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

An extension of Thwaites’ method for turbulent boundary layers

Rahul Agrawal\aff1 \corresp [email protected]    Sanjeeb T. Bose\aff2    \aff3 Kevin P. Griffin\aff4    Parviz Moin\aff1 \aff1 Center for Turbulence Research, Stanford University, CA \aff2 Cadence Design Systems, San Jose, CA \aff3 Institute for Computational and Mathematical Engineering, Stanford University, CA \aff4National Renewable Energy Laboratory, Golden, CO
Abstract

Thwaites (1949) developed an approximate method for determining the evolution of laminar boundary layers. The approximation follows from an assumption that the growth of a laminar boundary layer in the presence of pressure gradients could be parameterized solely as a function of a flow parameter, m=θ2/νdUedsm=\theta^{2}/\nu\frac{dU_{e}}{ds}, thus reducing the von Karman momentum integral to a first-order ordinary differential equation. This method is useful for the analysis of laminar flows, and in computational potential flow solvers to account for the viscous effects. However, for turbulent flows, a similar approximation for turbulent boundary layers subjected to pressure gradients does not yet exist. In this work, an approximate method for determining the momentum thickness of a two-dimensional, turbulent boundary layer is proposed. It is shown that the method provides good estimates of the momentum thickness, when compared to available high-fidelity simulation data, for multiple boundary layers including both favorable and adverse pressure gradient effects, up to the point of separation. In the limit of high Reynolds numbers, it is possible to derive a criterion for the onset of separation from the proposed model which is shown to be in agreement with prior empirical observations (Alber, 9th9^{th} Aerospace Sciences Meeting, 1971). The sensitivity of the separation location with respect to upstream perturbations is also analyzed through this model for the NASA/Boeing speed bump and the transonic Bachalo-Johnson bump.

keywords:

1 Introduction

Often in the early stages of the engineering design processes, viscous effects are estimated using boundary layer integral methods. Common elements in many such engineering flows are that boundary layers are turbulent and subjected to strong pressure gradients - both favorable and adverse. A pressure gradient effect can alter the transition location, or even cause the flow to separate, leading to significant changes in the flow patterns. Clauser (1954, 1956) studied turbulent boundary layers and defined a parameter, β\beta that quantifies the relative strength of the pressure gradient in relation to the skin friction across the boundary layer,

β=δρuτ2dPds\beta=\frac{\delta^{*}}{\rho u_{\tau}^{2}}\frac{dP}{ds} (1)

where δ\delta^{*} is the displacement thickness, dP/dsdP/ds is the pressure gradient and uτu_{\tau} is the friction velocity. This parameter has since been used in the analyses of equilibrium and non-equilibrium boundary layers. In fact, in the presence of strong, and/or prolonged favorable pressure gradients, the flow may relaminarize (Sreenivasan, 1982); thus regaining an ordered structure with a reduction in Reynolds stresses. On the other hand, in the presence of adverse pressure gradients, it is possible for the flow to eventually separate (Simpson, 1981, 1983), often associated with periods of rapid growth of the boundary layer. An adverse pressure gradient significantly energizes the outer flow (Knopp et al., 2021), and increases the momentum deficit compared to the freestream value. In fact, the effect of a pressure gradient is not only local but strongly dependent on its history (Bobke et al., 2017; Parthasarathy & Saxton-Fox, 2023; Prakash et al., 2023). Wei et al. (2017) derived the relationships between the turbulent shape factor and the quantities at the edge of the boundary layer of an adverse pressure gradient boundary layer and demonstrated that the boundary layer thicknesses, its growth rate, and the resulting skin friction are affected by upstream effects. Other experimental and computational investigations include the works of Harun et al. (2013); Skote et al. (1998); Knopp et al. (2015); Kitsios et al. (2017); Romero et al. (2022); these studies have indicated that the adverse pressure gradient energizes both the large-scale and the smaller-scale structures, while also modulating the impact of the large scales on the smaller scales.

In the limit of thin boundary layers, the growth rate of the momentum thickness is related to the pressure gradient as follows (via the Karman momentum integral, see White (2008) for more details),

dθds=Cf2+(2+H)θUe2dPeds\frac{d\theta}{ds}=\frac{C_{f}}{2}+(2+H)\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds} (2)

where θ\theta is the momentum thickness, and CfC_{f} is the skin-friction along the streamwise coordinate. HH is the shape factor which is the ratio of the displacement and the momentum thickness lengths of the boundary layer. However, this equation, in itself, is not particularly useful as a predictive tool and requires the knowledge of the skin friction and shape function distributions, which may be unknown a priori.

For laminar boundary layers, Thwaites (1949), and later Curle & Skan (1957), and Dey & Narasimha (1990) fitted data for various flows, including those with pressure gradients and flows undergoing separation, to derive an approximate expression for the growth of the momentum thickness for a given inviscid (freestream) velocity distribution. This analysis coupled with the Falkner & Skan (1931) analysis in the viscous layer provides a complete description of the mean velocity profile of the corresponding laminar flow. However, a similar analysis for turbulent boundary layers is more complex and lacking. Weber (1978) developed a semi-empirical method for calculating the thickness of turbulent boundary layers with arbitrary pressure gradients for the AFOSR-IFP-Stanford conference (Coles & Hirst, 1968; Kline et al., 1969). However, his approach relied on implied correlations between the skin friction, shape factor, and an inviscid pressure gradient parameter, some of which may be unknown, especially in complex engineering flows. Das (1987), Das & White (1986), Kalkhoran & Wilson (1986) developed integral approaches for measuring skin friction in turbulent boundary layers. A drawback of Das (1987)’s and Das & White (1986)’s approaches included the use of multiple correlations while also not accounting for history effects (through appropriate streamwise derivatives). Kalkhoran & Wilson (1986) made equilibrium assumptions for the wake and the pressure gradient even for non-equilibrium boundary layers. A comprehensive review of these approaches can be found in Das et al. (2004).

From a practical standpoint, the effects associated with pressure gradients pose significant modeling challenges for near-wall models that are employed to predict wall-shear stress in scale-resolving methods. In the past, various modeling efforts have been pursued (Balaras et al., 1996; Wang & Moin, 2002; Duprat et al., 2011; Hickel et al., 2013; Park & Moin, 2014) to include the non-equilibrium effects due to pressure gradients. However, these models have not shown significant differences in their predictive performance in comparison to the equilibrium wall model (Cabot & Moin, 2000; Kawai & Larsson, 2012) across a wide range of flows. Park & Moin (2014) developed a non-equilibrium wall model while including the full two-dimensional RANS equations, which adds to the computational complexity. This model was shown to have limited improvements over the equilibrium wall model in the flow past a hump (Park, 2017). Duprat et al. (2011) did not account for convective terms in their non-equilibrium wall model while adjusting the RANS eddy viscosity to account for pressure gradients. Hickel et al. (2013) accounts for both convective, and pressure gradient terms, but inconsistently uses an equilibrium flow-based eddy-viscosity model.

It is apparent from the aforementioned efforts that further developments in the understanding of the growth of turbulent boundary layers under the action of a pressure gradient are needed for improving potential flow solvers and for near-wall models to account for pressure gradient effects. The objective of this paper is to investigate whether a simple extension to the original Thwaites method for turbulent flows can be derived based on the assimilation of a limited set of high-fidelity simulation data. The rest of this article is organized as follows, Section II describes the original Thwaites method for laminar flows, and Section III provides the proposed extension to turbulent flows. Section IV describes the simulation database used in this study. Section V describes the model coefficients and their validation in several different flows. Section VI discusses an application of the model in terms of the sensitivity of the separation parameter to flow perturbations. Concluding remarks are offered in Section VII.

2 Thwaites method for laminar boundary layers

In this section, a brief summary of the method of Thwaites (1949) for approximating the momentum thickness of laminar boundary layers is provided. Recall, that for thin boundary layers in an incompressible flow, the Von Karman integral equation ((White, 2008)) is written as,

dθds+(2+H)θUedUeds=Cf2\frac{d\theta}{ds}+(2+H)\frac{\theta}{U_{e}}\frac{dU_{e}}{ds}=\frac{C_{f}}{2} (3)

where in the nsn-s coordinate system, ss is the streamwise coordinate, and nn is the wall-normal coordinate. δ\delta^{*} and θ\theta are the displacement and momentum thickness lengths of the boundary layer, respectively, and H=δ/θH=\delta^{*}/\theta is the shape factor. The following relations also hold under the thin boundary layer approximation,

ν2Un2|n=0=Pes=UeUes\nu\frac{\partial^{2}U}{\partial n^{2}}\bigg{|}_{n=0}=\frac{\partial P_{e}}{\partial s}=-U_{e}\frac{\partial U_{e}}{\partial s} (4)

where PeP_{e} and UeU_{e} are the values of the pressure, and the streamwise velocity at the edge of the boundary layer, and n=0n=0 refers to the wall. Using these relations, Thwaites derived that

2Reθdθds=2((2+H)m+χ)=L2Re_{\theta}\frac{d\theta}{ds}=2\left(\left(2+H\right)m+\chi\right)=L (5)

where mm is the Holstein-Bohlen pressure gradient parameter, m=θ2Ue2Un2|n=0m=\frac{\theta^{2}}{U_{e}}\frac{\partial^{2}U}{\partial n^{2}}\big{|}_{n=0}, and χ=θUeUn|n=0\chi=\frac{\theta}{U_{e}}\frac{\partial U}{\partial n}\big{|}_{n=0}. It should be noted that in deriving Equation 5, the assumption of a laminar flow has not yet been invoked. For laminar boundary layers, however, Thwaites further postulated that L=L(m)L=L(m), or that the effect of the pressure gradient primarily determines the growth rate of a laminar boundary layer. Specifically, the linear fit proposed in Thwaites’ original work is

LL(m)Thwaites0.45+6mL\approx L(m)^{Thwaites}\approx 0.45+6m (6)

Rearranging the model fit, an approximate closed-form expression of the momentum thickness can be written as,

θ2(s)0.45νUe60sUe5(r)𝑑r\theta^{2}(s)\approx 0.45\frac{\nu}{U^{6}_{e}}\int^{s}_{0}U^{5}_{e}(r)dr (7)

Further, Thwaites fitted experimental data to suggest that a laminar flow may separate when m0.09m\approx 0.09. Other similar empirical thresholds have been proposed by Stratford (1959), and by Curle & Skan (1957). More details can be found in Horton (1968). Curle & Skan (1957), and later Dey & Narasimha (1990) suggested slightly different model coefficients that resulted in a better agreement with the Blasius and Falkner-Skan type solutions. These methods have been used in potential flow solvers to account for viscous effects in laminar flows.

3 Extension to turbulent flows

Unlike in laminar flows, the inner and outer layers of a turbulent boundary layer respond differently to external pressure gradients. Further, a thin internal layer develops near the wall in turbulent flows due to the pressure gradient (Uzun & Malik, 2022; Parthasarathy & Saxton-Fox, 2023), and the amplitude of the small scales near the wall gets modulated due to the larger ones (Knopp et al., 2015; Romero et al., 2022). A general method for predicting the momentum thickness, θ\theta, similar to that described in the previous section is hence lacking for general, non-equilibrium turbulent flows.

Consider a zero pressure gradient turbulent boundary layer at a high Reynolds number. The growth rate of the momentum thickness of a turbulent boundary layer is higher than that of an equivalent Blasius laminar boundary layer (θ/sRes1/2\theta/s\sim Re_{s}^{-1/2}). Reconciling this fact with the fact that Equation 5 still holds for turbulent flows, the approximation made by Thwaites in Equation 6, L=L(m)L=L(m) must not hold for turbulent boundary layers. Since a boundary layer at a higher Reynolds number is less sensitive to the effects of pressure gradients (Vinuesa et al., 2018), the growth rate can not depend only on the pressure gradient (as it did in the laminar case). Therefore, it is proposed that the growth rate of the momentum thickness may also depend directly on the Reynolds number. Equation 5 can be re-expressed as,

2dθds=2(2+H)mReθ+2χReθ.2\frac{d\theta}{ds}=2\frac{(2+H)m}{Re_{\theta}}+\frac{2\chi}{Re_{\theta}}. (8)

A Taylor’s series expansion of 2dθ/ds2{d\theta}/{ds} in terms of m/Reθm/Re_{\theta} and 1/Reθ1/Re_{\theta} gives,

2dθds=CRe,+CmmReθ+Cc1Reθ+Cm,2(mReθ)2+Cc,2(1Reθ)2+Cm,c(1Reθ)(mReθ)+,2\frac{d\theta}{ds}=C_{Re,\infty}+C_{m}\frac{m}{Re_{\theta}}+C_{c}\frac{1}{Re_{\theta}}+C_{m,2}\left(\frac{m}{Re_{\theta}}\right)^{2}+C_{c,2}\left(\frac{1}{Re_{\theta}}\right)^{2}+C_{m,c}\left(\frac{1}{Re_{\theta}}\right)\left(\frac{m}{Re_{\theta}}\right)+\cdots, (9)

where C()C_{(\cdot)} are the constant coefficients of the Taylor expansion about some intermediate values of mm and ReθRe_{\theta}. In the limit of a large Reynolds number, ReθRe_{\theta}, such that m/Reθ, 1/Reθ1m/Re_{\theta},\;1/Re_{\theta}\ll 1, the linear truncation is given by

2dθdsCRe,+CmmReθ+Cc1Reθ.2\frac{d\theta}{ds}\approx C_{Re,\infty}+C_{m}\frac{m}{Re_{\theta}}+C_{c}\frac{1}{Re_{\theta}}. (10)

This model can be thought to be based on the assumptions that the growth of the boundary layer and that the effect of the pressure gradient decreases with the Reynolds number. An interpretation of the linear model is that the effect of the pressure gradient, and the Reynolds number are only weakly coupled. Rewriting Equation 10 in terms of the definition of LL in Equation 5, the model form for LL can be written as,

L=L(m,Re)Cc+Cmm+CRe,ReθL=L(m,Re)\approx C_{c}+C_{m}m+C_{Re,\infty}Re_{\theta} (11)

With this approximation, the growth rate of the momentum thickness is

dds(UeCmθ2)νCcUeCm1+CRe,UeCmθ.\begin{split}\frac{d}{ds}(U^{C_{m}}_{e}\theta^{2})\approx\nu C_{c}U^{C_{m}-1}_{e}+C_{Re,\infty}{U^{C_{m}}_{e}\theta}.\end{split} (12)

Unlike laminar boundary layers, this system does not have a closed-form solution for arbitrary pressure gradients. However, the equation is an ordinary differential equation that can be integrated numerically along the streamwise direction if the inflow and the freestream conditions are known.

3.1 Zero pressure gradient boundary layers

The special case of a zero pressure gradient (ZPG) boundary layer is now considered (Ue=U_{e}= constant, m=0m=0). Under this limit

dθdsCc21Reθ+CRe,2\begin{split}\frac{d\theta}{ds}\approx\frac{C_{c}}{2}\frac{1}{Re_{\theta}}+\frac{C_{Re,\infty}}{2}\end{split} (13)

The analytical solutions to this system suggest that the growth rate is slightly sublinear (and higher than s0.5s^{0.5}) which is in line with experimental evidence. From this model equation, it may be hypothesized that the turbulent contribution (contribution from the CRe,C_{Re,\infty} term) becomes relatively dominant when Reθ>Cc/CRe,600Re_{\theta}>C_{c}/C_{Re,\infty}\approx 600 (the values of CcC_{c} and CRe,C_{Re,\infty} are provided later in Section 5), which implies Reτzpg200Re^{zpg}_{\tau}\geq 200.

3.2 Constant pressure gradient boundary layers

In the case of a constant (or a slowly varying) externally applied constant pressure gradient, the Bernoulli equation evaluated in the inviscid freestream gives

dPds=UedUeds=Λ\frac{dP}{ds}=-U_{e}\frac{dU_{e}}{ds}=-\Lambda (14)

Then at a far downstream distance from the inflow, the freestream velocity can be approximated as Ue(s)U022ΛsU_{e}(s)\sim\sqrt{U^{2}_{0}-2\Lambda s} where U0U_{0} is the freestream velocity at s=0s=0. Invoking Equation 12 gives

dθdsCc2νUeθ+CRe,2+Cm2θUe2dPeds.\frac{d\theta}{ds}\approx\frac{C_{c}}{2}\frac{\nu}{U_{e}\theta}+\frac{C_{Re,\infty}}{2}+\frac{C_{m}}{2}\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds}. (15)

The third term scales as θ/s\theta/s, thus the effect of a constant pressure gradient is to affect the growth of the boundary layer linearly.

.

4 Database of pressure gradient flows

It has been well established that for turbulent boundary layers, the presence of strong pressure gradients leads to “history effects”, or that two boundary layers experiencing different upstream pressure gradients may respond differently to the action of the same pressure gradient even when the Reynolds number (ReθRe_{\theta} or ReτRe_{\tau}) is held constant (Bobke et al., 2017; Devenport & Lowe, 2022; Vinuesa et al., 2017; Nickels, 2004). This is because under the influence of the pressure gradient, the near-wall, and outer layer scales of motion respond at different spatial, and temporal scales (Knopp et al., 2021). To account for many of these effects, a comprehensive database of turbulent boundary layers is used in this work. All of the data used is from numerical simulations for ease of accessing continuous data, for computing streamwise derivatives and integrals. Relevant details of the datasets are given below. Table 1 summarizes the variation in the Reynolds number and the Clauser parameter for the boundary layers considered. It is noteworthy that Goc et al. (2023) reported similar skin-friction Reynolds numbers across the wing of a transonic aircraft model as those observed in this database.

Dataset Range of ReτRe_{\tau} Range of β\beta
ZPG (Eitel-Amor et al. (2014)) [500,2700] 0
APG (Bobke et al. (2017)) [300,1000] [1, 5]
Airfoils (Vinuesa et al. (2018), Tanarro et al. (2020)) [100,700] [1, 15]
FPG/APG SBSE (Uzun & Malik (2022)) [400,1200] [-2, 15]
FPG/APG BJ (Uzun & Malik (2019)) [1500,4000] [-6, 6]
Table 1: The variation in friction-based Reynolds number and Clauser parameters in the datasets considered in this work. For the flows that experience separation (FPG/APG SBSE and FPG/APG BJ), only pre-separation data is considered. Since both favorable and adverse pressure gradient boundary layers are included, the Clauser parameter ranges from a negative to a positive value.
  • ZPG Eitel: Eitel-Amor et al. (2014) performed wall-resolved large eddy simulations (LES) of a spatially developing zero pressure-gradient boundary layer up to Reθ=8300Re_{\theta}=8300.

  • APG Bobke: Bobke et al. (2017) performed wall-resolved LES of five different boundary layers at varying Reynolds numbers and strength of pressure gradients. Two of these five boundary layers were maintained at nearly constant Clauser parameters, β=1,and 2\beta=1,\;\mathrm{and}\;2 while the other three boundary layers followed a power law for the streamwise velocity at the edge of the boundary layer, Ue(xx0)mU_{e}\sim(x-x_{0})^{m} with m=0.13,0.16,and0.18m=-0.13,-0.16,\;\mathrm{and}\;-0.18.

  • Airfoils: Vinuesa et al. (2018) and Tanarro et al. (2020) simulated the flow over two different NACA airfoils, namely NACA 4412 and 0012 respectively at varying chord-length-based Reynolds number and angles of attack. In this work, we consider three particular cases, NACA 0012 airfoil at Rec=0.4MilRe_{c}=0.4\mathrm{Mil}, α=0.0o\alpha=0.0^{o} deg., and NACA 4412 airfoil at Rec=0.1Mil, 1MilRe_{c}=0.1\mathrm{Mil},\;1\mathrm{Mil}, α=5.0o\alpha=5.0^{o} deg.

  • FPG/APG SBSE: The problem of predicting smooth body separation using both Reynolds averaged methods (RANS) and scale-resolving methods in CFD has remained a challenge. Uzun & Malik (2022) performed a quasi-DNS of the flow over a smooth Gaussian bump as it experienced both a favorable pressure gradient, followed by an adverse pressure gradient, and thereafter a turbulent separation. These results were in agreement with the experiments of Williams et al. (2020). Later, Agrawal et al. (2022); Whitmore et al. (2022) performed wall-modeled large-eddy simulations of this flow and accurately predicted the flow separation.

  • FPG/APG BJ: In aerospace applications, such as the flow over the transonic Common Research Model (CRM) aircraft (Rivers & Dittberner, 2010), weak compressibility effects affect the flow pattern due to the shock-formation, and the consequent flow separation at high Mach numbers. Uzun & Malik (2019) performed high-fidelity simulations of the experiments of Bachalo & Johnson (1986) that studied the flow over a hump on a cylinder experiencing shock and adverse pressure-gradient induced flow separation.

The undetermined model coefficients, Cc,Cm,andCRe,C_{c},\;C_{m},\;\mathrm{and}\;C_{Re,\infty}, are ascertained from fitting the simulation data from the zero pressure gradient and adverse pressure gradients of Eitel-Amor et al. (2014) and Bobke et al. (2017), respectively. The remainder of the simulation database is used to assess the generality of the proposed model and fits.

5 Model Coefficients and Validation

5.1 Fitting model coefficients

The model coefficients Cc,Cm,andCRe,C_{c},\;C_{m},\;\mathrm{and}\;C_{Re,\infty} are determined in two stages. Since m=θ2Ue2Un2|n=0=0m=\frac{\theta^{2}}{U_{e}}\frac{\partial^{2}U}{\partial n^{2}}\big{|}_{n=0}=0 for a zero pressure-gradient boundary layer, the coefficients CcC_{c} and CRe,C_{Re,\infty} are determined from the data of Eitel-Amor et al. (2014) alone using a least-squares fit (with bisquare weights). Then CmC_{m} is fitted using the wall-resolved LES data of boundary layers under slightly dissimilar pressure gradients (Bobke et al., 2017). The numerical values of the coefficients obtained in this way are, Cc=1.45C_{c}=1.45, CRe,=0.0024C_{Re,\infty}=0.0024 and Cm=7.20C_{m}=7.20 with the corresponding 95%95\% confidence intervals Cc[1.43,1.47]C_{c}\in[1.43,1.47], CRe,[0.0023,0.0025]C_{Re,\infty}\in[0.0023,0.0025] and Cm[7.18,7.21]C_{m}\in[7.18,7.21] respectively. Once the three coefficients are fitted, they are kept constant as the fit is tested on other flows. Figures 1 and 2 show the variation in the Reynolds number and the Clauser parameter for the zero-pressure gradient boundary layer, and the five adverse pressure gradient boundary layers. The adverse pressure gradient flows have slightly different pressure-gradient distributions which lead to different growth rates of the boundary layers as the outer scales are energized differently.

Refer to caption
Figure 1: The distribution of the (a) Clauser parameter, β\beta, and the (b) friction Reynolds number, ReτRe_{\tau}, for the boundary layers considered in Bobke et al. (2017)

.

Refer to caption
Figure 2: The distribution of the skin-friction based Reynolds number, ReτRe_{\tau}, for the zero-pressure gradient boundary layer studied in Eitel-Amor et al. (2014).

Using Equation 12, the variation in θ\theta vs. ss is compared against the values obtained from mean velocity profiles using the method of Griffin et al. (2021). Figure 3 shows the model fit for the zero-pressure gradient boundary layer, and the adverse pressure gradient boundary layers, clearly the growth rate of θ\theta is reasonably predicted with minor discrepancies near the outlet of the domain.

Refer to caption
Figure 3: Fit of θ\theta vs the streamwise distance x/δ0x/\delta^{*}_{0}. Subfigure (a) is a model fit for the zero pressure gradient boundary layer in Eitel-Amor et al. (2014) and subfigure (b) is fit for the adverse pressure gradient boundary layers in Bobke et al. (2017). In subfigure (b) solid lines denote the values obtained from simulations of Bobke et al. (2017) and dotted lines refer to the model fit.

It is worth noting that unlike laminar boundary layers, turbulent boundary layers do not exhibit a universal relation between the shape factor, HH, and the Holstein-Bohlen pressure gradient parameter, mm. Figure 4 shows that for the boundary layers in Bobke et al. (2017), the shape factor varies significantly across the cases, and hence a universal fit of HH with mm is not possible. Although not shown, it was also verified that HH also does not follow a linear fit in (m,Reθ)(m,Re_{\theta}) space. Laminar potential flow solvers iteratively solve for θ\theta from Thwaites Method and find δ\delta^{*} from the H=H(m)H=H(m) relationship to determine the “effective body shape” due to flow acceleration. For turbulent flows, since H=H(m)H=H(m) is nonuniversal, this approach is not workable. Appendix A provides an alternate approach for finding the growth rate of the displacement thickness using boundary layer edge variables.

Refer to caption
Figure 4: The variation of the shape factor HH and Holstein-Bohlen parameter mm for the five adverse pressure gradient boundary layers in Bobke et al. (2017).

5.2 Testing on APG Wing data

Refer to caption
Figure 5: The distribution of the (a) Clauser parameter β\beta and the (b) skin-friction based Reynolds number, ReτRe_{\tau}, for a NACA 0012 airfoil at Rec=0.4×106Re_{c}=0.4\times 10^{6}, angle of attack α=0o\alpha=0^{o} (Tanarro et al. (2020)) and for a NACA 4412 airfoil at Rec=0.1,1×106Re_{c}=0.1,1\times 10^{6}, angle of attack α=5o\alpha=5^{o} (Vinuesa et al. (2018)) respectively.
Refer to caption
Figure 6: A fit of θ/θ0\theta/\theta_{0} vs the streamwise distance x/cx/c for a NACA 0012 airfoil at Rec=0.4×106Re_{c}=0.4\times 10^{6}, angle of attack α=0o\alpha=0^{o} (Tanarro et al. (2020)) and for a NACA 4412 airfoil at Rec=0.1,1×106Re_{c}=0.1,1\times 10^{6}, angle of attack α=5o\alpha=5^{o} (Vinuesa et al. (2018)) respectively.

Now, the model fit is evaluated for two airfoil flows at low angles of attack published in Vinuesa et al. (2018); Tanarro et al. (2020). Two NACA airfoils at different chord-based Reynolds numbers (RecRe_{c}) and angles of attack (α\alpha). A NACA 0012 airfoil at Rec=0.4×106,α=0oRe_{c}=0.4\times 10^{6},\;\alpha=0^{o} and NACA 4412 airfoil at Rec=0.1×106Re_{c}=0.1\times 10^{6} and Rec=1.0×106Re_{c}=1.0\times 10^{6}, α=5o\alpha=5^{o} are considered. Note that the dataset available for these airfoils only provides detailed statistics beyond x/c=15%x/c=15\% as the flow is tripped to become turbulent at x/c=0.1x/c=0.1.

For these cases, the pressure gradient distributions are quite different, as quantified by the Clauser parameter in Figure 5(a). Goc et al. (2023) recently evaluated that ReτRe_{\tau} over a transonic common research model aircraft (at α4o\alpha\approx 4^{o}) is approximately Reτ800Re_{\tau}\approx 800 at 1020%10-20\% chord length (similar to those experienced on the presently considered airfoil flows). Near the trailing edge, the pressure gradient for NACA 4412 airfoils varies quite significantly and is much stronger than effects of the wall-shear stress (β15\beta\approx 15 at the trailing edge) suggesting a strongly non-equilibrium flow. Nonetheless, the fit of the momentum thickness in Figure 6 is still reasonable in this region, suggesting a simple linear model is capturing the history effects from the pressure gradients. The increased Reynolds number is responsible for the lowered β\beta between the two NACA 4412 airfoils; as the lower Reynolds number boundary layer is more strongly affected by the pressure gradient (Vinuesa et al., 2017).

5.3 Testing on SBSE and BJ data

Smooth body separation refers to the flow condition where neither the separation nor reattachment points of a separation bubble are geometrically imposed. As a result, the flow development history (i.e. the influence of pressure gradients and/or compressibility) affects the tendency of the flow to separate. Uzun & Malik (2022) recently performed a quasi-DNS of flow over a spanwise-periodic Gaussian bump in which the flow separates on the leeward side of the bump due to the mild, adverse pressure gradient. Similarly, the wall-resolved LES of Uzun & Malik (2019) is an example of shock-induced separation in a transonic flow over a bump. Figure 7 shows a schematic of the bump geometries and the case-setup for both flows. Figure 8 shows the distribution of the Clauser parameter β\beta and the friction Reynolds number ReτRe_{\tau} for these cases. For both flows, an incoming zero-pressure gradient turbulent boundary layer experiences a favorable pressure gradient, which accelerates the flow, before eventually experiencing an adverse pressure gradient that leads to flow separation. The Clauser parameter, Reynolds numbers, and their “histories” until the point of separation (x/L0.1x/L\sim 0.1 for the SBSE, and x/c0.65x/c\sim 0.65 for the BJ) are significantly different.

Refer to caption
Figure 7: A schematic of the surface geometry over which the turbulent flow separates in the (a) subsonic Boeing speed bump case of Uzun & Malik (2022) and (b) transonic flow case of Uzun & Malik (2019). LL and cc are the characteristic lengths for the two cases respectively. UeU_{e} is the freestream flow velocity and the arrows next to it point along the positive streamwise direction.
Refer to caption
Figure 8: The streamwise distribution of the Clauser parameter β\beta (subfigures a, b) and the skin-friction based Reynolds number ReτRe_{\tau} (subfigures c, d) for the Boeing speed bump (subfigures a, c) in Uzun & Malik (2022) and the transonic Bachalo-Johnson bump (subfigures b, d) in Uzun & Malik (2019). For the transonic flow, the density on the wall is used for defining the Reynolds number. Note that x/L=0x/L=0 is the position of the apex of the Boeing speed bump, whereas x/c=0.5x/c=0.5 is the position of the apex of the transonic bump.
Refer to caption
Figure 9: Fit of θ\theta vs the streamwise distance xx non-dimensionalized by the respective relevant length scales (inlet displacement thickness of the laminar flow) as defined in the reference simulations. Subfigure (a) is a model fit for the Boeing speed bump in Uzun & Malik (2022) and subfigure (b) is a model fit for the transonic Bachalo-Johnson bump in Uzun & Malik (2019). The solid vertical lines denote the separation point as obtained from the reference data. Note that x/L=0x/L=0 is the position of the apex of the Boeing speed bump, whereas x/c=0.5x/c=0.5 is the position of the apex of the transonic bump.

As seen in Figure 9, the proposed model predicts the momentum thickness for both these flows reasonably up to the point of separation (for the transonic flow, Uzun & Malik (2019) performed their simulations at Ma=0.875Ma=0.875 at which the location of the shock, and the separation nearly coincide (Bachalo & Johnson, 1986)). The proposed model was derived for incompressible flows, and as such neglects any variations in density in the flow. The transonic flow in Uzun & Malik (2019) exhibited strong density variations; the ratio of the mean density at the edge of the boundary layer and the wall 23\approx 2-3. Finally, it is not surprising for the proposed, linear model to not predict the growth inside the separation bubble as upstream effects of the blockage of the separation bubble are not comprehended by the ODE model employed (see Simpson (1981) for a more comprehensive overview) and the thin boundary layer approximation utilized also breaks down at the point of separation.

However, the proposed model can be used to help predict the conditions under which the separation of the boundary layer is imminent. To date, there have been numerous attempts to characterize this point of separation empirically. Alber (1971) proposed that separation is imminent when θUe2dPeds>0.004\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds}>0.004. The exact value of the threshold was determined empirically for some weakly compressible flows and has been shown to be relatively robust in the presence of weak shocks (Alber et al., 1973; Adair, 1987). As seen in Figure 10 the value of θUe2dPeds\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds} is approximately 0.0040.004 at the location of separation for both the flow over the Boeing speed bump and the flow in the Bachalo-Johnson experiments. For laminar flows, Thwaites also proposed a fit between the shape factor, HH, and the Holstein-Bohlen parameter, mm that allowed computation of the skin friction using the local value of mm. Using this analysis, the separation boundary for a laminar flow was suggested to be m=0.09m=0.09. Although not shown, we have observed that the two flows considered in this work exhibited separation for different values of mm. For this reason, the parameter, mm, can not be used alone to predict flow separation in turbulent flows.

Refer to caption
Figure 10: The variation in the non-dimensional group θρeUe2dPds\frac{\theta}{\rho_{e}U^{2}_{e}}\frac{dP}{ds} along the streamwise direction for the (a) Boeing speed bump in Uzun & Malik (2022) and the (b) transonic Bachalo-Johnson bump in Uzun & Malik (2019). The dashed vertical lines denote the separation point as obtained from the reference data. Note that x/L=0x/L=0 is the position of the apex of the Boeing speed bump, whereas x/c=0.5x/c=0.5 is the position of the apex of the transonic bump.

Alternatively, it is possible to estimate the value of the Alber separation parameter by using the proposed model. For a locally constant and finite pressure gradient, the proposed model suggests that dθds\frac{d\theta}{ds} is a linear function of θUe2dPeds\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds}. Using the von Karman integral equation, this equation can be further simplified as,

Cf2Cc2νUeθ+CRe,2+(Cm2(2+H))θUe2dPeds\frac{C_{f}}{2}\approx\frac{C_{c}}{2}\frac{\nu}{U_{e}\theta}+\frac{C_{Re,\infty}}{2}+\left(\frac{C_{m}}{2}-\left(2+H\right)\right)\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds} (16)

In the limit of an asymptotically large Reynolds number (ReθRe_{\theta}\rightarrow\infty) flow nearing separation, an estimate of H2.32.5H\approx 2.3-2.5 (see Castillo et al. (2004) yields that the separation is imminent when

θUe2dPedsCRe,2((Cm2(2+H)))0.003\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds}\approx\frac{C_{Re,\infty}}{2((\frac{C_{m}}{2}-(2+H)))}\approx 0.003 (17)

which is of the same order of magnitude as the threshold proposed by Alber (1971). Equation 16 also describes a separation boundary (considered to be the point where the skin-friction crosses zero) for turbulent flows over smooth surfaces for a range of Reynolds numbers as,

Cc2νUeθ+(Cm2(2+H))θUe2dPeds=CRe,2\frac{C_{c}}{2}\frac{\nu}{U_{e}\theta}+\left(\frac{C_{m}}{2}-\left(2+H\right)\right)\frac{\theta}{U^{2}_{e}}\frac{dP_{e}}{ds}=-\frac{C_{Re,\infty}}{2} (18)

This describes a nonlinear boundary for predicting turbulent separation and reattachment in a space spanned by 1/Reθ,H,andδUe2dPeds1/Re_{\theta},\;H,\;\mathrm{and}\;\frac{\delta^{*}}{U^{2}_{e}}\frac{dP_{e}}{ds}.

6 Sensitivity analysis of separation using Alber’s parameter

The accurate prediction of flows exhibiting separation from mild, adverse pressure gradients remains a pacing item for the development of closure models for both Reynolds-averaged Navier-Stokes (RANS) and large-eddy simulations (LES). While it is known that many of these flows exhibit history effects (potentially originating from the inflow conditions), there are not many computationally efficient ways to quantify the impact of these history effects. Moreover, for a given flow, it is unclear which spatial locations are particularly sensitive with respect to the prediction of the separation point. The following subsections explore the use of the proposed model to quantify these sensitivities. The following assumes that the point of separation is governed by Alber’s parameter, as was motivated in the preceding section.

6.1 Sensitivity to inflow conditions

The sensitivity of the Alber parameter (denoted by J(θ)=θUedUedsJ(\theta)=-\frac{\theta}{U_{e}}\frac{dU_{e}}{ds} here) to the inflow conditions is given by dJdθ0\frac{dJ}{d\theta_{0}}.

Upon rewriting the ODE model in terms of Alber’s parameter, the following expression can be written,

θ(s)=θ0+Ccν20s1Ueθ𝑑rCm20sJ(θ)𝑑r+CRe,2s\theta(s)=\theta_{0}+\frac{C_{c}\nu}{2}\int^{s}_{0}\frac{1}{U_{e}\theta}dr-\frac{C_{m}}{2}\int^{s}_{0}J(\theta)dr+\frac{C_{Re,\infty}}{2}s (19)

The sensitivity of the momentum thickness due to a perturbation in the inlet momentum thickness (θ0\theta_{0}) can be then measured as

dθ(s)dθ0=1Ccν20s1Ueθ2dθdθ0𝑑r+Cm20s1UedUedrdθdθ0𝑑r\frac{d\theta(s)}{d\theta_{0}}=1-\frac{C_{c}\nu}{2}\int^{s}_{0}\frac{1}{U_{e}\theta^{2}}\frac{d\theta}{d\theta_{0}}dr+\frac{C_{m}}{2}\int^{s}_{0}\frac{1}{U_{e}}\frac{dU_{e}}{dr}\frac{d\theta}{d\theta_{0}}dr (20)

where it has been assumed that the inlet is sufficiently far upstream such that the growth of the inlet boundary layer thickness is negligible compared to other terms in the equation. The resulting system can be recast into a separable second-order differential equation relating θandθ0\theta\;\mathrm{and}\;\theta_{0} as

d2θ(s)dsdθ0+(Ccν2Ueθ2Cm21UedUeds)dθdθ0=0\frac{d^{2}\theta(s)}{dsd\theta_{0}}+\left(\frac{C_{c}\nu}{2U_{e}\theta^{2}}-\frac{C_{m}}{2}\frac{1}{U_{e}}\frac{dU_{e}}{ds}\right)\frac{d\theta}{d\theta_{0}}=0 (21)

Assuming that Alber’s threshold on J(θ)J(\theta) is approximately accurate (as also shown in Figure 10), the sensitivity of J(θ)J(\theta) at a location of interest ss^{*} (region close to the “true” point of separation) to inflow conditions is given as

dJdθ0|s=[1Ue2dPedseCm20s1Ue2dPeds𝑑r][eCcν20sdrUeθ2]\frac{dJ}{d\theta_{0}}\bigg{|}_{s^{*}}=\left[\frac{1}{U^{2}_{e}}\frac{dP_{e}}{ds}e^{-\frac{C_{m}}{2}\int^{s^{*}}_{0}\frac{1}{U^{2}_{e}}\frac{dP_{e}}{ds}dr}\right]\;\left[e^{-\frac{C_{c}\nu}{2}\int^{s^{*}}_{0}\frac{dr}{U_{e}\theta^{2}}}\right] (22)

This implies that the value of Alber’s parameter is exponentially sensitive (in a Lagrangian sense) to both the pressure gradient and the non-dimensional group sReθθ\frac{s^{*}}{Re_{\theta}\theta}. The latter carries the group sθ=Ue2sUe2θ\frac{s}{\theta}=\frac{U^{2}_{e}s}{U^{2}_{e}\theta} which is a measure of the integrated momentum deficit in the viscous and the momentum transport in an inviscid region. Further, as ss^{*}\rightarrow\infty, the asymptotic limit of dJdθ0\frac{dJ}{d\theta}\rightarrow 0 or that the parameter is insensitive to the inflow. Secondly, in the limit of infinitely large Reynolds number, ReθRe_{\theta}, the parameter (or the location of the separation) becomes purely a function of the pressure gradient as expected.

Figure 11 shows the sensitivity of JJ to a perturbation in θ0\theta_{0} for the SBSE and BJ flows considered herein. Traversing along the x-axis of the plot, the local (in space), relative sensitivity of JJ to a given perturbation in θ0\theta_{0} can be quantified (θ0/θ0\theta_{0}/\theta_{0} represents the number on the ordinate on the left side the figures). Alternatively, Figure 11 can be interpreted as follows: for a given location on the x-axis, the contours are suggestive of the change in JJ due to a change in θ0\theta_{0} (plotted on the ordinate on the left side). For the Boeing speed bump, inflow errors (in θ\theta) as large as 50%\approx 50\% do not appreciably change the value of the Alber’s parameter. However, for the Bachalo-Johnson transonic bump, the inflow conditions can affect the momentum thickness far downstream.

Refer to caption
Figure 11: The relative sensitivity of J=θUedUedsJ=-\frac{\theta}{U_{e}}\frac{dU_{e}}{ds} to the relative change in inlet momentum thickness (Δθ0θ0\frac{\Delta\theta_{0}}{\theta_{0}}) pre-separation for the (a) Boeing speed bump (Uzun & Malik, 2022) and for (b) the transonic Bachalo-Johnson bump (Uzun & Malik, 2019). Note that x/L=0x/L=0 is the position of the apex of the Boeing speed bump, whereas x/c=0.5x/c=0.5 is the position of the apex of the transonic bump.

6.2 Sensitivity to perturbations in the Holstein-Bohlen parameter

In scale-resolving simulation paradigms such as wall-modeled large eddy simulations, it is difficult to isolate the regions of deficiency of the near-wall model (for predicting the wall shear stress) in non-equilibrium flows, especially in flows undergoing turbulent smooth body separation. The parameter mJsepdJsepdm\frac{m}{J_{sep}}\frac{dJ_{sep}}{dm} is a relative measure of the sensitivity of flow separation to upstream modeling deficiencies (the perturbations in mm can be thought of as arising from the near-wall errors in the prediction of wall shear stress). This parameter can be re-expressed as,

mJsepdJsepdm=mJsepdJsepdJ(θ)dJ(θ)dm=mJsepdJsepdJ(θ)12Reθ=12θθsepdθsepdθ=12θθsep1dθdθsep\frac{m}{J_{sep}}\frac{dJ_{sep}}{dm}=\frac{m}{J_{sep}}\frac{dJ_{sep}}{dJ(\theta)}*\frac{dJ(\theta)}{dm}=\frac{m}{J_{sep}}\frac{dJ_{sep}}{dJ(\theta)}*\frac{1}{2Re_{\theta}}=\frac{1}{2}\frac{\theta}{\theta_{sep}}\frac{d\theta_{sep}}{d\theta}=\frac{1}{2}\frac{\theta}{\theta_{sep}}\dfrac{1}{\frac{d\theta}{d\theta_{sep}}} (23)
Refer to caption
Figure 12: The streamwise variation (non-dimensionalized by the boundary layer thickness at the point of separation) of the relative sensitivity of Jsep=θUedUedsJ_{sep}=-\frac{\theta}{U_{e}}\frac{dU_{e}}{ds} evaluated at the point of separation to the upstream perturbations in the Holstein-Bohlen parameter, mm, for (a) Boeing speed bump and (b) transonic Bachalo-Johnson bump.

By perturbing the momentum thickness at the true point of separation, and advancing the proposed model fit for θ\theta upstream, the sensitivity parameter can be completely determined. Figure 12 suggests that two different mechanisms are important for two separating flows considered above. For the Boeing speed bump, the sensitivity of Alber’s parameter to a perturbation in the Holstein-Bohlen parameter is non-local. Perturbations in the region after the apex (the adverse pressure gradient region) are most important, in that the imminently separating flow responds to those the most. The curve is increasing in the zero pressure gradient and the favorable pressure gradient region, which implies that the closer the perturbations are made to the separation point, the more responsive the tendency of the flow to actually separate in predictive simulations. These results are in agreement with Devenport & Lowe (2022) who suggest that the “history effects” of pressure gradients in a turbulent boundary layer may extend up to 50δ\delta. However, for the transonic bump, the sensitivity is primarily localized in a small region after the apex (in the adverse pressure gradient region), just before the shock. This suggests that the prediction of separation is primarily governed by the shock and the adverse pressure gradient just upstream of it; or that the upstream history or modeling errors are not that significant. These contrasting mechanisms provide clear evidence that flow separation can occur due to different flow development patterns, and future modeling efforts may need to account for these effects.

7 Concluding remarks

This work presents an extension of the method of Thwaites for determining the momentum thickness for a turbulent boundary layer under the action of pressure gradients. In the limit of large Reynolds numbers, a linear model is hypothesized, such that the Reynolds number and pressure gradient dependence on the growth rate of the momentum thickness can be superposed. A fit for the model coefficients is found using recent high-fidelity simulation data for various boundary layers ranging from Reynolds number, Reτ[100,4000]Re_{\tau}\approx[100,4000] and Clauser parameter, β,[6,10]\beta,\approx[-6,10]. The model fit predicts the growth of the momentum thickness, even up to the point of separation. The fits utilized in the model can then be used to derive a condition for imminent separation at high Reynolds numbers, with similar threshold values as compared to previously proposed values based on empirical data. An application of this extension of Thwaites method is demonstrated by estimating the importance of history effects on separation. Specifically, the importance of upstream history effects for the Boeing speed-bump flow is quantified. For a transonic bump flow, it is shown that only local perturbations just upstream of the shock location cause significant changes to the separation location.

Appendix A Relation between of δ\delta, δ,andθ\delta^{*},\;\mathrm{and}\;\theta

For laminar flows, the solution to Thwaites method for θ\theta and the universal correlation between mm and the shape factor HH provides a δ\delta^{*} which is useful for iteratively updating the “inviscid geometry” that is used for computing the freestream profiles. The proposed model in this work provides a good fit for θ\theta, but a fit for δ\delta^{*} is needed as well for iterative deployment with a potential flow solver. The analytical expression for the displacement thickness can be derived from the continuity equation as,

dδds=VeUe+1Ue2dUeds0δU𝑑n\frac{d\delta^{*}}{ds}=\frac{V_{e}}{U_{e}}+\frac{1}{U^{2}_{e}}\frac{dU_{e}}{ds}\int^{\delta}_{0}Udn (24)

where δ\delta is the thickness of the boundary layer. For general flows, the integral in Equation 24 is the measure of the mass flow rate inside the boundary layer, and is dependent on the local flow conditions (such as ReτRe_{\tau} and pressure gradient dPeds\frac{dP_{e}}{ds}) and is unknown “a priori”. For given values of δ\delta and the flow variables at the edge of the boundary layer UeU_{e}, VeV_{e}, and PeP_{e}, the growth rate of the displacement thickness is given as

dδds=VeUe1Ue2dUedss0sVe(r)Ue(r)Ue(r)𝑑r,\frac{d\delta^{*}}{ds}=\frac{V_{e}}{U_{e}}-\frac{1}{U^{2}_{e}}\frac{dU_{e}}{ds}\int^{s}_{s_{0}}\frac{V_{e}(r)}{U_{e}(r)}U_{e}(r)dr, (25)

where s0s_{0} is the streamwise location at which the flow can be first considered fully turbulent, within the boundary layer, and ss is the location of interest. From geometrical arguments, the ratio Ve/UeV_{e}/U_{e}, can be approximately related to the growth of the boundary layer thickness as

VeUedδds\frac{V_{e}}{U_{e}}\approx\frac{d\delta}{ds} (26)

Thus, the growth of the boundary layer affects the growth rate of the displacement thickness linearly as follows,

dδds=dδds1Ue2dUedss0sdδdrUe(r)𝑑r\frac{d\delta^{*}}{ds}=\frac{d\delta}{ds}-\frac{1}{U^{2}_{e}}\frac{dU_{e}}{ds}\int^{s}_{s_{0}}\frac{d\delta}{dr}U_{e}(r)dr (27)

Finally, a linear relationship between Ve/UeV_{e}/U_{e}, LL and mm is fitted from the simulation data as

L(m,Re)=Ueνdθ2ds5+8m200dδds+200dδzpg,corrdsL(m,Re)=\frac{U_{e}}{\nu}\frac{d\theta^{2}}{ds}\approx 5+8m-200\frac{d\delta}{ds}+200\frac{d\delta{zpg,corr}}{ds} (28)

which implies that,

dδds5200+8200mUe200νdθ2ds+dδzpg,corrds\frac{d\delta}{ds}\approx\frac{5}{200}+\frac{8}{200}m-\frac{U_{e}}{200\nu}\frac{d\theta^{2}}{ds}+\frac{d\delta{zpg,corr}}{ds} (29)

where δzpg,corr/s=0.16Res1/7=0.16(Ue(s)s/ν)1/7\delta_{zpg,corr}/s=0.16Re^{-1/7}_{s}=0.16(U_{e}(s)s/\nu)^{-1/7}. The quality of this fit is verified in Figure 13 by comparing the exact values of L(m,Re)L(m,Re) obtained from the datasets considered in this work, and from Equation 29. With the two proposed fits in this work, the displacement thickness can be determined along the streamwise coordinate using Equations 25 and 29.

Refer to caption
Figure 13: The quality of the fit between the exact value of L(m,Re)L(m,Re) and that obtained from the proposed Equation29. The red symbols contain data from the zero-pressure gradient boundary layer of Eitel-Amor et al. (2014), the five adverse pressure gradient boundary layers of Bobke et al. (2017), and the three boundary layers from the NACA airfoils (Vinuesa et al., 2017; Tanarro et al., 2020).

Acknowledgments

This work was supported by NASA’s Transformational Tools and Technologies project under Grant No. #80NSSC20M0201 and by Boeing Research & Technology. K.P.G. is supported by the Exascale Computing Project (Grant17-SC-20SC).

This work was authored in part by the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DEAC3608GO28308. The views expressed in the article do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work or allow others to do so, for the U.S. Government purposes.

Declaration of Interests: The authors report no conflict of interest.

References

  • Adair (1987) Adair, D 1987 Characteristics of a trailing flap flow with small separation. Experiments in fluids 5 (2), 114–128.
  • Agrawal et al. (2022) Agrawal, Rahul, Whitmore, Michael P., Griffin, Kevin P., Bose, Sanjeeb T. & Moin, Parviz 2022 Non-Boussinesq subgrid-scale model with dynamic tensorial coefficients. Physical Review Fluids 7, 074602.
  • Alber (1971) Alber, Irwin 1971 Similar solutions for a family of separated turbulent boundary layers. In 9th Aerospace Sciences Meeting, p. 203.
  • Alber et al. (1973) Alber, Irwin E, Bacon, John W, Masson, Bruce S & Collins, Donald J 1973 An experimental investigation of turbulent transonic viscous-inviscid interactions. AIAA Journal 11 (5), 620–627.
  • Bachalo & Johnson (1986) Bachalo, WDf & Johnson, DA 1986 Transonic, turbulent boundary-layer separation generated on an axisymmetric flow model. AIAA Journal 24 (3), 437–443.
  • Balaras et al. (1996) Balaras, Elias, Benocci, Carlo & Piomelli, Ugo 1996 Two-layer approximate boundary conditions for large-eddy simulations. AIAA Journal 34 (6), 1111–1119.
  • Bobke et al. (2017) Bobke, Alexandra, Vinuesa, Ricardo, Örlü, Ramis & Schlatter, Philipp 2017 History effects and near equilibrium in adverse-pressure-gradient turbulent boundary layers. Journal of Fluid Mechanics 820, 667–692.
  • Cabot & Moin (2000) Cabot, William & Moin, Parviz 2000 Approximate wall boundary conditions in the large-eddy simulation of high Reynolds number flow. Flow, Turbulence and Combustion 63 (1), 269–291.
  • Castillo et al. (2004) Castillo, Luciano, Wang, Xia & George, William K 2004 Separation criterion for turbulent boundary layers via similarity analysis. J. Fluids Eng. 126 (3), 297–304.
  • Clauser (1954) Clauser, Francis H 1954 Turbulent boundary layers in adverse pressure gradients. Journal of the Aeronautical Sciences 21 (2), 91–108.
  • Clauser (1956) Clauser, Francis H 1956 The turbulent boundary layer. Advances in Applied Mechanics 4, 1–51.
  • Coles & Hirst (1968) Coles, DE & Hirst, EA 1968 Computation of Turbulent Boundary Layers–1968 AFOSR-IFP-Stanford Conference: Proceedings, , vol. 2. Thermosciences Division, Stanford University.
  • Curle & Skan (1957) Curle, N & Skan, SW 1957 Approximate methods for predicting separation properties of laminar boundary layers. Aeronautical Quarterly 8 (3), 257–268.
  • Das (1987) Das, DK 1987 A numerical study of turbulent separated flows. In Forum on Turbulent Flows-1987, pp. 85–90.
  • Das & White (1986) Das, DK & White, FM 1986 Integral skin friction prediction for turbulent separated flows. J. Fluids Eng. .
  • Das et al. (2004) Das, Debendra K, Naganathan, Surrandar, Kono, Toshifumi & Kulkarni, Devdatta 2004 Comparison of three inner-variable approaches to establish an accurate method for turbulent boundary layers. Naval Engineers Journal 116 (2), 85–102.
  • Devenport & Lowe (2022) Devenport, William J & Lowe, K Todd 2022 Equilibrium and non-equilibrium turbulent boundary layers. Progress in Aerospace Sciences 131, 100807.
  • Dey & Narasimha (1990) Dey, J & Narasimha, R 1990 An extension of the thwaites method for calculation of incompressible laminar boundary layers. Journal of the Indian Institute of Science 70 (1), 1.
  • Duprat et al. (2011) Duprat, Cédric, Balarac, Guillaume, Métais, Olivier, Congedo, Pietro Marco & Brugière, Olivier 2011 A wall-layer model for large-eddy simulations of turbulent flows with/out pressure gradient. Physics of fluids 23 (1).
  • Eitel-Amor et al. (2014) Eitel-Amor, Georg, Örlü, Ramis & Schlatter, Philipp 2014 Simulation and validation of a spatially evolving turbulent boundary layer up to Reθ=8300Re_{\theta}=8300. International Journal of Heat and Fluid Flow 47, 57–69.
  • Falkner & Skan (1931) Falkner, VM & Skan, Sylvia W 1931 Lxxxv. Solutions of the boundary-layer equations. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 12 (80), 865–896.
  • Goc et al. (2023) Goc, Konrad, Agrawal, Rahul, Moin, Parviz & Bose, Sanjeeb 2023 Studies of transonic aircraft flows and prediction of initial buffet onset using large-eddy simulations. In AIAA Aviation 2023 Forum.
  • Griffin et al. (2021) Griffin, Kevin Patrick, Fu, Lin & Moin, Parviz 2021 General method for determining the boundary layer thickness in nonequilibrium flows. Physical Review Fluids 6 (2), 024608.
  • Harun et al. (2013) Harun, Zambri, Monty, Jason P, Mathis, Romain & Marusic, Ivan 2013 Pressure gradient effects on the large-scale structure of turbulent boundary layers. Journal of Fluid Mechanics 715, 477–498.
  • Hickel et al. (2013) Hickel, Stefan, Touber, Emile, Bodart, Julien & Larsson, Johan 2013 A parametrized non-equilibrium wall-model for large-eddy simulations. In Eighth International Symposium on Turbulence and Shear Flow Phenomena. Begel House Inc.
  • Horton (1968) Horton, Harry Philip 1968 Laminar separation bubbles in two and three dimensional incompressible flow. PhD thesis, Queen Mary University of London.
  • Kalkhoran & Wilson (1986) Kalkhoran, Iraj M & Wilson, Donald R 1986 An integral method to investigate two-dimensional incompressible turbulent boundary layer. Integral methods in Science and Engineering pp. 221–226.
  • Kawai & Larsson (2012) Kawai, Soshi & Larsson, Johan 2012 Wall-modeling in large eddy simulation: Length scales, grid resolution, and accuracy. Physics of Fluids 24 (1), 015105.
  • Kitsios et al. (2017) Kitsios, Vassili, Sekimoto, Atsushi, Atkinson, Callum, Sillero, Juan A, Borrell, Guillem, Gungor, A Gul, Jiménez, Javier & Soria, Julio 2017 Direct numerical simulation of a self-similar adverse pressure gradient turbulent boundary layer at the verge of separation. Journal of Fluid Mechanics 829, 392–419.
  • Kline et al. (1969) Kline, Stephen Jay, Coles, Donald Earl & Hirst, EA 1969 Computation of Turbulent Boundary Layers–1968 AFOSR-IFP-Stanford Conference: Proceedings, , vol. 1. Thermosciences Division, Stanford University.
  • Knopp et al. (2015) Knopp, Tobias, Buchmann, NA, Schanz, Daniel, Eisfeld, Bernhard, Cierpka, Christian, Hain, Rainer, Schröder, Andreas & Kähler, Chr J 2015 Investigation of scaling laws in a turbulent boundary layer flow with adverse pressure gradient using piv. Journal of Turbulence 16 (3), 250–272.
  • Knopp et al. (2021) Knopp, Tobias, Reuther, Nico, Novara, Matteo, Schanz, Daniel, Schülein, Erich, Schröder, Andreas & Kähler, CJ 2021 Experimental analysis of the log law at adverse pressure gradient. Journal of Fluid Mechanics 918, A17.
  • Nickels (2004) Nickels, TB 2004 Inner scaling for wall-bounded flows subject to large pressure gradients. Journal of Fluid Mechanics 521, 217–239.
  • Park (2017) Park, George Ilhwan 2017 Wall-modeled large-eddy simulation of a high Reynolds number separating and reattaching flow. AIAA Journal 55, 3709–3721.
  • Park & Moin (2014) Park, George Ilhwan & Moin, Parviz 2014 An improved dynamic non-equilibrium wall-model for large eddy simulation. Physics of Fluids 26 (1), 37–48.
  • Parthasarathy & Saxton-Fox (2023) Parthasarathy, Aadhy & Saxton-Fox, Theresa 2023 A family of adverse pressure gradient turbulent boundary layers with upstream favourable pressure gradients. Journal of Fluid Mechanics 966, A11.
  • Prakash et al. (2023) Prakash, Aviral, Balin, Riccardo, Evans, John A & Jansen, Kenneth E 2023 Turbulent boundary layer with strong favorable pressure gradient and curvature effects: Streamline coordinate and scaling analysis. arXiv preprint arXiv:2306.05972 .
  • Rivers & Dittberner (2010) Rivers, Melissa & Dittberner, Ashley 2010 Experimental investigations of the NASA Common Research Model. In 28th AIAA Applied Aerodynamics Conference, p. 4218.
  • Romero et al. (2022) Romero, Sylvia, Zimmerman, Spencer, Philip, Jimmy, White, Christopher & Klewicki, Joseph 2022 Properties of the inertial sublayer in adverse pressure-gradient turbulent boundary layers. Journal of Fluid Mechanics 937, A30.
  • Simpson (1981) Simpson, R. L. 1981 A Review of Some Phenomena in Turbulent Flow Separation. Journal of Fluids Engineering 103 (4), 520–533.
  • Simpson (1983) Simpson, Roger L 1983 A model for the backflow mean velocity profile. AIAA Journal 21 (1), 142–143.
  • Skote et al. (1998) Skote, Martin, Henningson, Dan S & Henkes, Ruud AWM 1998 Direct numerical simulation of self-similar turbulent boundary layers in adverse pressure gradients. Flow, turbulence and combustion 60 (1), 47–85.
  • Sreenivasan (1982) Sreenivasan, KR 1982 Laminarescent, relaminarizing and retransitional flows. Acta Mechanica 44, 1–48.
  • Stratford (1959) Stratford, BS 1959 The prediction of separation of the turbulent boundary layer. Journal of Fluid Mechanics 5 (1), 1–16.
  • Tanarro et al. (2020) Tanarro, Álvaro, Vinuesa, Ricardo & Schlatter, Philipp 2020 Effect of adverse pressure gradients on turbulent wing boundary layers. Journal of Fluid Mechanics 883, A8.
  • Thwaites (1949) Thwaites, Bryan 1949 Approximate calculation of the laminar boundary layer. Aeronautical Quarterly 1 (3), 245–280.
  • Uzun & Malik (2019) Uzun, Ali & Malik, Mujeeb R 2019 Wall-resolved large-eddy simulations of transonic shock-induced flow separation. AIAA Journal 57 (5), 1955–1972.
  • Uzun & Malik (2022) Uzun, Ali & Malik, Mujeeb R. 2022 High-fidelity simulation of turbulent flow past Gaussian bump. AIAA Journal 60 (4), 2130–2149.
  • Vinuesa et al. (2018) Vinuesa, Ricardo, Negi, Prabal Singh, Atzori, Marco, Hanifi, Ardeshir, Henningson, Dan S & Schlatter, Philipp 2018 Turbulent boundary layers around wing sections up to Rec=1,000,000{R}e_{c}=1,000,000. International Journal of Heat and Fluid Flow 72, 86–99.
  • Vinuesa et al. (2017) Vinuesa, Ricardo, Örlü, Ramis, Sanmiguel Vila, Carlos, Ianiro, Andrea, Discetti, Stefano & Schlatter, Philipp 2017 Revisiting history effects in adverse-pressure-gradient turbulent boundary layers. Flow, Turbulence and Combustion 99, 565–587.
  • Wang & Moin (2002) Wang, Meng & Moin, Parviz 2002 Dynamic wall modeling for large-eddy simulation of complex turbulent flows. Physics of Fluids 14 (7), 2043–2051.
  • Weber (1978) Weber, HE 1978 Boundary layer calculation for analysis and design. J. Fluids Eng 100, 232–236.
  • Wei et al. (2017) Wei, Tie, Maciel, Yvan & Klewicki, Joseph 2017 Integral analysis of boundary layer flows with pressure gradient. Physical Review Fluids 2 (9), 092601.
  • White (2008) White, Frank M 2008 Fluid mechanics. The McGraw Hill Companies.
  • Whitmore et al. (2022) Whitmore, Michael P., S.T., Bose & Moin, P. 2022 Progress on slip wall modeled LES for predicting smooth body separation. Center for Turbulence Research, Annual Research Briefs pp. 59–70.
  • Williams et al. (2020) Williams, Owen, Samuell, Madeline, Sarwas, E Sage, Robbins, Matthew & Ferrante, Antonino 2020 Experimental study of a CFD validation test case for turbulent separated flows. In AIAA Scitech 2020 Forum, p. 0092.