Hostname: page-component-7bb8b95d7b-dvmhs Total loading time: 0 Render date: 2024-09-27T02:01:07.315Z Has data issue: false hasContentIssue false

Dual scaling and the n-thirds law in grid turbulence

Published online by Cambridge University Press:  16 November 2023

S.L. Tang*
Affiliation:
Center for Turbulence Control, Harbin Institute of Technology, Shenzhen 518055, PR China
R.A. Antonia
Affiliation:
Discipline of Mechanical Engineering, College of Engineering, Science and Environment, University of Newcastle, Newcastle 2308, NSW, Australia
L. Djenidi
Affiliation:
Department of Mechanical Engineering, Indian Institute of Technology – Bombay, Powai, Mumbai 400076, India School of Electrical and Mechanical Engineering, Faculty of Sciences, Engineering and Technology, The University of Adelaide, Adelaide 5005, WA, Australia
*
Email address for correspondence: shunlin.tang88@gmail.com

Abstract

A dual scaling of the turbulent longitudinal velocity structure function $\overline {{(\delta u)}^n}$, i.e. a scaling based on the Kolmogorov scales ($u_K$, $\eta$) and another based on ($u'$, $L$) representative of the large scale motion, is examined in the context of both the Kármán–Howarth equation and experimental grid turbulence data over a significant range of the Taylor microscale Reynolds number $Re_\lambda$. As $Re_\lambda$ increases, the scaling based on ($u'$, $L$) extends to increasingly smaller values of $r/L$ while the scaling based on ($u_K$, $\eta$) extends to increasingly larger values of $r/\eta$. The implication is that both scalings should eventually overlap in the so-called inertial range as $Re_\lambda$ continues to increase, thus leading to a power-law relation $\overline {{(\delta u)}^n} \sim r^{n/3}$ when the inertial range is rigorously established. The latter is likely to occur only when $Re_\lambda \to \infty$. The use of an empirical model for $\overline {{(\delta u)}^n}$, which complies with $\overline {{(\delta u)}^n} \sim r^{n/3}$ as $Re_\lambda \to \infty$, shows that the finite Reynolds number effect may differ between even- and odd-orders of $\overline {{(\delta u)}^n}$. This suggests that different values of $Re_\lambda$ may be required between even and odd values of $n$ for compliance with $\overline {{(\delta u)}^n} \sim r^{n/3}$. The model describes adequately the dependence on $Re_\lambda$ of the available experimental data for $\overline {{(\delta u)}^n}$ and supports indirectly the extrapolation of these data to infinitely large $Re_\lambda$.

Type
JFM Papers
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1. Introduction

For the second- and third-order velocity structure functions in the inertial range (IR), the second similarity hypothesis of Kolmogorov (Reference Kolmogorov1941a,Reference Kolmogorovb) (or K41) predicts

(1.1)\begin{equation} \overline{{{(\delta u)}^2}} = {A_2}{(\bar{\varepsilon}r)^{2/3}}, \end{equation}

where $u$ is the longitudinal velocity fluctuation in the direction $x$, $\delta u = u(x+r) - u(x)$, $r$ being the separation in the direction $x$; the overbar denotes time averaging; ${A_2}$ is a universal constant; and $\bar {\varepsilon }$ is the mean dissipation rate of the turbulent kinetic energy. Also in the same range, one has

(1.2)\begin{equation} \overline{{(\delta u)}^3} ={-} \tfrac{4}{5} \bar{\varepsilon} r, \end{equation}

a result known as the $4/5$ law, derived rigorously by K41 from the Kármán–Howarth (or KH) equation after assuming that viscosity and the influence of the large scales can be neglected. These assumptions seem plausible when the Reynolds number is infinitely large. The extent of the IR is described by $\eta \ll r\ll L$, where $\eta = {({\nu ^3}/\bar {\varepsilon } )^{1/4}}$ is the Kolmogorov length scale and $L$ is the integral length scale, defined by $L = {{u'}^{ - 2}}\int _0^\infty {\overline {u(x + r)u(x)} }\,{\rm d}r$, where ${\overline {u(x + r)u(x)} }$ is the longitudinal correlation function. Within the framework of K41, a general expression for $\overline {{(\delta u)}^n}$ in the IR is

(1.3)\begin{equation} \overline{{{(\delta u)}^n}} = {A_n}{(\bar{\varepsilon} r)^{n/3}}, \end{equation}

(hereafter denoted $n$-thirds law), where ${A_n}$ are universal constants. Kolmogorov (Reference Kolmogorov1962) or K62 (see also Oboukhov Reference Oboukhov1962) made, in a response to Landau's remark (Landau & Lifshitz Reference Landau and Lifshitz1987), an important modification to (1.3), arguably to account for the effect of the large-scale motion on the IR. This has since been widely attributed to the intermittency of $\epsilon$ (Sreenivasan & Antonia Reference Sreenivasan and Antonia1997). Equation (1.3) was replaced by

(1.4)\begin{equation} \frac{{\overline{{{(\delta u)}^n}} }}{{{u'^n}}} \sim {\left(\frac{r}{L}\right)^{{\zeta _n}}}, \end{equation}

where $u' \equiv \overline {u^2}^{1/2}$. Like (1.1) and (1.2), both (1.3) and (1.4) are expected to be valid only when $Re_\lambda \to \infty$. Except when $n=3$, the premultipliers in (1.4) should not depend on the Reynolds number but may be affected by the macrostructure of the flow. As in (1.3), the exponents $\zeta _n$ are universal. The value of $\zeta _3$ remains equal to 1, in compliance with (1.2). For $n \ne 3$, the magnitude of $\zeta _n$ can now depart from $n/3$. This departure or anomalous scaling, which seems to be relatively well described by the relation ${\zeta _n} = {n}/{9} + 2\left [{1 - {{(2/3)}^{n/3}}} \right ]$ (She & Leveque Reference She and Leveque1994), has received fairly strong support from both experimental and numerical data in a wide range of turbulent flows. In their review, Sreenivasan & Antonia (Reference Sreenivasan and Antonia1997) described the support as ‘overwhelming’ but underlined that there are several uncertainties associated with the estimation of $\zeta _n$, not the least of which is the issue of ‘how one defines the scaling range and obtains scaling exponents from power-laws of modest quality.’ For example, an extended self-similarity approach (Benzi et al. Reference Benzi, Ciliberto, Tripiccione, Baudet, Massaioli and Succi1993) was frequently used to extract the scaling exponents of ${\overline {{{(\delta u)}^n}} }$ in the literature and the resulting exponents appear to agree with the predictions of She & Leveque (Reference She and Leveque1994). However, as demonstrated by Tang et al. (Reference Tang, Antonia, Djenidi, Danaila and Zhou2017), this approach masks the finite Reynolds number effect on the scaling exponents of ${\overline {{{(\delta u)}^n}} }$ (see the discussion in the text of figure 17 of Tang et al. Reference Tang, Antonia, Djenidi, Danaila and Zhou2017). A major consideration is whether an IR is actually realizable for all available experimental and numerical data, for which the Reynolds number is finite.

The studies of Qian (Reference Qian1997), Qian (Reference Qian1999), Lundgren (Reference Lundgren2002), Lundgren (Reference Lundgren2003), Antonia & Burattini (Reference Antonia and Burattini2006), Tchoufag, Sagaut & Cambon (Reference Tchoufag, Sagaut and Cambon2012), Antonia et al. (Reference Antonia, Tang, Djenidi and Zhou2019) and Meldi, Djenidi & Antonia (Reference Meldi, Djenidi and Antonia2021) have indicated that (1.2) is in fact approached very slowly, so that a very large value of $Re_\lambda$ ($=u'\lambda /\nu$, where $\lambda = u'/ (\partial u/\partial x)'$ and $\nu$ is the fluid kinematic viscosity) may be required before the IR is unambiguously established (see figure 4.6 of Sagaut & Cambon (Reference Sagaut and Cambon2018) which summarizes different predictions for $\overline {{(\delta u)}^3}$). In particular, the eddy damped quasinormal Markovian numerical simulations in both decaying and forced homogeneous isotropic turbulence (Meldi et al. Reference Meldi, Djenidi and Antonia2021) support Antonia & Burattini (Reference Antonia and Burattini2006) and Antonia et al. (Reference Antonia, Tang, Djenidi and Zhou2019). This seriously calls into question the validity of the published estimates of $\zeta _n$. There is compelling evidence (Tang et al. Reference Tang, Antonia, Djenidi, Danaila and Zhou2017) to suggest that the finite Reynolds number (FRN) effect needs to be taken into account, since $\overline {{{(\delta u)}^n}}$, regardless of how it is normalized, continues to evolve with $Re_\lambda$ in the scaling range.

One expects that, as described by Kolmogorov's first similarity hypothesis, the behaviour of the small scales is characterized by $\bar {\varepsilon }$ and $\nu$. The relevant scales are $\eta$ and $u_K$ ($= {(\nu \bar {\varepsilon } )^{1/4}}$). In the energy containing range of separations, the relevant scales are expected to be $u'$ and $L$. This dual scaling is expected to apply, albeit in an approximate fashion, at finite values of $Re_\lambda$. As $Re_\lambda$ continues to increase, there should be a region of overlap between the two different normalizations. When $Re_\lambda \to \infty$, the IR should be established rigorously and the two sets of scales should become interchangeable in this range.

Gamard & George (Reference Gamard and George2000) used this dual scaling approach to describe how the longitudinal velocity spectrum, as measured by Mydlarski & Warhaft (Reference Mydlarski and Warhaft1996) in decaying turbulence downstream of an active grid, evolves with $Re_\lambda$ in the overlap region. In this region, the spectrum was assumed to vary, albeit approximately, as $\alpha k_1^{ - 5/3 + \beta }$, where $\alpha$ and $\beta$ depend on $Re_\lambda$. In the limit $Re_\lambda \to \infty$, $\beta \to 0$ and $\alpha \to$ constant, a result (Obukhov Reference Obukhov1941) which corresponds to (1.1). The latter equation was derived, starting with the Kármán–Howarth equation, by Lundgren (Reference Lundgren2002) using matched asymptotic expansions. In essence, this approach is not dissimilar to using dual scaling, since outer and inner expansions satisfy similarity based on ($u'$, $L$) and ($u_K$, $\eta$), respectively. Lundgren (Reference Lundgren2002) noted that this derivation of the $2/3$ law is similar to Millikan's derivation of the log law in turbulent wall flows (Millikan Reference Millikan1939). This analogy was also noted by Barenblatt & Goldenfeld (Reference Barenblatt and Goldenfeld1995), Barenblatt & Chorin (Reference Barenblatt and Chorin1998) and McKeon & Morrison (Reference McKeon and Morrison2007). Djenidi, Antonia & Tang (Reference Djenidi, Antonia and Tang2019) showed that scale invariance leads to the ‘$n$-thirds’ law when $Re_\lambda \to \infty$. More recently, Djenidi, Antonia & Tang (Reference Djenidi, Antonia and Tang2022) and Djenidi, Antonia & Tang (Reference Djenidi, Antonia and Tang2023) found that constraints imposed by the Cauchy–Schwarz inequality require $\zeta _n$ to be equal to $n/3$ when $Re_\lambda \to \infty$. Equation (1.3) conforms with this requirement but (1.4) violates it if $\zeta _n \ne n/3$.

Grid turbulence, which represents a close, though imperfect, approximation to homogeneous isotropic turbulence is expected to satisfy dual scaling. This has already been verified by Mydlarski & Warhaft (Reference Mydlarski and Warhaft1996) and Gamard & George (Reference Gamard and George2000) in the context of the $u$ spectrum, which corresponds to $\overline {{{(\delta u)}^2}}$, for values of $Re_\lambda$ extending to approximately 470. These authors found, by extrapolation of $Re_\lambda$ to infinity, that the data comply with (1.1). The appropriateness of this result is all the more compelling, in the context of the dual scaling, when it is recognized that the KH equation satisfies similarity at all scales as $Re_\lambda \to \infty$ (e.g. Speziale & Bernard Reference Speziale and Bernard1992). It is clearly important to test the dual scaling approach to larger values of both $Re_\lambda$ and $n$. This is the main objective of this paper. For this purpose, we make use of the published data for $\overline {{{(\delta u)}^n}}$ obtained in the variable density turbulent tunnel (VDTT) facility at the Max Planck Institute in Göttingen (Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014). The distributions of $\overline {{{(\delta u)}^n}}$ as a function of $r$ have appeared in several publications (e.g. Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014; Sinhuber, Bodenschatz & Bewley Reference Sinhuber, Bodenschatz and Bewley2015; Sinhuber, Bewley & Bodenschatz Reference Sinhuber, Bewley and Bodenschatz2017; Birnir Reference Birnir2019; Kaminsky et al. Reference Kaminsky, Birnir, Bewley and Sinhuber2020) for $2 \leq n \leq ~8$ and values of $Re_\lambda$ as large as approximately 1500.

In § 2, we first examine the appropriateness of the dual scaling approach in the context of the Kármán–Howarth equation (i.e. the transport equation of $\overline {{{(\delta u)}^2}}$) for two cases: first, when the contribution of the large-scale term is neglected, and second, when the contribution from the viscous term is neglected. In § 3, we briefly recall the basic parameters for the VDTT grid turbulence data. In § 4, we test the VDTT data in the context of the dual scaling. We next present a model for $\overline {{{(\delta u)}^n}}$ which complies with the dual scaling. In particular, the model is constrained to yield a power-law behaviour ($\sim r^{\zeta _n}$, with $\zeta _n=n/3$) when $Re_\lambda \to \infty$. The adequate agreement between the model and the data (§ 5) allows extrapolation via the model to values of $Re_\lambda$ that are several orders of magnitude larger than in the experiment. Conclusions are given in § 6.

2. Dual scaling in the context of the Kármán–Howarth equation

We consider first the applicability of the inner (Kolmogorov scales, $\eta$ and $u_K$) and outer ($u'$ and $L$) scaling in the context of the Kármán–Howarth equation which governs the transport of $\overline {{(\delta u)}^2}$:

(2.1)\begin{equation} \frac{4}{5} \bar{\epsilon} r ={-} \overline{{(\delta u)}^3} + 6\nu \frac{\partial }{{\partial r}}\overline{{(\delta u)}^2} + I_u(r), \end{equation}

where $I_u(r)$ is the term which reflects the contribution from the large scales to the transport of $\overline {{(\delta u)}^2}$ in homogeneous and isotropic turbulence (e.g. Danaila et al. Reference Danaila, Anselmet, Zhou and Antonia1999; Hill Reference Hill2001). We will focus on decaying grid turbulence, partly because this flow can satisfy isotropy, at least for small scales, to a reasonable approximation and also because moderately large values of $Re_\lambda$ can be achieved, e.g. by using an active grid (e.g. Mydlarski & Warhaft Reference Mydlarski and Warhaft1996, Reference Mydlarski and Warhaft1998; Gylfason, Ayyalasomayajula & Warhaft Reference Gylfason, Ayyalasomayajula and Warhaft2004) or by varying the density of the fluid (e.g. Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014; Sinhuber et al. Reference Sinhuber, Bodenschatz and Bewley2015, Reference Sinhuber, Bewley and Bodenschatz2017; Kaminsky et al. Reference Kaminsky, Birnir, Bewley and Sinhuber2020). The departure from local isotropy at all scales can be quantified by the ratio between calculated and measured second-order spanwise velocity structure functions, i.e. $\overline {{{(\delta w)}^2}}_{iso}/\overline {{{(\delta w)}^2}}$, where $\overline {{{(\delta w)}^2}}_{iso}$ is given by (Monin & Yaglom Reference Monin and Yaglom2007)

(2.2)\begin{equation} \overline{{{(\delta w)}^2}}_{iso} = \left( {1 + \frac{r}{2}\frac{{\rm d}}{{{\rm d}r}}} \right)\overline{{{(\delta u)}^2}}. \end{equation}

At $r = \lambda$, a separation which resides near the lower end of the scaling range (Tang et al. Reference Tang, Antonia, Djenidi, Danaila and Zhou2017), the departure from local isotropy is relatively small ($\lesssim 10\,\%$, see figure 10 of Lavoie, Djenidi & Antonia (Reference Lavoie, Djenidi and Antonia2007) at $Re_\lambda =36\unicode{x2013}43$ in grid turbulence with and without a secondary contraction). Further, it was shown that the ratio $\overline {{{(\delta w)}^2}}_{iso}/\overline {{{(\delta w)}^2}}$ satisfies isotropy within approximately $10\,\%$ at all separations when a secondary contraction is introduced. This result is consistent with that of Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1966) and Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971) who used a contraction to improve the isotropy of grid turbulence, with $u'$ being nearly equal to $v'$. It is worth mentioning that since the magnitude of $Re_\lambda$ in the present paper is much larger than that of Lavoie et al. (Reference Lavoie, Djenidi and Antonia2007), we thus expect a further improved local isotropy, compared with that of Lavoie et al. (Reference Lavoie, Djenidi and Antonia2007). Also, since isotropy requires homogeneity, and isotropy at all scales is approximately satisfied in grid turbulence, the weak spatial decay along $x$, the mean flow direction, should not affect the behaviour of $\overline {{{(\delta u)}^n}}$. In other words, the influence of any small inhomogeneity in the $x$ direction on the behaviour of $\overline {{{(\delta u)}^n}}$ should be negligible. Further, it is well established, based on the KH equation, that the effect of the spatial decay along the $x$ direction in the scaling range weakens as $Re_\lambda$ increases (Antonia & Burattini Reference Antonia and Burattini2006; Tang et al. Reference Tang, Antonia, Djenidi, Danaila and Zhou2017; Antonia et al. Reference Antonia, Tang, Djenidi and Zhou2019). A further attractive feature of grid turbulence is that the energy budget is relatively simple and can hence be measured fairly accurately thus yielding a reliable estimation of $\bar {\varepsilon }$ (e.g. Antonia, Zhou & Zhu Reference Antonia, Zhou and Zhu1998).

The two main items of interest here are the conditions for which inner (or Kolmogorov) scaling ($u_K$, $\eta$) and outer scaling ($u'$, $L$) satisfy the similarity of (2.1).

(i) When the effect of the large scale term $I_u(r)$ is neglected, the similarity of (2.1) is readily satisfied by

(2.3)$$\begin{gather} \frac{{\overline{{{(\delta u)}^2}} }}{{u_0^2}} = f\left( {\frac{r}{{{l_0}}}} \right), \end{gather}$$
(2.4)$$\begin{gather}\frac{{\overline{{{(\delta u)}^3}} }}{{u_0^3}} = g\left( {\frac{r}{{{l_0}}}} \right), \end{gather}$$

when $u_0 \equiv u_K$ and $l_0 \equiv \eta$, viz. the ($u_K$, $\eta$) scaling applies. This was discussed in some detail by Antonia, Djenidi & Danaila (Reference Antonia, Djenidi and Danaila2014). Equation (2.1) can be rewritten as

(2.5)\begin{equation} \frac{4}{5}\frac{r}{{{l_0}}} = \frac{{6\nu u_0^2}}{{\bar{\epsilon} l_0^2}}f' - \frac{{u_0^3}}{{\bar{\epsilon} {l_0}}}g\left( {\frac{r}{{{l_0}}}} \right) \end{equation}

(here, the prime signifies a derivative with respect to $r/l_0$), so that $\bar {\varepsilon } {l_0}/u_0^3$ and $\bar {\varepsilon } l_0^2/\nu u_0^2$ (or $\bar {\varepsilon } {l_0}/u_0^3$, ${u_0}{l_0}/\nu$) are constants. Clearly, ($u_K$, $\eta$) is a possible solution since

(2.6a,b)\begin{equation} \frac{\bar{\varepsilon} {\eta}}{u_K^3}=1 \quad {\rm and} \quad \frac{{u_K}{\eta}}{\nu}=1. \end{equation}

Scaling based on ($u'$, $L$) is also possible provided

(2.7a,b)\begin{equation} \frac{\bar{\epsilon} {L}}{u'^3}=C_\epsilon={\rm constant} \quad {\rm and} \quad \frac{{u'}{L}}{\nu}={\rm constant}. \end{equation}

The first requirement in (2.7a,b) has been validated in several different flows (e.g. Sreenivasan Reference Sreenivasan1984; Burattini, Lavoie & Antonia Reference Burattini, Lavoie and Antonia2005b; Ishihara, Gotoh & Kaneda Reference Ishihara, Gotoh and Kaneda2009; Mi, Xu & Zhou Reference Mi, Xu and Zhou2013; Vassilicos Reference Vassilicos2015; McComb et al. Reference McComb, Berera, Yoffe and Linkmann2015), at least when $Re_\lambda$ is sufficiently large. The second can, to our knowledge, hold only in the far field of a circular jet (Antonia, Satyaprakash & Hussain Reference Antonia, Satyaprakash and Hussain1980; Burattini, Antonia & Danaila Reference Burattini, Antonia and Danaila2005a) when $Re_\lambda$ is finite. It should however be satisfied in decaying grid turbulence when $Re_\lambda \to \infty$ (e.g. George Reference George1992; Speziale & Bernard Reference Speziale and Bernard1992; Antonia et al. Reference Antonia, Smalley, Zhou, Anselmet and Danaila2003). The above observations lead to the conclusion that the scaling ($u_K$, $\eta$) should be effective regardless of the flow since the two dimensionless parameters in (2.6a,b) are indeed universal (with a value of 1). Nevertheless, the neglect of the large-scale term (2.5) for values of $r/\eta$ that extend into the IR is unlikely to be fully justifiable unless $Re_\lambda \to \infty$. Neither of the parameters in (2.7a,b) is likely to be flow independent; indeed, there is adequate evidence (e.g. Sreenivasan Reference Sreenivasan1984; Burattini et al. Reference Burattini, Lavoie and Antonia2005b; Mi et al. Reference Mi, Xu and Zhou2013; Vassilicos Reference Vassilicos2015) that the magnitude of $C_\epsilon$ depends on the flow.

(ii) $I_u(r)$ is now retained and the viscous term in (2.1) is neglected. The focus now is on the larger scales of motion. Equation (2.1) can be rewritten as

(2.8)\begin{equation} \frac{4}{5}\frac{r}{{{l_0}}} ={-} \frac{{u_0^3}}{{\bar{\varepsilon} {l_0}}}g\left( {\frac{r}{{{l_0}}}} \right) + \frac{{{I_u}(r)}}{{\bar{\varepsilon} {l_0}}}. \end{equation}

We concentrate on grid turbulence, for which

(2.9)\begin{equation} I_u(r)={-} \frac{3}{{{r^4}}}\int_0^r {{s^4}}\left[ U\frac{{\partial \overline{(\delta u)^2} }}{{\partial x}}\right]\,{\rm d}s, \end{equation}

where $s$ is a dummy variable, identifiable with the separation along $x$, and $U$ is the (constant) mean velocity in the $x$ direction (e.g. Danaila et al. Reference Danaila, Anselmet, Zhou and Antonia1999; Antonia et al. Reference Antonia, Zhou, Danaila and Anselmet2000; Danaila, Anselmet & Antonia Reference Danaila, Anselmet and Antonia2002). Using (2.9), ${I_u}/\bar {\varepsilon } {l_0}$ becomes

(2.10)\begin{equation} \frac{{3}}{{{{(r/{l_0})}^4}}}\left( {\int_0^{r/{l_0}} {{{\left( {\frac{s}{{{l_0}}}} \right)}^4}f\left( {\frac{r}{{{l_0}}}} \right)\frac{{{\rm d}s}}{{{l_0}}}} } \right)\left[ {\frac{U}{{\bar{\varepsilon} }}\frac{{\partial u_0^2}}{{\partial x}} - \frac{{Uu_0^2}}{{\bar{\varepsilon} }}{{\left( {\frac{r}{{{l_0}}}} \right)}^2}\frac{f'}{{{l_0}}}\frac{{{\rm d}{l_0}}}{{{\rm d}\kern0.06em x}}} \right]. \end{equation}

The expression within square brackets must be constant for similarity to be satisfied. The first term of this expression is indeed constant since $\bar \epsilon =-\tfrac {3}{2}U({{\partial {{u_0^2}} }}/{{\partial x}})$ is the turbulent energy budget (when $u_0 \equiv u'$). The second term is also constant with the ($u'$, $L$) scaling since $\bar {\varepsilon } {L}/u'^3$ is constant and ${({1}/{{{u'}}})({{{\rm d}{L}}}/{{{\rm d}\kern0.06em x}})}$ is also constant as both $u'$ and ${\rm d}L/{{\rm d}\kern0.06em x}$ vary as $x^{n/2}$ when the energy decays according to a power-law, viz. $\overline {u^2}\sim x^{n}$ and $\bar {\epsilon }\sim x^{n-1}$; the latter two relations, together with $C_\varepsilon =\bar {\epsilon } L/u'^3={\rm const}$, lead to $L \sim x^{n/2+1}$. It is worth mentioning that $U$ in the second term within the square brackets does not affect the above analysis and the corresponding results since we can replace $U/{{\rm d}\kern0.06em x}$ with $1/{\rm d}t$ in grid turbulence. The previous similarity requirements were previously outlined by Kármán & Von Howarth (Reference Von Kármán and Howarth1938) and Townsend (Reference Townsend1976). We conclude that the similarity of (2.8), i.e. the transport equation of ${\overline {{{(\delta u)}^2}} }$ when $r$ is sufficiently large to allow the viscous term to be neglected, is satisfied by the ($u'$, $L$) scaling. We should note however that the ($u_K$, $\eta$) scaling can also satisfy similarity of (2.8) when $n=-1$ (in this case, both ${({1}/{{\bar {\varepsilon } }})({{\partial u_K^2}}/{{\partial x}})}$ and ${({1}/{{u_K }})({{\partial \eta }}/{{\partial x}})}$ are constant). This rate of decay was shown by Speziale & Bernard (Reference Speziale and Bernard1992) to be the asymptotically consistent high-Reynolds-number solution. The condition ${{u'}{L}}/{\nu }={\rm constant}$ in (2.7a,b) is immediately satisfied in this case so that the scales ($u'$, $L$), just like ($u_K$, $\eta$), would satisfy complete similarity (or self-preservation), i.e. similarity for all scales of motion. The Taylor microscale $\lambda$ should then be a valid replacement for $L$ since $Re_\lambda$ should no longer depend on $x$ and $\lambda$ should be proportional to $L$. The general self-preservation analysis of Djenidi & Antonia (Reference Djenidi and Antonia2015) confirmed the dependence of $n$ on the initial conditions and showed how this dependence is related to the initial length and velocity scales. These authors suggested that it may be possible to have initial conditions that lead to $n=-1$ with a finite $Re_\lambda$. However, $Re_\lambda$ must remain constant throughout the decay.

The available grid turbulence data in the literature indicate that $n$ is typically smaller than $-1$, usually with a value ranging between $-1.1$ and $-1.5$ (e.g. George Reference George1992; Lavoie et al. Reference Lavoie, Djenidi and Antonia2007; Lee et al. Reference Lee, Djenidi, Antonia and Danaila2013; Sinhuber et al. Reference Sinhuber, Bodenschatz and Bewley2015). In particular, Sinhuber et al. (Reference Sinhuber, Bodenschatz and Bewley2015) showed that $n$ is independent of the Reynolds number, with a value of $n = - 1.18 \pm 0.02$, over a wide range of Reynolds numbers; this value is close to the prediction of Saffman (Reference Saffman1967) ($n =-1.2$). Lavoie et al. (Reference Lavoie, Djenidi and Antonia2007) showed experimentally that even when the turbulence is nearly perfectly isotropic, $n$ is not equal to $-1$. Our interest is to assess the validity of the dual scaling and the range of $r/l_0$ over which it applies. The scaling based on ($u_K$, $\eta$) is expected to extend to increasingly larger values of $r/\eta$ as $Re_\lambda$ increases. Correspondingly, the scaling based on ($u'$, $L$) is expected to extend to increasingly smaller values of $r/L$ as $Re_\lambda$ increases. Since both scalings must eventually overlap as $Re_\lambda \to \infty$, the overlap region should include the inertial range. Regardless of the scaling used, this overlap region should satisfy similarity and hence be independent of $Re_\lambda$. All of the above analysis is focused on the grid turbulence where the influence of large scales on $-\overline {{{(\delta u)}^3}}$ can be quantified by the $I_u(r)$ term in (2.9). If one considers different flows with different large-scale forcing, the influence of the large-scale motion is likely to differ from that in grid turbulence. Accordingly, in those situations, the flow may locally reflect the large-scale anisotropy at finite $Re_\lambda$ imposed by the forcing. For example, when anisotropic forcing is used in direct numerical simulations of box turbulence at $Re_{q\lambda }=20\unicode{x2013}32$ ($=\overline {u_iu_i}^{1/2}\lambda /\nu$), it induces small-scale anisotropy which is traced to the dynamics of non-local triads with one low wavenumber in the directly forced wavenumber band (Yeung & Brasseur Reference Yeung and Brasseur1991). Also, Kurien & Sreenivasan (Reference Kurien and Sreenivasan2000) observed that the structure functions of order two to six exhibit a strong anisotropy at moderate scales in the atmospheric surface layer at $Re_\lambda =870\unicode{x2013}2100$.

3. Basic flow parameters in grid turbulence

The grid turbulence data we used for our study were digitized from figures presented by Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) and Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) ($Re_\lambda =110\unicode{x2013}1620$). Since all the experimental data reported here have already been published, the reader can find detailed descriptions of the experimental conditions and measurement techniques in the original papers. Here we only describe and discuss the basic flow parameters associated with those two datasets and explain the treatment we have had to apply to these data.

Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) estimated $\bar {\varepsilon }$ from $\overline {(\delta u)^3}$ by assuming $\bar {\varepsilon }=\max (-5/4\overline {(\delta u)^3}/r)$, i.e. assuming that (1.2) is tenable. This would result in $\bar {\varepsilon }$ being underestimated since the maximum value of $\overline {(\delta u)^3}/ \bar {\varepsilon } r$ has not yet reached the value of 4/5 for the available data in grid turbulence (see for example figure 1a of Antonia et al. (Reference Antonia, Tang, Djenidi and Zhou2019)). Indeed, their magnitudes of $C_\epsilon$ ($\equiv \bar {\varepsilon } L/u'^3=0.58$, 0.71, 0.83, 0.77, 0.77 and 0.75 at $Re_\lambda=150$, 400, 549, 730, 1370 and 1620, respectively, Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) are smaller than those of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) ($C_\varepsilon=1.06$, 1.12, 0.91, 0.91 and 0.81 at $Re_\lambda=110$, 264, 508, 1000 and 1450, respectively), in which $\bar {\varepsilon }$ was estimated from the turbulent energy budget $\bar {\epsilon }=-\tfrac {3}{2}U({{\partial {{u'^2}} }}/{{\partial x}})$. Also, the spectra of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) in the dissipative range do not collapse, see figure 1(b), which is reproduced from their figure 30. To obtain a more accurate estimation of $\bar {\epsilon }$, the spectral chart method of Djenidi & Antonia (Reference Djenidi and Antonia2012) has been applied to the spectra of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) (their figure 30, or figure 1(b) in the present paper). The ‘new’ estimates of $\bar {\varepsilon }$ are denoted by $\bar {\varepsilon }_{spec}$. This method is based on the observation that there is collapse in the upper part of the dissipative range not only for grid turbulence, but also for many other turbulent flows; see for example figure 9 of Saddoughi & Veeravalli (Reference Saddoughi and Veeravalli1994), figure 6.14 of Pope (Reference Pope2000), figure 3 of Gotoh, Fukayama & Nakano (Reference Gotoh, Fukayama and Nakano2002), figure 5 of Larssen & Devenport (Reference Larssen and Devenport2011), figures 1–2 of Antonia et al. (Reference Antonia, Djenidi and Danaila2014) and figure 1(a) of Tang et al. (Reference Tang, Antonia, Djenidi and Zhou2020). Therefore, the values of $\bar {\varepsilon }_{spec}$ ensure that there is collapse in the upper part of the dissipative range, as illustrated in figure 1(a) where the distributions are normalized by $\bar {\varepsilon }_{spec}$ and $\nu$. For comparison, we have included the grid turbulence data of Larssen & Devenport (Reference Larssen and Devenport2011) at $Re_\lambda= 124$ and 1360. It is worth mentioning that all the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) were obtained in passive grid turbulence. However, the data of Larssen & Devenport (Reference Larssen and Devenport2011) at $Re_\lambda =124$ and 1362 were measured in passive and active grid turbulence, respectively. The distribution of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) is in almost perfect agreement with that of Larssen & Devenport (Reference Larssen and Devenport2011) at $Re_\lambda =124$ at all scales. This is not too surprising since both distributions were measured in passive grid turbulence at the same $Re_\lambda$. We can observe that there is a large spike at low wavenumbers (around $k_1\eta \approx 0.0002$) for the data of Larssen & Devenport (Reference Larssen and Devenport2011) at $Re_\lambda =1362$. This is due to the rotation of the active grid which ‘introduces additional energy into the spectra’ (Larssen & Devenport Reference Larssen and Devenport2011). It is plausible that this spike, which reflects the presence of a strong coherent motion, may affect the medium wavenumber range, i.e. $0.002\lesssim k_1\eta \lesssim 0.04$, over which the magnitude of the energy spectrum at $Re_\lambda =1362$ is larger than all others. We recall that Larssen & Devenport (Reference Larssen and Devenport2011) calculated $\bar {\varepsilon }$ by integrating the longitudinal dissipation spectrum, i.e. $\bar {\varepsilon } = 15\nu \int {k_1^2{\phi _u}({k_1})d} {k_1}$. Local isotropy requires that the integration of the Kolmogorov-normalized longitudinal dissipation spectrum is equal to 1/15 ($\equiv 0.0667$). Although not shown here, we should stress that the dissipation spectrum at $Re_\lambda =1362$ is in good agreement with all other distributions at all scales for $Re_\lambda \leq 701$, confirming the accuracy of the dissipation spectrum at $Re_\lambda =1362$. More importantly, the integration across all scales at $Re_\lambda =1362$ leads to 0.069, which is only 3.5 % larger than the isotropic value of 1/15. This further indicates that the magnitude of $\bar {\varepsilon }$ at $Re_\lambda =1362$ should be reasonably accurate. More importantly, there is reasonable agreement between the ‘$\bar {\varepsilon }_{spec}$’-normalized spectra and the spectra of Larssen & Devenport (Reference Larssen and Devenport2011) for $k_1\eta \gtrsim 0.3$ at all $Re_\lambda$. In contrast, as seen in figure 1(b), there is no collapse in the upper part of the dissipative range since $\bar {\epsilon }$ was underestimated by using the maximum value of $-\overline {(\delta u)^3}/ \bar {\varepsilon } r$ at finite $Re_\lambda$. In particular, as $Re_\lambda$ increases, the approach to a $-$5/3 scaling of the spectra, as reflected by the nearly horizontal distributions, is similar to that of Mydlarski & Warhaft (Reference Mydlarski and Warhaft1996). The magnitudes of $C_\varepsilon$, based on $\bar {\varepsilon }_{spec}$, are close to those of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) (figure 2), confirming that $\bar {\varepsilon }_{spec}$ should be very close to the true value $\bar {\varepsilon }$. To further confirm the magnitudes of $\bar {\varepsilon }_{spec}$, we have also included the grid turbulence data of Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971) (${Re}_{\lambda }=61\unicode{x2013}72$, ) and Larssen & Devenport (Reference Larssen and Devenport2011) (${Re}_{\lambda }=101\unicode{x2013}1362$, )) (see figure 2). We can observe that the magnitudes of $C_\varepsilon$, based on $\bar {\varepsilon }_{spec}$, are also close to those of Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971) and Larssen & Devenport (Reference Larssen and Devenport2011), especially for ${Re}_{\lambda }\lesssim 300$ and ${Re}_{\lambda }\gtrsim 900$; the scatter for the data of Larssen & Devenport (Reference Larssen and Devenport2011) in the range $300 \lesssim {Re}_{\lambda }\lesssim 900$ is relatively large. We recall that $C_\varepsilon$ in stationary forced periodic box turbulence follows a functional form (Donzis, Sreenivasan & Yeung Reference Donzis, Sreenivasan and Yeung2005):

(3.1)\begin{equation} {C_\varepsilon } = A(1 + \sqrt {1 + {{(B/{Re_\lambda })}^2}} ), \end{equation}

where $A=0.2$ and $B=92$ (see green curve in figure 2). When using (3.1) to fit the grid turbulence data in figure 2, we observe that the asymptotic value of ${C_\varepsilon }$ in grid turbulence is approximately 0.8, which corresponds to $A=0.4$. Here, $B$ is essentially constant which reflects the finite Reynolds number effect on ${C_\varepsilon }$ in this flow. We take $B=60$ which appears to fit all the grid turbulence data reasonably well. We finally compare the $\overline {(\delta u)^2}/u_K^2$ distributions, calculated from the Kolmogorov-normalized energy spectra (figure 1a) using the following relation (e.g. Dickey & Mellor Reference Dickey and Mellor1979; Monin & Yaglom Reference Monin and Yaglom2007):

(3.2)\begin{equation} \overline{{{(\delta {u})}^2}} /u_K^2 = 2\int_0^\infty {{\phi _{{u}}}(k_1\eta)} (1 - \cos (k_1r))\,{\rm d}(k_1\eta), \end{equation}

with the other grid turbulence data (Comte-Bellot & Corrsin Reference Comte-Bellot and Corrsin1971; Malecot Reference Malecot1998; Zhou & Antonia Reference Zhou and Antonia2000; Larssen & Devenport Reference Larssen and Devenport2011) (see figure 3). Evidently, the distributions of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) are quite close to the other grid turbulence data at all scales when ${Re_\lambda }$ is comparable and in the dissipative range, regardless of ${Re_\lambda }$. Based on the above analysis, we can conclude that the magnitudes of $\bar {\varepsilon }_{spec}$, obtained from the spectral chart method, are reasonably accurate and reliable. The values of $Re_\lambda$ are then re-estimated from the isotropic relation $\overline {u^2}/u_K^2=Re_\lambda /15^{1/2}$. The final values are 124, 379, 495, 701, 1310 and 1558, which are slightly smaller than the values of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014), i.e. $Re_\lambda =150, 400, 549, 730, 1370$ and 1620, respectively. For reference, we have also included the data for stationary forced periodic box turbulence. We can observe from figure 2 that $C_\varepsilon$ is close to 0.8 at large $Re_\lambda$ for the grid turbulence and to 0.4 for the stationary forced periodic box turbulence. This difference is consistent with the difference observed in $L/\eta$ since ${C_\varepsilon }$ can be written as

(3.3)\begin{equation} {C_\varepsilon } = \frac{{\bar{\varepsilon} L}}{{u'{^3}}} = \frac{{L/\eta }}{{u'{^3}/u_K^3}} = \frac{{L/\eta }}{{Re_\lambda ^{3/2}/{{15}^{3/4}}}}, \end{equation}

where the isotropic relation $u{'^2}/u_K^2 = {Re_\lambda }/{15^{1/2}}$ has been used. A plausible reason for the difference in the magnitude of ${C_\varepsilon }$ between grid and box turbulence is the difference in $L/\eta$ at any given $Re_\lambda$, which reflects, at least partially, the different types of large-scale forcing between these two flows. For example, $L/\eta =2522$ at $Re_\lambda =1300$ in box turbulence (Iyer, Sreenivasan & Yeung Reference Iyer, Sreenivasan and Yeung2020), whereas $L/\eta =5516$ at $Re_\lambda =1360$ in grid turbulence (Larssen & Devenport Reference Larssen and Devenport2011). Tsuji (Reference Tsuji2009) observed a difference of a factor of two in $L/\eta$ between the DNS data and experimental data in most other flows (see their figure 1b). This difference in $L/\eta$ is equivalent to the factor of two difference in ${C_\varepsilon }$ between grid turbulence and box turbulence.

Figure 1. (a) Kolmogorov normalized compensated spectra of $u$ in grid turbulence (Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) ($Re_\lambda=$ 1101450), normalized by $\bar {\epsilon }_{spec}$ and $\nu$ (the spectral chart method). For comparison, we have included the grid turbulence data of Larssen & Devenport (Reference Larssen and Devenport2011) (dashed curves) at $Re_\lambda = 124$ and 1360. (b) Corresponding grid turbulence data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) ($Re_\lambda = 110\unicode{x2013}1450$), normalized by $\bar {\epsilon }$, estimated from $\overline {(\delta u)^3}$ (see text), and $\nu$; note that to see the dissipative range more clearly, a smaller range $k_1\eta =0.01\unicode{x2013}1$ is used for the vertical axis.

Figure 2. Distributions of $C_\epsilon = \bar {\epsilon } L /u'^3$ versus ${Re}_{\lambda }$ in grid turbulence. ( red), (Kaminsky et al. Reference Kaminsky, Birnir, Bewley and Sinhuber2020); ( blue), re-estimated $C_\epsilon$, based on $\bar {\epsilon }_{spec}$ (the spectral chart method), versus $Re_\lambda$ for the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) (see text). For comparison, we have included the grid turbulence data of Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971) () and Larssen & Devenport (Reference Larssen and Devenport2011) (). Black curve, (3.1) with $A=0.4$ and $B=60$. Also included are the data in stationary forced periodic box turbulence (($\bigcirc$ black), Wang et al. (Reference Wang, Chen, Brasseur and Wyngaard1996); ($\bigcirc$ blue), Ishihara & Kaneda (Reference Ishihara and Kaneda2002); (${\ast }$ black), Jimenez et al. (Reference Jimenez, Wray, Saffman and Rogallo1993); ( green), Ishihara et al. (Reference Ishihara, Gotoh and Kaneda2009); ($+$ black), Cao, Chen & Doolen (Reference Cao, Chen and Doolen1999); ( blue), Yeung & Zhou (Reference Yeung and Zhou1997); ( blue), Gotoh et al. Reference Gotoh, Fukayama and Nakano2002). Green curve, (3.1) with $A=0.2$ and $B=92$.

Figure 3. Distributions of $\overline {(\delta u)^2}/u_K^2$ versus $r/\eta$ in grid turbulence for the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) at ${Re_\lambda }=124$, 379, 495, 701, 1310 and 1558 (green curves). For comparison, other grid turbulence data are also shown; red curves, Zhou & Antonia (Reference Zhou and Antonia2000) (${Re_\lambda }=27$, 51, 62, 75, 83, 89 and 100); cyan curves, Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971) (${Re_\lambda }=61$ and 65); black curves, Larssen & Devenport (Reference Larssen and Devenport2011) (${Re_\lambda }=124$ and 1360); blue curve, Malecot (Reference Malecot1998) (${Re_\lambda }=144$); pink curve, $(r/\eta )^2/15$. The arrow indicates the direction of increasing ${Re_\lambda }$. Note that the data of Comte-Bellot & Corrsin (Reference Comte-Bellot and Corrsin1971), Larssen & Devenport (Reference Larssen and Devenport2011) and Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) are calculated from their spectra, based on (3.2).

Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) have shown distributions of $\overline {(\delta u)^2}$, $\overline {|\delta u|^3}$, $\overline {(\delta u)^4}$, $\overline {(\delta u)^6}$ and $\overline {(\delta u)^8}$ in grid turbulence at $Re_\lambda = 110$, 264, 508, 1000 and 1450, respectively; the $\overline {(\delta u)^2}{/u_K^2}$ distributions are shown in figure 4(a) (green curves). Figures 1–5 of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) were digitized to obtain $\overline {{(\delta u)^2}}$ distributions. We then divided $\overline {{(\delta u)^2}}$ by $u_K^2$ ($u_K$ was calculated from $\eta$ and $\nu$ shown in their table 1). For comparison, also included are the other grid turbulence data, same as in figure 3. It can be observed that the distributions of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) strongly depart from the other distributions. This is because they actually plotted $\overline {{(\delta u)^n}}/u_K^n$ versus $r/(\eta M)$, instead of $\overline {{(\delta u)^n}}/u_K^n$ versus $r/\eta$. We recall that there is an overwhelming amount of data confirming that the $\overline {(\delta u)^2}/u_K^2$ distributions collapse in the dissipative range over a large range of ${Re_\lambda }$ in various flows (e.g. Pearson & Antonia Reference Pearson and Antonia2001; Gotoh et al. Reference Gotoh, Fukayama and Nakano2002; Ishihara et al. Reference Ishihara, Gotoh and Kaneda2009; McComb et al. Reference McComb, Yoffe, Linkmann and Berera2014). The same is observed in grid turbulence (figure 3). It is then expected that the distributions of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) should collapse with those reported in the literature, at least in the small-scale range, if one removes the factor $M$ from the ratio $r/(\eta M)$. Since the value of $M$ was not provided by Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020), we replotted their data with $r/(\eta M)$ multiplied by different values of $M$ so the distributions align with with other grid turbulence data (figure 4b). The values of $M$ are 29, 6.9, 2.96, 1 and 0.54 at $Re_\lambda =110$, 264, 508, 1000 and 1450, respectively; see also table 1. These values will be used to plot $\overline {(\delta u)^4}/u_K^4$ and $\overline {(\delta u)^6}/u_K^6$ versus $r/\eta$, which will be presented and discussed later. In other words, we do not use ‘$M$’ to correct $\eta$ (or equivalently $u_K$); instead, it is simply a means of forcing agreement with the other grid turbulence data in the dissipative range. As discussed above, the universality of this range is supported by an overwhelming amount of data over a large range of ${Re_\lambda }$ in various flows, as expected on the basis of the first similarity hypothesis of K41. Finally, we compare the distributions of $\overline {(\delta u)^4}/u_K^4$ at ${Re_\lambda }=110$ and 508 with those of Gylfason et al. (Reference Gylfason, Ayyalasomayajula and Warhaft2004) at comparable ${Re_\lambda }$ (168 and 656) in grid turbulence (figure 5a). It is worth mentioning that $\overline {(\delta u)^4}/u_K^4$ leads to $15^{-2}F_4(r/\eta )^4$ as $r/\eta \to 1$. Also, we can observe (later from figure 16) that $F_4$ at ${Re_\lambda }=550$ and 656 are close to each other; the same can be observed for ${Re_\lambda }=110$ and 168. This implies that the distributions at ${Re_\lambda }=550$ and 656 (and also those at ${Re_\lambda }=110$ and 168) should be close to each other at small $r/\eta$. These two expectations are indeed observed in figure 5(a) in the separation range $r/\eta \lesssim 20\unicode{x2013}30$. Figure 5(a) further justifies our use of the numerical factor $M$ on the data of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020). As mentioned above, $\overline {(\delta u)^4}/u_K^4$ approaches $15^{-2}F_4(r/\eta )^4$ as $r/\eta \to 1$. Also, since $F_4$ increases systematically with increasing $Re_\lambda$ for ${Re_\lambda }\lesssim 500$ (see figure 16 later), the $\overline {(\delta u)^4}/u_K^4$ distribution at $Re_\lambda =264$ at small $r/\eta$ must lie between the distributions of $Re_\lambda=110$ and 508. Yet, we can observe from figure 5(b) that, when $M=6.9$, the distribution at $Re_\lambda =264$ is smaller than that at $Re_\lambda =110$; this is incorrect and should be rectified. The value $M=10$, which ensures that the distribution at $Re_\lambda =264$ at small $r/\eta$ lies between the distributions of $Re_\lambda = 110$ and 508, as illustrated in figure 5(b), is a reasonable value for plotting $\overline {(\delta u)^4}/u_K^4$ at $Re_\lambda =264$ (see figure 5b). Also, the data for $\overline {(\delta u)^8}$ are not presented in the present paper since their behaviour is similar to that of $\overline {(\delta u)^6}$. As mentioned above, Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) have also shown the $\overline {|\delta u|^3}$ distributions. In the present paper, we will only discuss the behaviour of $\overline {(\delta u)^3}$ using the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) in the same flow, instead of $\overline {|\delta u|^3}$, since there is an exact equation for the former quantity, as discussed in § 1. It is worth noting that the data mentioned above were measured by a hot-wire anemometry and Taylor's hypothesis was used to calculate $\overline {(\delta u)^n}$. It is generally accepted that this hypothesis is reliable when the mean rate of shear and turbulence intensity are small. Indeed, according to He, Jin & Yang (Reference He, Jin and Yang2017), in a flow such as grid turbulence where the mean shear rate is zero, the conversion from time to space can be written as

(3.4)\begin{equation} r = \sqrt {1 + (u'/U)^2} Ut. \end{equation}

For the VDTT grid turbulence data, the turbulence intensity $u'/U$ is only in the range $2.35\,\%\sim 3.7\,\%$ (see table II of Bodenschatz et al. Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014), and thus justifies the use of Taylor's hypothesis. Pope (Reference Pope2000) pointed out that ‘in grid turbulence with $u'/U\ll 1$, it [Taylor's hypothesis] is quite accurate.’

Figure 4. (a) Distributions of $\overline {(\delta u)^2}/u_K^2$ versus $r/\eta$ in grid turbulence for the data of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) at ${Re_\lambda }=110$, 264, 508, 1000 and 1450 (green curves). Panel (b) corresponds to the same data, after applying a numerical factor $M$ to their data (see text). For comparison, other grid turbulence data, same as in figure 3, are also shown in (a,b). Pink curve, $(r/\eta )^2/15$. The arrow indicates the direction of increasing ${Re_\lambda }$.

Figure 5. (a) Distributions of $\overline {(\delta u)^4}/u_K^4$ versus $r/\eta$ in grid turbulence (Kaminsky et al. Reference Kaminsky, Birnir, Bewley and Sinhuber2020) after applying a numerical factor $M(=29$ and 2.96 for $Re_\lambda=110$ and 508) to their data (see text). For comparison, the grid turbulence data of Gylfason et al. (Reference Gylfason, Ayyalasomayajula and Warhaft2004) (without shear) at $Re_\lambda=168$ (pink curve) and 656 (cyan curve) are shown; they are reproduced from figure 5(c,d) of Gylfason et al. (Reference Gylfason, Ayyalasomayajula and Warhaft2004). (b) Distributions of $\overline {(\delta u)^4}/u_K^4$ versus $r/\eta$ in grid turbulence (Kaminsky et al. Reference Kaminsky, Birnir, Bewley and Sinhuber2020) at $Re_\lambda =264$ after applying a numerical factor $M(=6.9$ and 10) to their data (see text). The dotted curve, $15^{-2}F_4(r/\eta )^4$ with $F_4=9.6$. Other curves are the same as (a).

Table 1. Values of $M$ used to plot $\overline {(\delta u)^2}/u_K^2$, $\overline {(\delta u)^4}/u_K^4$ and $\overline {(\delta u)^6}/u_K^6$ versus $r/\eta$ at each $Re_\lambda$ for the data of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020) (see text). Note that a value of 10, instead of 6.9, is used to plot $\overline {(\delta u)^4}/u_K^4$ at $Re_\lambda =264$ since its distribution should lie between the distributions of $Re_\lambda =110$ and 508 in the dissipative range (see figure 5).

4. Dual scaling and its consequences

We show in figures 6–9 the distributions of $\overline {(\delta u)^n}$ ($n=2,3,4,6$) normalized by ($u_K$, $\eta$) and ($u'$, $L$) in grid turbulence at $Re_\lambda = 110\unicode{x2013}1558$, using the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) (after reevaluating $\bar {\varepsilon }$, as discussed in § 3) and Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020). We can observe from figures 6–9 that the scaling based on ($u'$, $L$) extends to increasingly smaller values of $r/L$ as $Re_\lambda$ increases. Further, the scaling based on ($u_K$, $\eta$) extends to increasingly larger values of $r/\eta$ as $Re_\lambda$ increases. These results indicate that both scalings should eventually overlap at larger values of $Re_\lambda$. Another feature of figures 6–9 is that the range over which the scaling based on ($u'$, $L$) is tenable appears to be larger than that over which the scaling based on ($u_K$, $\eta$) holds over the present $Re_\lambda$ range, especially for $n>3$. One possible reason for this behaviour is that $\overline {(\delta u)^n}/u'^n$ as $r\to L$ approaches the appropriate Gaussian values, which are $Re_\lambda$-independent, while $\overline {(\delta u)^n}/u_K^n$ as $r\to \eta$ leads to the normalized moments of the velocity derivative ($F_n$), which are likely to be constant only at large $Re_\lambda$ for $n>3$ (note that $F_2=1$ by definition and $|F_3|\approx 0.53$ for $Re_\lambda \gtrsim 300$ (Antonia et al. Reference Antonia, Djenidi, Danaila and Tang2017)); the behaviour of $F_n$ will be discussed briefly later in the context of figure 16. However, as $r\to \infty$, we can estimate that, for $n=2, 4$ and 6, $\overline {(\delta u)^n}/u'^n$ approach the values of 2, 12 and 120, respectively, when (3.1), (3.7) and (3.10) of Pearson & Antonia (Reference Pearson and Antonia2001) are used. The $Re_\lambda$ dependence of the limiting values of $\overline {(\delta u)^n}/u_K^n$ ($n=2,3,4,6$) as $r\to \infty$ has been discussed by Pearson & Antonia (Reference Pearson and Antonia2001) (see their figure 6). They concluded that the assumption that $\delta u$ is ‘Gaussian when $r$ is large is, experimentally, almost impossible to disprove’.

Figure 6. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^2}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020).

Figure 7. Effect of $Re_\lambda$ on the distributions of $-\overline {(\delta u)^3}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014); note that for (a), $u_K$ and $\eta$ were re-evaluated as explained in the text.

Figure 8. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^4}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020).

Figure 9. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^6}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020).

The considerations in § 2 indicate that similarity based on ($u_K$, $\eta$) can satisfy the Kármán–Howarth equation even when $Re_\lambda$ is finite provided $r$ is sufficiently small. Equally, the Kármán–Howarth equation admits a similarity solution based on ($u'$, $L$) when $r$ is sufficiently large. It is therefore reasonable to assume that there are similarity solutions of the form

(4.1)$$\begin{gather} \frac{{\overline{{{(\delta u)}^2}} }}{{u_K^2}} = f_1\left( {\frac{r}{{{\eta}}}} \right), \end{gather}$$
(4.2)$$\begin{gather}\frac{{\overline{{{(\delta u)}^2}} }}{\overline{u^2}} = f_2\left( {\frac{r}{{{L}}}} \right), \end{gather}$$

for sufficiently small and large $r$, respectively. As $Re_\lambda$ increases, the range of $r/\eta$ and $r/L$ over which (4.1) and (4.2) are valid also increases (figure 6). The expectation that these ranges must eventually overlap over the inertial range is highly plausible (see Gamard and George's discussion of the overlap in the spectra of $u$, when normalized by ($u_K$, $\eta$) and ($u'$, $L$) Gamard & George Reference Gamard and George2000), especially as $Re_\lambda \to \infty$. Note that we are dealing with the same distributions of $\overline {{{(\delta u)}^2}}$, after subjecting it to two different types of normalization. Matching (4.1) with (4.2) requires that

(4.3)\begin{equation} u_K^2{f_1}\left( {\frac{r}{\eta }} \right) = \overline{{u^2}} {f_2}\left( {\frac{r}{L}} \right). \end{equation}

If it is further assumed that, in the inertial range, ${f_1}(r/\eta )$ exhibits a power-law form, viz.

(4.4)\begin{equation} {f_1}\left( {\frac{r}{\eta }} \right) \sim {\left( {\frac{r}{\eta }} \right)^\alpha }, \end{equation}

then

(4.5)\begin{equation} {f_2}\left( {\frac{r}{L}} \right) \sim \frac{{u_K^2}}{{\overline{{u^2}} }}{\left( {\frac{L}{\eta }} \right)^\alpha }{\left( {\frac{r}{L}} \right)^\alpha }. \end{equation}

For isotropic turbulence, ${{u_K^2}}/{{\overline {{u^2}} }} = {{{{15}^{1/2}}}}/{Re_\lambda }$ and ${L}/{\eta } = {Re_\lambda ^{3/2}}/{{{{15}^{3/4}}}}$ if $C_\epsilon =1$. The plots in figure 6 show that (4.1) and (4.2) are closely satisfied for significant ranges of $r/\eta$ and $r/L$. In particular, at the two largest values of $Re_\lambda$, the upper value of $r/\eta$ exceeds 50 whilst the lower value of $r/L$ lies below 0.01. This implies that an overlap region is starting to conform with the expected limits of the inertial range, i.e. $\eta \ll r \ll L$. The Reynolds number independence displayed by figure 6 for $r/\eta \lesssim 50$ and $r/L\gtrsim 0.01$ is expected to be approached in the overlap region. As discussed in §§ 1 and 2, the two scalings should become interchangeable as $Re_\lambda \to \infty$ so that the inertial range should also be independent of $Re_\lambda$. The formulation (4.5) or ${f_2}\sim Re_\lambda ^{3/2\alpha - 1}{(r/L)^\alpha }$ is independent of $Re_\lambda$ when $3/2\alpha -1=0$, i.e. when $\alpha =2/3$, which corresponds to the K41 prediction. This result was also derived by Lundgren (Reference Lundgren2002) from the Kármán–Howarth equation using matched asymptotic expansions when $Re_\lambda \to \infty$. A similar result was derived earlier by Gamard & George (Reference Gamard and George2000) in the context of the $u$ spectrum; the approach adopted in this paper is consistent with the asymptotic invariance principle and the methodology of near-asymptotics, introduced by George (Reference George1989, Reference George1994). The previous approach for $\overline {{{(\delta {u})}^2}}$ can be extended to $\overline {{{(\delta {u})}^n}}$ (see for example Djenidi et al. Reference Djenidi, Antonia and Tang2019).

Figures 7–9 indicate that the dual scaling applies to $\overline {{{(\delta {u})}^3}}$, $\overline {{{(\delta {u})}^4}}$ and $\overline {{{(\delta {u})}^6}}$. One would expect that, in general, the dual scaling will also apply to $\overline {{{(\delta {u})}^n}}$, if the probability distribution function (p.d.f.) of $\delta {u}$ complies with the two types of normalization. For simplicity, we next focus on a parametrized form of $\overline {{{(\delta {u})}^n}}$ (Dhruva Reference Dhruva2000; Kurien & Sreenivasan Reference Kurien and Sreenivasan2000), viz.

(4.6) \begin{equation} \dfrac{{\overline{{{(\delta u)}^n}} }}{{u_K^n}} = \dfrac{1}{{{{15}^{n/2}}}}{F_n}\dfrac{{{{\Big(\dfrac{r}{\eta }\Big)}^n}{{\Big({1 + {D_n}\Big(\dfrac{r}{L}\Big)} \Big)}^{2{C_n} - n}}}}{{{{\Big( {1 + {B_n}{{\Big(\dfrac{r}{\eta }\Big)}^2}}\Big)}^{{C_n}}}}}, \end{equation}

where $B_n$, $C_n$, $D_n$ and $F_n$ ($\equiv \overline {{{(\partial u/\partial x)}^n}} /{\overline {{{(\partial u/\partial x)}^2}} ^{n/2}}$) are constants. Note that, for $n>2$, $F_n$ is only expected to approach a constant value at increasingly larger values of $Re_\lambda$ as $n$ increases. An earlier version of (4.6), without the large-scale term, viz. ${{{\left ( {1 + {D_n}({r}/{L})} \right )}^{2{C_n} - n}}}$, was proposed by Stolovitzky, Sreenivasan & Juneja (Reference Stolovitzky, Sreenivasan and Juneja1993) who tested it against laboratory boundary layer data for a moderate Reynolds number ($Re_\lambda \approx 200$). The stated objective was to obtain reliable estimates of the scaling exponents (i.e. $C_n$). One can however query this objective given that the power-law behaviour of $\overline {{{(\delta {u})}^n}}$ is rigorous only when the inertial range is established, i.e. $Re_\lambda \to \infty$. As will be shown later, ‘$\zeta _n$’ for $\overline {{{(\delta {u})}^n}}$ cannot be defined even at the largest $Re_\lambda$ ($\approx 1500$) for the present grid turbulence data; this is not surprising since the 4/5 law has yet to be established (see figure 12a later). Dhruva (Reference Dhruva2000) found that (4.6) provided a good fit to atmospheric surface layer data for $\overline {{{(\delta {u})}^2}}$, $\overline {{{(\delta {u})}^4}}$ and $\overline {{{(\delta {u})}^6}}$ over the complete range of $r$. Our main interest is that (4.6) should represent as adequate a fit as possible to the experimental data for $\overline {{{(\delta {u})}^n}}$ irrespectively of $n$ and $Re_\lambda$. In particular, the constants in (4.6) should satisfy the asymptotic case ($Re_\lambda \to \infty$). This is really the context in which (4.6) was used by Antonia & Burattini (Reference Antonia and Burattini2006) and Antonia et al. (Reference Antonia, Tang, Djenidi and Zhou2019) with $n=2$. In other words, we shall not use (4.6) to estimate $C_n$ for different values of $Re_\lambda$, as was done, for example, by Stolovitzky et al. (Reference Stolovitzky, Sreenivasan and Juneja1993). When $Re_\lambda$ is finite, at least at $Re_\lambda \sim 10^3$, (4.6) does not reduce to a simple power-law form ($\sim r^\alpha$) in the scaling range in various turbulent flows such as the stationary forced periodic box turbulence and along the axes of axisymmetric and plane jets (Tang et al. Reference Tang, Antonia, Djenidi, Danaila and Zhou2017).

Equation (4.6) should be interpreted as a composite model for $\overline {{{(\delta {u})}^n}}$, which satisfies the dual scaling at small and large $r$. In particular, in the inertial range ($\eta \ll r \ll L$), (4.6) reduces to

(4.7)\begin{equation} \frac{{\overline{{{(\delta u)}^n}} }}{{u_K^n}} = \frac{1}{{{{15}^{n/2}}}}{F_n}B_n^{ - {C_n}}{\Big(\dfrac{r}{\eta }\Big)^{n - 2{C_n}}}. \end{equation}

Using the isotropic relations ${L/\eta } = {C_\epsilon }{15^{ - 3/4}}Re_\lambda ^{3/2}$ and $\overline {u^2}/u_K^2=Re_\lambda /15^{ 1/2}$, we have

(4.8)\begin{equation} \frac{{\overline{{{(\delta u)}^n}} }}{{{{\overline{{u^2}} }^{n/2}}}} = \frac{{{{15}^{n/4}}Re_\lambda ^{3/2(n - 2{C_n})}}}{{{{15}^{n/2}}Re_\lambda ^{n/2}}}{F_n}B_n^{ - {C_n}}{\left(\frac{r}{L}\right)^{n - 2{C_n}}}C_\varepsilon ^{^{n - 2{C_n}}}{15^{ - 3/4(n - 2{C_n})}}. \end{equation}

Note that (4.8) is the generalization of (4.5) to $\overline {{{(\delta {u})}^n}}$. It is also the starting point of nearly all intermittency models, viz. ${{\overline {{{(\delta u)}^n}} }}/{{{u'^n}}} \sim {({r}/{L})^{{\zeta _n}}}$, with a prefactor that is independent of $Re_\lambda$ but possibly dependent on the macrostructure of the flow. We underline that (4.8) is equivalent to (4.7) since only a transformation of variables has been applied. If (4.8) is $Re_\lambda$-independent, we immediately obtain

(4.9)\begin{equation} \tfrac{3}{2}(n - 2{C_n}) - n/2 = 0, \end{equation}

which leads to

(4.10)\begin{equation} C_n= n/3. \end{equation}

We will use this value of $C_n$ in (4.6) in the next section. Substituting $C_n= n/3$ into (4.8), we can obtain

(4.11)\begin{equation} \frac{{\overline{{{(\delta u)}^n}} }}{{{{\overline{{u^2}} }^{n/2}}}} = \frac{1}{{{{15}^{n/2}}}}{F_n}B_n^{ - {C_n}}{\left(\frac{r}{L}\right)^{n - 2{C_n}}}C_\varepsilon ^{n - 2C_n}. \end{equation}

Note that (4.7) and (4.11) are identical if $C_\varepsilon =1$. The use of $C_n= n/3$ in (4.6) is fully justifiable since (4.6) should comply with complete similarity, i.e. similarity at all values of $r/\eta$ (or $r/L$) as $Re_\lambda \to \infty$. For finite values of $Re_\lambda$, (4.6) should describe the incomplete similarity inherent in the dual scaling scheme (this will be illustrated in figures 18–21). For $n=2$, (4.6) was, as noted earlier, used previously and shown to describe adequately the dependence of $\overline {{{(\delta u)}^2}}$ on $Re_\lambda$ in various flows.

Figure 10. Distributions of $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ based on (4.6) at ${Re_\lambda }=124$, 379, 495, 701, 1310 and 1558. Symbols with the same colour correspond to the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) shown in figure 3.

Figure 11. (a) Distributions of $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 6(a). (b) Local slope ($LS_2= {\rm d}\log (\overline {{{(\delta {u})}^2}}) /{\rm d}\log r$). Dashed horizontal line, $2/3$.

Figure 12. (a) Distributions of $-\overline {{{(\delta u)}^3}} /\bar {\epsilon } r$ based on (4.6) at ${Re_\lambda }=117$, 352, 535, 679, 1323, 1548, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 7(a). (b) Local slope ($LS_3= {\rm d}\log (-\overline {{{(\delta {u})}^3}}) /{\rm d}\log r$). Dashed horizontal line, $1$.

Figure 13. (a) Distributions of $\overline {{{(\delta u)}^4}} /(\bar {\epsilon } r)^{4/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 8(a). (b) Local slope ($LS_4= {\rm d}\log (\overline {{{(\delta {u})}^4}}) /{\rm d}\log r$) corresponding to the distributions in (a). Dashed horizontal line, $4/3$.

Figure 14. (a) Distributions of $\overline {{{(\delta u)}^6}} /(\bar {\epsilon } r)^{6/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 9(a). (b) Local slope ($LS_6= {\rm d}\log (\overline {{{(\delta {u})}^6}}) /{\rm d}\log r$). Dashed horizontal line, $6/3$.

It is worth mentioning that, as $r\to \infty$, (4.6) is proportional to $Re_\lambda ^{n/2}$ after taking $C_n= n/3$ and using the relation ${L/\eta } = {C_\epsilon }{15^{ - 3/4}}Re_\lambda ^{3/2}$. This is consistent with the limiting values as $r\to \infty$ for $n=2$, 4 and 6 (see (3.1), (3.7) and (3.10) of Pearson & Antonia Reference Pearson and Antonia2001).

5. Comparison between (4.6) and experimental data

Figures 10, 11(a)–14(a) show a comparison between the experimental data in figures 3, 6(a)–9(a) and (4.6). Note that we have used $C_n= n/3$; the values for $B_n$, $C_\epsilon$, $D_n$ and $F_n$ are listed in tables 2 and 3. It is worth mentioning that $B_2=0.0056$ and $F_2=1$ lead to a Kolmogorov constant for ${\overline {{{(\delta u)}^2}} }$ of $A_2=$ 2.11. Also, $B_3=0.0114$ and $-F_3=0.53$ lead to a Kolmogorov constant for ${\overline {{{(\delta u)}^3}} }$ of $A_3=$ 4/5. The magnitudes of $F_3$, $F_4$ and $F_6$ used in (4.6) are consistent with the available data in grid turbulence and along the axis of the plane jet (see figure 16 later). The magnitude of $C_\epsilon\ (=0.8)$ is consistent with the VDTT data at large $Re_\lambda$ (figure 2). Here, $D_2$ and $D_3$ are selected so that (4.6) fits the data for $\overline {(\delta u)^2}/u_K^2$ and $\overline {(\delta u)^3}/u_K^3$ reasonably well at all $Re_\lambda$. The magnitudes of $D_4$ and $B_4$ are determined by trial and error until (4.6) fits the data for $\overline {(\delta u)^4}/u_K^4$ at all $Re_\lambda$; the magnitudes of $D_6$ and $B_6$ are determined in the same way. It should be pointed out that all the constants $B_n$, $C_n$, $D_n$ and $F_n$ are $Re_\lambda$-independent except for $F_4$ in the range $Re_\lambda <1000$ and for $F_6$ in the range $Re_\lambda <1450$. Here, $F_4=10$ for $Re_\lambda \geq 1000$ and $B_4=0.011$ lead to a Kolmogorov constant for ${\overline {{{(\delta u)}^4}} }$ of $A_4=18.18$. Also, $F_6=440$ for $Re_\lambda \geq 1450$ and $B_6=0.0178$ lead to a Kolmogorov constant for ${\overline {{{(\delta u)}^6}} }$ of $A_6=413.33$. In particular, we can obtain two ratios: $A_4/A_2^2=4.1$ and $A_6/A_2^3=43.7$. Their magnitudes are consistent with the predictions of Qian (Reference Qian2000) who used a non-Gaussian p.d.f. of $\delta u$ to obtain $A_4/A_2^2=4.1\unicode{x2013}5.4$ and $A_6/A_2^3=37.4\unicode{x2013}69.0$. Here, the main objective of using (4.6), with the above magnitudes for $B_n$, $C_n$, $D_n$ and $F_n$ in grid turbulence, is to describe the dependence on $Re_\lambda$ of the available experimental data for $\overline {(\delta u)^n}/u_K^n$ which allows us to extrapolate the $\overline {(\delta u)^n}/u_K^n$ distributions to infinitely large $Re_\lambda$. The local slope $LS_n$ ($= {\rm d}\log |\overline {{{(\delta {u})}^n}}| /{\rm d}\log r$), based on (4.6), is shown in figures 11(b)–14(b). Note that the local slope $LS_2$ of the distributions in figure 10 is similar to that in figure 11(b) and thus not shown. We can observe from these figures that the agreement between curves and symbols is satisfied approximately at almost all scales over a significant range of ${Re_\lambda }$, confirming the adequacy of the fit used to emulate the available data for ${\overline {{{(\delta u)}^2}} }$, ${\overline {{{(\delta u)}^3}} }$, ${\overline {{{(\delta u)}^4}} }$ and ${\overline {{{(\delta u)}^6}} }$. It should be pointed out that the agreement between the model and the data for some of the distributions at large $Re_\lambda$ and at moderate scales is not so good, as seen for example for the distributions of $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ at ${Re_\lambda }= 1310$ and 1558 in the range $r/\eta \approx 20\unicode{x2013}100$, which correspond to the range $k_1\eta \approx 0.06\unicode{x2013}0.31$ in the energy spectra. We can indeed observe from figure 1(a) that the magnitude of the Kolmogorov-normalized energy spectra in the range $k_1\eta \approx 0.06\unicode{x2013}0.31$ at ${Re_\lambda }= 1310$ and 1558 is smaller than that for all the other distributions, including those of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014) at ${Re_\lambda }=124$, 379, 495 and 701 and Larssen & Devenport (Reference Larssen and Devenport2011) at ${Re_\lambda }=124$ and 1360. This corresponds to $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ being underestimated at ${Re_\lambda }=$ 1310 and 1558 in the range $r/\eta \approx 20\unicode{x2013}100$ which is most likely due to the measurement uncertainty. We next use the Kármán–Howarth equation (2.1) to further examine the data of Bodenschatz et al. (Reference Bodenschatz, Bewley, Nobach, Sinhuber and Xu2014). The distributions of $-\overline {{{(\delta u)}^3}} /\bar {\epsilon } r$ in figure 15 are calculated using (2.1) and (4.6) with $n=2$ at ${Re_\lambda }=117$, 352, 535, 679, 1323 and 1548. We emphasize that using (4.6) with $n=2$ is equivalent to using of the experimental data for $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ in figure 10; here we used the former since the calculation of the integral term in (2.1), i.e. $I_u(r)$ term, based on (4.6) with $n=2$ should be more accurate than that based on the discrete experimental data. Also shown in figure 15 are the grid turbulence data (symbols) at the same $Re_\lambda$, reproduced from figure 12(a). The agreement between curves and symbols in figure 15 is satisfactory. This agreement not only confirms the fit (i.e. (4.6) with $n=2$) used to emulate the available data for $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ (figure 10), but also vindicates the choice of $M$ for $\overline {{{(\delta u)}^2}} /u_K^2$ in figure 4. Finally, figure 15 further indicates that the magnitude of the model at ${Re_\lambda }= 1310$ and 1558 in the range $r/\eta \approx 20\unicode{x2013}100$ being larger than that of the data most likely reflects a measurement uncertainty in the experimental data (figure 10). Correspondingly, the magnitude of the energy spectrum distributions in the range $k_1\eta \approx 0.06\unicode{x2013}0.31$ at ${Re_\lambda }= 1310$ and 1558 in figure 1(a) being smaller than that of all other distributions also most likely reflects a measurement uncertainty. Further, there is no well-defined plateau for all the experimental data (figures 10, 11a14a), even at $Re_\lambda \sim 1500$, underlining the importance of the FRN effect in the context of establishing the inertial range. This effect can be also observed in the $LS_n$ distributions for all the experimental data; as an example, we have included in figure 13(b) the local slope $LS_4$ for the corresponding VDTT grid turbulence data of Kaminsky et al. (Reference Kaminsky, Birnir, Bewley and Sinhuber2020). In particular, $LS_3$ for $Re_\lambda =1558$ is far from exhibiting a plateau (figure 12b), implying that (1.2) is actually far from being satisfied by the VDTT grid turbulence data, which underlines the absence of an IR. Note that the calculated distributions indicate that a plateau is likely to emerge as $Re_\lambda$ increases to an infinitely large value. Both the VDTT grid turbulence data and $LS_n$ distributions based on (4.6) in figures 10, 11–14 show an evolution of ${\overline {{{(\delta u)}^n}} }$ and the corresponding local slopes with ${Re_\lambda }$ with a slow approach to $n/3$. Such a trend is not only consistent with all the analysis in the present paper, but also consistent with the plane jet data of Tang et al. (Reference Tang, Antonia, Djenidi, Danaila and Zhou2017). Finally, we should also stress that the local slopes $LS_n$ are not affected by the values of $\bar {\varepsilon }$. Obviously, they are not affected by the choice of the value of $M$ used in replotting the $\overline {(\delta u)^n}/u_K^n$ distributions.

Table 2. Values of $B_n$, $C_\epsilon$, $D_n$, $F_2$ and $F_3$ used in figures 10–21. Note that the magnitudes of $F_3$ are compared with data in grid turbulence and along the axis of the plane jet in figure 16.

Table 3. Values of $F_4$ and $F_6$ used in figures 13–17 and 20–21. Note that the magnitudes of $F_4$ and $F_6$ are compared with data in grid turbulence and along the axis of the plane jet in figure 16.

Figure 15. Analytical predictions of $-\overline {{{(\delta u)}^3}} /\bar {\epsilon } r$ (curves) based on (2.1) and (4.6) with $n=2$ at ${Re_\lambda }=117$, 352, 535, 679, 1323 and 1548. The details of how the integral term in (2.1), i.e. $I_u(r)$ term, was calculated are discussed elsewhere (see for example (4.9) of Tang et al. (Reference Tang, Antonia, Djenidi, Danaila and Zhou2017) where the power-law decay rate for $\overline {u^2}$, viz. $\overline {u^2}\sim x^{n}$ with $n\approx -1.2$ (Sinhuber et al. Reference Sinhuber, Bodenschatz and Bewley2015), was used here). Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 12(a). The dashed horizontal line in each plot corresponds to the value of 4/5.

Figures 11–14 indicate that $Re_\lambda$ may need to be of order $10^5$ before an inertial range is established unequivocally. To our knowledge, this should be the first time that plausible evidence has been presented for the approach to the IR based on values of $n$ extending to six. We next quantify, albeit approximately, the values of $Re_\lambda$ required for an inertial range to be established for $\overline {(\delta u)^n}$. Figure 17 shows the range of $r/\lambda$ over which the distributions, based on (4.6), of $\overline {{{(\delta {u})}^2}}/ {(\bar {\epsilon } r)^{ {2/3}}}$, $-\overline {{{(\delta {u})}^3}} /{\bar {\epsilon } r}$, $\overline {{{(\delta {u})}^4}} /{(\bar {\epsilon } r)^{ {4/3}}}$ and $\overline {{{(\delta {u})}^6}}/ {(\bar {\epsilon } r)^{ {6/3}}}$ depart from 2.11, 4/5, 18.16 and 413 (these values correspond to the plateaus at $Re_\lambda =10^5$ in figures 11(a)–14a), respectively, by no more than 2.5 %. It should be pointed out that another possibility would be to use $r=\lambda$ as the ‘base’ line for IR. This is consistent with Obligado & Vassilicos (Reference Obligado and Vassilicos2019) and Meldi & Vassilicos (Reference Meldi and Vassilicos2021) who showed, using the eddy damped quasinormal Markovian, that the interscale energy flux is closest to $\bar {\epsilon }$ when $r$ is close to $\lambda$ while the larger $r$, compared to $\lambda$, the larger the departure from equilibrium in which the interscale energy flux is equal to $\bar {\epsilon }$. This would result in slightly larger values of $Re_\lambda$ being required to establish the IR. We can observe from figure 17 that, when $n$ is even, the larger $n$ is being required, the slower is the rate at which $C_n = n/3$ is established convincingly. Interestingly, the $r/\lambda$ range for $\overline {{{(\delta {u})}^6}}/ {(\bar {\epsilon } r)^{ {6/3}}}$ is close to that for $-\overline {{{(\delta {u})}^3}} /{\bar {\epsilon } r}$. Finally, we can observe that $C_2 = 2/3$ is attained at a smaller $Re_\lambda$ than $C_3 = 1$. This is consistent with the observation of Antonia et al. (Reference Antonia, Tang, Djenidi and Zhou2019) where a similar model for $\overline {{{(\delta {u})}^2}}$ with a slightly different constant ($B_2=0.0061$) was used and $\overline {{{(\delta {u})}^3}} /{(\bar {\epsilon } r)}$ was calculated from the Kármán–Howarth equation.

Figure 16. The magnitudes (diamonds) of $|F_{3}|$, $F_{4}$ and $F_{6}$ used to plot figures 12–14. For reference, the data along the axis of the plane jet, reproduced from Antonia et al. (Reference Antonia, Djenidi, Danaila and Tang2017) (triangles), and in grid turbulence, reproduced from Gylfason et al. (Reference Gylfason, Ayyalasomayajula and Warhaft2004) (squares), are also shown; for the latter, only the laboratory shearless grid turbulence data are shown. The horizontal dashed lines indicate the mean values of each of the plotted quantities for the plane jet data.

Figure 17. The range of $r/\lambda$ over which $\overline {{{(\delta {u})}^2}} /{(\bar {\epsilon } r)^{{2/3}}}$ (blue line), $\overline {{{(\delta {u})}^3}} /{(\bar {\epsilon } r)}$ (red line), $\overline {{{(\delta {u})}^4}} /{(\bar {\epsilon } r)^{{4/3}}}$ (green line) and $\overline {{{(\delta {u})}^6}} /{(\bar {\epsilon } r)^{6/3}}$ (pink line) depart from 2.11, 4/5, 18.16 and 413 (these values correspond to the plateaus at $Re_\lambda =10^5$ in figures 11a14a), respectively, by no more than 2.5 %; they are calculated based on (4.6), using $C_n= n/3$ and the constants for $B_n$, $C_\epsilon /D_n$ and $F_n$ in table 2. The locations of the vertical arrows give an approximate indication of the values of $Re_\lambda$ needed to attain an IR, of two decades in extent, for $\overline {{{(\delta {u})}^2}} /{(\bar {\epsilon } r)^{{2/3}}}$ (blue arrow), $\overline {{{(\delta {u})}^3}} /{(\bar {\epsilon } r)}$ (red arrow), $\overline {{{(\delta {u})}^4}} /{(\bar {\epsilon } r)^{{4/3}}}$ (green arrow) and $\overline {{{(\delta {u})}^6}} /{(\bar {\epsilon } r)^{6/3}}$ (pink arrow).

Figure 18. Distributions of $\overline {(\delta u)^2}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 19. Distributions of $-\overline {(\delta u)^3}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 20. Distributions of $\overline {(\delta u)^4}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 21. Distributions of $\overline {(\delta u)^6}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

It is of interest to examine the dual scaling using (4.6) at large $Re_\lambda$. Figures 18–21 show the distributions of $\overline {(\delta u)^n}$ ($n=2,3,4,6$) normalized by ($u_K$, $\eta$) and ($u'$, $L$) in grid turbulence at $Re_\lambda = 1600\unicode{x2013}10^5$. In conformity with the observation based on figures 6–9, the scaling based on ($u'$, $L$) extends to increasingly smaller values of $r/L$ while the scaling based on ($u_K$, $\eta$) extends to increasingly larger values of $r/\eta$ as $Re_\lambda$ increases. In particular, at the two largest values of $Re_\lambda$, the upper value of $r/\eta$ exceeds $10^4$ for all $\overline {(\delta u)^n}$ whilst the lower value of $r/L$ lies below $5\times 10^{-4}$. Taking $Re_\lambda = 2\times 10^4$ and using the isotropic relation ${L}/{\eta } = C_\epsilon ({Re_\lambda ^{3/2}}/{{{{15}^{3/4}}}})$ with $C_\epsilon =0.8$, we can estimate that the overlap region is given by $10^2\lesssim r/\eta \lesssim 10^4$. Namely, a distinct range of overlap between the ($u'$, $L$) and ($u_K$, $\eta$) scalings can be observed at $Re_\lambda =2\times 10^4$, suggesting the onset of an inertial range for $\overline {(\delta u)^n}$ ($n=2,3,4,6$) (see also figure 17). These results further confirm that the dual scaling applies to $\overline {(\delta u)^n}$ and an overlap range begins to emerge when $Re_\lambda >10^4$.

6. Conclusions

A dual scaling analysis has been carried out in the context of the Kármán–Howarth equation, or transport equation for $\overline {(\delta u)^2}$, for grid turbulence. When the effect of the large-scale term is neglected, the scaling ($u_K$, $\eta$) should be effective, even at moderate $Re_\lambda$, since the two dimensionless parameters in (2.6a,b) are universal. Also, when the viscous term is neglected, the scaling ($u'$, $L$) should be also tenable since both $\bar {\varepsilon } {L}/u'^3$ and ${({1}/{{{u'}}})({{{\rm d}{L}}}/{{{\rm d}\kern0.06em x}})}$ should approach constant values as $Re_\lambda$ increases. At finite $Re_\lambda$, the dual scaling approach is consistent with incomplete similarity. As $Re_\lambda \to \infty$, it becomes compatible with complete similarity. The dual scaling approach is subsequently tested in the context of $\overline {(\delta u)^n}$ ($n \geq 2$) using the VDTT grid turbulence data at $Re_\lambda =110\unicode{x2013}1558$ (figures 6–9). It is found that the scaling based on ($u'$, $L$) extends to increasingly smaller values of $r/L$ as $Re_\lambda$ increases. Further, the scaling based on ($u_K$, $\eta$) extends to increasingly larger values of $r/\eta$ as $Re_\lambda$ increases. These observations are consistent with the dual scaling analysis in the context of the Kármán–Howarth equation. More importantly, the inference from these observations is that both scalings should eventually overlap, as $Re_\lambda$ increases further, and conform with the power-law relation $\overline {{(\delta u)}^n} \sim r^{n/3}$ as $Re_\lambda \to \infty$. The extent of the overlap should become significant as $Re_\lambda \to \infty$ since the KH equation has been shown (George Reference George1992; Speziale & Bernard Reference Speziale and Bernard1992; Djenidi & Antonia Reference Djenidi and Antonia2015; Djenidi et al. Reference Djenidi, Antonia and Tang2019) to admit a similarity solution based on a single set of scales; in this case, ($u'$, $L$) can be interchanged with ($u_K$, $\eta$).

An empirical model for $\overline {{(\delta u)}^n}$ ($n=2,3,4,6$), i.e. (4.6), has been compared with the VDTT grid turbulence data (figures 10, 11a14a). The model, which is consistent with $\overline {{(\delta u)}^n} \sim r^{n/3}$ as $Re_\lambda \to \infty$, is in reasonable accord with the data for values of $Re_\lambda$ up to approximately $1500$, thus allowing extrapolation of the model-based results to larger values of $Re_\lambda$. The major conclusions with respect to the comparison and extrapolation can be summarized as follows.

  1. (i) There is strictly no power-law range for the experimental data of $\overline {{(\delta u)}^n}$, even at $Re_\lambda \sim 1500$. It is evident that for all the values of $n$ considered here, the local slope $LS_n$ continues to evolve with $Re_\lambda$ (figures 11b14b) and begins to exhibit a small plateau only when $Re_\lambda$ exceeds $10^4$. An important inference from this trend is that the VDTT data, like the spectral data of Mydlarski & Warhaft (Reference Mydlarski and Warhaft1996) and subsequent analysis by Gamard & George (Reference Gamard and George2000), are consistent with (1.3) and do not support (1.4) when $\zeta _n \ne n/3$.

  2. (ii) The FRN effect may differ between even- and odd-order moments of $\delta u$. Different values of $Re_\lambda$ may therefore be required, between even- and odd-order moments of $\delta u$, for compliance with $\zeta _n = n/3$. This effect, which, as noted in item (i), is evident in the VDTT data, has, by and large, been either ignored or underestimated in the literature. The approach to $\zeta _n = n/3$ is consistent with the inequality constraint discussed by Djenidi et al. (Reference Djenidi, Antonia and Tang2023) and with the dual scaling approach outlined in this paper.

  3. (iii) The dual scaling can be used to adequately describe $\overline {{(\delta u)}^n}$. A larger $Re_\lambda$ results in wider $r/\eta$ and $r/L$ ranges over which the scalings ($u_K$, $\eta$) and ($u'$, $L$) are tenable. We should stress that the empirical model for $\overline {{(\delta u)}^n}$, i.e. (4.6), does not provide a perfect description of the data. However, it adequately describes the dependence, irrespectively of $n$, on $Re_\lambda$. Furthermore, it fully complies with the dual scaling methodology and satisfies the limiting behaviour of $\overline {{(\delta u)}^n}$ as $r\to \eta$ and $r\to L$. We should also point out that (4.6) reflects the strongly non-Gaussian character of the small scale turbulence when $r\to 0$ and they transition to quasi Gaussian properties when $r$ exceeds $L$. The compliance of (4.6) with $\overline {{(\delta u)}^n} \sim r^{n/3}$ when $Re_\lambda \to \infty$ is consistent with the observation by Qian (Reference Qian2000) and Qian (Reference Qian2001) who used a non-Gaussian p.d.f. of $\delta u$ with stretched exponential tails together with a ‘quasi-closure’ scheme to show that $\overline {{(\delta u)}^n} \sim r^{n/3}$ as $Re_\lambda \to \infty$.

Funding

This research was supported by the National Natural Science Foundation of China (project no. 91952109), Guangdong Basic and Applied Basic Research Foundation (project no. 2023B1515020069) and Shenzhen Science and Technology Program (project nos. RCYX20210706092046085 and GXWD20220817171516009).

Declaration of interests

The authors report no conflict of interest.

References

Antonia, R.A., Tang, S.L., Djenidi, L. & Zhou, Y. 2019 Finite Reynolds number effect and the 4/5 law. Phys. Rev. Fluids 4 (8), 084602.10.1103/PhysRevFluids.4.084602CrossRefGoogle Scholar
Antonia, R.A. & Burattini, P. 2006 Approach to the 4/5 law in homogeneous isotropic turbulence. J. Fluid Mech. 550, 175184.10.1017/S0022112005008438CrossRefGoogle Scholar
Antonia, R.A., Djenidi, L. & Danaila, L. 2014 Collapse of the turbulent dissipation range on Kolmogorov scales. Phys. Fluids 26, 045105.10.1063/1.4869305CrossRefGoogle Scholar
Antonia, R.A., Djenidi, L., Danaila, L. & Tang, S.L. 2017 Small scale turbulence and the finite Reynolds number effect. Phys. Fluids 29 (2), 020715.10.1063/1.4974323CrossRefGoogle Scholar
Antonia, R.A., Satyaprakash, B.R. & Hussain, A.K.M.F. 1980 Measurements of dissipation rate and some other characteristics of turbulent plane and circular jets. Phys. Fluids 23, 695700.10.1063/1.863055CrossRefGoogle Scholar
Antonia, R.A., Smalley, R.J., Zhou, T., Anselmet, F. & Danaila, L. 2003 Similarity of energy structure functions in decaying homogeneous isotropic turbulence. J. Fluid Mech. 487, 245269.10.1017/S0022112003004713CrossRefGoogle Scholar
Antonia, R.A., Zhou, T., Danaila, L. & Anselmet, F. 2000 Streamwise inhomogeneity of decaying grid turbulence. Phys. Fluids 12, 3086.10.1063/1.1314336CrossRefGoogle Scholar
Antonia, R.A., Zhou, T. & Zhu, Y. 1998 Three-component vorticity measurements in a turbulent grid flow. J. Fluid Mech. 374, 2957.10.1017/S0022112098002547CrossRefGoogle Scholar
Barenblatt, G.I. & Chorin, A.J. 1998 New perspectives in turbulence: scaling laws, asymptotics, and intermittency. SIAM Rev. 40 (2), 265291.10.1137/S0036144597320047CrossRefGoogle Scholar
Barenblatt, G.I. & Goldenfeld, N. 1995 Does fully developed turbulence exist? Reynolds number independence versus asymptotic covariance. Phys. Fluids 7, 30783082.10.1063/1.868685CrossRefGoogle Scholar
Benzi, R., Ciliberto, S., Tripiccione, R., Baudet, C., Massaioli, F. & Succi, S. 1993 Extended self-similarity in turbulent flows. Phys. Rev. E 48, 2932.10.1103/PhysRevE.48.R29CrossRefGoogle ScholarPubMed
Birnir, B. 2019 The turbulence problem. Nonlinear Stud. 26, 767781.Google Scholar
Bodenschatz, E., Bewley, G.P., Nobach, H., Sinhuber, M. & Xu, H. 2014 Variable density turbulence tunnel facility. Rev. Sci. Instrum. 85 (9), 093908.10.1063/1.4896138CrossRefGoogle ScholarPubMed
Burattini, P., Antonia, R.A. & Danaila, L. 2005 a Similarity in the far field of a turbulent round jet. Phys. Fluids 17, 025101.10.1063/1.1833414CrossRefGoogle Scholar
Burattini, P., Lavoie, P. & Antonia, R. 2005 b On the normalized turbulent energy dissipation rate. Phys. Fluids 17, 98103.10.1063/1.2055529CrossRefGoogle Scholar
Cao, N., Chen, S. & Doolen, G.D. 1999 Statistics and structures of pressure in isotropic turbulence. Phys. Fluids 11, 22352250.10.1063/1.870085CrossRefGoogle Scholar
Comte-Bellot, G. & Corrsin, S. 1966 The use of a contraction to improve the isotropy of grid-generated turbulence. J. Fluid Mech. 25, 657682.10.1017/S0022112066000338CrossRefGoogle Scholar
Comte-Bellot, G. & Corrsin, S. 1971 Simple Eulerian time correlation of full-and narrow-band velocity signals in grid-generated, ‘isotropic’ turbulence. J. Fluid Mech. 48 (2), 273337.10.1017/S0022112071001599CrossRefGoogle Scholar
Danaila, L., Anselmet, F. & Antonia, R.A. 2002 An overview of the effect of large-scale inhomogeneities on small-scale turbulence. Phys. Fluids 14, 24752482.10.1063/1.1476300CrossRefGoogle Scholar
Danaila, L., Anselmet, F., Zhou, T. & Antonia, R.A. 1999 A generalization of Yaglom's equation which accounts for the large-scale forcing in heated decaying turbulence. J. Fluid Mech. 391, 359372.10.1017/S0022112099005418CrossRefGoogle Scholar
Dhruva, B. 2000 An Experimental Study of High Reynolds Number Turbulence in the Atmosphere. PhD thesis, Yale University.Google Scholar
Dickey, T.D. & Mellor, G.L. 1979 The Kolmogoroff $r^ 2/3$ law. Phys. Fluids 22, 10291032.10.1063/1.862707CrossRefGoogle Scholar
Djenidi, L., Antonia, R.A. & Tang, S.L. 2019 Scale invariance in finite Reynolds number homogeneous isotropic turbulence. J. Fluid Mech. 864, 244272.10.1017/jfm.2019.28CrossRefGoogle Scholar
Djenidi, L. & Antonia, R.A. 2012 A spectral chart method for estimating the mean turbulent kinetic energy dissipation rate. Exp. Fluids 53, 10051013.10.1007/s00348-012-1337-xCrossRefGoogle Scholar
Djenidi, L. & Antonia, R.A. 2015 A general self-preservation analysis for decaying homogeneous isotropic turbulence. J. Fluid Mech. 773, 345365.10.1017/jfm.2015.250CrossRefGoogle Scholar
Djenidi, L., Antonia, R.A. & Tang, S.L. 2022 Scaling of velocity structure functions in isotropic turbulence. In Proceedings 23rd Australasian Fluid Mechanics Conference, Sydney, Australia.Google Scholar
Djenidi, L., Antonia, R.A. & Tang, S.L. 2023 Scaling of turbulent velocity structure functions: plausibility constraints. J. Fluid Mech. 965, A14.10.1017/jfm.2023.416CrossRefGoogle Scholar
Donzis, D.A., Sreenivasan, K.R. & Yeung, P.K. 2005 Scalar dissipation rate and dissipative anomaly in isotropic turbulence. J. Fluid Mech. 532, 199216.10.1017/S0022112005004039CrossRefGoogle Scholar
Gamard, S. & George, W.K. 2000 Reynolds number dependence of energy spectra in the overlap region of isotropic turbulence. Flow Turbul. Combust. 63, 443477.10.1023/A:1009988321057CrossRefGoogle Scholar
George, W.K. 1989 The self-preservation of turbulent flows and its relation to initial conditions and coherent structures. In Advances in Turbulence (ed. W.K. George & R. Arndt), pp. 39–73. Hemisphere.Google Scholar
George, W.K. 1992 The decay of homogeneous isotropic turbulence. Phys. Fluids 4, 14921509.10.1063/1.858423CrossRefGoogle Scholar
George, W.K. 1994 Some new ideas for similarity of turbulent shear flows. In Proceedings of Turbulence, Heat and Mass Transfer Symposium Lisbon, Portugal (ed. K. Hanjalic & J. Pereira), pp. 24–49. Begell House.Google Scholar
Gotoh, T., Fukayama, D. & Nakano, T. 2002 Velocity field statistics in homogeneous steady turbulence obtained using a high-resolution direct numerical simulation. Phys. Fluids 14, 10651081.10.1063/1.1448296CrossRefGoogle Scholar
Gylfason, A., Ayyalasomayajula, S. & Warhaft, Z. 2004 Intermittency, pressure and acceleration statistics from hot-wire measurements in wind-tunnel turbulence. J. Fluid Mech. 501, 213229.10.1017/S002211200300747XCrossRefGoogle Scholar
He, G.W., Jin, G.D. & Yang, Y. 2017 Space–time correlations and dynamic coupling in turbulent flows. Annu. Rev. Fluid Mech. 49, 5170.10.1146/annurev-fluid-010816-060309CrossRefGoogle Scholar
Hill, R.J. 2001 Equations relating structure functions of all orders. J. Fluid Mech. 434, 379388.10.1017/S0022112001003949CrossRefGoogle Scholar
Ishihara, T., Gotoh, T. & Kaneda, Y. 2009 Study of high-Reynolds number isotropic turbulence by direct numerical simulation. Annu. Rev. Fluid Mech. 41, 165–80.10.1146/annurev.fluid.010908.165203CrossRefGoogle Scholar
Ishihara, T. & Kaneda, Y. 2002 High resolution DNS of incompressible homogeneous forced turbulence: time dependence of the statistics. In Proceedings of the International Workshop on Statistical Theories and Computational Approaches to Turbulence (ed. Y. Kaneda & T. Gotoh), pp. 177–188. Springer.10.1007/978-4-431-67002-5_11CrossRefGoogle Scholar
Iyer, K.P., Sreenivasan, K.R. & Yeung, P.K. 2020 Scaling exponents saturate in three-dimensional isotropic turbulence. Phys. Rev. Fluids 5 (5), 054605.10.1103/PhysRevFluids.5.054605CrossRefGoogle Scholar
Jimenez, J., Wray, A.A., Saffman, P.G. & Rogallo, R.S. 1993 The structure of intense vorticity in isotropic turbulence. J. Fluid Mech. 255, 6590.10.1017/S0022112093002393CrossRefGoogle Scholar
Kaminsky, J., Birnir, B., Bewley, G.P. & Sinhuber, M. 2020 Reynolds number dependence of the structure functions in homogeneous turbulence. J. Nonlinear Sci. 30 (3), 10811114.10.1007/s00332-019-09602-yCrossRefGoogle Scholar
Von Kármán, T. & Howarth, L. 1938 On the statistical theory of isotropic turbulence. Proc. R. Soc. Lond. A 164, 192215.10.1098/rspa.1938.0013CrossRefGoogle Scholar
Kolmogorov, A.N. 1941 a The local structure of turbulence in incompressible viscous fluid for very large Reynolds number. Dokl. Akad. Nauk SSSR 30, 299–303 (see also Proc. R. Soc. Lond. A 434, 9–13).Google Scholar
Kolmogorov, A.N. 1941 b Dissipation of energy in the locally isotropic turbulence. Dokl. Akad. Nauk SSSR 32, 19–21 (see also Proc. R. Soc. Lond. A (1991), 434, 15–17).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, 8285.10.1017/S0022112062000518CrossRefGoogle Scholar
Kurien, S. & Sreenivasan, K.R. 2000 Anisotropic scaling contributions to high-order structure functions in high-Reynolds-number turbulence. Phys. Rev. E 62, 22062212.10.1103/PhysRevE.62.2206CrossRefGoogle ScholarPubMed
Landau, L.D. & Lifshitz, E.M. 1987 Fluid Mechanics: Vol 6 of course of theoretical physics, Second English Edition, Revised. Pergamon.Google Scholar
Larssen, J.V. & Devenport, W.J. 2011 On the generation of large-scale homogeneous turbulence. Exp. Fluids 50, 12071223.10.1007/s00348-010-0974-1CrossRefGoogle Scholar
Lavoie, P., Djenidi, L. & Antonia, R.A. 2007 Effects of initial conditions in decaying turbulence generated by passive grids. J. Fluid Mech. 585, 395420.10.1017/S0022112007006763CrossRefGoogle Scholar
Lee, S.K., Djenidi, L., Antonia, R.A. & Danaila, L. 2013 On the destruction coefficients for slightly heated decaying grid turbulence. Intl J. Heat Fluid Flow 43, 129136.10.1016/j.ijheatfluidflow.2013.03.009CrossRefGoogle Scholar
Lundgren, T.S. 2002 Kolmogorov two-thirds law by matched asymptotic expansion. Phys. Fluids 14 (2), 638642.10.1063/1.1429965CrossRefGoogle Scholar
Lundgren, T.S. 2003 Kolmogorov turbulence by matched asymptotic expansions. Phys. Fluids 15 (4), 10741081.10.1063/1.1558332CrossRefGoogle Scholar
Malecot, Y. 1998 Intermittence en turbulence 3D: statistiques de la vitesse et de la vorticité. PhD thesis, University of Grenoble.Google Scholar
McComb, W.D., Yoffe, S.R., Linkmann, M.F. & Berera, A. 2014 Spectral analysis of structure functions and their scaling exponents in forced isotropic turbulence. Phys. Rev. E 90 (5), 053010.10.1103/PhysRevE.90.053010CrossRefGoogle ScholarPubMed
McComb, W.D., Berera, A., Yoffe, S.R. & Linkmann, M.F. 2015 Energy transfer and dissipation in forced isotropic turbulence. Phys. Rev. E 91, 043013.10.1103/PhysRevE.91.043013CrossRefGoogle ScholarPubMed
McKeon, B.J. & Morrison, J.F. 2007 Asymptotic scaling in turbulent pipe flow. Phil. Trans. R. Soc. A 365 (1852), 771787.10.1098/rsta.2006.1945CrossRefGoogle ScholarPubMed
Meldi, M., Djenidi, L. & Antonia, R.A. 2021 Sensitivity analysis of the second and third-order velocity structure functions to the Reynolds number in decaying and forced isotropic turbulence using the EDQNM model. Eur. J. Mech. (B/Fluids) 88, 229242.10.1016/j.euromechflu.2021.04.003CrossRefGoogle Scholar
Meldi, M. & Vassilicos, J.C. 2021 Analysis of Lundgren's matched asymptotic expansion approach to the Kármán-Howarth equation using the eddy damped quasi-normal Markovian turbulence closure. Phys. Rev. Fluids 6, 064602.10.1103/PhysRevFluids.6.064602CrossRefGoogle Scholar
Mi, J., Xu, M. & Zhou, T. 2013 Reynolds number influence on statistical behaviors of turbulence in a circular free jet. Phys. Fluids 25, 075101.10.1063/1.4811403CrossRefGoogle Scholar
Millikan, C.B. 1939 A critical discussion of turbulent flows in channels and tubes. In Proceedings of the 5th International Congress on Applied Mechanics (ed. J.P. Den Hartog & H. Peters), pp. 386–392. Wiley.Google Scholar
Monin, A.S. & Yaglom, A.M. 2007 Statistical Fluid Dynamics, vol. 2. MIT.Google Scholar
Mydlarski, L. & Warhaft, Z. 1996 On the onset of high-Reynolds-number grid-generated wind tunnel turbulence. J. Fluid Mech. 320, 331368.10.1017/S0022112096007562CrossRefGoogle Scholar
Mydlarski, L. & Warhaft, Z. 1998 Passive scalar statistics in high-Péclet-number grid turbulence. J. Fluid Mech. 358, 135175.10.1017/S0022112097008161CrossRefGoogle Scholar
Obligado, M. & Vassilicos, J.C. 2019 The non-equilibrium part of the inertial range in decaying homogeneous turbulence. Europhys. Lett. 127, 64004.10.1209/0295-5075/127/64004CrossRefGoogle Scholar
Oboukhov, A.M. 1962 Some specific features of atmospheric turbulence. J. Fluid Mech. 13 (1), 7781.10.1017/S0022112062000506CrossRefGoogle Scholar
Obukhov, A.M. 1941 On the distribution of energy in the spectrum of turbulent flow. Dokl. Akad. Nauk SSSR 32, 2224.Google Scholar
Pearson, B.R. & Antonia, R.A. 2001 Reynolds-number dependence of turbulent velocity and pressure increments. J. Fluid Mech. 444, 343382.10.1017/S0022112001005511CrossRefGoogle Scholar
Pope, S.B. 2000 Turbulent Flows. Cambridge University Press.10.1017/CBO9780511840531CrossRefGoogle Scholar
Qian, J. 1997 Inertial range and the finite Reynolds number effect of turbulence. Phys. Rev. E 55, 337342.10.1103/PhysRevE.55.337CrossRefGoogle Scholar
Qian, J. 1999 Slow decay of the finite Reynolds number effect of turbulence. Phys. Rev. E 60, 3409.10.1103/PhysRevE.60.3409CrossRefGoogle ScholarPubMed
Qian, J. 2000 Closure approach to high-order structure functions of turbulence. Phys. Rev. Lett. 84, 646649.10.1103/PhysRevLett.84.646CrossRefGoogle ScholarPubMed
Qian, J. 2001 Quasi-closure and scaling of turbulence. Intl J. Mod. Phys. B 15, 10851116.10.1142/S0217979201004514CrossRefGoogle Scholar
Saddoughi, S.G. & Veeravalli, S.V. 1994 Local isotropy of turbulent boundary layers at high Reynolds number. J. Fluid Mech. 268, 333372.10.1017/S0022112094001370CrossRefGoogle Scholar
Saffman, P.G. 1967 Note on the decay of homogeneous turbulence. Phys. Fluids 10, 13491352.10.1063/1.1762284CrossRefGoogle Scholar
Sagaut, P. & Cambon, C. 2018 Homogeneous Turbulence Dynamics. Springer.10.1007/978-3-319-73162-9CrossRefGoogle Scholar
She, Z.S. & Leveque, E. 1994 Universal scaling laws in fully developed turbulence. Phys. Rev. Lett. 72 (3), 336.10.1103/PhysRevLett.72.336CrossRefGoogle ScholarPubMed
Sinhuber, M., Bewley, G.P. & Bodenschatz, E. 2017 Dissipative effects on inertial-range statistics at high Reynolds numbers. Phys. Rev. Lett. 119 (13), 134502.10.1103/PhysRevLett.119.134502CrossRefGoogle ScholarPubMed
Sinhuber, M., Bodenschatz, E. & Bewley, G.P. 2015 Decay of turbulence at high Reynolds numbers. Phys. Rev. Lett. 114 (3), 034501.10.1103/PhysRevLett.114.034501CrossRefGoogle ScholarPubMed
Speziale, C.G. & Bernard, P.S. 1992 The energy decay in self-preserving isotropic turbulence revisited. J. Fluid Mech. 241, 645667.10.1017/S0022112092002180CrossRefGoogle Scholar
Sreenivasan, K. & Antonia, R.A. 1997 The phenomenology of small-scale turbulence. Annu. Rev. Fluid Mech. 29, 435472.10.1146/annurev.fluid.29.1.435CrossRefGoogle Scholar
Sreenivasan, K.R. 1984 On the scaling of the turbulence energy dissipation rate. Phys. Fluids 27, 10481051.10.1063/1.864731CrossRefGoogle Scholar
Stolovitzky, G., Sreenivasan, K.R. & Juneja, A. 1993 Scaling functions and scaling exponents in turbulence. Phys. Rev. E 48 (5), R3217R3220.10.1103/PhysRevE.48.R3217CrossRefGoogle ScholarPubMed
Tang, S.L., Antonia, R.A., Djenidi, L., Danaila, L. & Zhou, Y. 2017 Finite Reynolds number effect on the scaling range behavior of turbulent longitudinal velocity structure functions. J. Fluid Mech. 820, 341369.10.1017/jfm.2017.218CrossRefGoogle Scholar
Tang, S.L., Antonia, R.A., Djenidi, L. & Zhou, Y. 2020 Scaling of the turbulent energy dissipation correlation function. J. Fluid Mech. 891, A26.10.1017/jfm.2020.171CrossRefGoogle Scholar
Tchoufag, J., Sagaut, P. & Cambon, C. 2012 Spectral approach to finite Reynolds number effects on Kolmogorov's 4/5 law in isotropic turbulence. Phys. Fluids 24 (1), 015107.10.1063/1.3678334CrossRefGoogle Scholar
Townsend, A.A. 1976 The Structure of Turbulent Shear Flow. Cambridge University Press.Google Scholar
Tsuji, Y. 2009 High-Reynolds-number experiments: the challenge of understanding universality in turbulence. Fluid Dyn. Res. 41, 064003.10.1088/0169-5983/41/6/064003CrossRefGoogle Scholar
Vassilicos, J.C. 2015 Dissipation in turbulent flows. Annu. Rev. Fluid Mech. 47, 95114.10.1146/annurev-fluid-010814-014637CrossRefGoogle Scholar
Wang, L.-P., Chen, S., Brasseur, J.G. & Wyngaard, J.C. 1996 Examination of hypotheses in Kolmogorov refined turbulence theory through high-resolution simulations. Part 1. Velocity field. J. Fluid Mech. 309, 113156.10.1017/S0022112096001589CrossRefGoogle Scholar
Yeung, P.K. & Brasseur, J.G. 1991 The response of isotropic turbulence to isotropic and anisotropic forcing at the large scales. Phys. Fluids 3, 884897.10.1063/1.857966CrossRefGoogle Scholar
Yeung, P.K. & Zhou, Y. 1997 Universality of the Kolmogorov constant in numerical simulations of turbulence. Phys. Rev. E 56, 17461752.10.1103/PhysRevE.56.1746CrossRefGoogle Scholar
Zhou, T. & Antonia, R.A. 2000 Reynolds number dependence of the small-scale structure of grid turbulence. J. Fluid Mech. 406, 81107.10.1017/S0022112099007296CrossRefGoogle Scholar
Figure 0

Figure 1. (a) Kolmogorov normalized compensated spectra of $u$ in grid turbulence (Bodenschatz et al.2014) ($Re_\lambda=$ 110–1450), normalized by $\bar {\epsilon }_{spec}$ and $\nu$ (the spectral chart method). For comparison, we have included the grid turbulence data of Larssen & Devenport (2011) (dashed curves) at $Re_\lambda = 124$ and 1360. (b) Corresponding grid turbulence data of Bodenschatz et al. (2014) ($Re_\lambda = 110\unicode{x2013}1450$), normalized by $\bar {\epsilon }$, estimated from $\overline {(\delta u)^3}$ (see text), and $\nu$; note that to see the dissipative range more clearly, a smaller range $k_1\eta =0.01\unicode{x2013}1$ is used for the vertical axis.

Figure 1

Figure 2. Distributions of $C_\epsilon = \bar {\epsilon } L /u'^3$ versus ${Re}_{\lambda }$ in grid turbulence. ( red), (Kaminsky et al.2020); ( blue), re-estimated $C_\epsilon$, based on $\bar {\epsilon }_{spec}$ (the spectral chart method), versus $Re_\lambda$ for the data of Bodenschatz et al. (2014) (see text). For comparison, we have included the grid turbulence data of Comte-Bellot & Corrsin (1971) () and Larssen & Devenport (2011) (). Black curve, (3.1) with $A=0.4$ and $B=60$. Also included are the data in stationary forced periodic box turbulence (($\bigcirc$ black), Wang et al. (1996); ($\bigcirc$ blue), Ishihara & Kaneda (2002); (${\ast }$ black), Jimenez et al. (1993); ( green), Ishihara et al. (2009); ($+$ black), Cao, Chen & Doolen (1999); ( blue), Yeung & Zhou (1997); ( blue), Gotoh et al.2002). Green curve, (3.1) with $A=0.2$ and $B=92$.

Figure 2

Figure 3. Distributions of $\overline {(\delta u)^2}/u_K^2$ versus $r/\eta$ in grid turbulence for the data of Bodenschatz et al. (2014) at ${Re_\lambda }=124$, 379, 495, 701, 1310 and 1558 (green curves). For comparison, other grid turbulence data are also shown; red curves, Zhou & Antonia (2000) (${Re_\lambda }=27$, 51, 62, 75, 83, 89 and 100); cyan curves, Comte-Bellot & Corrsin (1971) (${Re_\lambda }=61$ and 65); black curves, Larssen & Devenport (2011) (${Re_\lambda }=124$ and 1360); blue curve, Malecot (1998) (${Re_\lambda }=144$); pink curve, $(r/\eta )^2/15$. The arrow indicates the direction of increasing ${Re_\lambda }$. Note that the data of Comte-Bellot & Corrsin (1971), Larssen & Devenport (2011) and Bodenschatz et al. (2014) are calculated from their spectra, based on (3.2).

Figure 3

Figure 4. (a) Distributions of $\overline {(\delta u)^2}/u_K^2$ versus $r/\eta$ in grid turbulence for the data of Kaminsky et al. (2020) at ${Re_\lambda }=110$, 264, 508, 1000 and 1450 (green curves). Panel (b) corresponds to the same data, after applying a numerical factor $M$ to their data (see text). For comparison, other grid turbulence data, same as in figure 3, are also shown in (a,b). Pink curve, $(r/\eta )^2/15$. The arrow indicates the direction of increasing ${Re_\lambda }$.

Figure 4

Figure 5. (a) Distributions of $\overline {(\delta u)^4}/u_K^4$ versus $r/\eta$ in grid turbulence (Kaminsky et al.2020) after applying a numerical factor $M(=29$ and 2.96 for $Re_\lambda=110$ and 508) to their data (see text). For comparison, the grid turbulence data of Gylfason et al. (2004) (without shear) at $Re_\lambda=168$ (pink curve) and 656 (cyan curve) are shown; they are reproduced from figure 5(c,d) of Gylfason et al. (2004). (b) Distributions of $\overline {(\delta u)^4}/u_K^4$ versus $r/\eta$ in grid turbulence (Kaminsky et al.2020) at $Re_\lambda =264$ after applying a numerical factor $M(=6.9$ and 10) to their data (see text). The dotted curve, $15^{-2}F_4(r/\eta )^4$ with $F_4=9.6$. Other curves are the same as (a).

Figure 5

Table 1. Values of $M$ used to plot $\overline {(\delta u)^2}/u_K^2$, $\overline {(\delta u)^4}/u_K^4$ and $\overline {(\delta u)^6}/u_K^6$ versus $r/\eta$ at each $Re_\lambda$ for the data of Kaminsky et al. (2020) (see text). Note that a value of 10, instead of 6.9, is used to plot $\overline {(\delta u)^4}/u_K^4$ at $Re_\lambda =264$ since its distribution should lie between the distributions of $Re_\lambda =110$ and 508 in the dissipative range (see figure 5).

Figure 6

Figure 6. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^2}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (2020).

Figure 7

Figure 7. Effect of $Re_\lambda$ on the distributions of $-\overline {(\delta u)^3}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Bodenschatz et al. (2014); note that for (a), $u_K$ and $\eta$ were re-evaluated as explained in the text.

Figure 8

Figure 8. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^4}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (2020).

Figure 9

Figure 9. Effect of $Re_\lambda$ on the distributions of $\overline {(\delta u)^6}$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively. These data are reproduced from Kaminsky et al. (2020).

Figure 10

Figure 10. Distributions of $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ based on (4.6) at ${Re_\lambda }=124$, 379, 495, 701, 1310 and 1558. Symbols with the same colour correspond to the data of Bodenschatz et al. (2014) shown in figure 3.

Figure 11

Figure 11. (a) Distributions of $\overline {{{(\delta u)}^2}} /(\bar {\epsilon } r)^{2/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 6(a). (b) Local slope ($LS_2= {\rm d}\log (\overline {{{(\delta {u})}^2}}) /{\rm d}\log r$). Dashed horizontal line, $2/3$.

Figure 12

Figure 12. (a) Distributions of $-\overline {{{(\delta u)}^3}} /\bar {\epsilon } r$ based on (4.6) at ${Re_\lambda }=117$, 352, 535, 679, 1323, 1548, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 7(a). (b) Local slope ($LS_3= {\rm d}\log (-\overline {{{(\delta {u})}^3}}) /{\rm d}\log r$). Dashed horizontal line, $1$.

Figure 13

Figure 13. (a) Distributions of $\overline {{{(\delta u)}^4}} /(\bar {\epsilon } r)^{4/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 8(a). (b) Local slope ($LS_4= {\rm d}\log (\overline {{{(\delta {u})}^4}}) /{\rm d}\log r$) corresponding to the distributions in (a). Dashed horizontal line, $4/3$.

Figure 14

Figure 14. (a) Distributions of $\overline {{{(\delta u)}^6}} /(\bar {\epsilon } r)^{6/3}$ based on (4.6) at ${Re_\lambda }=110$, 264, 508, 1000, 1450, $10^4$ and $10^5$. Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 9(a). (b) Local slope ($LS_6= {\rm d}\log (\overline {{{(\delta {u})}^6}}) /{\rm d}\log r$). Dashed horizontal line, $6/3$.

Figure 15

Table 2. Values of $B_n$, $C_\epsilon$, $D_n$, $F_2$ and $F_3$ used in figures 10–21. Note that the magnitudes of $F_3$ are compared with data in grid turbulence and along the axis of the plane jet in figure 16.

Figure 16

Table 3. Values of $F_4$ and $F_6$ used in figures 13–17 and 20–21. Note that the magnitudes of $F_4$ and $F_6$ are compared with data in grid turbulence and along the axis of the plane jet in figure 16.

Figure 17

Figure 15. Analytical predictions of $-\overline {{{(\delta u)}^3}} /\bar {\epsilon } r$ (curves) based on (2.1) and (4.6) with $n=2$ at ${Re_\lambda }=117$, 352, 535, 679, 1323 and 1548. The details of how the integral term in (2.1), i.e. $I_u(r)$ term, was calculated are discussed elsewhere (see for example (4.9) of Tang et al. (2017) where the power-law decay rate for $\overline {u^2}$, viz.$\overline {u^2}\sim x^{n}$ with $n\approx -1.2$ (Sinhuber et al.2015), was used here). Symbols with the same colour are the corresponding grid turbulence data, reproduced from figure 12(a). The dashed horizontal line in each plot corresponds to the value of 4/5.

Figure 18

Figure 16. The magnitudes (diamonds) of $|F_{3}|$, $F_{4}$ and $F_{6}$ used to plot figures 12–14. For reference, the data along the axis of the plane jet, reproduced from Antonia et al. (2017) (triangles), and in grid turbulence, reproduced from Gylfason et al. (2004) (squares), are also shown; for the latter, only the laboratory shearless grid turbulence data are shown. The horizontal dashed lines indicate the mean values of each of the plotted quantities for the plane jet data.

Figure 19

Figure 17. The range of $r/\lambda$ over which $\overline {{{(\delta {u})}^2}} /{(\bar {\epsilon } r)^{{2/3}}}$ (blue line), $\overline {{{(\delta {u})}^3}} /{(\bar {\epsilon } r)}$ (red line), $\overline {{{(\delta {u})}^4}} /{(\bar {\epsilon } r)^{{4/3}}}$ (green line) and $\overline {{{(\delta {u})}^6}} /{(\bar {\epsilon } r)^{6/3}}$ (pink line) depart from 2.11, 4/5, 18.16 and 413 (these values correspond to the plateaus at $Re_\lambda =10^5$ in figures 11a–14a), respectively, by no more than 2.5 %; they are calculated based on (4.6), using $C_n= n/3$ and the constants for $B_n$, $C_\epsilon /D_n$ and $F_n$ in table 2. The locations of the vertical arrows give an approximate indication of the values of $Re_\lambda$ needed to attain an IR, of two decades in extent, for $\overline {{{(\delta {u})}^2}} /{(\bar {\epsilon } r)^{{2/3}}}$ (blue arrow), $\overline {{{(\delta {u})}^3}} /{(\bar {\epsilon } r)}$ (red arrow), $\overline {{{(\delta {u})}^4}} /{(\bar {\epsilon } r)^{{4/3}}}$ (green arrow) and $\overline {{{(\delta {u})}^6}} /{(\bar {\epsilon } r)^{6/3}}$ (pink arrow).

Figure 20

Figure 18. Distributions of $\overline {(\delta u)^2}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 21

Figure 19. Distributions of $-\overline {(\delta u)^3}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 22

Figure 20. Distributions of $\overline {(\delta u)^4}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.

Figure 23

Figure 21. Distributions of $\overline {(\delta u)^6}$ at large $Re_\lambda$ in grid turbulence. Panels (a,b) are normalized by ($u_K$, $\eta$) and ($u'$, $L$), respectively.