1. Introduction
Like most turbulent flows, turbulent wall-bounded flows are characterised by a wide range of scales of motion. Interscale interactions, within or between the inner and outer regions, lead to energy transfer that regulate flow dynamics. The inner region of wall-bounded flows is mainly governed by the near-wall cycle (Jiménez & Pinelli Reference Jiménez and Pinelli1999), proved to be Reynolds-number-independent (Hutchins et al. Reference Hutchins, Nickels, Marusic and Chong2009). But as the Reynolds number is increased, the presence of large-scale (LS) motions in the outer region is accompanied by the footprint of these large scales in the inner region, which is reflected in both wall-parallel velocity fluctuations (Hoyas & Jiménez Reference Hoyas and Jiménez2006) and wall shear stress (Abe, Kawamura & Choi Reference Abe, Kawamura and Choi2004). In addition to this phenomenon, known as superposition, amplitude modulation of the small scales of near-wall turbulence is also observed. The intensity of small-scale (SS) fluctuations is modulated by the superimposed LS signal. As a result, the near-wall peak of streamwise velocity fluctuations no longer follows the viscous scaling (Monkewitz & Nagib Reference Monkewitz and Nagib2015), and its intensity becomes dependent on the friction Renyolds number. These observations were extensively studied in the last 15 years (Mathis, Hutchins & Marusic Reference Mathis, Hutchins and Marusic2009; Marusic, Mathis & Hutchins Reference Marusic, Mathis and Hutchins2010; Ganapathisubramani et al. Reference Ganapathisubramani, Hutchins, Monty, Chung and Marusic2012; Talluru et al. Reference Talluru, Baidya, Hutchins and Marusic2014; Baars et al. Reference Baars, Talluru, Hutchins and Marusic2015; Marusic, Baars & Hutchins Reference Marusic, Baars and Hutchins2017; Dogan et al. Reference Dogan, Örlü, Gatti, Vinuesa and Schlatter2019) mainly because of the scientific challenges and unresolved questions raised by these phenomena occurring at high Reynolds number. But the interscale and inner–outer interactions observed in canonical boundary layers also share connections with other important phenomenon, such as adverse pressure gradients, wall roughness (Lozier, Marusic & Deshpande Reference Lozier, Marusic and Deshpande2024) or freestream turbulence (Dogan, Hanson & Ganapathisubramani Reference Dogan, Hanson and Ganapathisubramani2016). From a practical point of view, having a modelling framework, based on a Reynolds-averaged Navier–Stokes (RANS) approach, that takes into account these interactions then appears very desirable to predict these important features of wall-bounded flows.
Recent efforts by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) provide a RANS two-scale Reynolds stress model (RSM) whose principle lies in the transport of the energy carried by small and large scales, and which aims to describe the energy transfer process occurring between the inner and outer regions of wall-bounded flows. Unlike standard RANS turbulence models that impose a constant-rate energy transfer from the production region at large scales down to the small scales where energy is dissipated, multi-scale models (Hanjalić et al. Reference Hanjalić, Launder and Schiestel1979; Schiestel Reference Schiestel1987) offer the possibility of considering several interscale energy transfer rates. The two-scale RSM developed by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) was specifically designed to capture SS and LS contributions in wall-bounded flows that can be discriminated by their Reynolds number dependences. To make it simple, the SS contribution in the two-scale model may be assimilated to the self-sustained near-wall cyle contribution (Jiménez & Pinelli Reference Jiménez and Pinelli1999), and is shown to be independent of the Reynolds number
$\textit{Re}_{\tau }$
, while the LS contribution is meant to represent the attached eddies (Marusic & Monty Reference Marusic and Monty2019), LS motions and very-large-scale motions (or superstructures) (see e.g. Smits, McKeon & Marusic Reference Smits, McKeon and Marusic2011). The multi-scale (Schiestel Reference Schiestel1987) framework from which the two-scale RSM is derived offers the possibility of taking into account interscale energy transfers within and between the inner and outer regions of wall-bounded flows. The disctinction between the small and large scales is inspired from the scale separation proposed by Lee & Moser (Reference Lee and Moser2019) and applied to direct numerical simulations (DNS) of turbulent channel flows. The cut-off wavelength was arbitrarily fixed at
$\lambda ^+=1000$
, where superscript
$+$
refers to quantities express in wall units using the friction velocity
$u_{\tau }$
and the kinematic viscosity
$\nu$
. This threshold value was already evidenced in other studies addressing the issue of interscale interactions in wall-bounded flows (e.g. Jiménez Reference Jiménez2024). In the two-scale RSM, the cut-off wavelength is not directly fixed but results from the ratio of energy carried by the two transported scales. Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) showed that the cut-off wavelength in the two-scale RSM was approximately
$1000$
wall units in the inner region, and increases in the outer region, but most importantly, that it clearly separates the SS and LS contributions. The energy transfer from the large scales, marked with superscript
$^{(1)}$
, to the small scales, marked
$^{(2)}$
, is modelled by a tensor, denoted
${\tilde {\varepsilon }_{\textit{ij}}}^{(1)}$
, corresponding to the rate of transfer of the
$ij$
-component of the LS contribution of the Reynolds stress to the corresponding SS counterpart. The tensor trace
${\tilde {\varepsilon }}^{(1)}$
is then one of the transported variables in the two-scale RSM. Comparisons with the DNS results from Lee & Moser (Reference Lee and Moser2019) have proved the relevance of the approach to catch the SS and LS contributions. The only remaining problem is that of modelling energy transfer from the small to the large scales observed in DNS in the near-wall region, for which the proposed ad hoc closure is only suited to the channel cases dealt with at the time. To extend the two-scale RSM to very high Reynolds number boundary layer configurations, the modelling needs to be revised, and this requires an in-depth study of energy transfers in this region based on the available DNS data.
The LS contributions to the wall-parallel components of the Reynolds stress tensor observed in wall-bounded flows around the near-wall peak are commonly attributed to the superposition phenomenon (Marusic et al. Reference Marusic, Mathis and Hutchins2010; Andreolli et al. Reference Andreolli, Gatti, Vinuesa, Örlü and Schlatter2023) that corresponds to the signatures (or footprints) in the near-wall region of the large scales (Hutchins & Marusic Reference Hutchins and Marusic2007) originating from the outer region. Additionally, large structures influence the smallest near-wall structures by modulating their amplitude at large scales (Mathis et al. Reference Mathis, Hutchins and Marusic2009). Since it is now well established that the near-wall cycle is Reynolds number independent (Hutchins et al. Reference Hutchins, Nickels, Marusic and Chong2009), the near-wall peak dependence on
$\textit{Re}_{\tau }$
may be attributed to superposition and amplitude modulation, superposition being by far the main contributor to this dependence (Marusic et al. Reference Marusic, Baars and Hutchins2017). These effects were introduced in the two-scale RSM using an ad hoc formulation that was shown to exhibit several limitations despite obvious merits concerning the increase of the near-wall peak of streamwise component of the diagonal of the Reynolds stress tensor. First, the model was not invariant by rotation and was thus not suitable to computational fluid dynamics code requirements. But far more importantly, Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) pointed out that the near-wall peak Reynolds number dependence involves backscattering energy transfer as long as we consider the scale separtion introduced by Lee & Moser (Reference Lee and Moser2019). This echoes the work of Andreolli et al. (Reference Andreolli, Gatti, Vinuesa, Örlü and Schlatter2023), who have shown, using numerical simulations with specific forcings, that in the absence of amplitude modulation, the imprint of large scales near the wall disappears. They thus suggest that there are energy transfer mechanisms from small to large scales in the near-wall region (see figure 14 of Andreolli et al. Reference Andreolli, Gatti, Vinuesa, Örlü and Schlatter2023). A close look at the DNS results of Lee & Moser (Reference Lee and Moser2019) indicated that the LS contributions to the streamwise and spanwise diagonal components of the Reynolds stress, say
$\overline {u^{\prime 2}}^{(1)}$
and
$\overline {w^{\prime 2}}^{(1)}$
– where the
$ (\overline {\phantom {x}} )$
symbol denotes Reynolds averaging, and
$ (\,' )$
denotes fluctuating quantities – in the inner region result from the balance between viscous dissipation and diffusion on one hand, and turbulent diffusion and backscattering on the other (see figures 7 and 8 of Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024). The two source terms being turbulent diffusion, or more exactly turbulent transport, and backscattering, and since turbulent transport contribution is a direct consequence of the LS contribution to the near-wall peak that tends to spread and lower the peak, the main driver of the LS contribution in the inner region is the energy transferred from small scale to large scale. Thus around the near-wall peak, the transfer rate
${\tilde {\varepsilon }}^{(1)}$
was shown to have negative values indicating transfer from the small scales to the largest. The formulation proposed by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) was unable to reproduce such behaviour for reasons inherent to the model closures. This limitation was pointed out as the main weakness of the proposed model, and we intend here to extend the two-scale RSM to make it suitable to account for the interscale energy transfers in an appropriate manner, respecting the observations gained from the DNS of Lee & Moser (Reference Lee and Moser2019).
The objective of this study is then twofold. First, a systematic and comprehensive analysis of the high Reynolds number channel flow configurations provided by Lee & Moser (Reference Lee and Moser2019) should identify energy transfer processes that occur between the LS and SS contributions in the inner and outer regions. Second, the term controlling the energy transfers from small to large scales in the two-scale RSM will be reformulated with respect to the interscale contribution based on the observations made in the aforementioned analysis. The paper is then organised as follows. Section 2 presents an analysis of the DNS data obtained in channel flow configurations by Lee & Moser (Reference Lee and Moser2019) focusing on interscale energy transfers. In § 3, the conclusions drawn from the analysis are leveraged to revisit the modelling of energy transfers between large and small scales. The new version of the two-scale RSM is then tested in § 4 against DNS and experimental data obtained in high Reynolds number boundary layer configurations, demonstrating the relevance of the model used to describe the energy transfers responsible for high Reynolds number effects observed in wall-bounded flows.
2. Interscale energy transfers
The DNS computations in turbulent channel flows at moderate to high Reynolds numbers, namely
$\displaystyle \textit{Re}_{\tau }\in \{550,1000,2000,5200\}$
, conducted by Lee & Moser (Reference Lee and Moser2019) are thoroughly analysed in this section in order to shed light on the interscale energy transfer processes occurring in both the inner and outer regions of wall-bounded flows. For this purpose, energy spectra are first scrutinised. We denote
$\varPsi _{\textit{ij}} (\kappa _x,y,\kappa _z )$
as the two-dimensional spectrum tensor corresponding to the Fourier transform of the two-point correlation tensor
$R_{\textit{ij}} (r_x,y,r_z )$
, where
$r_x$
and
$r_z$
are the separations between the two points in the streamwise and spanwise directions, respectively, and
$\kappa _x$
and
$\kappa _z$
are the corresponding wavenumbers. By partitioning spectra
$\varPsi _{\textit{ij}}$
in two parts for a given cut-off wavenumber
$\kappa ^+=\sqrt {\kappa _x^{+^2}+\kappa _z^{+^2}}=\kappa _c^+$
, Lee & Moser (Reference Lee and Moser2019) demonstrated that the mean Reynolds stress components can be split into two distinct contributions with respect to the Reynolds number dependency. The first, mainly concentrated in the inner region, is independent of the Reynolds number and related to the near-wall cycle (Jiménez & Pinelli Reference Jiménez and Pinelli1999). This first contribution corresponds to the action of the small scales of turbulence, i.e. those located above the chosen cut-off wavenumber. The second contribution, which strongly depends on the Reynolds number, is twofold. In the outer region, LS structures promote the growth of the Reynolds stress component as the Reynolds number increases. In the meantime, a LS contribution also exists in the inner region for the streamwise and the spanwise diagonal Reynolds stress components. This specific LS contribution was proved to be correlated to the LS structures of the outer region via the superposition phenomenon, which ultimately leads to the increase of the near-wall peak of the turbulent kinetic energy as the Reynolds number is increased. In their analysis, Lee & Moser (Reference Lee and Moser2019) applied a scale separation to the two-dimensional spectra at a fixed wavelength
$\displaystyle \lambda _c^+={\displaystyle 2\pi }/{\displaystyle \kappa _c^+}=1000$
. In figure 2 of Lee & Moser (Reference Lee and Moser2019), showing results obtained at
$\textit{Re}_{\tau }=5200$
, there is a clear evidence that large scales imprint throughout the whole channel height for
$R_{11}=\overline {{u^{\prime }}^2}$
and
$R_{33}=\overline {{w^{\prime }}^2}$
components of the two-point velocity correlation tensor. The energy on
$\overline {{u^{\prime }}^2}$
is mainly carried out by streamwise elongated structures (streak-like structures) with
$\displaystyle \lambda _x^+ \gg \lambda _z^+$
, while the energy for
$\overline {{v^{\prime }}^2}$
– and in a lesser manner for
$\overline {{w^{\prime }}^2}$
– is more spread over a large range of wavelengths. But more interestingly, a remarkable behaviour is observed on the spanwise component
$\overline {{w^{\prime }}^2}$
in the logarithmic layer, i.e.
$y^+\in [100,1000]$
in the Lee & Moser (Reference Lee and Moser2019) figure, where the energy is being transferred between streamwise elongated structures at large scale and more isotropic structures (
$\displaystyle \lambda _x^+ \approx \lambda _z^+$
) at lower scale,
$\lambda ^+$
changing for approximately a decade. This behaviour is almost absent on the
$R_{22}=\overline {{v^{\prime }}^2}$
component. This observation points out the specific role played by the spanwise component
${w^{\prime }}^2$
in the energy transfer from large to small scale acting in the logarithmic layer.

Figure 1. One-dimensional premultiplied spectra
$\displaystyle \varPhi _{\textit{ij}}^z(\kappa _x,y)$
(red line contours) and
$\displaystyle \varPhi _{\textit{ij}}^x(y,\kappa _z)$
(grey filled contours) of the two-point velocity correlations (a)
$\overline {{u^{\prime }}^2}$
, (b)
$\overline {{v^{\prime }}^2}$
and (c)
$\overline {{w^{\prime }}^2}$
) obtained from DNS of channel flows of Lee & Moser (Reference Lee and Moser2019). Each spectrum is normalised by its maximum value. Equidistant isocontour values are
$\{1/6,1/3,1/2,2/3,5/6\}$
, and are represented with increasingly darker colours.
To further confirm these observations, we consider the one-dimensional premultiplied energy spectra
$\displaystyle \varPhi _{\textit{ij}}^z(\kappa _x,y)=\smallint \kappa _z\varPsi _{\textit{ij}}\, {\textrm{d}}\kappa _z$
and
$\displaystyle \varPhi _{\textit{ij}}^x(y,\kappa _z)=\smallint \kappa _x\varPsi _{\textit{ij}}\, {\textrm{d}}\kappa _x$
. Three of their components are plotted in figure 1 for the three diagonal components of the two-point velocity correlation tensor for the channel configuration at the highest friction Reynolds number
$\textit{Re}_{\tau }=5200$
. The one-dimensional spectra for the streamwise velocity cover wavelength ranges separated by approximately an order of magnitude, indicating there again the elongated nature of the structures bearing the
$\overline {{u^{\prime }}^2}$
component, whereas spectra for
$\overline {{v^{\prime }}^2}$
and
$\overline {{w^{\prime }}^2}$
show more comparable ranges of
$\lambda _x$
and
$\lambda _z$
values for a given isocontour value. The Reynolds number effects in the inner region, which lead to the failure of viscous scaling for wall-parallel Reynolds stress components (Monkewitz & Nagib Reference Monkewitz and Nagib2015), are related to the emergence at high Reynolds numbers of an outer peak, visible in figure 1(a). The location of the peak on the spectrum
$\displaystyle \varPhi _{11}^z$
is
$\textit{Re}_{\tau }$
-dependent (Mathis et al. Reference Mathis, Hutchins and Marusic2009), and at very high
$\textit{Re}_{\tau }$
, the location of the outer peak marks the start of a broad plateau (Vallikivi, Hultmark & Smits Reference Vallikivi, Hultmark and Smits2015; Samie et al. Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018). The high wavelength energy in the near-wall region in spectra
$\displaystyle \varPhi _{11}^z$
and
$\displaystyle \varPhi _{11}^x$
is the footprint of this outer energy site. The picture is different on spectra
$\displaystyle \varPhi _{33}^z$
and
$\displaystyle \varPhi _{33}^x$
, for which no clear outer peak can be identified. However, a Reynolds number dependence in the outer and inner regions is also reported (see e.g. Lee & Moser Reference Lee and Moser2019; Jiménez Reference Jiménez2024) for the spanwise fluctuations intensity
$\overline {{w^{\prime }}^2}$
, while spectra
$\displaystyle \varPhi _{22}^z$
and
$\displaystyle \varPhi _{22}^x$
do not show any
$\textit{Re}_{\tau }$
-dependence in the inner region (Lee & Moser Reference Lee and Moser2019). In figure 1(c) for spectra
$\displaystyle \varPhi _{33}^z$
, energy around the inner peak is spread over a large range of wavelength, with amplitudes at large scales almost constant from
$y^+=30$
up to several hundreds of wall units for the most elongated structures.

Figure 2. One-dimensional premultiplied spectra (a)
$\displaystyle \varPhi _{11}^x(y,\kappa _z)$
and (b)
$\displaystyle \varPhi _{33}^x(y,\kappa _z)$
for the four considered Reynolds numbers
$\textit{Re}_{\tau }=550,1000,2000,5200$
in channel flow DNS from Lee & Moser (Reference Lee and Moser2019). Isocontour values are
$1/6,1/3,1/2,2/3,5/6$
, and are represented with increasingly darker grey lines: dotted for
$\textit{Re}_{\tau }=550$
, dashed for
$\textit{Re}_{\tau }=1000$
, dash-dotted for
$\textit{Re}_{\tau }=2000$
, and solid for
$\textit{Re}_{\tau }=5200$
. Square symbols denote the maximum
$y^+$
location reached by each contour. Darker red colour indicates increased
$\textit{Re}_{\tau }$
values.
To illustrate the determining role of the
$\overline {{w^{\prime }}^2}$
component in the Reynolds dependence of the inner region, figure 2 compares one-dimensional spectra
$\displaystyle \varPhi ^x_{11}$
and
$\displaystyle \varPhi ^x_{33}$
for the four considered Reynolds numbers. Below
$\lambda _z^+=1000$
, spectra are almost independent of the Reynolds number. Above, Reynolds number effects on the
$\overline {{u^{\prime }}^2}$
component tend to increase the LS contributions in the outer region following a linear evolution in the
$ (y^+,\lambda _z^+ )$
plane, independent of the Reynolds number, marked in the figure by the locations of the maximum
$y^+$
value reached by each isocontour. This line is very close to that found by Jiménez (Reference Jiménez2024) when plotting the maxima for each wavelength in the outer region, which approximately follows a line
$\lambda _z^+=5y^+$
. In the inner region, the near-wall peak of
$\overline {{u^{\prime }}^2}$
spreads over larger scales as
$\textit{Re}_{\tau }$
is increased. The effect of the Reynolds number, and therefore of the increase of the LS contributions, on the
$\overline {{u^{\prime }}^2}$
correlation appears to be localised in the near-wall peak region. Important differences can be noticed when looking at the
$\varPhi ^x_{33}$
spectra. Below
$\lambda _z^+=1000$
, if spectra are still independent of the Reynolds number, then the distribution of the increasing contribution of the large scale in the
$ (y^+,\lambda _z^+ )$
plane is dependent on the Reynolds number, as indicated by the different lines drawn by the maximum locations of the isocontours. Isocontours of
$\varPhi ^x_{33}$
are significantly modified in the region
$y^+\in [100,400]$
as
$\textit{Re}_{\tau }$
is increased. Unlike component
$\overline {{u^{\prime }}^2}$
, the Reynolds number dependence is not restricted to the near-wall peak region, but spreads over a large range of
$y^+$
values.

Figure 3. Opposite of the energy flux rates (a)
$\displaystyle {\tilde {\varepsilon }_{11}}^{(1)}$
and (b)
$\displaystyle {\tilde {\varepsilon }_{33}}^{(1)}$
obtained by DNS (Lee & Moser Reference Lee and Moser2019) for
$\textit{Re}_{\tau }=550,1000,2000,5200$
. Darker symbols indicate increased
$\textit{Re}_{\tau }$
values.
In order to see how the observations made on spectra
$\varPhi ^x_{11}$
and
$\varPhi ^x_{33}$
translate in terms of interscale energy transfers, the evolutions across the channel height of the energy flux rate components
$\displaystyle {\tilde {\varepsilon }_{11}}^{(1)}$
and
$\displaystyle {\tilde {\varepsilon }_{33}}^{(1)}$
for the four considered friction Reynolds numbers are plotted in figure 3. The energy flux rate tensor
${\tilde {\varepsilon }_{\textit{ij}}}^{(1)}$
, following notations introduced by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) and recalled in § 3, is computed from the DNS data as the rest of the turbulent term due to triple correlations once the wall-normal turbulent transport is removed. Therefore, by definition, this term is symmetrical between the Reynolds stress budgets of the small and large scales, and is exactly zero if integrated over the whole spectra. Negative values of
$\displaystyle -{\tilde {\varepsilon }_{11}}^{(1)}$
or
$\displaystyle -{\tilde {\varepsilon }_{33}}^{(1)}$
in figure 3 indicate energy transfers from large to small scales. The energy flux rate tensor must be seen as a sink term in the budget of the LS contribution to the Reynolds stress tensor, and as a source term for the SS contribution respectively (see § 3). In figure 3,
$\displaystyle -{\tilde {\varepsilon }_{11}}^{(1)}$
and
$\displaystyle -{\tilde {\varepsilon }_{33}}^{(1)}$
are shown to have positive values in the inner region. In the framework of the scale separation performed by Lee & Moser (Reference Lee and Moser2019), these positive values indicate backscatter with energy transfer from small to large scale on each of the components
$\overline {{u^{\prime }}^2}$
and
$\overline {{w^{\prime }}^2}$
. Only negative values are observed for
$\displaystyle -{\tilde {\varepsilon }_{22}}^{(1)}$
, confirming that backscattering only concerns the streamwise and spanwise components of the Reynolds stress tensor. Backscatter fluxes are very localised for the
${u^{\prime }}^2$
component, and centred about
$y^+=6.5$
and vanishing at
$y^+=15$
, while they cover a region ranging from
$y^+=1$
up to a value at approximately
$y^+=100$
that drops as
$\textit{Re}_{\tau }$
increases. The locations of these energy transfers from small to large scales are in line with the energy distributions observed in figure 2. Transfers to large scales are more distributed across the inner region, up to the logarithmic region, for the
$\overline {{w^{\prime }}^2}$
component. The LS contribution to
$\overline {{w^{\prime }}^2}$
is thus increased with
$\textit{Re}_{\tau }$
in the whole inner region, and due to pressure distribution, so is the LS contribution to
$\overline {{u^{\prime }}^2}$
. The peak location of
$\displaystyle {\tilde {\varepsilon }_{33}}^{(1)}$
tends to lower the
$y^+$
value as
$\textit{Re}_{\tau }$
increases, whereas it remains fixed for
$\displaystyle {\tilde {\varepsilon }_{11}}^{(1)}$
. Interestingly, the amplitude of
$\displaystyle -{\tilde {\varepsilon }_{11}}^{(1)}$
scales with
$\textit{Re}_{\tau }^{3/5}$
, while
$\displaystyle -{\tilde {\varepsilon }_{33}}^{(1)}$
scales with
$\textit{Re}_{\tau }^{1/3}$
, showing a more rapid increase on
$\overline {{u^{\prime }}^2}$
than on
$\overline {{w^{\prime }}^2}$
of the LS contribution in the near-wall region. In the present study, we do not discuss the asymptotic behaviour of the near-wall intensity, whether it is of logarithmic type (Monkewitz & Nagib Reference Monkewitz and Nagib2015; Hwang Reference Hwang2024) or following a power law (Chen & Sreenivasan Reference Chen and Sreenivasan2020; Pirozzoli Reference Pirozzoli2024). A thorough discussion on this point is reported by Jiménez (Reference Jiménez2024). The observed scalings in figure 3 are limited to a small range of Reynolds numbers, and no conclusion can be drawn from that concerning the asymptotic behaviour of the near-wall peak. However, these observations will benefit the energy transfer modelling described in § 3, in particular to characterise evolutions with respect to
$\textit{Re}_{\tau }$
for intermediate values, i.e.
$\textit{Re}_{\tau } \in [500,4000]$
.
3. Modelling energy transfers from small to large scales
Before revising the two-scale RSM of Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024), the foundations of the model are briefly recalled. A thorough explanation of the model’s principles can be found in Schiestel (Reference Schiestel1987). To model the LS and SS contributions resulting from the partitioning of spectra of two-point velocity correlations, Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024), relying on ideas introduced by Schiestel (Reference Schiestel1974), developed a two-scale RSM whose underlying basis for each of the two transported scales is that of the RANS RSM proposed by Manceau (Reference Manceau2015). The spectral partitioning consists of two slices covering the wavenumber range
$[0,\kappa _c]\times [\kappa _c,\kappa _\eta ]$
(see figure 1 in Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024). The first slice,
$m=1$
, corresponding to large scale, is delimited by the cut-off wavenumber
$\kappa _c$
. The second slice,
$m=2$
, representing the SS contribution, ranges from
$\kappa _c$
to
$\kappa _\eta$
, the Kolmogorov scale at which energy is disspated. The cut-off wavenumber
$\kappa _c$
is not imposed but results from the energy balance between the two spectral slices provided by the model. Practically, it was shown (Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024) to adequately discriminate the SS and LS contributions in a similar manner to that of Lee & Moser (Reference Lee and Moser2019), where
$\lambda ^+=1000$
was imposed. The transport equations of partially integrated Reynolds stress tensors, over spectral slices
$m=1$
and
$m=2$
are considered, resulting in a set of
$16$
equations, i.e.
$2\times 6$
equations for the partial Reynolds stresses,
$2$
equations for the associated outgoing energy fluxes from each spectral bandwidth, and
$2$
equations to manage wall damping according to the elliptic blending approach of Manceau & Hanjalić (Reference Manceau and Hanjalić2002). The outgoing energy flux for the SS contribution, namely
${\tilde {\varepsilon }}^{(2)}$
, corresponds to dissipation at Kolmogorov scale, whereas
${\tilde {\varepsilon }}^{(1)}$
is the rate of energy transferred from large to small scale. The near-wall dissipations for each of the two spectral slices, denoted
${\varepsilon }^{(1)}_w$
and
${\varepsilon }^{(2)}_w$
, respectively, are treated separately using the algebraic expressions
where
${k}^{(m)}$
is the turbulent kinetic energy in the spectral slice
$m$
, and
$y_0$
is a parameter controlling the extent of the dissipation due to the presence of the wall. On each spectral slice, the total dissipation
${\varepsilon }^{(m)}$
is therefore decomposed into a homogeneous part
${\tilde {\varepsilon }}^{(m)}$
corresponding to the outgoing energy flux, and an inhomogeneous part
${\varepsilon }^{(m)}_w$
due to wall. This is a key feature of the model, enabling us to isolate the energy flux
${\tilde {\varepsilon }}^{(1)}$
that characterises the energy transfers from large to small scales, and which can be then be treated directly as a transported variable.
The value of
$y_0^+$
was slighly adjusted compared to that used in Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024), where
$y_0^+=2.5$
was retained to match the near-wall behaviour obtained by the reference model of Manceau (Reference Manceau2015). Here, we retain
$y_0^+=4$
for both slices
$m=1$
and
$m=2$
. The validity of (3.1) is tested against the DNS data provided by Lee & Moser (Reference Lee and Moser2019) in figure 4. In the DNS, the dissipation is not separated into homogeneous and inhomogeneous parts; only the energy flux
${\tilde {\varepsilon }}^{(1)}$
has been explicitly computed. The matching with DNS data is then very good on the LS contribution insofar as the dissipation obtained in the DNS is only due to near-wall effects, the energy behind essentially transferred to small scale without viscous dissipation away from the walls. The matching is nonethless very satisfying on the SS contribution of the dissipation in the vicinity of the wall for all the considered
$\textit{Re}_{\tau }$
numbers. A slightly better agreement can be obtained using the standard form
$\displaystyle {\varepsilon }^{(2)}_w={2\nu {k}^{(2)}}/{y^2}$
, but for consistency reasons, and also for modelling reasons detailed below, we decided to keep the form given in (3.1).

Figure 4. Wall dissipations
${\varepsilon }^{(1)}$
and
${\varepsilon }^{(2)}$
for the LS and SS contributions, respectively. Darker grey colours show increasing
$\textit{Re}_{\tau }$
values. Symbols are DNS data (Lee & Moser Reference Lee and Moser2019), and solid lines are the expressions given in (3.1). Dashed lines are obtained with the expression
$\displaystyle {\varepsilon }^{(2)}_w={2\nu {k}^{(2)}}/{y^2}$
.
The two-scale approach relies on a generalised form of the Reynolds decomposition for each instantaneous velocity component
$u_i=\overline {u_i} + {u^{\prime}_i}^{(1)}+{u^{\prime}_i}^{(2)}$
. To lighten notations,
$R_{\textit{ij}}=\overline {u'_iu'\!_j}$
now designates the one-point velocity correlation, i.e. the
$ij$
-component of the Reynolds stress tensor. The two-scale RSM is governed by the following system composed of sixteen equations:
\begin{equation} \begin{array}{l} \kern-7pt \displaystyle \frac {{\textrm{D}}{R}^{(1)}_{\textit{ij}}}{{\textrm{D}}t} \!=\! - {R}^{(1)}_{\textit{ik}}\frac {\partial \overline {u\!_j}}{\partial x_k} - {R}^{(1)}_{jk}\frac {\partial \overline {u_i}}{\partial x_k} + {\varPhi }^{(1)}_{\textit{ij}} - {\varepsilon }^{(1)}_{\textit{ij}} + {M}^{(1)}_{\textit{ij}} + \frac {\partial }{\partial x_l} \! \left [\! \left (\nu +\frac {c_s}{{\sigma _k\!}^{(1)}}{R}^{(1)}_{lm}{t}^{(1)}_t\right )\frac {\partial\! {R}^{(1)}_{\textit{ij}}}{\partial x_m}\! \right ]\!, \\[0.5cm] \kern-7pt \displaystyle \frac {{\textrm{D}}{R}^{(2)}_{\textit{ij}}}{{\textrm{D}}t} \!=\! \displaystyle -{R}^{(2)}_{\textit{ik}}\frac {\partial \overline {u\!_j}}{\partial x_k}-{R}^{(2)}_{jk}\frac {\partial \overline {u_i}}{\partial x_k} + {\tilde {\varepsilon }}^{(1)}_{\textit{ij}} + {\varPhi }^{(2)}_{\textit{ij}} - {\varepsilon }^{(2)}_{\textit{ij}} + \frac {\partial }{\partial x_l}\left [\! \left (\nu +\frac {c_s}{{\sigma _k\!}^{(2)}}{R}^{(2)}_{lm}{t}^{(2)}_t\right )\frac {\partial {R}^{(2)}_{\textit{ij}}}{\partial x_m}\! \right ]\!, \\[0.5cm] \kern-7pt \displaystyle \frac {{\textrm{D}}{\tilde {\varepsilon }}^{(1)}}{{\textrm{D}}t} \!=\! \displaystyle \frac {{C'\!_{\varepsilon\! _1}}^{(1)}{P_k\!}^{(1)}-{C_{\varepsilon _2}\!}^{(1)}{\tilde {\varepsilon }}^{(1)}}{{t}^{(1)}_t} +\displaystyle \frac {\partial }{\partial x_l}\left [\! \left (\nu +\frac {c_s}{{\sigma _{\varepsilon }\!}^{(1)}}{R}^{(1)}_{lm}{t}^{(1)}_t \!\right )\frac {\partial {\tilde {\varepsilon }}^{(1)}}{\partial x_m}\right ]\!, \\[0.5cm] \kern-7pt \displaystyle \frac {{\textrm{D}}{\tilde {\varepsilon }}^{(2)}}{{\textrm{D}}t} \!=\! \displaystyle \frac {{C'_{\varepsilon _1}\!\!}^{(2)}{P_k\!}^{(2)}-{C'\!_{\varepsilon _2}\!\!\!}^{(2)}{\tilde {\varepsilon }}^{(2)} +{C_{\varepsilon _3}\!\!\!}^{(2)}{\tilde {\varepsilon }}^{(1)}}{{t}^{(2)}_t} +\frac {\partial }{\partial x_l}\left [\left (\nu +\frac {c_s}{{\sigma _{\varepsilon }\!}^{(2)}}{R}^{(2)}_{lm}{t}^{(2)}_t\right )\frac {\partial {\tilde {\varepsilon }}^{(2)}}{\partial x_m}\right ]\!, \\[0.3cm] \kern-5pt \displaystyle {\alpha }^{(1)} - \displaystyle {{{l}_t^{(1)}}^2}{\nabla} ^2 {\alpha }^{(1)} = 1,\quad{\alpha }^{(2)} - \displaystyle {{l}^{(2)}_t}^2{\nabla} ^2 {\alpha }^{(2)} = 1. \end{array} \end{equation}
Here,
$\varPhi ^{(m)}_{\textit{ij}}$
is the pressure–strain correlation corresponding to the redistribution term in the slice. It is composed of a homogeneous part given by the Spezial–Sarkar–Gatski model (Speziale, Sarkar & Gatski Reference Speziale, Sarkar and Gatski1991) and an inhomogeneous contribution provided by Manceau & Hanjalić (Reference Manceau and Hanjalić2002). Also,
$\displaystyle {t}^{(m)}_t$
is the turbulent time scale in slice
$m$
, and is defined in (3.4) below following Durbin (Reference Durbin1991). Each partial dissipation tensor
$\displaystyle {\varepsilon }^{(m)}_{\textit{ij}}$
is computed from its partial energy dissipation
$\displaystyle {\varepsilon }^{(m)}=({1}/{2}){\varepsilon }^{(m)}_{ii}$
following the approach of Manceau (Reference Manceau2015). The same is applied to the partial rate of transfer
${\tilde {\varepsilon }}^{(1)}_{\textit{ij}}$
:
\begin{eqnarray} && {\varepsilon }^{(m)}_{\textit{ij}} = \displaystyle \big (1-{f_w\!\!}^{(m)}\big )\frac {{R}^{(m)}_{\textit{ij}}}{{k}^{(m)}}{\varepsilon }^{(m)} + {f_w\!\!}^{(m)}\left (C_{\varepsilon }^{(m)}\frac {{R}^{(m)}_{\textit{ij}}}{{k}^{(m)}}{\varepsilon }^{(m)}+\big (1-C_{\varepsilon }^{(m)}\big )\frac {2}{3}{\tilde {\varepsilon }}^{(m)}\delta _{\textit{ij}}\right )\!, \nonumber\\ && {\tilde {\varepsilon }}^{(1)}_{\textit{ij}} = \displaystyle \big (1-{f_w\!\!}^{(2)}\big )\frac {{R}^{(1)}_{\textit{ij}}}{{k}^{(1)}}{\tilde {\varepsilon }}^{(1)} + {f_w\!\!}^{(2)}\left ( C_{\tilde {\varepsilon }}\frac {{R}^{(1)}_{\textit{ij}}}{{k}^{(1)}}{\tilde {\varepsilon }}^{(1)} +\left (1-C_{\tilde {\varepsilon }}\right )\frac {2}{3}{\tilde {\varepsilon }}^{(1)}\delta _{\textit{ij}}\right )\!. \end{eqnarray}
The blending function
$\displaystyle {f_w\!\!}^{(m)}$
, evolving from
$0$
at the wall to
$1$
far from walls, and linking the homogeneous and inhomogeneous parts of
${\varPhi }^{(m)}_{\textit{ij}}$
,
$\displaystyle {\varepsilon }^{(m)}_{\textit{ij}}$
and
$\displaystyle {\tilde {\varepsilon }}^{(1)}_{\textit{ij}}$
, are calculated directly from functions
$\displaystyle {\alpha }^{(m)}$
and are taken as
$\displaystyle {f_w\!}^{(1)}=\displaystyle {{\alpha }^{(1)}}^2$
and
$\displaystyle {f_w\!}^{(2)}=\displaystyle {{\alpha }^{(2)}}^3$
. Functions
${\alpha }^{(m)}$
are controlled by the turbulent length scale
${l}^{(m)}_t$
, given in
\begin{equation} \begin{array}{l} \displaystyle {t}^{(m)}_t = \displaystyle \max \left (\frac {{k}^{(m)}}{{\varepsilon }^{(m)}},{C_T\!\!\kern1.5pt}^{(m)}\left (\frac {\nu }{{\varepsilon }^{(m)}}\right )^{\frac {1}{2}}\right ), \nonumber\\ \displaystyle {l}^{(m)}_t = \displaystyle {C_L\!\!\kern1.5pt}^{(m)}\max \left (\frac {{{k}^{(m)}}^{\frac {3}{2}}}{{\varepsilon }^{(m)}},{C_\eta\!\!}^{(m)}\left (\frac {\nu ^{3}}{{\varepsilon }^{(m)}}\right )^{\frac {1}{4}}\right ). \end{array} \end{equation}
By construction, the length scale
${l}^{(2)}_t$
associated with small scale is almost independent of
$\textit{Re}_{\tau }$
, and so is the blending function
$\displaystyle {f_w\!\!}^{(2)}$
. On the contrary,
$\displaystyle {f_w\!\!}^{(1)}$
is shifted up towards higher
$y^+$
values when
$\textit{Re}_{\tau }$
is increased.
To account for the Reynolds number dependence of
$\overline {{u^{\prime }}^2}$
and
$\overline {{w^{\prime }}^2}$
in the inner region and due to LS contributions, the
${M}^{(1)}_{\textit{ij}}$
term was introduced into the model. Because
${\tilde {\varepsilon }}^{(1)}_{\textit{ij}}$
is given by (3.3), it is strictly positive, and instead of searching for a formulation allowing
${\tilde {\varepsilon }}^{(1)}_{\textit{ij}}$
to have negative values, it is more convenient to consider an additional term to balance the budget of
${R}^{(1)}_{\textit{ij}}$
in the inner region. In this regard, according to the decomposition introduced by Lee & Moser (Reference Lee and Moser2019), the rate of transfer
${\tilde {\varepsilon }}^{(1)}_{\textit{ij}}$
computed in the DNS results from multiple contributions, and there is no contradiction in modelling energy transfers between large and small scales from several terms. Term
${M}^{(1)}_{\textit{ij}}$
is thus the key term on which efforts are concentrated below, in order to reproduce the observations made in § 2.
The analysis presented in § 2 supports the idea of two separated contributions for interscale energy transfer in the near-wall region, i.e. a first one acting on energy transfers for
$\overline {{u^{\prime }}^2}$
, and a second one for
$\overline {{w^{\prime }}^2}$
, with different Reynolds number dependences. Let
$\displaystyle n_i$
be the wall-normal vector, and let
$t_i^x$
and
$t_i^z$
be the streamwise and spanwise components of the wall-tangent vector. We begin with the contribution acting in the spanwise direction
${t_i^z}$
. According to figure 3, the contribution should be spread over the buffer layer and have a weak dependence on
$\textit{Re}_{\tau }$
. We choose to make this latter be carried by the damping function
$ {f_w\!}^{(1)}$
, and to use the
$\textit{Re}_{\tau }$
-independent function
$\displaystyle {{\tilde {\varepsilon }}^{(2)}{\overline {{w^{\prime }}^2}}^{(2)}\!}/{{k}^{(2)}}$
to spatially distribute the contribution. Concerning the streamwise contribution, we make use of the function
$ (1-{f}^{(2)}_w )( {{\varepsilon }^{(2)}_w{\overline {{u^{\prime }}^2}}^{(2)}\!}/{{k}^{(2)}})$
to locate the action around
$y^+=6.5$
. This choice was made in conjunction with that for the expression of
${\varepsilon }^{(2)}_w$
in (3.1). The dependence to
$\textit{Re}_{\tau }^{3/5}$
is hardly reproducible from any of the transported variables. Instead, we introduce a function of
$\textit{Re}_{\theta }$
, the Reynolds number based on the momentum thickness
$\theta$
. The
$\displaystyle {M}^{(1)}_{\textit{ij}}$
term finally reads
\begin{equation} \displaystyle {M}^{(1)}_{\textit{ij}}= C_m^x \tanh \left (\frac {\textit{Re}_{\theta }}{\textit{Re}_{\theta _0}}\right ) \big (1-{f}^{(2)}_w\big )\frac {{\varepsilon }^{(2)}_w{\overline {{u^{\prime }}^2}}^{(2)}}{{k}^{(2)}} t_i^xt_j^x + C_m^z {f}^{(1)}_w \frac {{\tilde {\varepsilon }}^{(2)}{\overline {{w^{\prime }}^2}}^{(2)}}{{k}^{(2)}} t_i^zt_j^z, \end{equation}
with
$\textit{Re}_{\theta _0}=16\,000$
. It must be noticed that the
$\textit{Re}_{\theta }$
-dependence of the streamwise contribution in (3.5) suggests a finite value for the near-wall peak intensity
$\overline {{u^{\prime }}^2}$
. However, as the contribution on
$\overline {{w^{\prime }}^2}$
is not bounded with
$\textit{Re}_{\theta }$
since it depends on
${l_t}^{(1)}$
through
${f_w}^{(1)}$
, the near-wall peak of
$\overline {{u^{\prime }}^2}$
may still tend to infinity for large Reynolds number values because of the action of redistribution. As a reminder, the purpose of the work is not to discuss the theoretical behaviour of the near-wall peak intensity, but to focus on the practical modelling of the
$\textit{Re}_{\tau }$
-dependence of the Reynolds tensor components, which will remain in this context limited to a restricted range of the Reynolds numbers. The introduction of the integral parameter
$\textit{Re}_{\theta }$
is not suited to most actual computational fluid dynamics solvers – although some modern solvers are able to cope with integral parameters – since it makes the expression in (3.5) non-local, but the two-scale RSM is not intended to become an industrial RANS model due to its inherent complexity. Above all, this should be considered as proof of concept that will enable the development of industrialisable models at a later stage, particularly with a reduced number of transported variables of the first-order RANS model type.
To further enrich the original two-scale RSM, a Reynolds-number dependence was also applied to two other terms. First, the amplitude of
${\tilde {\varepsilon }}^{(1)}$
in the outer region was shown (Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024) to be overestimated for low
$\textit{Re}_{\tau }$
values, leading to overpredicted Reynolds stress components of the LS contribution. Therefore, constant
${C_{\varepsilon _3}\!\!\!}^{(2)}$
was changed for
$\displaystyle {{C_{\varepsilon _3}\!\!\!}^{(2)}}^1 + {{C_{\varepsilon _3}\!\!\!}^{(2)}}^2\tanh ( {\textit{Re}_{\theta }}/{\textit{Re}_{\theta _0}} )$
such that
$\displaystyle {{C_{\varepsilon _3}\!\!\!}^{(2)}}^1 + {{C_{\varepsilon _3}\!\!\!}^{(2)}}^2={C_{\varepsilon _3}\!\!\!}^{(2)}$
and
${{C_{\varepsilon _3}\!\!\!}^{(2)}}^1 \approx 5{{C_{\varepsilon _3}\!\!\!}^{(2)}}^2$
. The second modification concerns the damping function
${f}^{(1)}_w$
, and more specifically
${\alpha }^{(1)}_w$
. At low to moderate Reynolds numbers,
${\alpha }^{(1)}$
could not reach its maximal value
$1$
at
$y^+=\textit{Re}_{\tau }$
, which also contributed to degraded predictions of the LS contribution at low Reynolds numbers. Instead of applying a Reynolds dependence to
${l}^{(1)}_t$
via the coefficient
${C_L\!}^{(1)}$
, we preferred to rearrange the elliptic equation in the form
${\alpha }^{(1)}- {{l}^{(1)}_t}^2{\nabla} ^2 {\alpha }^{(1)} = \min (0.5+0.5( {\textit{Re}_{\theta }}/{2\textit{Re}_{\theta _0}}),1 )$
and to normalise
${\alpha }^{(1)}$
by the factor
$\displaystyle \min (0.5+0.5( {\textit{Re}_{\theta }}/{2\textit{Re}_{\theta _0}}),1 )$
so that it remains bounded between
$0$
and
$1$
. These two modifications have virtually no effect, or at least very limited effect, on results for
$\textit{Re}_{\theta }\gt 16\,000$
, and only serve to accommodate some terms at low Reynolds number conditions, smoothing the transition towards the solution, independent of
$\textit{Re}_{\tau }$
, as given by any standard turbulence models having a single scale.
Model constants are recapped in Appendix A. Minor modifications to constants were made to adapt to the introduction of the new formulation of term
${M}^{(1)}_{\textit{ij}}$
in (3.5), and to the use of
$\textit{Re}_{\theta }$
-dependent functions.
4. High Reynolds number boundary layer applications
To go beyond the validation previously carried out on turbulent channel flows (Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024), which remains satisfied with modifications presented in § 3, we are turning to zero pressure gradient (ZPG) turbulent boundary layer configurations. The two-scale RSM was implemented in a RANS boundary layer code, developed at ONERA for several decades and called CLICET. The parabolic system of equations is solved at each station and marched in the streamwise direction. The code uses a second-order finite volume scheme together with adaptive grids to ensure convergence.
In order to challenge the model and reach very high Reynolds numbers, we are using data from the University of Melbourne and, notably, the measurements undertaken by Samie et al. (Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018). But before considering these data, for which only streamwise mean and fluctuating velocity are available, DNS data provided by Sillero, Jiménez & Moser (Reference Sillero, Jiménez and Moser2013) at
$\textit{Re}_{\tau }\approx 2000$
(
$\textit{Re}_{\theta }\approx 6500$
) are thoroughly examined. One-dimensional pre-multiplied spectra
$\varPhi ^x_{\textit{ij}}$
are available in this dataset, but for a limited number of locations in the boundary layer thickness. Each spectrum is split into two parts with a cut-off wavelength
$\lambda _z^+=1000$
, allowing an assessment of the SS and LS contributions, although it is not similar to the decomposition proposed by Lee & Moser (Reference Lee and Moser2019), and therefore not strictly coherent with respect to the model. Additionally, the power law used to discretise the spectra only leaves
$15$
points above
$\lambda _z^+=1000$
over the
$1364$
available, introducing some uncertainty on the assessment of the LS contribution. Figure 5 shows the mean velocity profile and the three profiles of the diagonal stresses obtained with the two-scale RSM and compared to the DNS data. The agreement is satisfying given that the Reynolds number
$\textit{Re}_{\tau }$
is still limited. The results support the choices made to include Reynolds number dependent functions to accommodate the model to low to moderate Reynolds number configurations. When plotting the SS and LS contributions (see figure 6), the agreement is all the more remarkable, with an energy splitting well distributed in the boundary layer thickness for all Reynolds stress components. A too abrupt separation between inner and outer contributions on the large scale can nevertheless be pointed out on
$\overline {{u^{\prime }}^2}$
and
$\overline {{w^{\prime }}^2}$
, without being able to say whether deviations are coming from the model or are due to the use of one-dimensional spectra. In any case, this has no major consequences on the complete profiles. Figure 6 also permits us to assess the weak impact of the spectra discretisation when comparing Reynolds stress profiles obtained from the integration over the whole spectra and those directly computed by Sillero et al. (Reference Sillero, Jiménez and Moser2013), proving the overall good estimation of the LS contributions, even though some errors may be observed in the outer region, particularly on
$\overline {{w^{\prime }}^2}$
.

Figure 5. Profiles of the velocity (black) and the streamwise (red), wall-normal (blue) and spanwise (green) Reynolds stress components for a ZPG boundary layer at
$\textit{Re}_{\tau }\approx 2000$
. Symbols are DNS data from Sillero et al. (Reference Sillero, Jiménez and Moser2013), and solid lines are results of the two-scale RSM.

Figure 6. The SS (triangle symbols) and LS (diamond symbols) contributions of the diagonal components of the Reynolds stress tensor for a ZPG boundary layer at
$\textit{Re}_{\tau } \approx 2000$
obtained from partitioned one-dimensional spectra of Sillero et al. (Reference Sillero, Jiménez and Moser2013). Circle symbols correspond to diagonal components computed over the whole spectra. Dashed black lines are DNS results from Sillero et al. (Reference Sillero, Jiménez and Moser2013) as in figure 5. Solid lines are SS and LS contributions in the results of the two-scale model.
Experiments conducted by Samie et al. (Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018) are simulated using the two-scale RSM to assess the performances of the model at high Reynolds number. These measurements, acquired using the nanoscale thermal anemometry probes (NSTAPs) manufactured at Princeton University (Vallikivi & Smits Reference Vallikivi and Smits2014), were obtained at four stations corresponding to dimensionless boundary layer thickness
$\delta ^+=\textit{Re}_{\tau }\in \{6000,10\,000,14\,500,20\,000\}$
(
$\textit{Re}_{\theta }\in [14\,000,55\,000]$
). Over this range of Reynolds numbers, high Reynolds number effects are clearly identifiable on the fluctuating velocity profiles. The inner and outer contributions of the large scale are indeed distinctly spatially separated at such high
$\textit{Re}_{\tau }$
. In figure 7, the mean streamwise velocity profiles are plotted for the four considered cases.

Figure 7. Experimental (solid lines) mean velocity profiles in a ZPG boundary layer (Samie et al. Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018) at
$\textit{Re}_{\tau }=6000,10\,000,14\,500,20\,000$
. Darker purple lines indicate increasing
$\textit{Re}_{\tau }$
values. Corresponding purple dashed lines are results from the two-scale RSM. Mean velocity profiles are shifted up by
$2$
units for each increasing value of
$\textit{Re}_{\tau }$
.

Figure 8. Streamwise Reynolds stress component profiles (solid lines) from (Samie et al. Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018) at
$\textit{Re}_{\tau }=6000,10\,000,14\,500,20\,000$
. Darker purple lines indicate increasing
$\textit{Re}_{\tau }$
values. Corresponding (a) green and (b) purple dashed lines are results from the previous version of the two-scale RSM and the present one, respectively. Red dashed lines are results obtained with the EBRSM (Manceau Reference Manceau2015).
Mean profiles are well reproduced by the two-scale RSM although some differences are visible in the wake regions. The agreement reflects the good predictions of the SS and LS contributions to
$\overline {u'v'}$
for all Reynolds numbers. In figure 8, fluctuating streamwise profiles are drawn for the previous version of the two-scale RSM (Chedevergne et al. Reference Chedevergne, Coroama, Gleize and Bézard2024) and for the one described above in order to highlight the improvements made by the new closure of the interscale energy fluxes. The most remarkable achievement of the present model is the agreement across all the
$\overline {{u^{\prime }}^2}$
profiles. The first, trivial, observation that can be made from these results is that two-scale modelling enables behaviours in the outer region to be recovered, which is not possible with a single-scale RANS model. To exemplify this problem, results obtained with the elliptic blending RSM (EBRSM) of Manceau (Reference Manceau2015), which is a reference RANS model for calculating Reynolds stress profiles of wall-bounded flows, are plotted in figure 8. As the EBRSM uses only one length scale to characterise energy transfer, a single profile for
$\overline {{u^{\prime }}^2}$
is obtained up to
$y^+=1000$
, irrespective of the Reynolds number, which prevents the effects of high Reynolds numbers from being restored. Second, the
$\textit{Re}_{\tau }$
-dependence at the
$\overline {{u^{\prime }}^2}$
peak is fairly well captured by the new version of the two-scale RSM. However, there are a slight shift in the position of the peak, and some differences in amplitude. The dynamics with respect to the Reynolds number is nevertheless fairly faithful to the experimental data. Finally, a particularly noteworthy feature of the results obtained is the contribution of
$\overline {{w^{\prime }}^2}$
in (3.5) on
$\overline {{u^{\prime }}^2}$
profiles in the region covering values of
$y^+$
ranging from
$80$
to a few hundreds. Figure 8 clearly shows the improvements achieved by modifying the
${M}^{(1)}_{\textit{ij}}$
term (3.5) compared to the previous version of the two-scale RSM. The
$\overline {{u^{\prime }}^2}$
profiles are strongly affected by the increase in Reynolds number in this region of the boundary layer. This effect on
$\overline {{u^{\prime }}^2}$
is due to the redistribution term, since it is in fact the contribution on
$\overline {{w^{\prime }}^2}$
that acts in this region. The modelling strategy, inspired by the analysis presented in § 2, therefore proves to be entirely relevant. Most interestingly, at these Reynolds numbers, the dependence functions introduced in the modelling (see § 3) have practically no effect. Thus the Reynolds evolution obtained is the result of the balance between the different terms of the
$\overline {{u^{\prime }}^2}$
budget – more precisely, the budget of
${\overline {{u^{\prime }}^2}}^{(1)}$
– involving in particular interscale energy transfers within the inner and outer regions, and some pressure redistribution in the logarithmic zone due to the evolution of the Reynolds stress component
$\overline {{w^{\prime }}^2}$
at large scales.
5. Conclusion
High Reynolds number effects in wall-bounded flows involve interscale energy transfers between the inner and outer regions, and despite all data acquired over the years, a comprehensive description of these energy transfers is not fully achieved. Debates are also still very active about the expected asymptotic behaviours of the near-wall peak intensity (Jiménez Reference Jiménez2024). From a modelling perspective, the Reynolds number dependence of the wall-parallel fluctuation intensities in the inner region is a challenging phenomenon since it undermines the standard formalism used in turbulence models to describe the near-wall cycle, with a solution independent of the friction Reynolds number
$\textit{Re}_{\tau }$
. To account for this dependency it is thus necessary to consider at least two transported turbulent length scales, characteristics of small-scale (SS) and large-scale (LS) contributions. This formalism was used succesfully by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024) to develop a two-scale RSM that reproduced DNS results of Lee & Moser (Reference Lee and Moser2019) obtained in turbulent channel flows, who considered a scale separation at a fixed wavelength. However, the LS contributions in the inner region was modelled with an ad hoc expression that did not reflect some flow physics features observed in the DNS. An extensive analysis of energy spectra, along with observations made on the energy flux between large and small scales in the near-wall region, emphasised the role played by the spanwise component of the Reynolds stress tensor, namely
$\overline {{w^{\prime }}^2}$
, in the interscale energy transfers. The term accounting for the energy flux from small to large scale in the two-scale RSM was then reformulated, incorporating these observations in the derivation of a consistent expression. The results obtained are in excellent agreement with DNS data in a ZPG boundary layer configuration at a moderate Reynolds number (Sillero et al. Reference Sillero, Jiménez and Moser2013) and especially with Samie et al. (Reference Samie, Marusic, Hutchins, Fu, Fan, Hultmark and Smits2018) data at high Reynolds numbers. These results both demonstrate the effectiveness of the model adopted and above all confirm the importance of the role of
$\overline {{w^{\prime }}^2}$
in interscale transfers. This echoes the outcomes of Zhou, Xu & Jiménez (Reference Zhou, Xu and Jiménez2022), who partly confirmed the inner–outer model of Toh & Itano (Reference Toh and Itano2005) by clearly showing that near-wall streaks drift in the spanwise direction under the influence of the outer LS motions. This preferential movement in the spanwise direction suggests that
$\overline {{w^{\prime }}^2}$
may play a special role in the footprint left by large scales on smaller ones. Additional analysis of DNS data involving multi-scale decomposition are required to further dig in this direction. Causality analysis (Andreolli et al. Reference Andreolli, Gatti, Vinuesa, Örlü and Schlatter2023) may also bring a different insight on this issue. Numerical experiments where modulation was suppressed suggest that the superposition phenomenon may not be responsible for the LS contributions in the near wall, but that interaction of SS modes may transfer a significant amount of energy to the large scale. In the two-scale RSM too, the LS contribution in the inner region is modelled via energy transfer from the small scale, in accordance with results provided by the scale separation applied to the DNS data of Lee & Moser (Reference Lee and Moser2019), although no mechanism has yet been identified to explain these observations.
Acknowledgements
The authors are very grateful to M. Lee and R. Moser for sharing their DNS data on turbulent channel flow (Lee & Moser Reference Lee and Moser2019).
Funding
This work has been funded within the frame of the Clean Aviation Joint Undertaking, Multi-MW Hybrid-Electric Propulsion System for Regional Aircraft, being part of the Horizon Europe research and innovation funding programme of the European Commission. A CC-BY public copyright licence has been applied by the authors to the present document and will be applied to all subsequent versions up to the Author Accepted Manuscript arising from this submission, in accordance with the grant’s open access conditions.
Declaration of interests
The authors report no conflict of interest.
Appendix A. Two-scale RSM constants
In (3.2), the following relations are employed:
\begin{equation} \begin{array}{l} \displaystyle {C'_{\varepsilon _1}\!}^{(1)} = \displaystyle {C_{\varepsilon _1}\!\!\!}^{(1)}\left (1+{A_1\!}^{(1)}\left (1-{{f_w\!}^{(1)}}\right )\frac {{P_k\!}^{(2)}}{{\varepsilon }^{(2)}}\right )\!, \\[18pt] \displaystyle {C'_{\varepsilon _1}\!}^{(2)} = \displaystyle {C_{\varepsilon _1}\!\!\!}^{(2)}\left (1-{f_w}^{(2)}+0.8{{f_w}^{(1)}}+{A_1}^{(2)}\big (1-{{f_w\!}^{(2)}}\big )\frac {{P_k\!}^{(2)}}{{\varepsilon }^{(2)}}\right )\!, \\[18pt] \displaystyle {C'_{\varepsilon _2}\!}^{(2)} = \displaystyle {C_{\varepsilon _2}\!\!\!}^{(2)}\left (1-{A_2}^{(2)}{{f_w}^{(2)}}\frac {{P_k\!}^{(2)}}{{\varepsilon }^{(2)}}\right )\!. \end{array} \end{equation}
The constants of the two-scale RSM are
\begin{equation} \begin{array}{c} \displaystyle {C_{\varepsilon _1}\!\!}^{(1)}=1.6,\quad {C_{\varepsilon _1}\!\!}^{(2)}=1.45,\\[2pt] \displaystyle {C_{\varepsilon _2}\!\!}^{(1)}=1.95,\quad {C_{\varepsilon _2}\!}^{(2)}=1.7,\\[2pt] \displaystyle {{C_{\varepsilon _3}\!}^{(2)}}^1=1.44,\quad {{C_{\varepsilon _3}\!}^{(2)}}^2=0.27,\\[2pt] \displaystyle {A_1\!}^{(1)}=1.3,\quad {A_1\!}^{(2)}=0.22,\quad {A_2\!}^{(2)}=0.5,\\[2pt] \displaystyle c_s=0.21,\\[2pt] \displaystyle {\sigma _k\!}^{(1)}=0.9,\quad {\sigma _k\!}^{(2)}=1.0,\\[2pt] \displaystyle {\sigma _{\varepsilon }\!}^{(1)}=1.1,\quad {\sigma _{\varepsilon }\!}^{(2)}=1.0,\\[2pt] \displaystyle {C_T\!}^{(1)}=6.0,\quad {C_T\!}^{(2)}=6.0,\\[2pt] \displaystyle {C_L\!}^{(1)}=0.275,\quad {C_L\!}^{(2)}=0.112,\\[2pt] \displaystyle {C_\eta }^{(1)}=150,\quad {C_\eta }^{(2)}=78,\\[2pt] \displaystyle C_m^x = 15.5,\quad C_m^z=0.1. \end{array} \end{equation}
Details concerning the generation of the two-scale RSM can be found in the original article by Chedevergne et al. (Reference Chedevergne, Coroama, Gleize and Bézard2024).



































