Hostname: page-component-848d4c4894-xfwgj Total loading time: 0 Render date: 2024-06-20T04:19:02.397Z Has data issue: false hasContentIssue false

Universal relative scaling of longitudinal structure functions in shear-dominated turbulence

Published online by Cambridge University Press:  01 April 2024

K.R. Maryada*
Affiliation:
Department of Mechanical and Mechatronics Engineering, The University of Auckland, Auckland 1010, New Zealand
S.W. Armfield
Affiliation:
School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, New South Wales 2006, Australia
M. MacDonald
Affiliation:
Department of Mechanical and Mechatronics Engineering, The University of Auckland, Auckland 1010, New Zealand
P. Dhopade
Affiliation:
Department of Mechanical and Mechatronics Engineering, The University of Auckland, Auckland 1010, New Zealand
S.E. Norris
Affiliation:
Department of Mechanical and Mechatronics Engineering, The University of Auckland, Auckland 1010, New Zealand
*
Email address for correspondence: kmar699@aucklanduni.ac.nz

Abstract

Shear significantly influences turbulence in the energy-containing range of shear-dominated flows, and the longitudinal structure functions do not have a universal form as they do in homogeneous isotropic turbulence. Despite this, the relative scaling of structure functions exhibits universal sub-Gaussian behaviour in shear-dominated flows, in particular for turbulent boundary layers, channels and Taylor–Couette flows. Our investigation of a turbulent vertical buoyancy layer at $Pr = 0.71$ using direct numerical simulation shows this universality even in moderate-Reynolds-number buoyancy-driven but shear-dominated boundary layers. It is demonstrated that the universality is related to the energy density of the eddies, which attains a hierarchical equilibrium in the energy-containing range of shear-dominated turbulence. We conjecture that the universal sub-Gaussian behaviour of the energy density of the energy-containing range, which was considered to be non-trivial in prior studies, is related to the universal anomalous scaling exponents of the inertial subrange turbulence. Based on this conjecture, we propose a hypothesis that relates large-scale eddies and the intermittent dissipation field in shear-dominated turbulence, highlighting a relationship between large and small scales. A phenomenological model is also developed to predict the scaling, which is verified using data from a turbulent boundary layer, half-channel and vertical buoyancy layer at friction Reynolds numbers spanning four orders of magnitude. Excellent agreement is observed.

Type
JFM Papers
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press.

1. Introduction

The scaling of turbulent flows has intrigued researchers for over a century, as understanding the scaling would pave the way for a unified theory of turbulence. One of the most influential works in this regard is the theory of Kolmogorov (Reference Kolmogorov1941) on homogeneous isotropic turbulence. In large-Reynolds-number ($Re$) flows, at scales where viscosity and integral length scales are insignificant (the inertial subrange), eddies with length scale $l$ only depend on the turbulent dissipation rate $\epsilon$. These eddies are significantly larger than the Kolmogorov microscale; therefore, the dissipation effects are insignificant. According to Kolmogorov (Reference Kolmogorov1941), at these scales, in homogeneous isotropic turbulence, the structure functions (a proxy for turbulent eddies) of different orders scale as

(1.1)\begin{equation} S_n = \langle |u(x+r) - u(x)|^n \rangle = C_n (\epsilon r)^{n/3}, \end{equation}

where $u$ is the velocity field subtracted from the mean, $x$ is the longitudinal coordinate, $r$ is the separation distance and is representative of length scale $l$, $C_n$ is a constant, which is not necessarily universal for different flows (Sreenivasan & Antonia Reference Sreenivasan and Antonia1997), and $n$ is the order of the structure function. All the relevant quantities are ensemble-averaged unless specified. Throughout this paper, we use $x, y$ and $z$ to represent the streamwise, spanwise and wall-normal directions, respectively.

In homogeneous isotropic turbulence, it has since been known that $S_n$ significantly deviates from (1.1), especially for $n > 3$, and this has been referred to as intermittency. This is due to the nature of $\epsilon$, which is not uniform but has intermittent regions of $\epsilon$ significantly different from the mean. Essentially, $S_n \propto r^{\zeta _n}$ instead of $S_n \propto r^{n/3}$ as in (1.1). Here, $\zeta _n$ is often called the anomalous scaling exponent. Consequently, several phenomenological models have been proposed to predict $\zeta _n$ (Kolmogorov Reference Kolmogorov1962; Frisch, Sulem & Nelkin Reference Frisch, Sulem and Nelkin1978; Meneveau & Sreenivasan Reference Meneveau and Sreenivasan1987; She & Leveque Reference She and Leveque1994). A major assumption in all these models is that turbulence in the inertial subrange can be modelled as a random multiplicative process.

Even in real-world shear-driven flows having significant large-scale anisotropy, the inertial subrange scaling ($S_n \propto r^{\zeta _n}$) is recovered at small scales, i.e. at scales smaller than the Corrsin length scale (Corrsin Reference Corrsin1958) but larger than the Kolmogorov microscale (Arneodo et al. Reference Arneodo1996; Toschi et al. Reference Toschi, Amati, Succi, Benzi and Piva1999; Toschi, Leveque & Ruiz-Chavarria Reference Toschi, Leveque and Ruiz-Chavarria2000; Gualtieri et al. Reference Gualtieri, Casciola, Benzi, Amati and Piva2002; Casciola et al. Reference Casciola, Gualtieri, Jacob and Piva2005; Attili & Bisetti Reference Attili and Bisetti2012). This implies that the small scales (inertial subrange) in shear-driven flows scale similarly to homogeneous isotropic turbulence despite large-scale anisotropy. This result was also verified by de Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015) using high-$Re$ datasets of atmospheric and turbulent boundary layers.

However, the scaling of large-scale eddies is less certain. The exponents $\zeta _n$ of the inertial subrange are not expected to hold at scales where the outer length scales of a system are comparable to $r$. Even external forces such as shear, buoyancy and stratification are expected to alter the scaling properties of the integral-scale motions. In most engineering flows, shear is almost always present and cannot be ignored. The variation of $\zeta _n$ is significant when shear is present, and several corrections have been proposed to predict the scaling of the structure functions under the assumption that a power-law scaling holds (Benzi et al. Reference Benzi, Amati, Casciola, Toschi and Piva1999; Toschi et al. Reference Toschi, Amati, Succi, Benzi and Piva1999, Reference Toschi, Leveque and Ruiz-Chavarria2000; Casciola et al. Reference Casciola, Benzi, Gualtieri, Jacob and Piva2001, Reference Casciola, Gualtieri, Benzi and Piva2003, Reference Casciola, Gualtieri, Jacob and Piva2005; Gualtieri et al. Reference Gualtieri, Casciola, Benzi, Amati and Piva2002). However, the power-law scaling of large-scale eddies is not unequivocal. Recent studies of turbulent boundary layers, channels and pipes at high $Re$ have shown that the power-law scaling of structure functions is not always observed at large scales (Davidson, Krogstad & Nickels Reference Davidson, Krogstad and Nickels2006a; Davidson, Nickels & Krogstad Reference Davidson, Nickels and Krogstad2006b; Davidson & Krogstad Reference Davidson and Krogstad2009, Reference Davidson and Krogstad2014; Chung et al. Reference Chung, Marusic, Monty, Vallikivi and Smits2015b; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017; Pan & Chamecki Reference Pan and Chamecki2016; Yang, Marusic & Meneveau Reference Yang, Marusic and Meneveau2016; Agostini & Leschziner Reference Agostini and Leschziner2017; Chamecki et al. Reference Chamecki, Dias, Salesky and Pan2017; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017; Ghannam et al. Reference Ghannam, Katul, Bou-Zeid, Gerken and Chamecki2018; Xie et al. Reference Xie, de Silva, Baidya, Yang and Hu2021).

A different approach is to assume that at large scales, as the correlation between the eddies is weak, turbulence can be approximated as an additive process rather than multiplicative as assumed at small scales (Townsend Reference Townsend1976; Perry, Henbest & Chong Reference Perry, Henbest and Chong1986; Jiménez Reference Jiménez1998; Mouri et al. Reference Mouri, Takaoka, Hori and Kawashima2006; Mouri, Hori & Takaoka Reference Mouri, Hori and Takaoka2009). This assumption regarding turbulence at energy-containing scales, especially for turbulent boundary layers and channels (Davidson et al. Reference Davidson, Nickels and Krogstad2006b; Yang et al. Reference Yang, Marusic and Meneveau2016; Agostini & Leschziner Reference Agostini and Leschziner2017; Marusic & Monty Reference Marusic and Monty2019), has received renewed interest in the past few years. Under this assumption, the turbulence at scale $r$ is statistically modelled as increments of eddies (addends) that are identically and independently distributed. Conceptually, this is similar to the model of Tennekes & Lumley (Reference Tennekes and Lumley1972) and Jiménez (Reference Jiménez1998), which the authors formalised in the spectral domain. This assumption regarding turbulence at large scales can be used to explain scaling laws in the energy-containing range that are different from power-law scaling (Davidson et al. Reference Davidson, Nickels and Krogstad2006b; Chung et al. Reference Chung, Marusic, Monty, Vallikivi and Smits2015b; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015; Yang et al. Reference Yang, Marusic and Meneveau2016; Agostini & Leschziner Reference Agostini and Leschziner2017; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017; Marusic & Monty Reference Marusic and Monty2019).

As the large scales can be approximated as a random additive process, turbulence at the energy-containing scales is expected to be Gaussian under the central limit theorem. However, it is well known that the energy-containing eddies in shear-dominated turbulence exhibit sub-Gaussian scaling behaviour (Jiménez Reference Jiménez1998; Meneveau & Marusic Reference Meneveau and Marusic2013; Stevens, Wilczek & Meneveau Reference Stevens, Wilczek and Meneveau2014; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017). Recently, based on experimental and numerical data, it has been demonstrated that the even-order structure functions in turbulent boundary layers, channels and Taylor–Couette flows exhibit non-trivial but universal sub-Gaussian scaling behaviour (de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017). However, it is not clear why this particular form of scaling is observed. The physical basis for the sub-Gaussian scaling is not understood, and it is not apparent how intermittency affects the scaling (Davidson & Krogstad Reference Davidson and Krogstad2014; de Silva et al. Reference de Silva, Krug, Lohse and Marusic2017).

It is evident from the above discussion that our understanding of the scaling of large-scale eddies in shear-dominated turbulence, especially structure functions of different orders, is far from complete, which forms the primary motivation of the current study. Here, we demonstrate that the energy ratio of the large-scale eddies in shear-dominated turbulence is universal, which is different from the universality observed in the scaling exponents of the inertial subrange. We then show that the universal scaling of the energy ratio of the energy-containing range is related to the scaling of the intermittent dissipation field in shear-dominated flows, which has been hitherto undetected and is one of the novel aspects of the current study. Notably, we establish that the sub-Gaussian scaling of the structure functions, which has been thought to be non-trivial in previous studies (Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017; de Silva et al. Reference de Silva, Krug, Lohse and Marusic2017), is related to the anomalous scaling of inertial subrange turbulence, providing a crucial link between large-scale eddies and the highly intermittent turbulent dissipation. Further, we propose a phenomenological model that sheds light on large-scale/small-scale interactions in shear-dominated turbulence while accurately predicting the energy ratio for low-order and high-order structure functions. The current results can form the basis for a theoretical framework that unites small and large scales in shear-dominated turbulence.

The paper is organised as follows. The hypothesis and the phenomenological model are described in § 2. An overview of the different numerical and experimental datasets used in the current study is given in § 3. To ensure that the hypothesis is not limited by low-Reynolds-number behaviour, we have investigated numerical and experimental datasets of three different shear-dominated flows across friction Reynolds numbers ($Re_\tau = u_\tau \delta / \nu$, where $u_\tau$ is the friction velocity, $\delta$ is the outer length scale and $\nu$ is the kinematic viscosity) spanning four orders of magnitude. The proposed phenomenological model is validated in § 4 by investigating low-order and high-order structure functions, which account for the intense and rare fluctuations and those closer to the mean. Excellent agreement is observed between the data and the proposed model, with the percentage deviation being less than $10\,\%$. The significance of the current theory with respect to canonical wall turbulence is also highlighted in this section. The conclusions are given in § 5.

2. Hierarchical energy equilibrium of the eddies

2.1. Preliminaries

We begin our analysis by considering eddies having scale $L_I < r < L_{R}$, where $L_I$ ($L_I = \int ^\infty _0{R_{uu}\, {{\rm d} x}}$, with $R_{uu}$ being the two-point correlation in the streamwise direction) is the integral length scale and $L_{R}$ is the location at which $R_{uu} \to 0$. While $L_{R}$ imposes a strict upper limit on the range of statistically significant scales, it cannot be said that $L_I$ is a strict lower bound. We assume that $L_I$ imposes an upper bound on the multiplicative process, beyond which the additive process dominates (Mouri et al. Reference Mouri, Takaoka, Hori and Kawashima2006, Reference Mouri, Hori and Takaoka2009). Nevertheless, the exact values are not of concern as $L_{I}$ and $L_{R}$ are only characteristic scales. We use the terms energy-containing range and large scales to refer to this scaling range.

For $L_I < r < L_{R}$, we hypothesise that longitudinal structure functions of streamwise velocity fluctuations $S_{2p}$ scale according to the ansatz

(2.1)\begin{equation} S_{2p}^{1/p} = K_{2p} \chi f(r) + M_{2p}. \end{equation}

Here, $p = n/2$ is the order, $S_{2p}^{1/p}$ is the normalised structure function, $\chi$ is a quantity of interest representing the eddies (elaborated below), $K_{2p}$ and $M_{2p}$ are constants (not universal for all turbulent flows) and $r$ now corresponds to the non-dimensional separation distance. The $1/p$ power is chosen to ensure that $S_{2p}^{1/p}$ would always have the units of $S_2$, i.e. energy per unit mass. Equation (2.1) differs from (1.1) in that a power-law scaling is not explicitly assumed as $f(r)$ is an arbitrary function, which can vary for different flows. The exact functional form of $f(r)$ is unnecessary for the proposed scaling and is not predicted.

Equation (2.1) ensures the relation $S_2 = 2 \langle u u \rangle (1 - R_{uu})$ is satisfied if $f(r) = 1 - R_{uu}$, $\chi = \langle u u \rangle$, $K_{2} = 2$ and $M_{2} = 0$ (Davidson et al. Reference Davidson, Nickels and Krogstad2006b). The energy dissipation and streamwise velocity variance are related as $\langle u u \rangle \propto (\epsilon L_I)^{2/3}$ (Frisch Reference Frisch1995). In shear-dominated turbulence, $\langle u u \rangle \propto \langle uw \rangle$, where $\langle uw \rangle$ is the Reynolds shear stress. In the logarithmic region of turbulent boundary layers, $\langle uw \rangle \sim u_\tau ^2$ (Townsend Reference Townsend1976; Perry et al. Reference Perry, Henbest and Chong1986; Marusic & Monty Reference Marusic and Monty2019). Equation (2.1) reduces to the following $\ln (r)$ law in turbulent boundary layers if $f(r) = \ln (r)$ and $\chi = u_\tau ^2$ (Davidson et al. Reference Davidson, Nickels and Krogstad2006b):

(2.2)\begin{equation} S_{2p}^{1/p} = A_{p} u_\tau^2 \ln(\hat{r}) + B_{p}, \end{equation}

where $A_{p}$ and $B_p$ are constants and $\hat {r}$, in this case, is normalised using the distance-from-the-wall scaling or the dissipation length scale (Townsend Reference Townsend1976; Perry et al. Reference Perry, Henbest and Chong1986; Davidson & Krogstad Reference Davidson and Krogstad2014; Pan & Chamecki Reference Pan and Chamecki2016; Chamecki et al. Reference Chamecki, Dias, Salesky and Pan2017; Marusic & Monty Reference Marusic and Monty2019).

Therefore, $\chi$ in (2.1) can be replaced with any correlated quantity. Hence, for the rest of this paper, we focus on $\chi$ instead of making explicit references to $\epsilon$, $u_\tau$, $\langle uu \rangle$ and $\langle uw \rangle$.

We restrict ourselves to positive structure functions in this study. Also, we only work with absolute moments of structure functions as most phenomenological models for inertial subrange turbulence revolve around absolute moments (Kolmogorov Reference Kolmogorov1962; Meneveau & Sreenivasan Reference Meneveau and Sreenivasan1987; She & Leveque Reference She and Leveque1994). This also implies that the odd-order moments do not vanish even if the velocity fluctuations in the energy-containing range are assumed to be symmetric.

Along with structure functions, the ‘energy-density’ at scale $r$, which indicates the energy of eddies having size $r$, is another representative quantity of the eddies (Townsend Reference Townsend1976; Davidson et al. Reference Davidson, Nickels and Krogstad2006b). It is defined as

(2.3)\begin{equation} E^r_{p} = r \frac{\partial S_{2p}^{1/p} }{\partial r}. \end{equation}

This relies on the observation that the ensembled averaged $S_{2p}^{1/p}$ is differentiable even if the eddies are modelled as discrete objects.

2.2. Phenomenological model

Turbulence, especially in the inertial subrange, is often conceptualised as eddies of different sizes/intensities organised hierarchically (Kolmogorov Reference Kolmogorov1941, Reference Kolmogorov1962; Meneveau & Sreenivasan Reference Meneveau and Sreenivasan1987; She & Leveque Reference She and Leveque1994; Frisch Reference Frisch1995; Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996; Ching et al. Reference Ching, She, Su and Zou2002; Jiang et al. Reference Jiang, Gong, Liu, Zhou and She2006). Several studies of shear turbulence have postulated and observed hierarchical structures even in the energy-containing range (Townsend Reference Townsend1976; Perry et al. Reference Perry, Henbest and Chong1986; Yang et al. Reference Yang, Marusic and Meneveau2016; Dong et al. Reference Dong, Lozano-Durán, Sekimoto and Jiménez2017; Marusic & Monty Reference Marusic and Monty2019; Motoori & Goto Reference Motoori and Goto2021). This paper also adopts the notion that the energy-containing eddies in shear turbulence exhibit hierarchical properties.

Also, let us assume that a normalised structure function (as shown in (2.1)) of order $p$ indicates eddies having ‘intensity’ $p$ (She & Leveque Reference She and Leveque1994; Frisch Reference Frisch1995). This would imply that high-order structure functions are indicative of rare and intense fluctuations, while low-order structure functions indicate fluctuations closer to the mean. This assumption is reasonable as high-order structure functions are related to the tails of the probability density functions (p.d.f.s) of velocity increments. In contrast, the low-order structure functions are related to the core of the p.d.f.s. Such an assumption was also made in the hierarchical structural model (HSM) of turbulence (She & Leveque Reference She and Leveque1994; She & Waymire Reference She and Waymire1995).

In a HSM, which is typically used to describe inertial subrange, it is assumed that the structure functions of different orders attain a hierarchical equilibrium (She & Leveque Reference She and Leveque1994; She & Waymire Reference She and Waymire1995; Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996; Ching et al. Reference Ching, She, Su and Zou2002; Jiang et al. Reference Jiang, Gong, Liu, Zhou and She2006). In this work, instead of focusing on structure functions themselves, we hypothesise that in shear-dominated turbulence, at the large scales where the additive process approximation holds, the energy of the eddies attains a ‘hierarchical equilibrium’ such that the energy of eddies of size $r$ having intensity $p>0$ is related to the energy of eddies of size $r$ having intensity $m>0$ as

(2.4)\begin{equation} \frac{E^r_p}{E^r_m} = F_{pm}, \end{equation}

where $F_{pm}$ depends on $p$ and $m$. Here, $p$ and $m$ refer to the order of the structure function ($p \neq m$). Here, $F_{pm}$ indicates an underlying process in a statistical sense, and the instantaneous energy ratio need not be equal to $F_{pm}$. It is to be observed that the energy ratio of the eddies attains a hierarchical equilibrium and not the eddies themselves.

For Gaussian statistics, $F_{pm}$ can be written as

(2.5)\begin{equation} F_{pm} = \frac{E^r_p}{E^r_m} = G_{pm} = {\rm \pi}^{(({1/m-1/p})/{2})} \frac{[\varGamma((2p+1)/2)]^{1/p}}{[\varGamma((2m+1)/2)]^{1/m}}, \end{equation}

where $\varGamma$ is the Gamma function (Winkelbauer Reference Winkelbauer2012). For even-order moments, this reduces to

(2.6)\begin{equation} F_{pm} = \frac{E^r_p}{E^r_m} = G_{pm} = \frac{[(2p-1)!!]^{1/p}}{[(2m-1)!!]^{1/m}}, \end{equation}

where $!!$ is the double factorial (Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017).

We contend that $F_{pm}$ is governed by processes that are not strictly Gaussian. This is based on several observations that revealed the sub-Gaussian nature of shear-dominated turbulence (Jiménez Reference Jiménez1998; Meneveau & Marusic Reference Meneveau and Marusic2013; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017). The essence of this argument relies on the assumption that the non-Gaussian scaling of shear-dominated turbulence observed in previous studies is related to the non-Gaussian scaling of the energy ratio (verified in §§ 4.1 and 4.2, and discussed in § 4.4).

To account for this behaviour, we assume that two dominant processes govern the eddies’ energy. One is a Gaussian process, and the other is a rare ‘defect’ event, which modulates turbulence. The phenomenology of defect events is qualitatively similar to the defect events of inertial subrange turbulence, which are responsible for the deviations from Kolmogorov (Reference Kolmogorov1941) scaling. In the energy-containing range, the primary effect of the defect events is to reduce the energy of the fluctuations and is responsible for deviations from Gaussian scaling. This, on phenomenological grounds, is expressed as $F_{pm} = G_{pm} W_{pm}$, with $W_{pm}$ governing the defect events. In this formulation, $W_{pm}$ is a correction term responsible for the non-Gaussian scaling. The rest of this section is dedicated to examining $W_{pm}$, as a model for $W_{pm}$ allows us to quantify the deviation from the Gaussian scaling.

Kolmogorov (Reference Kolmogorov1941) postulated that the small scales are universal and are completely independent of the large scales. However, several studies since then have shown that the large scales of turbulence influence the small scales (e.g. Kolmogorov Reference Kolmogorov1962; Obukhov Reference Obukhov1962; Kraichnan Reference Kraichnan1974, Reference Kraichnan1991; Landau & Lifshitz Reference Landau and Lifshitz1987; Frisch Reference Frisch1995; Yeung, Brasseur & Wang Reference Yeung, Brasseur and Wang1995; Sreenivasan & Antonia Reference Sreenivasan and Antonia1997; Mouri et al. Reference Mouri, Takaoka, Hori and Kawashima2006, Reference Mouri, Hori, Kawashima and Hashimoto2012). In inertial subrange turbulence, the influence of large scales on small scales is observed in the coefficients and not in the scaling exponents (Kraichnan Reference Kraichnan1974; Frisch et al. Reference Frisch, Sulem and Nelkin1978; Frisch Reference Frisch1995; Sreenivasan & Antonia Reference Sreenivasan and Antonia1997), suggesting that despite the scaling exponents of the structure functions being universal, the large-scale structures affect the small scales. The influence of large scales on small scales has also been demonstrated in shear-driven flows (e.g. Mathis, Hutchins & Marusic Reference Mathis, Hutchins and Marusic2009; Ganapathisubramani et al. Reference Ganapathisubramani, Hutchins, Monty, Chung and Marusic2012; Baars, Hutchins & Marusic Reference Baars, Hutchins and Marusic2016; Jacobi et al. Reference Jacobi, Chung, Duvvuri and McKeon2021; Andreolli et al. Reference Andreolli, Gatti, Vinuesa, Örlü and Schlatter2023).

As large scales influence small scales, it is reasonable to assume that the large and small scales share the same signatures. This does not imply that the small scales dictate the properties of large scales but implies that large and small scales are related. In shear-dominated turbulence, we assume this signature is evident in the energy ratio of large-scale eddies. The inertial subrange scaling with intermittency corrections gives us information about small scales. Statistically, we postulate that the defect events in the energy-containing range are related to the defect events in the inertial subrange. Consequently, we include the anomalous scaling of the inertial subrange to model the large scales. Specifically, the distribution of the defect events in small scales is related to the distribution across large scales, implying that $W_{pm}$ is a function of the anomalous scaling exponents of the inertial subrange. Of course, the validity of this crucial assumption requires verification using numerical and experimental data, which we address in § 4.

Note that the inertial subrange scaling is recovered in shear-dominated turbulence at scales larger than the Kolmogorov microscale and smaller than the Corrsin length scale (Toschi et al. Reference Toschi, Amati, Succi, Benzi and Piva1999; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015), suggesting that the above assumption on using the anomalous scaling of inertial subrange for small scales in shear-dominated turbulence is reasonable.

Based on the above arguments, we conjecture that (2.5) takes the following modified form:

(2.7)\begin{equation} F_{pm} = G_{pm} W_{pm} = G_{pm} \frac{\xi_p}{\xi_m} = {\rm \pi}^{(({1/m-1/p})/{2})} \frac{[\varGamma((2p+1)/2)]^{1/p}}{[\varGamma((2m+1)/2)]^{1/m}} \frac{\xi_p}{\xi_m}, \end{equation}

where $\xi _p$ and $\xi _m$ are the anomalous scaling exponents of normalised structure functions of inertial subrange. Here, $\xi _p = \zeta _n/p$ and $\xi _m = \zeta _n/m$.

In (2.7), we use the HSM to calculate the anomalous scaling exponents $\xi _q$ ($q = p$ or $m$) (She & Leveque Reference She and Leveque1994):

(2.8) \begin{equation} \xi_q = \zeta_n/q = \Big(\frac{n}{9} + 2\Big[ 1 - \Big(\frac{2}{3}\Big)^{n/3}\Big]\Big) \Big/ q, \end{equation}

where $\zeta _n$ is the formulation given by She & Leveque (Reference She and Leveque1994). Also, $\xi _q$ in (2.8) does not violate Novikov's inequality (Frisch Reference Frisch1995). This ensures $\xi _p/\xi _m < 1$ for all $p$ and $m$ so long as $p > m$. Therefore, $F_{pm}$ in (2.7) would always exhibit sub-Gaussian behaviour.

Further, (2.8) can be decomposed as

(2.9a)$$\begin{gather} \xi_q = \left.\left(\frac{n}{3} + \tau_{n/3} \right) \right/ q, \end{gather}$$
(2.9b)$$\begin{gather}\tau_n ={-}\frac{2}{3}n + 2\left[1 - \left(\frac{2}{3}\right)^{n} \right], \end{gather}$$

with $n/3$ being the Kolmogorov (Reference Kolmogorov1941) prediction and $\tau _n$ being the intermittency correction (She & Leveque Reference She and Leveque1994; Frisch Reference Frisch1995).

From (2.5), (2.7) and (2.9), we see that Gaussian scaling is recovered if $\tau _n = 0$ (without any intermittency corrections) or if the dissipation is assumed to be uniform, as was done by Kolmogorov (Reference Kolmogorov1941). Note that it is the intermittent dissipation field that leads to anomalous scaling exponents in the inertial subrange (Kolmogorov Reference Kolmogorov1962; Obukhov Reference Obukhov1962; Frisch et al. Reference Frisch, Sulem and Nelkin1978; Meneveau & Sreenivasan Reference Meneveau and Sreenivasan1987; Dubrulle Reference Dubrulle1994; She & Leveque Reference She and Leveque1994; Frisch Reference Frisch1995; She & Waymire Reference She and Waymire1995; Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996; Sreenivasan & Antonia Reference Sreenivasan and Antonia1997). This result is significant as it demonstrates that the departure from the Gaussian profile of the energy ratio of large-scale eddies in shear-dominated flows is directly related to the intermittency observed in inertial subrange turbulence of the same flow. Crucially, we use the anomalous scaling exponents of the inertial subrange to model the energy ratio of the large scales, highlighting a relationship between large and small scales. Note that we only use the characteristic scales (mentioned in § 2.1) to identify the range at which (2.7) holds. These are not exact numerical values as they depend on $Re$, the type of flow and the order of the structure function.

A natural consequence of the above conjecture and (2.7) and (2.9) is that the defect events in the energy-containing range are related to a Poisson distribution as $\xi _q$ can be modelled using a Poisson process (Dubrulle Reference Dubrulle1994; She & Waymire Reference She and Waymire1995). This does not state that the inertial subrange and the energy-containing range defect events are the same. It only suggests that these are related. Also, it is emphasised that this does not imply that the scaling of the energy ratio of large-scale eddies is similar to isotropic turbulence, as (2.7) is distinct from (1.1) with or without intermittency corrections.

Now, we comment on the signature of intermittency and universality at large scales in shear-dominated turbulence. In inertial subrange turbulence, it is known that the presence of hierarchical moments of turbulent dissipation is a signature of intermittency and universality (Dubrulle Reference Dubrulle1994; She & Leveque Reference She and Leveque1994; She & Waymire Reference She and Waymire1995; Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996). At large scales, the signature of intermittency and universality is the hierarchical energy equilibrium of the eddies, leading to sub-Gaussian scaling. The present picture of sub-Gaussian scaling holds irrespective of the ‘physical shape’ of the eddies but only in the relative distribution of the energy across eddies of different intensities (the definition of energy ratio).

Note that when we refer to the intermittency of large scales, it is not related to the external intermittency observed in shear turbulence. In shear turbulence, external intermittency is often referred to as the turbulent/non-turbulent interface between the fully turbulent flow and the stationary/irrotational ambient (Chauhan, Philip & Marusic Reference Chauhan, Philip and Marusic2014; Gauding et al. Reference Gauding, Bode, Brahami, Varea and Danaila2021; Gustenyov et al. Reference Gustenyov, Egerer, Hultmark, Smits and Bailey2023).

The present scaling is not related to the SO(3) decomposition or the integral structure function scaling of shear turbulence. The SO(3) decomposition is primarily used to isolate turbulence's isotropic and anisotropic components (Arad, L'vov & Procaccia Reference Arad, L'vov and Procaccia1999; Biferale et al. Reference Biferale, Lohse, Mazzitelli and Toschi2002; Casciola et al. Reference Casciola, Gualtieri, Jacob and Piva2005). The integral structure function scaling predicts the scaling exponents by assuming that a power-law scaling exists (Toschi et al. Reference Toschi, Leveque and Ruiz-Chavarria2000), which is unrelated to the present study. In this study, we work with normalised structure functions (2.1), and $f(r)$ is assumed to be an arbitrary function (see § 2.1 for details) and does not always have a power-law scaling (Davidson et al. Reference Davidson, Nickels and Krogstad2006b; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015; Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017). It is stressed that this study investigates the energy ratio and does not examine the scaling exponents of inertial subrange turbulence. Also, this hypothesis differs from what is observed in the inertial subrange, where the structure functions exhibit generalised extended self-similarity, hierarchy and universality (Arneodo et al. Reference Arneodo1996; Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996; Ching et al. Reference Ching, She, Su and Zou2002; Jiang et al. Reference Jiang, Gong, Liu, Zhou and She2006). Equation (2.7) differs from the (generalised) extended self-similarity formulation. In the present case, the energy ratio of the large-scale eddies of different orders attains universality even if the eddies themselves are not assumed to be universal.

3. Description of the numerical and experimental datasets

To verify the conjectured equation (2.7), we investigate the longitudinal structure functions in turbulent boundary layers (SLTEST, MBL1 and MBL2) using data from Kunkel & Marusic (Reference Kunkel and Marusic2006), Hutchins et al. (Reference Hutchins, Nickels, Marusic and Chong2009), Baars et al. (Reference Baars, Talluru, Hutchins and Marusic2015, Reference Baars, Hutchins and Marusic2016), de Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017), turbulent half-channel flow (C590) of MacDonald (Reference MacDonald2022) and vertical buoyancy layers (N580 and N395). The datasets used in the current study are summarised in table 1.

Table 1. Summary of the experimental and numerical datasets.

SLTEST (Kunkel & Marusic Reference Kunkel and Marusic2006) corresponds to a field test, while MBL1 (Hutchins et al. Reference Hutchins, Nickels, Marusic and Chong2009) and MBL2 (Baars et al. Reference Baars, Talluru, Hutchins and Marusic2015, Reference Baars, Hutchins and Marusic2016) correspond to wind tunnel experiments at the University of Melbourne. The measurements were obtained using hot-wire anemometry, and further details are available in the relevant publications.

The relative scaling of longitudinal structure functions of SLTEST and MBL1 are from de Silva et al. (Reference de Silva, Krug, Lohse and Marusic2017). For the MBL2 dataset (Baars et al. Reference Baars, Talluru, Hutchins and Marusic2015, Reference Baars, Hutchins and Marusic2016), the longitudinal structure functions were calculated using Taylor's frozen turbulent hypothesis, akin to what is done by de Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017) and Krug et al. (Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017).

We use direct numerical simulation (DNS) data for the half-channel flow and vertical buoyancy layer. The half-channel flow (C590) was simulated using a fourth-order staggered-grid finite-difference scheme described by Chung et al. (Reference Chung, Chan, MacDonald, Hutchins and Ooi2015a). The reader is referred to MacDonald (Reference MacDonald2022) for additional details on numerical methods and boundary conditions. DNS of the vertical buoyancy layer (N395 and N580) was performed using an in-house second-order finite volume code (Armfield et al. Reference Armfield, Morgan, Norris and Street2003), which has been previously used to simulate turbulent natural convection (Maryada et al. Reference Maryada, Armfield, Dhopade and Norris2023). Numerical simulations were performed at $Pr = 0.71$ using the non-dimensional governing equations and boundary conditions described by Maryada et al. (Reference Maryada, Armfield, Dhopade and Norris2023), and we refer the reader to the same paper for further details. The vertical buoyancy layer is a simplified representation of a vertical natural convection boundary layer immersed in a stably stratified medium (Giometto et al. Reference Giometto, Katul, Fang and Parlange2017; Maryada et al. Reference Maryada, Armfield, Dhopade and Norris2022, Reference Maryada, Armfield, Dhopade and Norris2023). It is a buoyancy-driven flow (a model for vertical convection) in contrast to turbulent flat plate and channel flow boundary layers, which are driven by the momentum of the free stream or pressure differential.

To date, the structure functions of vertical buoyancy layers (vertical convection problems in general) have not been examined extensively, and it is not evident how similar or different these flows are when compared with canonical wall turbulence. In the general context of vertical convection, the vertical buoyancy layer investigated in this study is in the ‘classical regime’ and not the ‘ultimate regime’ (Ng et al. Reference Ng, Ooi, Lohse and Chung2017; Ke et al. Reference Ke, Williamson, Armfield and Komiya2023; Maryada et al. Reference Maryada, Armfield, Dhopade and Norris2023). In the classical regime, most of the turbulence is observed in the outer layer (wall-normal locations beyond the velocity maximum) and not close to the wall. This implies that the vertical buoyancy layer does not exhibit all the essential features of canonical wall turbulence (Ng et al. Reference Ng, Ooi, Lohse and Chung2017; Ke et al. Reference Ke, Williamson, Armfield and Komiya2023), making it distinct from turbulent boundary layers and channels. Hence, it is examined in this study. We shall return to this flow configuration in § 4.3 where it is shown that the vertical buoyancy layer, despite its differences, exhibits the energy ratio universality of (2.7).

It is well known that high-order structure functions are prone to errors due to poor convergence. For SLTEST and MBL1, the convergence was verified until $2p = 10$ (de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015). For MBL2 and DNS data, the convergence of the structure functions was checked using the methodology suggested by Huisman, Lohse & Sun (Reference Huisman, Lohse and Sun2013), Meneveau & Marusic (Reference Meneveau and Marusic2013) and de Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017), and the data reported correspond to converged statistics (until $2p = 8$ for N580 and N395, and $2p=6$ for C590 and MBL2 datasets). As a representative case, figure 1(a) shows the premultiplied p.d.f. of the velocity increment at $r^+ \approx 7000$ ($r^+ > L_I^+$) for the MBL2 dataset. The area under the curve in the figure is the moment, and the smoothness of the curve suggests convergence (Huisman et al. Reference Huisman, Lohse and Sun2013; Meneveau & Marusic Reference Meneveau and Marusic2013; Stevens et al. Reference Stevens, Wilczek and Meneveau2014; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017). Figure 1(b) shows the normalised second-, fourth- and sixth-order structure functions for the same dataset, highlighting its smoothness. Similar results were observed for N580, N395 and C590 datasets and are not shown for brevity.

Figure 1. (a) Premultiplied p.d.f. of the velocity increment ($\varDelta _r u^+ = [u^+(x+r) - u^+(x)]$) at $r^+ \approx 7000$ and $z^+ \approx 825$ for the MBL2 dataset. All the curves in panel (a) are normalised by an arbitrary factor $J_p$ such that the maximum is one for all orders. (b) Normalised second-, fourth- and sixth-order structure function for the same dataset as panel (a) shows the smoothness of the computed structure function. In panel (b), $r^+$ is normalised using $z^+$ and $S^{1/p}_{2p}$ is normalised using $u_\tau ^2$ to facilitate easy comparison with de Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015, Reference de Silva, Krug, Lohse and Marusic2017). This normalisation is not essential for the present scaling (see § 2).

4. Results and discussion

The energy ratio is calculated by computing the slope between normalised structure functions of different orders using a least squares method, similar to the methodology employed by de Silva et al. (Reference de Silva, Krug, Lohse and Marusic2017) and Krug et al. (Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017, Reference Krug, Zhu, Chung, Marusic, Verzicco and Lohse2018). The authors do not explicitly state it as the energy ratio, but from § 4.4, it is evident that their studies are equivalent to the present case. This differs from the extended self-similarity observed in inertial subrange turbulence, where the relative slope of the logarithm of the structure functions exhibits similarity and universality (Benzi et al. Reference Benzi, Ciliberto, Tripiccione, Baudet, Massaioli and Succi1993; Grossmann, Lohse & Reeh Reference Grossmann, Lohse and Reeh1997). In the present case, the slope is calculated directly, which is the energy ratio.

The rest of this section is dedicated to analysing high-order and low-order structure functions across different flows and at different wall-normal locations. The symbols and labels in the various plots below are summarised in table 2.

Table 2. The wall-normal locations where the longitudinal structure functions were calculated for the different datasets. Here, $z^+$ corresponds to the wall-normal location in viscous units, while $z^+/Re_\tau$ corresponds to the wall-normal location in outer units. The labels and symbols are the same as in figures 2 and 3. See table 1 for details of different cases.

Figure 2. The energy ratio of longitudinal structure functions for different flows. See text and table 1 for details on the different symbols. (a) $E^r_p/E^r_1$ of different flows at wall-normal locations where the effect of viscosity is negligible. (b) Effect of wall-normal location on $E^r_p/E^r_1$. Equation (2.5) is the Gaussian scaling and (2.7) is the proposed scaling. See table 2 for a description of the different labels.

Figure 3. $E^r_p/E^r_1$ for low-order fractional p. See table 2 for a description of the different labels.

4.1. High-order structure functions

First, we examine the scaling of the structure functions at locations where the viscous effects are negligible. The viscous effects are negligible in the outer layer (wall-normal locations beyond the velocity maximum) of the vertical buoyancy layer (Maryada et al. Reference Maryada, Armfield, Dhopade and Norris2023), and at $z^+ > 2.6 \sqrt {Re_\tau }$ for turbulent boundary layers and channels (Wei et al. Reference Wei, Fife, Klewicki and McMurtry2005).

Figure 2(a) shows $E^r_p/E^r_1$ for all the different datasets. The error bars for SLTEST and MBL1 represent the $95\,\%$ confidence interval (de Silva et al. Reference de Silva, Krug, Lohse and Marusic2017). For the rest of the data, the standard deviation of the fit is better or comparable to that of SLTEST and MBL1 and, hence, not shown for clarity. It is evident from the figure that, for all the flows considered, the scaling deviates significantly from (2.5) at higher orders ($2p > 3$), which represents the Gaussian scaling. However, it closely follows the scaling in (2.7), which is related to the anomalous scaling of the inertial subrange, demonstrating the efficacy of the proposed scaling.

Having established that the scaling at high orders is universal at moderate $Re_\tau$ and at locations where the effects of the wall are minimal, we now investigate the effect of wall-normal location. Figure 2(b) shows the scaling of energy ratio for C590, MBL2 and N580 datasets. Viscosity is significant for C590B and MBL2W. It is evident that (2.7) correctly predicts the scaling at different wall-normal locations for all the flows considered, even in regions where viscosity is dominant.

The hierarchical scaling is observed for cases N395, N580 and C590, where $Re_\tau$ is significantly smaller than that of MBL1, MBL2 and SLTEST, demonstrating that the hierarchical scaling develops at moderate $Re_\tau$ even if the asymptotic high-$Re_\tau$ behaviour is absent. This was observed by de Silva et al. (Reference de Silva, Krug, Lohse and Marusic2017) and Krug et al. (Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017) in shear-driven wall turbulence for even-order structure functions, but here, we have shown that the scaling is also valid for odd-order absolute structure functions. It is valid even in a buoyancy-driven flow, which is discussed in § 4.3.

Here, it is worth mentioning that (2.7) also holds for C590B, demonstrating that the dynamics of the large scales ($r > L_I$) in the buffer layer in wall turbulence can also be approximated using an additive process. This sub-Gaussian scaling was also observed in Taylor–Couette turbulence (Krug et al. Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017), suggesting that (2.7) would correctly predict the scaling even in those situations.

4.2. Low-order fractional structure functions

High-order structure functions investigated in § 4.1 represent the tails of the p.d.f.s of the velocity increments. The tails of the p.d.f.s correspond to rare events, and it is not certain that the tails always contain universal effects (Chen et al. Reference Chen, Dhruva, Kurien, Sreenivasan and Taylor2005). The low-order structure functions, however, correspond to the core of the p.d.f.s (common/frequent events). The effects of the rare events, which dominate high-order moments, are insignificant for low-order moments. Therefore, we investigate the scaling of the energy ratio of the fractional moments for $0.1 \leq 2p \leq 1.5$ as it acts as a more stringent measure to test the universality of energy ratio and the validity of (2.7). The scaling of the energy ratio is shown in figure 3, where it is clear that the data follow (2.7) better than (2.5). The deviation from Gaussian scaling for low-order moments is evident for all the flows considered, demonstrating that the anomalous scaling also characterises the fluctuations close to the mean and does not only correspond to rare events. This strongly suggests that the non-Gaussian scaling is present for all the energy-containing fluctuations in shear-dominated flows. Like the inertial subrange, where a transitional behaviour for $\zeta _n$ is absent for structure functions of order above and below unity (Sreenivasan et al. Reference Sreenivasan, Vainshtein, Bhiladvala, San Gil, Chen and Cao1996), transitional behaviour is also not present for the energy ratio.

It should be noted that due to the unavailability of data (precise numerical values) for the MBL1 dataset, it is not shown in figures 2(b) and 3. The $Re_\tau$ of MBL2 ($Re_\tau \approx 14\, 750$) is comparable to MBL1 ($Re_\tau \approx 19\, 000$) and, therefore, one should expect quantitatively similar results for the MBL1 dataset.

The scaling of the energy ratio of high-order and low-order structure functions indicates that the energy-containing eddies attain a form of universality in different shear-dominated flows, with the scaling being related to turbulent dissipation. It is again stressed that the present similarity concerns the energy ratio of the eddies and not the eddies themselves.

4.3. Universality of the energy ratio in a turbulent vertical buoyancy layer

In figures 2 and 3, it is evident that the universality of the energy ratio observed in turbulent boundary layers and channels is also observed in a turbulent vertical buoyancy layer. As discussed in § 3, the vertical buoyancy layer investigated in this study is a buoyancy-driven flow and is substantially different from canonical wall turbulence. In this case, the streamwise velocity variance in the vertical buoyancy layer (vertical convection in general) is produced by both shear and buoyancy (Giometto et al. Reference Giometto, Katul, Fang and Parlange2017; Maryada et al. Reference Maryada, Armfield, Dhopade and Norris2022, Reference Maryada, Armfield, Dhopade and Norris2023; Ke et al. Reference Ke, Williamson, Armfield and Komiya2023). The presence of shear and buoyancy fluxes results in the large-scale eddies being governed by two different processes, in contrast to canonical wall turbulence where only shear is present (Tennekes & Lumley Reference Tennekes and Lumley1972). Therefore, it is not trivial that we observe the present scaling. Despite this, we observe shear-dominated scaling, with the similarity not limited to low-order statistics but also governing intense and rare fluctuations.

Crucially, the strong distance-from-the-wall scaling observed in the logarithmic regions of turbulent boundary layers is not observed in the vertical buoyancy layer (especially in the ‘classical regime’, i.e. at moderate $Re_\tau$), indicating that the scaling proposed in this study is not limited to attached eddies alone. This, combined with the scaling observed in the viscosity-dominated regions of canonical wall turbulence (see figure 2b), again suggests that the physical shape of the eddies is not essential for the energy ratio scaling to hold.

This striking observation of shear-dominated scaling in the vertical buoyancy layer shows that certain vertical convection problems in the ‘classical regime’ can be analysed using a framework similar to shear-driven flows.

4.4. Significance of the present hypothesis

It is worthwhile to briefly discuss the present hypothesis's significance and applicability. First, we examine how the present hypothesis relates to works of de Silva et al. (Reference de Silva, Krug, Lohse and Marusic2017) and Krug et al. (Reference Krug, Yang, de Silva, Ostilla-Mónico, Verzicco, Marusic and Lohse2017). In those studies, the authors focused on $A_p/A_m$ instead of the energy ratio (the authors specifically focused on $A_p/A_1$, but the same conclusions hold for different $m$). From § 2.1, we can see that $E^r_{p} = A_p u_\tau ^2$ for the $\ln (r)$ law (2.2), which is valid in the logarithmic region of turbulent boundary layers (Davidson et al. Reference Davidson, Nickels and Krogstad2006b; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015; Chamecki et al. Reference Chamecki, Dias, Salesky and Pan2017). From (2.4), we can write $F_{pm} = A_p/A_m = E^r_{p} / E^r_{m}$. Therefore, $A_p/A_m$ and $F_{pm}$ are mathematically equivalent. In figure 2(a), SLTEST, MBL1 and MBL2L correspond to the logarithmic region of the turbulent boundary layer, demonstrating that the above relation holds. A consequence of the equivalence of $A_p/A_m$ and $F_{pm}$ is that the present model also provides an elegant solution to the fundamental issue of relating $A_p/A_m$ of wall turbulence to the universal intermittency exponents (de Silva et al. Reference de Silva, Krug, Lohse and Marusic2017).

The previous paragraph implies that the coefficient $A_p$ in the $\ln (r)$ law is related to the energy of the eddies, which is a function of the anomalous scaling exponents of the inertial subrange. De Silva et al. (Reference de Silva, Marusic, Woodcock and Meneveau2015) have shown that the structure functions and the one-point moments of streamwise velocity fluctuations are related in the logarithmic region of turbulent boundary layers. Therefore, it is trivial to see that the sub-Gaussian scaling of one-point moments of streamwise velocity fluctuations is also related to the energy ratio, which is related to the intermittent scaling of dissipation. This again highlights a relationship between large and small scales.

5. Concluding remarks and outlook

In summary, we have shown that shear-dominated turbulence attains a hierarchical equilibrium state at large scales ($L_I < r < L_R$) such that the energy ratio of the eddies takes a universal non-Gaussian form. We establish that this is related to the anomalous scaling observed in inertial subrange turbulence (at least for $0.1 \leq 2p \leq 10$), demonstrating a relationship between the large and small scales in shear-dominated flows. A phenomenological model is also proposed to explain this behaviour. By examining data from numerical simulations and experiments, it is shown that the high-order ($2p \geq 2$) and low-order ($0.1 \leq 2p < 2$) energy ratios deviate from Gaussian scaling but closely follow the proposed model, implying that the anomalous scaling is apparent for both rare and frequent events characterising turbulence. The scaling is also valid at different wall-normal locations and under diverse flow conditions, implying that the hierarchical universality of the energy ratio is evident across various turbulent flows so long as shear is dominant. The universality observed is remarkable as it demonstrates that the hierarchical energy equilibrium, as postulated in (2.7), is present and that some form of universality is observed in large scales of shear-dominated turbulence, which are often considered to have significant non-universal effects.

The above observations concerning the correlation between the sub-Gaussian statistics of large scales and the anomalous scaling exponents of inertial subrange turbulence would aid in developing empirical turbulence models that assume a superposition of eddies due to a hierarchical additive process (Marusic & Monty Reference Marusic and Monty2019). Another potential application of the present work concerns validating numerical simulations and experiments, which mostly rely on spectra. However, investigating the high-order properties, such as the relative energy density, would be a much more rigorous test to assess the accuracy (Meneveau & Marusic Reference Meneveau and Marusic2013; de Silva et al. Reference de Silva, Marusic, Woodcock and Meneveau2015). Further, this theory can be used to develop scaling theories that relate the generalised scaling of small scales (Benzi et al. Reference Benzi, Biferale, Ciliberto, Struglia and Tripiccione1996) to the energy ratio of large scales. This is particularly interesting as it would allow us to develop a more refined picture of intermittency in shear-dominated turbulence, including small and large scales.

Acknowledgements

We thank New Zealand eScience Infrastructure (NeSI) for providing high-performance computational resources. We thank the Fluid Mechanics Research Group at the University of Melbourne for making the MBL2 dataset publicly available.

Funding

This research received no specific grant from any funding agency, commercial or not-for-profit sectors.

Declaration of interest

The authors report no conflict of interest.

References

Agostini, L. & Leschziner, M. 2017 Spectral analysis of near-wall turbulence in channel flow at $Re_\tau = 4200$ with emphasis on the attached-eddy hypothesis. Phys. Rev. Fluids 2 (1), 014603.CrossRefGoogle Scholar
Andreolli, A., Gatti, D., Vinuesa, R., Örlü, R. & Schlatter, P. 2023 Separating large-scale superposition and modulation in turbulent channels. J. Fluid Mech. 958, A37.CrossRefGoogle Scholar
Arad, I., L'vov, V.S. & Procaccia, I. 1999 Correlation functions in isotropic and anisotropic turbulence: the role of the symmetry group. Phys. Rev. E 59 (6), 6753.CrossRefGoogle ScholarPubMed
Armfield, S.W., Morgan, P., Norris, S. & Street, R. 2003 A parallel non-staggered Navier–Stokes solver implemented on a workstation cluster. In Computational Fluid Dynamics 2002 (ed. S.W. Armfield, P. Morgan & K. Srinivas), pp. 30–45. Springer.CrossRefGoogle Scholar
Arneodo, A., et al. 1996 Structure functions in turbulence, in various flow configurations, at Reynolds number between 30 and 5000, using extended self-similarity. Europhys. Lett. 34 (6), 411.CrossRefGoogle Scholar
Attili, A. & Bisetti, F. 2012 Statistics and scaling of turbulence in a spatially developing mixing layer at $Re_\lambda = 250$. Phys. Fluids 24 (3), 035109.CrossRefGoogle Scholar
Baars, W.J., Hutchins, N. & Marusic, I. 2016 Spectral stochastic estimation of high-Reynolds-number wall-bounded turbulence for a refined inner-outer interaction model. Phys. Rev. Fluids 1 (5), 054406.CrossRefGoogle Scholar
Baars, W.J., Talluru, K.M., Hutchins, N. & Marusic, I. 2015 Wavelet analysis of wall turbulence to study large-scale modulation of small scales. Exp. Fluids 56, 115.CrossRefGoogle Scholar
Benzi, R., Amati, G., Casciola, C.M., Toschi, F. & Piva, R. 1999 Intermittency and scaling laws for wall bounded turbulence. Phys. Fluids 11 (6), 12841286.CrossRefGoogle Scholar
Benzi, R., Biferale, L., Ciliberto, S., Struglia, M.V. & Tripiccione, R. 1996 Generalized scaling in fully developed turbulence. Phys. D 96 (1–4), 162181.CrossRefGoogle Scholar
Benzi, R., Ciliberto, S., Tripiccione, R., Baudet, C., Massaioli, F. & Succi, S. 1993 Extended self-similarity in turbulent flows. Phys. Rev. E 48 (1), R29.CrossRefGoogle ScholarPubMed
Biferale, L., Lohse, D., Mazzitelli, I.M. & Toschi, F. 2002 Probing structures in channel flow through SO(3) and SO(2) decomposition. J. Fluid Mech. 452, 3959.CrossRefGoogle Scholar
Casciola, C.M., Benzi, R., Gualtieri, P., Jacob, B. & Piva, R. 2001 Double scaling and intermittency in shear dominated flows. Phys. Rev. E 65 (1), 015301.CrossRefGoogle ScholarPubMed
Casciola, C.M., Gualtieri, P., Benzi, R. & Piva, R. 2003 Scale-by-scale budget and similarity laws for shear turbulence. J. Fluid Mech. 476, 105114.CrossRefGoogle Scholar
Casciola, C.M., Gualtieri, P., Jacob, B. & Piva, R. 2005 Scaling properties in the production range of shear dominated flows. Phys. Rev. Lett. 95 (2), 024503.CrossRefGoogle ScholarPubMed
Chamecki, M., Dias, N.L., Salesky, S.T. & Pan, Y. 2017 Scaling laws for the longitudinal structure function in the atmospheric surface layer. J. Atmos. Sci. 74 (4), 11271147.CrossRefGoogle Scholar
Chauhan, K., Philip, J. & Marusic, I. 2014 Scaling of the turbulent/non-turbulent interface in boundary layers. J. Fluid Mech. 751, 298328.CrossRefGoogle Scholar
Chen, S.Y., Dhruva, B., Kurien, S., Sreenivasan, K.R. & Taylor, M.A. 2005 Anomalous scaling of low-order structure functions of turbulent velocity. J. Fluid Mech. 533, 183192.CrossRefGoogle Scholar
Ching, E.S.C., She, Z.-S., Su, W. & Zou, Z. 2002 Extended self-similarity and hierarchical structure in turbulence. Phys. Rev. E 65 (6), 066303.CrossRefGoogle ScholarPubMed
Chung, D., Chan, L., MacDonald, M., Hutchins, N. & Ooi, A. 2015 a A fast direct numerical simulation method for characterising hydraulic roughness. J. Fluid Mech. 773, 418431.CrossRefGoogle Scholar
Chung, D., Marusic, I., Monty, J.P., Vallikivi, M. & Smits, A.J. 2015 b On the universality of inertial energy in the log layer of turbulent boundary layer and pipe flows. Exp. Fluids 56, 110.CrossRefGoogle Scholar
Corrsin, S. 1958 Local isotropy in turbulent shear flow. Res. Memo. 58B11. NACA.Google Scholar
Davidson, P.A. & Krogstad, P.Å. 2009 A simple model for the streamwise fluctuations in the log-law region of a boundary layer. Phys. Fluids 21 (5), 055105.CrossRefGoogle Scholar
Davidson, P.A. & Krogstad, P.-Å. 2014 A universal scaling for low-order structure functions in the log-law region of smooth-and rough-wall boundary layers. J. Fluid Mech. 752, 140156.CrossRefGoogle Scholar
Davidson, P.A., Krogstad, P.-Å. & Nickels, T.B. 2006 a A refined interpretation of the logarithmic structure function law in wall layer turbulence. Phys. Fluids 18 (6), 065112.CrossRefGoogle Scholar
Davidson, P.A., Nickels, T.B. & Krogstad, P.Å. 2006 b The logarithmic structure function law in wall-layer turbulence. J. Fluid Mech. 550, 5160.CrossRefGoogle Scholar
Dong, S., Lozano-Durán, A., Sekimoto, A. & Jiménez, J. 2017 Coherent structures in statistically stationary homogeneous shear turbulence. J. Fluid Mech. 816, 167208.CrossRefGoogle Scholar
Dubrulle, B. 1994 Intermittency in fully developed turbulence: Log-Poisson statistics and generalized scale covariance. Phys. Rev. Lett. 73 (7), 959.CrossRefGoogle ScholarPubMed
Frisch, U. 1995 Turbulence: The Legacy of A. N. Kolmogorov. Cambridge University Press.CrossRefGoogle Scholar
Frisch, U., Sulem, P.-L. & Nelkin, M. 1978 A simple dynamical model of intermittent fully developed turbulence. J. Fluid Mech. 87 (4), 719736.CrossRefGoogle Scholar
Ganapathisubramani, B., Hutchins, N., Monty, J.P., Chung, D. & Marusic, I. 2012 Amplitude and frequency modulation in wall turbulence. J. Fluid Mech. 712, 6191.CrossRefGoogle Scholar
Gauding, M., Bode, M., Brahami, Y., Varea, É. & Danaila, L. 2021 Self-similarity of turbulent jet flows with internal and external intermittency. J. Fluid Mech. 919, A41.CrossRefGoogle Scholar
Ghannam, K., Katul, G.G., Bou-Zeid, E., Gerken, T. & Chamecki, M. 2018 Scaling and similarity of the anisotropic coherent eddies in near-surface atmospheric turbulence. J. Atmos. Sci. 75 (3), 943964.CrossRefGoogle Scholar
Giometto, M.G., Katul, G.G., Fang, J. & Parlange, M.B. 2017 Direct numerical simulation of turbulent slope flows up to Grashof number $Gr = 2.1 \times 10^{11}$. J. Fluid Mech. 829, 589620.CrossRefGoogle Scholar
Grossmann, S., Lohse, D. & Reeh, A. 1997 Application of extended self-similarity in turbulence. Phys. Rev. E 56 (5), 5473.CrossRefGoogle Scholar
Gualtieri, P., Casciola, C.M., Benzi, R., Amati, G. & Piva, R. 2002 Scaling laws and intermittency in homogeneous shear flow. Phys. Fluids 14 (2), 583596.CrossRefGoogle Scholar
Gustenyov, N., Egerer, M., Hultmark, M., Smits, A.J. & Bailey, S.C.C. 2023 Similarity of length scales in high-Reynolds-number wall-bounded flows. J. Fluid Mech. 965, A17.CrossRefGoogle Scholar
Huisman, S.G., Lohse, D. & Sun, C. 2013 Statistics of turbulent fluctuations in counter-rotating Taylor–Couette flows. Phys. Rev. E 88 (6), 063001.CrossRefGoogle ScholarPubMed
Hutchins, N., Nickels, T.B., Marusic, I. & Chong, M.S. 2009 Hot-wire spatial resolution issues in wall-bounded turbulence. J. Fluid Mech. 635, 103136.CrossRefGoogle Scholar
Jacobi, I., Chung, D., Duvvuri, S. & McKeon, B.J. 2021 Interactions between scales in wall turbulence: phase relationships, amplitude modulation and the importance of critical layers. J. Fluid Mech. 914, A7.CrossRefGoogle Scholar
Jiang, X.-Q., Gong, H., Liu, J.-K., Zhou, M.-D. & She, Z.-S. 2006 Hierarchical structures in a turbulent free shear flow. J. Fluid Mech. 569, 259286.CrossRefGoogle Scholar
Jiménez, J. 1998 Turbulent velocity fluctuations need not be Gaussian. J. Fluid Mech. 376, 139147.CrossRefGoogle Scholar
Ke, J., Williamson, N., Armfield, S.W. & Komiya, A. 2023 The turbulence development of a vertical natural convection boundary layer. J. Fluid Mech. 964, A24.CrossRefGoogle Scholar
Kolmogorov, A.N. 1941 The local structure of turbulence in incompressible viscous fluid for very large Reynolds number. In Doklady Akademii Nauk SSSR, vol. 30, pp. 301–303.Google Scholar
Kolmogorov, A.N. 1962 A refinement of previous hypotheses concerning the local structure of turbulence in a viscous incompressible fluid at high Reynolds number. J. Fluid Mech. 13 (1), 8285.CrossRefGoogle Scholar
Kraichnan, R.H. 1974 On Kolmogorov's inertial-range theories. J. Fluid Mech. 62 (2), 305330.CrossRefGoogle Scholar
Kraichnan, R.H. 1991 Turbulent cascade and intermittency growth. Proc. R. Soc. Lond. A 434 (1890), 6578.Google Scholar
Krug, D., Yang, X.I.A., de Silva, C.M., Ostilla-Mónico, R., Verzicco, R., Marusic, I. & Lohse, D. 2017 Statistics of turbulence in the energy-containing range of Taylor–Couette compared to canonical wall-bounded flows. J. Fluid Mech. 830, 797819.CrossRefGoogle Scholar
Krug, D., Zhu, X., Chung, D., Marusic, I., Verzicco, R. & Lohse, D. 2018 Transition to ultimate Rayleigh–Bénard turbulence revealed through extended self-similarity scaling analysis of the temperature structure functions. J. Fluid Mech. 851, R3.CrossRefGoogle Scholar
Kunkel, G.J. & Marusic, I. 2006 Study of the near-wall-turbulent region of the high-Reynolds-number boundary layer using an atmospheric flow. J. Fluid Mech. 548, 375402.CrossRefGoogle Scholar
Landau, L.D. & Lifshitz, E.M. 1987 Fluid Mechanics, 2nd edn. Pergamon Press.Google Scholar
MacDonald, M. 2022 Direct numerical simulation of momentum and scalar turbulent internal boundary layers. In 23rd Australasian Fluid Mechanics Conference (ed. C. Lei), p. 302.Google Scholar
Marusic, I. & Monty, J.P. 2019 Attached eddy model of wall turbulence. Annu. Rev. Fluid Mech. 51, 4974.CrossRefGoogle Scholar
Maryada, K.R., Armfield, S.W., Dhopade, P. & Norris, S.E. 2022 Oblique-mode breakdown of the vertical buoyancy layer. J. Fluid Mech. 953, A34.CrossRefGoogle Scholar
Maryada, K.R., Armfield, S.W., Dhopade, P. & Norris, S.E. 2023 Large-scale motions in a turbulent natural convection boundary layer immersed in a stably stratified environment. J. Fluid Mech. 967, A40.CrossRefGoogle Scholar
Mathis, R., Hutchins, N. & Marusic, I. 2009 Large-scale amplitude modulation of the small-scale structures in turbulent boundary layers. J. Fluid Mech. 628, 311337.CrossRefGoogle Scholar
Meneveau, C. & Marusic, I. 2013 Generalized logarithmic law for high-order moments in turbulent boundary layers. J. Fluid Mech. 719, R1.CrossRefGoogle Scholar
Meneveau, C. & Sreenivasan, K.R. 1987 Simple multifractal cascade model for fully developed turbulence. Phys. Rev. Lett. 59 (13), 1424.CrossRefGoogle ScholarPubMed
Motoori, Y. & Goto, S. 2021 Hierarchy of coherent structures and real-space energy transfer in turbulent channel flow. J. Fluid Mech. 911, A27.CrossRefGoogle Scholar
Mouri, H., Hori, A., Kawashima, Y. & Hashimoto, K. 2012 Large-scale length that determines the mean rate of energy dissipation in turbulence. Phys. Rev. E 86 (2), 026309.CrossRefGoogle ScholarPubMed
Mouri, H., Hori, A. & Takaoka, M. 2009 Large-scale lognormal fluctuations in turbulence velocity fields. Phys. Fluids 21 (6), 065107.CrossRefGoogle Scholar
Mouri, H., Takaoka, M., Hori, A. & Kawashima, Y. 2006 On Landau's prediction for large-scale fluctuation of turbulence energy dissipation. Phys. Fluids 18 (1), 015103.CrossRefGoogle Scholar
Ng, C.S., Ooi, A., Lohse, D. & Chung, D. 2017 Changes in the boundary-layer structure at the edge of the ultimate regime in vertical natural convection. J. Fluid Mech. 825, 550572.CrossRefGoogle Scholar
Obukhov, A.M. 1962 Some specific features of atmospheric turbulence. J. Geophys. Res. 67 (8), 30113014.CrossRefGoogle Scholar
Pan, Y. & Chamecki, M. 2016 A scaling law for the shear-production range of second-order structure functions. J. Fluid Mech. 801, 459474.CrossRefGoogle Scholar
Perry, A.E., Henbest, S. & Chong, M.S. 1986 A theoretical and experimental study of wall turbulence. J. Fluid Mech. 165, 163199.CrossRefGoogle Scholar
She, Z.-S. & Leveque, E. 1994 Universal scaling laws in fully developed turbulence. Phys. Rev. Lett. 72 (3), 336.CrossRefGoogle ScholarPubMed
She, Z.-S. & Waymire, E.C. 1995 Quantized energy cascade and log-Poisson statistics in fully developed turbulence. Phys. Rev. Lett. 74 (2), 262.CrossRefGoogle ScholarPubMed
de Silva, C.M., Krug, D., Lohse, D. & Marusic, I. 2017 Universality of the energy-containing structures in wall-bounded turbulence. J. Fluid Mech. 823, 498510.CrossRefGoogle Scholar
de Silva, C.M., Marusic, I., Woodcock, J.D. & Meneveau, C. 2015 Scaling of second-and higher-order structure functions in turbulent boundary layers. J. Fluid Mech. 769, 654686.CrossRefGoogle Scholar
Sreenivasan, K.R. & Antonia, R.A. 1997 The phenomenology of small-scale turbulence. Annu. Rev. Fluid Mech. 29 (1), 435472.CrossRefGoogle Scholar
Sreenivasan, K.R., Vainshtein, S.I., Bhiladvala, R., San Gil, I., Chen, S. & Cao, N. 1996 Asymmetry of velocity increments in fully developed turbulence and the scaling of low-order moments. Phys. Rev. Lett. 77 (8), 1488.CrossRefGoogle ScholarPubMed
Stevens, R.J.A.M., Wilczek, M. & Meneveau, C. 2014 Large-eddy simulation study of the logarithmic law for second-and higher-order moments in turbulent wall-bounded flow. J. Fluid Mech. 757, 888907.CrossRefGoogle Scholar
Tennekes, H. & Lumley, J.L. 1972 A First Course in Turbulence. MIT Press.CrossRefGoogle Scholar
Toschi, F., Amati, G., Succi, S., Benzi, R. & Piva, R. 1999 Intermittency and structure functions in channel flow turbulence. Phys. Rev. Lett. 82 (25), 5044.CrossRefGoogle Scholar
Toschi, F., Leveque, E. & Ruiz-Chavarria, G. 2000 Shear effects in nonhomogeneous turbulence. Phys. Rev. Lett. 85 (7), 1436.CrossRefGoogle ScholarPubMed
Townsend, A.A. 1976 The Structure of Turbulent Shear Flow, 2nd edn. Cambridge University Press.Google Scholar
Wei, T., Fife, P., Klewicki, J. & McMurtry, P. 2005 Properties of the mean momentum balance in turbulent boundary layer, pipe and channel flows. J. Fluid Mech. 522, 303327.CrossRefGoogle Scholar
Winkelbauer, A. 2012 Moments and absolute moments of the normal distribution. arXiv:1209.4340.Google Scholar
Xie, J.-H., de Silva, C., Baidya, R., Yang, X.I.A. & Hu, R. 2021 Third-order structure function in the logarithmic layer of boundary-layer turbulence. Phys. Rev. Fluids 6 (7), 074602.CrossRefGoogle Scholar
Yang, X.I.A., Marusic, I. & Meneveau, C. 2016 Hierarchical random additive process and logarithmic scaling of generalized high order, two-point correlations in turbulent boundary layer flow. Phys. Rev. Fluids 1 (2), 024402.CrossRefGoogle Scholar
Yeung, P.K., Brasseur, J.G. & Wang, Q. 1995 Dynamics of direct large-small scale couplings in coherently forced turbulence: concurrent physical-and Fourier-space views. J. Fluid Mech. 283, 4395.CrossRefGoogle Scholar
Figure 0

Table 1. Summary of the experimental and numerical datasets.

Figure 1

Figure 1. (a) Premultiplied p.d.f. of the velocity increment ($\varDelta _r u^+ = [u^+(x+r) - u^+(x)]$) at $r^+ \approx 7000$ and $z^+ \approx 825$ for the MBL2 dataset. All the curves in panel (a) are normalised by an arbitrary factor $J_p$ such that the maximum is one for all orders. (b) Normalised second-, fourth- and sixth-order structure function for the same dataset as panel (a) shows the smoothness of the computed structure function. In panel (b), $r^+$ is normalised using $z^+$ and $S^{1/p}_{2p}$ is normalised using $u_\tau ^2$ to facilitate easy comparison with de Silva et al. (2015, 2017). This normalisation is not essential for the present scaling (see § 2).

Figure 2

Table 2. The wall-normal locations where the longitudinal structure functions were calculated for the different datasets. Here, $z^+$ corresponds to the wall-normal location in viscous units, while $z^+/Re_\tau$ corresponds to the wall-normal location in outer units. The labels and symbols are the same as in figures 2 and 3. See table 1 for details of different cases.

Figure 3

Figure 2. The energy ratio of longitudinal structure functions for different flows. See text and table 1 for details on the different symbols. (a) $E^r_p/E^r_1$ of different flows at wall-normal locations where the effect of viscosity is negligible. (b) Effect of wall-normal location on $E^r_p/E^r_1$. Equation (2.5) is the Gaussian scaling and (2.7) is the proposed scaling. See table 2 for a description of the different labels.

Figure 4

Figure 3. $E^r_p/E^r_1$ for low-order fractional p. See table 2 for a description of the different labels.