Hostname: page-component-7c8c6479df-ph5wq Total loading time: 0 Render date: 2024-03-28T18:20:11.332Z Has data issue: false hasContentIssue false

Dissipative distinctions

Published online by Cambridge University Press:  22 December 2015

Ian P. Castro*
Affiliation:
Aeronautics, Faculty of Engineering and the Environment, University of Southampton, Southampton SO17 1BJ, UK
*
Email address for correspondence: i.castro@soton.ac.uk

Abstract

There have been numerous studies concerning the possibility of self-similar scaling laws in fully developed turbulent shear flows, driven over the past half-century or so by the early seminal work of Townsend (1956, The Structure of Turbulent Shear Flow. Cambridge University Press). His and nearly all subsequent analyses depend crucially on a hypothesis about the nature of the dissipation, ${\it\epsilon}$, of turbulence kinetic energy, $k$. It has usually been assumed (sometimes implicitly) that this is governed by the famous Kolmogorov relation ${\it\epsilon}=C_{{\it\epsilon}}k^{3/2}/L$, where $L$ is a length scale of the energy-containing eddies and $C_{{\it\epsilon}}$ is a constant. The paper by Dairay et al. (J. Fluid Mech. vol. 781, 2015, pp. 166–195) demonstrates, however, that, in the specific context of an axisymmetric wake, there can be regions where ${\it\epsilon}$ has a different behaviour, characterised by a $C_{{\it\epsilon}}$ that is not constant but depends on a varying local Reynolds number (despite the existence of a $-5/3$ region in the spectra). This leads to fundamentally different scaling laws for the wake.

Type
Focus on Fluids
Copyright
© 2016 Cambridge University Press 

1. Introduction

Townsend’s classical self-preserving approach to characterising the behaviour of turbulent shear flows rests on the hypothesis that, at sufficiently large Reynolds numbers, the flow will eventually (i.e. sufficiently far downstream) ‘forget’ anything about how it was created and thus have a universal form determined solely by the necessary integral constraints. Analysis then leads to the well-known scaling laws for the mean velocity and turbulence stress fields, whose details depend only on the type of flow. For an axisymmetric wake, the analysis leads to a wake width governed by ${\it\delta}\sim (x-x_{o})^{{\it\lambda}}$ and a decay in centreline velocity deficit governed by $u_{o}\sim (x-x_{o})^{-2{\it\lambda}}$ (with ${\it\lambda}=1/3$ ), where $x$ is the distance downstream from the wake-generating object and $x_{o}$ is a virtual origin. The turbulence statistics follow corresponding self-similar behaviour. Over the 60 years or so since Townsend’s early work, there have been many, largely experimental, studies of all the possible flow types, which have assessed the adequacy of the classical scaling laws. Such studies are never very straightforward, and this is particularly true for the axisymmetric wake. To be confident about both ${\it\lambda}$ and $x_{o}$ , it is necessary that the Reynolds numbers (both initial – set by the body geometry and upstream velocity – and local – set by $\sqrt{k}$ and ${\it\delta}$ , say) are large enough to ensure that Townsend’s ‘memory’ hypothesis and the Kolmogorov dissipation hypothesis are reasonable. In addition, there should be a sufficiently large range in $x$ downstream of the initial development region (a priori of unknown length). It is not easy to satisfy both requirements simultaneously.

There has thus been some controversy over whether the Townsend scalings hold. In fact, starting with the work of Bevilaqua & Lykoudis (Reference Bevilaqua and Lykoudis1978) over 35 years ago, there is mounting evidence that the geometry of the wake-generating body has a marked influence on the far-wake growth rate and turbulence even in regions where self-similarity is present, which leads to questions about whether the initial conditions really are ever forgotten. Much later, a similar result was found for plane wakes by Zhou & Antonia (Reference Zhou and Antonia1995). Johansson, George & Gourlay (Reference Johansson, George and Gourlay2003) undertook a more complete analysis and showed that, for small Reynolds number, an additional scaling was possible (for which ${\it\lambda}=1/2$ ) and that both solutions can indeed be dependent on initial conditions. A numerical study was first done for a high-Reynolds-number case by Gourlay et al. (Reference Gourlay, Arendt, Fritts and Werne2001). Both they and Redford, Castro & Coleman (Reference Redford, Castro and Coleman2012) studied the spatially homogeneous but time-developing equivalent of axisymmetric wakes using direct numerical simulation (DNS). The latter showed apparently unequivocally that, for late enough times (corresponding to very far downstream in the spatially developing case), the classical ${\it\lambda}=1/3$ universal behaviour occurs, in which the multiplying constants (e.g. in the growth-rate relation for ${\it\delta}$ ) are truly independent of the initial conditions and $C_{{\it\epsilon}}$ is essentially constant.

It is especially crucial to recognise that all the extant work has assumed the adequacy of Kolmogorov’s hypothesis (that the small-scale motions evolve much more rapidly than the time scale of the evolution of the whole flow), which led, with additional assumptions, to the famous equilibrium dissipation law, ${\it\epsilon}=C_{{\it\epsilon}}k^{3/2}/L$ (with $C_{{\it\epsilon}}=\text{const}.$ ). The major objective of Dairay, Obligado & Vassilicos (Reference Dairay, Obligado and Vassilicos2015) is to ‘establish the existence of a new non-equilibrium dissipation law’, which assumes that, given a global Reynolds number set by the initial conditions, $C_{{\it\epsilon}}\sim \mathit{Re}_{l}^{-n}$ , where $\mathit{Re}_{l}$ is a local Reynolds number. They do this for an axisymmetric wake, comparing their data with the scaling-law exponents that arise on the basis of this new dissipation law.

2. Overview

After a brief introduction to the arguments leading to the classical axisymmetric wake scalings in which ${\it\lambda}=1/3$ , not least their reliance on the assumption that $C_{{\it\epsilon}}=\text{const}.$ , Dairay et al. (Reference Dairay, Obligado and Vassilicos2015) introduce their alternative non-equilibrium dissipation law (discussed more fully by Vassilicos (Reference Vassilicos2015)), which states that $C_{{\it\epsilon}}\sim \mathit{Re}_{G}^{m}/\mathit{Re}_{l}^{n}$ , where $\mathit{Re}_{G}$ is a global Reynolds number set by the initial conditions and $\mathit{Re}_{l}$ is a local one, which for an axisymmetric wake (as for grid turbulence) falls with distance downstream. From the equations of motion and on the basis of the similarity arguments by George (Reference George, George and Arndt1989) (and see also Johansson et al. (Reference Johansson, George and Gourlay2003)), this leads to the wake scaling laws derived by Nedić, Vassilicos & Ganapathisubramani (Reference Nedić, Vassilicos and Ganapathisubramani2013), which can essentially be expressed as ${\it\delta}\sim (x-x_{o})^{1/(3-n)}$ and $u_{o}\sim (x-x_{o})^{-2/(3-n)}$ , with $n=m=1$ (cf. $n=m=0$ for ${\it\lambda}=1/3$ ). Note that this value of $n$ and $m$ is the same as for a laminar wake ( ${\it\lambda}=1/2$ ), although it arises for different reasons. However, it can also arise (and does) when the classical Kolmogorov law is assumed for a fully turbulent, high-Reynolds-number wake characterised by large ratios of turbulence to viscous stress, provided only that the eddy viscosity is constant (Redford et al. Reference Redford, Castro and Coleman2012). The presence of $n=1$ scaling is thus not necessarily associated with low Reynolds number.

Dairay et al. (Reference Dairay, Obligado and Vassilicos2015) examine their scalings by exploring the wake of an irregular (fractal-type) bluff plate with sharp edges, using both a wind tunnel experiment and a matching DNS. This is unusual, not only in that both approaches are used in the same work but also because the DNS is of a spatially developing wake (as in the experiment), rather than a time-developing one, with the generating body included in the computational domain. Whilst this allows capture of the near wake it has the inevitable consequence that the available downstream extent of the wake is somewhat limited. It is also unusual because of the use of a fractal-type wake-generating plate, which leads, in the very near field, to a (‘multi-scale’) flow having a mixture of wake-like and jet-like character. For the DNS, great care is taken to ensure that there is sufficient domain size, grid resolution and statistical convergence. For the experiments, the plate is placed in a low-turbulence wind tunnel at $\mathit{Re}_{G}=U_{\infty }L_{b}/{\it\nu}=40\,000$ , where $L_{b}=\sqrt{A}$ , with $A$ the plate area. The measurements are made using hot-wire anemometry and extend to $x\approx 50L_{b}$ , whereas the DNS has $\mathit{Re}_{G}=5000$ and reaches approximately $x=100L_{b}$ .

The results suggest that, over most of the extent of the wind tunnel wake ( $15\leqslant x/L_{b}\leqslant 50$ ), $C_{{\it\epsilon}}\mathit{Re}_{l}^{n}$ with $n=1$ is more closely constant than is $C_{{\it\epsilon}}$ . Actually, they show that a better fit requires $n\approx 0.77$ . Further downstream ( $55\leqslant x/L_{b}\leqslant 100$ ) the DNS data suggest a change to $n\approx 0.5$ . To derive theoretical scalings for wake width and velocity deficit that have $n\neq 1$ , the authors make a ‘constant anisotropy’ assumption – that the Reynolds shear stress and the turbulent kinetic energy profiles scale in the same way (but not with $u_{o}^{2}$ ). This is essentially a revised Townsend–George theory (Townsend Reference Townsend1976; George Reference George, George and Arndt1989), but includes the new ‘non-equilibrium dissipation’ law. (Note that only the latter is necessary for $n=1$ .) The variations of $u_{o}$ and ${\it\delta}$ along the wake are shown to conform quite well to the new scalings, $u_{o}\sim (x-x_{o})^{{\it\alpha}}$ and ${\it\delta}\sim (x-x_{o})^{{\it\beta}}$ , with ${\it\alpha}=-2{\it\beta}=-2(1+n)/(3+n)$ , albeit with different $n$ ( ${<}1$ ) for the upstream and downstream halves of the $x$ region studied. Given that the local Reynolds number ( $\sqrt{k}{\it\delta}/{\it\nu}$ ) only falls to around 230 by $x=100L_{b}$ in the DNS, and is very much higher in the wind tunnel, it is arguably difficult to claim that it is too small to expect the classical scaling to hold.

In addition to the increasing body of evidence that the value of $C_{{\it\epsilon}}$ can depend on initial conditions, Dairay et al. (Reference Dairay, Obligado and Vassilicos2015) (following Nedić et al. (Reference Nedić, Vassilicos and Ganapathisubramani2013), along with the same group’s work on grid turbulence) thus go much further and question the universality of Reynolds-number independence of $C_{{\it\epsilon}}$ . No physical explanation is offered for why the $C_{{\it\epsilon}}\mathit{Re}_{l}^{n}=\text{const}.$ dissipation law might apply, but it does provide revised scaling laws for axisymmetric wakes that fit the present data.

3. Future

Although there is considerable evidence for the adequacy of the classical dissipation relation, it is apparent that it may be too simplistic, at more than one level. (Actually, it has long been recognised that $C_{{\it\epsilon}}$ is unlikely to be universal (Taylor Reference Taylor1935).) It is already clear that, whilst at sufficiently large Reynolds number $C_{{\it\epsilon}}$ may become constant, its precise value can depend on initial conditions (e.g. Sreenivasan Reference Sreenivasan1998; Antonia & Pearson Reference Antonia and Pearson2000). Dairay et al.’s data seem even more revealing, in that, using only one wake-generating body, they show that $C_{{\it\epsilon}}$ is not even constant, but rather varies with local Reynolds number. The Kolmogorov law ${\it\epsilon}=C_{{\it\epsilon}}k^{3/2}/L$ , with $C_{{\it\epsilon}}=\text{const}.$ , is often seen as ‘one of the cornerstone assumptions of turbulence theory’ but Lumley (Reference Lumley1992) remarked that the ‘mechanism that sets the level of dissipation in a turbulent flow, particularly in changing circumstances’, is worthy of further study. If the results of the Dairay et al. (Reference Dairay, Obligado and Vassilicos2015) study can be shown to be typical of other high-Reynolds-number turbulent shear flows ‘in changing circumstances’, or indeed in axisymmetric wakes at significantly higher $\mathit{Re}_{l}$ than they could reach, the premonition implied by Lumley’s remark will prove to have been prophetic. The matter is certainly worthy of more extensive study, not least because the non-equilibrium dissipation law seems to break the link between the presence of $-5/3$ spectra and classical cascade arguments. This new law seems to hold over much of any wake region that is likely to exist in real applications, so, even without any physical explanation, it would seem to be important. One can expect further experiments exploring the issue, aimed not least at finding whether (and if so, why) there is a final transition to a more classical scaling at some greater distance downstream, as Dairay et al. (Reference Dairay, Obligado and Vassilicos2015) suggest, even though $\mathit{Re}_{l}$ must continue to fall.

Finally, note that little is known about how the very-near-wake flow transitions to the region explored by Dairay et al. (Reference Dairay, Obligado and Vassilicos2015). The process must surely be very dependent on the geometry of the generating body. The near-wake usually contains interesting and complex dynamics (e.g. the recent work of Rigas et al. (Reference Rigas, Morgans, Brackston and Morrison2015) and references therein). There would seem to be much scope for exploration of the various transition regions. And there remains the need for a physical explanation for the new non-equilibrium dissipation relation.

References

Antonia, R. A. & Pearson, B. R. 2000 Effect of initial conditions on the mean energy dissipation rate and the scaling exponent. Phys. Rev. E 62, 80868090.CrossRefGoogle ScholarPubMed
Bevilaqua, P. M. & Lykoudis, P. S. 1978 Turbulence memory in self-preserving wakes. J. Fluid Mech. 89, 589606.CrossRefGoogle Scholar
Dairay, T., Obligado, M. & Vassilicos, J. C. 2015 Non-equilibrium scaling laws in axisymmetric turbulent wakes. J. Fluid Mech. 781, 166195.CrossRefGoogle Scholar
George, W. K. 1989 The self-preservation of turbulent flows and its relation to initial conditions. In Advances in Turbulence (ed. George, W. K. & Arndt, R.), pp. 3973. Hemisphere.Google Scholar
Gourlay, M. J., Arendt, S. C., Fritts, D. C. & Werne, J. 2001 Numerical modelling of inititially turbulent wakes with net momentum. Phys. Fluids 13, 37823802.CrossRefGoogle Scholar
Johansson, P. B. V., George, W. K. & Gourlay, M. J. 2003 Equilibrium similarity, effects of initial conditions and local Reynolds number on the axisymmetric wake. Phys. Fluids 15, 603617.CrossRefGoogle Scholar
Lumley, J. L. 1992 Some comments on turbulence. Phys. Fluids A 4, 201211.CrossRefGoogle Scholar
Nedić, N., Vassilicos, J. C. & Ganapathisubramani, B. 2013 Drag and near wake characteristics of flat plates normal to the flow with fractal edge geometries. Phys. Rev. Lett. 111, 144503.Google Scholar
Redford, J. A., Castro, I. P. & Coleman, G. N. 2012 On the universality of turbulent axisymmetric wakes. J. Fluid Mech. 710, 419452.CrossRefGoogle Scholar
Rigas, G., Morgans, A. S., Brackston, P. D. & Morrison, J. F. 2015 Diffusive dynamics and stochastic models of turbulent axisymmetric wakes. J. Fluid Mech. 778, R2.CrossRefGoogle Scholar
Sreenivasan, K. R. 1998 An update on the energy dissipation rate in isotropic turbulence. Phys. Fluids 10, 528529.CrossRefGoogle Scholar
Taylor, G. I. 1935 Statistical theory of turbulence. Proc. R. Soc. Lond. A 151, 421444.CrossRefGoogle Scholar
Townsend, A. A. 1956 The Structure of Turbulent Shear Flow. Cambridge University Press.Google Scholar
Townsend, A. A. 1976 The Structure of Turbulent Shear Flow, 2nd edn. Cambridge University Press.Google Scholar
Vassilicos, J. C. 2015 Dissipation in turbulent flows. Annu. Rev. Fluid Mech. 47, 95114.CrossRefGoogle Scholar
Zhou, Y. & Antonia, R. A. 1995 Memory effects in a turbulent plane wake. Exp. Fluids 19, 112120.CrossRefGoogle Scholar