to consider only forcing from $\hat {f}_{v}$ to a response with $\hat {u}$ and $\hat {w}$. This mimics the conditions in (5.7) which captures only amplification from the lift-up mechanism. Here, it is found that despite the $\langle \beta _{\infty } \rangle$ variation in the amplification of $\mathcal {L}_{SQ}^{-1}$ and $\mathcal {L}_{OS}^{-1}$, the growth in $\sigma _{1,v\rightarrow (u,w)}$ is linear with $\| \bar {U}_{y} \|_{o}$. Fitting a line to $\sigma _{1,v\rightarrow (u,w)}$ as
where $T_{j}$ is a constant for each $j$ in figure 13(c), shows that the approximations made in this section can explain the amplification from the lift-up mechanism. Furthermore, it demonstrates that any $\bar {\boldsymbol {U}}$ dependence on $\mathcal {L}_{OS}^{-1}$ and $\mathcal {L}_{SQ}^{-1}$ does not affect the linear growth with $\| \bar {U}_{y} \|_{o}$. The differences in the mode shapes are not accounted for in (5.8), which may explain why the lines of best fit do not cross $\sigma _{1,v\rightarrow (u,w)}$ for the ZPG TBL.
The linear fit in (5.8) holds when $\hat {f}_{\omega } = 0$ such that the assumption in (5.7) holds. For a more general case, a new linear fit is proposed based on the upper bound in (5.5) such that
where $a_{j,i}(\lambda _{z},\lambda _{t})$ are scale-dependent constants for each $j$. The coefficients $a_{j,0}$ and $a_{j,1}$ are analogous to $\| \mathcal {L}^{-1}_{SQ} \|_{o}\| \;\hat {f}_{\omega } \|_{o}$ and $k_{z}\| \mathcal {L}^{-1}_{SQ} \|_{o}\| \mathcal {L}^{-1}_{OS}\;\hat {f}_{v} \|_{o}$, respectively. These linear fits are shown in figure 14 for the same $\sigma _{j}(\lambda _{z},\lambda _{t})$ plotted in figure 9 demonstrating that there is indeed linear growth in $\sigma _{j}$ with $\| \bar {U}_{y} \|_{o}$ for the APG TBLs as suggested by the upper bound. Furthermore, figure 12(c) shows that the APG TBL datasets used here have mean flow fields such that
where $d_{0}$ and $d_{1}$ were constants. Taking both (5.9) and (5.10) recovers the linear form in $\sigma _{j}$ with $\langle \beta _{\infty } \rangle$ from (4.1). Hence, it can be concluded that the linear growth of $\sigma _{j}$ in $\| \bar {U}_{y} \|_{o}$ for the large-scale structures can be predicted from the LNSE if the large-scale modes are supported in the outer region of the flow. Equation (5.9) stems from the dominant terms of the LNSE while the linear fit in (4.1) requires that (5.10) holds, which is not guaranteed for general APG TBLs. Although the linear fit for the APG TBL modes was motivated by neglecting the non-parallel terms by considering $\tilde {\sigma }_{1}$; $\sigma _{1}$ is significantly larger than $\tilde {\sigma }_{1}$. This indicates that the non-parallel terms are non-negligible. Next, the influence of the non-parallel terms on the amplification will be examined.
5.2. Influence of the non-parallel terms on the linear amplification
Although artificially neglecting the non-parallel terms in the LNSE can still explain the linear growth with $\langle \beta _{\infty } \rangle$, figure 11(f) demonstrates that the non-parallel terms are a non-negligible source of amplification. Here, the effect of the non-parallel terms will be examined to discuss their role in the linear amplification. Equations (5.2) and (5.3) are now reconsidered, except that the $\mathcal {L}_{SQ}$ operators are augmented to incorporate the $\bar {U}_{x}$ terms as $\tilde {\mathcal {L}}_{SQ} = \mathcal {L}_{SQ} + \bar {U}_{x}$ while the $\mathcal {L}_{OS}$ operators are left unchanged because their amplification was shown to be approximately independent of $\langle \beta _{\infty } \rangle$ when using the full LNSE in figure 13(b). Incorporating $\bar {U}_{x}$ also introduces off-diagonal terms in the full OSS, though they are negligible relative to the mean shear, $k_{z}\| \bar {U}_{y} \|_{o}$. Since only $\mathcal {L}_{SQ}$ has changed, an upper and lower bound for $\sigma _{1}$ can be constructed by following the steps in § 5.1 as
The effect of $\bar {U}_{x}$ is to augment the amplification of $\tilde {\mathcal {L}}_{SQ}$. This can be seen by computing the resolvent amplification of $\tilde {\mathcal {L}}_{SQ}$, $\tilde {\sigma }_{1,SQ}$ and comparing with the $\sigma _{j,SQ}$ of $\mathcal {L}_{SQ}$ in figure 15(a). It can be seen that $\tilde {\sigma }_{j,SQ}>\sigma _{j,SQ}$ for $j = 1,2,3,4,5$, indicating that the inclusion of $\bar {U}_{x}$ increases the amplification by $10\,\%$. This increase in amplification then manifests itself as an increase in the amplification via the lift-up effect. To show this, $\sigma _{j,(v \rightarrow u,w)}$ is compared with $\tilde {\sigma }_{j,(v \rightarrow u,w)}$, the resolvent amplification from $\hat {f}_{v}$ to $(\hat {u},\hat {w})$ for the LNSE neglecting $\bar {U}_{x},$ $\bar {V}_{x}$ and $\bar {V}$, in figure 15(b), demonstrating that $\sigma _{j,(v \rightarrow u,w)}>\tilde {\sigma }_{j,(v \rightarrow u,w)}$. From the lines of best fit, the rate of growth with $\| \bar {U}_{y} \|_{o}$ is larger for the full LNSE ($\sigma _{j,(v \rightarrow u,w)}$) than by neglecting the non-parallel terms ($\tilde {\sigma }_{j,(v \rightarrow u,w)}$). From the lower bound in (5.11), the increased rate of growth stems from the increase in amplification of $\tilde {\mathcal {L}}_{SQ}^{-1}$ since it was argued that $\mathcal {L}_{OS}^{-1}$ was hardly affected by the change in base flow. This increased rate of growth with $\| \bar {U}_{y} \|_{o}$ then manifests itself as a larger $a_{j,1}$ from (5.9). This can be seen in figure 15(c) where $\bar {\sigma }_{1}$ is computed for the LNSE where only $\bar {V}$ and $\bar {V}_{y}$ are neglected. By including $\bar {U}_{x}$, $\bar {\sigma }_{1} >\tilde {\sigma }_{1}$. Since $\mathcal {L}_{SQ}^{-1}$ also amplifies $\hat {f}_{\omega }$, the magnitude of $a_{j,0}$ is also increased when $\bar {U}_{x}$ is included.
The inclusion of the $\bar {V}$ terms provide wall-normal advection that is neither present in the ZPG TBL, nor in LNSE with artificially neglected $\bar {V}$. This can be seen in figure 16(a,b), where $| \psi _{1,u} |$ is compared using the full LNSE and the LNSE with $\bar {V}$ and $\bar {V}_{x}$ artificially set to $0$. It can be seen that the modes computed with the full LNSE reach their peak amplitude farther from the wall than the artificial modes. As the APG strength increases, the disparity between the two sets of modes increases, as can be seen by comparing the m13 and b1n modes. As expected, the ZPG modes are almost indistinguishable. Because the modes computed with stronger APG are farther from the wall, they experience a stronger streamwise varying $\bar {U}$. As a result, the modes computed with the full LNSE experience a stronger convective non-normality and thus increased amplification from the $\mathcal {L}_{SQ}^{-1}$ term. As a result, for $\sigma _{1}$, both $a_{j,1}$ and $| a_{j,0} |$ are the largest in figure 15(c), indicating a further increase in amplification when none of the non-parallel terms are neglected.
6. Discussion
Biglobal resolvent analysis in a long streamwise domain permits the study of pressure gradient history and the effects of streamwise growth (non-parallel effects) in the APG TBL on linear amplification. Because this analysis derives from the LNSE, it is possible to analytically track how these effects manifest themselves in the linear amplification. Through a scale-based analysis, the linear growth in $\sigma _{j}$ with $\langle \beta _{\infty } \rangle$ follows from the behaviour of the mean flow fields in the near-equilibrium APG TBLs through (5.10) and the role of the componentwise amplification in the upper and lower bounds in (5.5) and (5.7).
Even by including the non-parallel terms, the resolvent amplification was still shown to grow linearly with $\| \bar {U}_{y} \|_{o}$, for the APG TBLs rather than the ZPG TBL. This was likely due to the differences in the outer region of the mean flow fields (Wei & Knopp Reference Wei and Knopp2023), which in turn changed where the modes are localized in the ZPG case. For the APG TBLs, the mean flow fields did not significantly affect the resolvent mode shape.
Extending the linear growth of $\sigma _{j}$ to $\langle \beta _{\infty } \rangle$ in (4.1) requires that both (4.1) and (5.9) hold. It is expected that (4.1) is more sensitive to the choice of $\bar {\boldsymbol {U}}$ than (5.9) because the former requires that both (5.10) and (5.9) hold. The datasets studied herein were ‘near-equilibrium’ TBL (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017) at relatively similar $Re_{\tau }$ and similar free stream boundary conditions such that $U_{\infty } \sim (x-x_{d})^{m}$ with mild to moderate $\beta$. Due to the similarity in the APG TBLs, (4.1) held well for the APG TBL datasets. However, it is expected that for datasets with strong APG, the non-parallel terms may be non-negligible and create sources of amplification that can cause the linear fit with $\| \bar {U}_{y} \|_{o}$ to fail. Furthermore, it is expected that APG TBL with non-equilibrium conditions, like strongly varying $\beta (x)$ or flow separation, could make (4.1) not hold since $\bar {\boldsymbol {U}}$ would experience significant non-parallel terms which could cause (5.9) to fail.
While the explanations in § 5 were used to explain why $\sigma _{1}$ grows monotonically with $\langle \beta _{\infty } \rangle$, they can also be used to explain the secondary peaks of $\hat {E}^+$ that become more amplified with increasing $\beta (x_{r_{1}})$ present in figure 5. Since the datasets used in figure 5 were characterized by regions with nearly constant values of $\beta (x)$, this discussion will approximate $\langle \beta _{\infty } \rangle$ as $\beta (x_{r_{1}})$. The integrand of $\hat {E}^+$, $| \sigma _{1}^+\psi _{u,1}^+ |^{2}$, scales differently for the near-wall small-scale modes and the outer-region large-scale modes. First, due to the choice of normalization, $\sigma ^+_{j} = Re_{\tau }^{2}(x_{r_{1}})Re^{-1}(x_{r_{1}})\sigma _{j}$ and $\psi _{u,j}^+ = \sqrt {Re_{\tau }}\psi _{u,j}$. For the small-scale modes, it can be shown that if $\bar {U}^+$ is self-similar where $\boldsymbol {\psi }_{1}$ is supported, then $\sigma _{1}^+$ and $\psi _{u,1}^+$ are also self-similar (Moarref et al. Reference Moarref, Sharma, Tropp and McKeon2013; Gomez Reference Gomez2024). In the case of a mild APG TBL at low $Re_{\tau }$, the log-layer is affected by the APG, which affects the self-similarity of the near-wall modes, especially the higher-order modes (Gomez Reference Gomez2024). However, for the leading modes, these near-wall APG effects were small enough that the near-wall linear amplification of the APG TBL could still resemble that of the ZPG TBL. As a result, the integrand in $\hat {E}^+$, $| \sigma _{1}^+\psi _{1,u}^+ |^{2}$, is only slightly affected by $\beta$. While this study focused on mild to moderate APGs, strong APGs that can cause separation are expected to significantly affect the near-wall amplification.
For the large-scale modes, the small variation in $\| \psi _{u,1} \|_{r}$ with $\langle \beta _{\infty } \rangle$ in figure 12(a) allows the approximation that the amplitude of the large-scale modes, $| \psi _{u,1} |$, is independent of $\langle \beta _{\infty } \rangle$. Although figure 8 showed that the increased $\bar {V}$ in stronger APG TBL advected the $\psi _{u,1}$ farther from the wall, this effect will be neglected for this discussion since it is only focused on the magnitude of the modes rather than their shape. As a result, the amplitudes of the large-scale modes, $| \psi _{u,1}^+ |$, are approximated to be independent of $\langle \beta _{\infty } \rangle$ in the calculation of $E^+$. Following (4.1),
Although $Re(x_{r_{1}})$ increases with $\langle \beta _{\infty } \rangle$, it is negligible compared with the $1 + m_{j,1}(\lambda _{t},\lambda _{z})\langle \beta _{\infty } \rangle$ term such that $\sigma _{j}^+$ increases with $\langle \beta _{\infty } \rangle$. Thus, the increased amplification of large-scale modes in $\hat {E}^+$ can be explained solely by the increase in $\sigma _{1}$ with $\langle \beta _{\infty } \rangle$ for the large-scale modes.
The biglobal resolvent approach is able to qualitatively capture various effects seen in APG TBL. For example, the large-scale energization with increasing $\beta$ is reflected in the increased large-scale amplification with increasing $\beta$ along with the near-wall self-similarity of the small-scale structures. Of particular interest is how the history effect is shown to affect both the large-scale linear amplification and the outer region Reynolds stresses (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Tanarro et al. Reference Tanarro, Vinuesa and Schlatter2020). One way in which the monotonic growth with $\langle \beta _{\infty } \rangle$ enters the outer region turbulent fluctuations is through the linear growth in $\sigma _{j}$ with $\langle \beta _{\infty } \rangle$. This means that the linear amplification of large-scale nonlinear interactions by the APG mean flow fields increases with $\langle \beta _{\infty } \rangle$, which may be reflected in the large-scale turbulent fluctuations. The collapse of the Reynolds stresses with $u_{hyb}^{2}$, which is linear in $\langle \beta _{\infty } \rangle$, may point to a connection between the linear amplification studied herein and the true nonlinear nature of the APG TBLs.
7. Conclusion
A biglobal resolvent approach was used to study spatially developing TBL flows over flat plates and identify the relative influences of non-parallel componentwise non-normalities from $\bar {V}_{x}$, convective non-normalities, wall-normal advection and the pressure gradient history effect on linear amplification in near-equilibrium mild to moderate APG TBL (Bobke et al. Reference Bobke, Vinuesa, Örlü and Schlatter2017; Pozuelo et al. Reference Pozuelo, Li, Schlatter and Vinuesa2022) and high $Re_{\tau }$ ZPG TBL (Eitel-Amor et al. Reference Eitel-Amor, Örlü and Schlatter2014). The effects of $Re_{\tau }$, $\beta$ and $\langle \beta _{\infty } \rangle$ on the analysis were investigated.
Applying a spatial mask to measure the response amplification in the vicinity of $x_{r}$ allowed both small-scale and large-scale modes to have a non-zero response at $x_{r}$. This allowed for the study of local $\beta$ effects at fixed $Re_{\tau }$ and local $Re_{\tau }$ effects on the linear amplification. Furthermore, the premultiplied amplification, $\hat {E}(y,k_{z})$, is used as a rank-one approximation of the premultiplied energy spectra for a velocity fluctuation defined as $\hat {u} = \sigma _{1}\psi _{u,1}$. Although it does not have any nonlinear closure, $\hat {E}$ was shown to reflect certain characteristics seen in the premultiplied energy spectra in simulations. One feature is the presence of a near-wall peak using both the ZPG and APG TBL base flows. The inner peak was shown to be self-similar for $\bar {\boldsymbol {U}}$ with near-wall self-similarity such as the ZPG TBL datasets and the high $Re_{\tau }$ APG TBL of Pozuelo et al. (Reference Pozuelo, Li, Schlatter and Vinuesa2022). A secondary outer peak for large $\lambda _{z}$ is also present for the linear amplification of the low $Re_{\tau }$ APG TBLs of Bobke et al. (Reference Bobke, Vinuesa, Örlü and Schlatter2017) and is notably absent in the low $Re_{\tau }$ ZPG TBL. For this secondary peak, $\hat {E}^+$ was shown to increase with $\beta (x_{r_{1}})$. Similarly, for the ZPG TBL, $\hat {E}^+$ was shown to increase with $Re_{\tau }$ for the large $\lambda _{z}$ away from the wall and demonstrate evidence of a secondary peak for the largest $Re_{\tau }$ interrogated.
By applying a spatial mask to consider only the region upstream of $x_{r_{2}}$ in the inner products, the biglobal resolvent analysis was used to investigate how the linear amplification varies due to the upstream history. This history is measured through $\langle \beta _{\infty } \rangle$, the upstream average of $\beta _{\infty }$ over the same computational domain. It was found that the $\sigma _{j}$ increased linearly with $\langle \beta _{\infty } \rangle$. By drawing similarities to the linear variation in $\beta$ in the hybrid velocity scale of Romero et al. (Reference Romero, Zimmerman, Philip and Klewicki2022) and Sekimoto et al. (Reference Sekimoto, Kitsios, Atkinson and Soria2019), an augmented hybrid velocity scale, $u_{hyb}$ was posed that incorporates a measure of the history effect through a linear dependence with $\langle \beta _{\infty } \rangle$. Rescaling the Reynolds stresses with $u_{hyb}^{2}$ mitigated the APG-related outer peaks and improved the collapse of the Reynolds stresses compared with $u_{h}$. This linked the monotonic growth in the secondary outer peaks of $\overline {uu}^+$ with $\langle \beta _{\infty } \rangle$ with the linear growth of $\sigma _{j}$ with $\langle \beta _{\infty } \rangle$.
Finally, an explanation based on the lift-up mechanism is offered to explain the increase in $\sigma _{1}$. It was shown that neglecting the $\bar {U}_{x}$, $\bar {V}_{x}$ and $\bar {V}$ terms in the LNSE still allowed for linear growth with $\langle \beta _{\infty } \rangle$, albeit with reduced amplification than the full system. As a result, a simplified two-dimensional OSS system was constructed. By using the triangle inequality, it was shown that for this simplified system, $\sigma _{1}$ was bounded below and above by terms that grew linearly with $\| \bar {U}_{y} \|_{o}$, a measure of the mean shear in the outer region of the flow. Thus, the equations suggest the $\sigma _{1}$ grows linearly with $\| \bar {U}_{y} \|_{o}$. Furthermore, it was shown that for these datasets, $\| \bar {U}_{y} \|_{o}$ also grows linearly with $\langle \beta _{\infty } \rangle$. As a result, $\sigma _{1}$ was shown to grow linearly with $\langle \beta _{\infty } \rangle$ for the APG TBL flows from an analysis of the LNSE and a property of the datasets used herein. It was also shown that the non-parallel terms add additional amplification by increasing the role of the convective non-normality. The inclusion of the non-parallel shear terms only affected the linear fit with $\| \bar {U}_{y} \|_{o}$ by increasing the magnitude of the linear fit's coefficients.
The key result of the study is that several features associated with APG TBLs can be identified through study of the linear operator by consideration of the optimal amplification, i.e. using a rank-one approximations. Changes in free stream $\bar {\boldsymbol {U}}$ and streamwise pressure gradient $\bar {P}$ are encoded in the mean fields and thus change the amplification associated with the resolvent operator. It is encouraging for modellers that these effects can be identified without detailed knowledge of the nonlinear forcing, which would require a large amount of data to cover the range of APGs investigated here. However, any resolvent reconstruction will require nonlinear closure through scale-dependent coefficients and higher-order modes, as is well known for the ZPG and other flow cases.
Most of this study focused on APG TBL in regimes that can be described as near-equilibrium and at lower $Re_{\tau }$, in accordance with the data available in the literature. Extending these results to high $Re_{\tau }$ APG TBL would require an APG TBL database over a large streamwise domain. Future studies will be used to investigate APG TBL where the $\beta$ variation is more severe or where separation is approached.
Acknowledgements
We gratefully acknowledge S. Romero and J. Klewicki for discussions regarding the hybrid velocity scaling, and R. Vinuesa for sharing the TBL database.
Funding
This work is supported by the U.S. Office of Naval Research through grants N00014-17-1-2307 and N00014-17-1-3022.
Declaration of interests
The authors report no conflict of interest.
Appendix. Simplified biglobal Orr–Sommerfeld operator
For brevity, the sponge layer terms are ignored with their effect only used to enforce compactness in the forcing and response modes. It is noted that the optimal forcing modes are solenoidal vectors (Rosenberg Reference Rosenberg2018) such that
The wall-parallel velocities can be written in terms of $\hat {\omega }_{2}$ and $\hat {v}$ as
Assuming compact support in $\hat {\omega }_{2}$ and $\hat {v}$, in turn making $(\partial _{xx}-k_{z}^{2})^{-1}$ self-adjoint, the kinetic energy is then
An equation for $\hat {v}$ can be written by substituting (A1) for $\hat {p}$ in the wall-normal component of the LNSE as
The terms proportional to $\bar {V}_{x}$ are not expanded since they will be neglected later. Similarly, by taking the wall-normal component of the curl of the LNSE, an equation for $\hat {\omega }_{2}$ can be written as follows:
By using the observation that $| \bar {V}_{x} |\ll | \bar {U}_{x} |$, (A5) and (A6) can be written as
This linear system can be solved for explicitly as
Together, (A9) forms the OSS system, with $\bar {V}_{x}$ neglected. Due to the $\mathcal {P}_{s}$ term, there is an off-diagonal entry that introduces componentwise non-normality into the OSS system by allowing forcing from $\hat {f}_{2}$ to $\hat {v}$ caused by $\bar {U}_{x}$. However, for the APG TBL studied in this section, the presence of $\mathcal {P}_{s}$ does not cause $\| \psi _{1,v} \|$ to increase as the APG strength increases. This is likely because the off-diagonal component related to the mean shear scales with $\bar {U}_{y}$ while the other off-diagonal term scales with $\bar {U}_{x}$. Due to this, the $\mathcal {A}_{OS}^{-1}\mathcal {P}_{s}(\mathcal {M}_{s}^{-1}\mathcal {A} -\mathcal {A}_{OS}^{-1}\mathcal {P}_{s})^{-1}\mathcal {M}_{s}^{-1}$ term will be neglected. Next, it is noted that $\mathcal {M}_{s}^{-1}\mathcal {A}_{SQ} \sim O({\delta _{99}})$ while $\mathcal {A}_{OS}^{-1}\mathcal {P}_{s} \sim O({\delta _{99}U_{\infty } {\partial U_{\infty }}/{\partial x}})$. As a result, the $\mathcal {A}_{OS}^{-1}\mathcal {P}_{s}$ term can be neglected such that the following simplified OSS can be established:
Finally, by neglecting the $\bar {U}_{x}$ and $\bar {V}$ terms, the simplified OSS system in (5.2) and (5.7) can be recovered. For clarity,
and