Dimensionality influence on energy, enstrophy and passive scalar transport.

Similar documents
S.Di Savino, M.Iovieno, L.Ducasse, D.Tordella. EPDFC2011, August Politecnico di Torino, Dipartimento di Ingegneria Aeronautica e Spaziale

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

Characteristics of Linearly-Forced Scalar Mixing in Homogeneous, Isotropic Turbulence

Passive Scalars in Stratified Turbulence

Note the diverse scales of eddy motion and self-similar appearance at different lengthscales of the turbulence in this water jet. Only eddies of size

Energy and water vapor transport across a simplified cloud-clear air interface

Lecture 2. Turbulent Flow

Simulations of Three-Dimensional Turbulent Mixing for Schmidt Numbers of the Order 1000

CVS filtering to study turbulent mixing

Scale interactions and scaling laws in rotating flows at moderate Rossby numbers and large Reynolds numbers

Problem C3.5 Direct Numerical Simulation of the Taylor-Green Vortex at Re = 1600

Passive scalars in stratified turbulence

Before we consider two canonical turbulent flows we need a general description of turbulence.

meters, we can re-arrange this expression to give

Buoyancy Fluxes in a Stratified Fluid

Multiscale Computation of Isotropic Homogeneous Turbulent Flow

On the (multi)scale nature of fluid turbulence

Turbulent Flows. g u

Final abstract for ONERA Taylor-Green DG participation

Mixing in Isotropic Turbulence with Scalar Injection and Applications to Subgrid Modeling

Applied Computational Fluid Dynamics

Natalia Tronko S.V.Nazarenko S. Galtier

Evolution of the pdf of a high Schmidt number passive scalar in a plane wake

Cumulative distribution of the stretching and twisting of vortical structures in isotropic turbulence.

Lagrangian intermittency in drift-wave turbulence. Wouter Bos

O. A Survey of Critical Experiments

Turbulence (January 7, 2005)

Direct numerical simulation of a statistically stationary, turbulent reacting flow

Linearly forced isotropic turbulence

Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2)

Chapter 13. Eddy Diffusivity

Introduction Efficiency bounds Source dependence Saturation Shear Flows Conclusions STIRRING UP TROUBLE

Database analysis of errors in large-eddy simulation

Evolution of passive scalar statistics in a spatially developing turbulence. Abstract

Turbulent velocity fluctuations need not be Gaussian

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman

An evaluation of a conservative fourth order DNS code in turbulent channel flow

arxiv:nlin/ v2 [nlin.cd] 29 Jul 2001

Max Planck Institut für Plasmaphysik

Introduction to Turbulence and Turbulence Modeling

The Spatial Variation of the Maximum Possible Pollutant Concentration from Steady Sources

NONLINEAR FEATURES IN EXPLICIT ALGEBRAIC MODELS FOR TURBULENT FLOWS WITH ACTIVE SCALARS

Turbulent Mixing of Passive Tracers on Small Scales in a Channel Flow

Lagrangian statistics in turbulent channel flow

Logarithmic scaling in the near-dissipation range of turbulence

Homogeneous Rayleigh-Bénard convection

Validation of an Entropy-Viscosity Model for Large Eddy Simulation

6 VORTICITY DYNAMICS 41

Intermittency, Fractals, and β-model

Probability density function modeling of scalar mixing from concentrated sources in turbulent channel flow

An Introduction to Theories of Turbulence. James Glimm Stony Brook University

On the turbulent Prandtl number in homogeneous stably stratified turbulence

Turbulence Instability

THE DECAY OF HOMOGENEOUS TURBULENCE GENERATED BY MULTI-SCALE GRIDS

Turbulence: Basic Physics and Engineering Modeling

Spectral analysis of energy transfer in variable density, radiatively heated particle-laden flows

Logarithmically modified scaling of temperature structure functions in thermal convection

On the decay of two-dimensional homogeneous turbulence

arxiv: v1 [physics.flu-dyn] 1 Sep 2010

fluctuations based on the resolved mean flow

The behaviour of high Reynolds flows in a driven cavity

DNS STUDY OF TURBULENT HEAT TRANSFER IN A SPANWISE ROTATING SQUARE DUCT

Role of polymers in the mixing of Rayleigh-Taylor turbulence

Chapter (3) TURBULENCE KINETIC ENERGY

Acceleration and dissipation statistics of numerically simulated isotropic turbulence

Multi-Scale Modeling of Turbulence and Microphysics in Clouds. Steven K. Krueger University of Utah

Nonequilibrium Dynamics in Astrophysics and Material Science YITP, Kyoto

Turbulence - Theory and Modelling GROUP-STUDIES:

Locality of Energy Transfer

Angular Statistics of Lagrangian Trajectories in Turbulence. Wouter Bos, Benjamin Kadoch and Kai Schneider

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

P2.1 Scalar spectra in the near-dissipation range derived from structure functions

Schmidt number effects on turbulent transport with uniform mean scalar gradient

The Johns Hopkins Turbulence Databases (JHTDB)

Effects of Forcing Scheme on the Flow and the Relative Motion of Inertial Particles in DNS of Isotropic Turbulence

Turbulent energy density and its transport equation in scale space

Influence of high temperature gradient on turbulence spectra

Probability density function (PDF) methods 1,2 belong to the broader family of statistical approaches

Wall turbulence with arbitrary mean velocity profiles

Examination of hypotheses in the Kolmogorov refined turbulence theory through high-resolution simulations. Part 2. Passive scalar field

Dissipation Scales & Small Scale Structure

Journal of Fluid Science and Technology

Homogeneous Turbulence Generated by Multiscale

Turbulent Rankine Vortices

Direct numerical simulation of homogeneous turbulence with hyperviscosity

Effects of Forcing Scheme on the Flow and the Relative Motion of Inertial Particles in DNS of Isotropic Turbulence

Homogeneous Turbulence Dynamics

A simple subgrid-scale model for astrophysical turbulence

On modeling pressure diusion. in non-homogeneous shear ows. By A. O. Demuren, 1 M. M. Rogers, 2 P. Durbin 3 AND S. K. Lele 3

Experimental assessment of Taylor s hypothesis and its applicability to dissipation estimates in turbulent flows

Intermittency, pressure and acceleration statistics from hot-wire measurements in wind-tunnel turbulence

arxiv:nlin/ v1 [nlin.cd] 29 Jan 2004

Environmental Atmospheric Turbulence at Florence Airport

The viscous-convective subrange in nonstationary turbulence

Best Practice Guidelines for Combustion Modeling. Raphael David A. Bacchi, ESSS

2.3 The Turbulent Flat Plate Boundary Layer

Wind and turbulence experience strong gradients in vegetation. How do we deal with this? We have to predict wind and turbulence profiles through the

Concentration and segregation of particles and bubbles by turbulence : a numerical investigation

Fundamentals of Turbulence

Analysis of Turbulent Free Convection in a Rectangular Rayleigh-Bénard Cell

Transcription:

Dimensionality influence on energy, enstrophy and passive scalar transport. M. Iovieno, L. Ducasse, S. Di Savino, L. Gallana, D. Tordella 1 The advection of a passive substance by a turbulent flow is important in many natural and engineering contexts, e.g. turbulent mixing, combustion, pollution dispersal, in which the prediction of mixing and dispersion rates of a scalar contaminant is a problem of great interest. Although the concentration of a passive substance exhibits a complex behaviour that shows many phenomenological parallels with the behaviour of the turbulent velocity field, the statistical properties of passive scalar turbulence, strongly influenced by the Kolmogorov cascade phenomenology, are in part decoupled from those of the underlying velocity field [1]. The subject, in the last years, is in fact undergoing a reinterpretation as empirical evidence shows that local isotropy, both at the inertial and dissipation scales, is violated [2]. Moreover, dispersion usually occurs in time dependent inhomogeneous flows, which present a much more complicate behaviour than homogeneous flows and thus are beyond the reach of analytical models or even numerical simulations. In this work, we go one step beyond the homogeneous and isotropic turbulence and consider the simplest inhomogeneous and shearfree turbulent Navier-Stokes motion: the system where an energetic turbulent isotropic field is left to convectively diffuse into a low energy one. In this system the region where the two turbulences interact is preceded by a highly intermittent thin layer - where the energy flux is maximum - that propagates into the low energy region, see [3, 4, 5]. Transport and intermittency un shearless mixings side The initial conditions introduce a kinetic energy gradient in the direction of inhomogeneity. Outside this inhomogeneous region, the kinetic energy in the two turbulence fields decays as a power law, with an exponent approximately equal to -1.2, in the three-dimensional simulations, while in two dimensions there is no significant decay. In both cases, the initially imposed energy ratio is preserved during the time evolution of the flow. The mixing layer between the two homogeneous flows was observed to be highly intermittent, as shown by the large values of the velocity skewness and kurtosis, see [6, 3, 4, 5]. In this work, we focus on the scalar dispersion through the shearless mixing layer. The scalar interface is spread by the turbulent eddies and a scalar mixing region with high scalar variance is generated in between the two homogeneous flows. The width of this region can be measured by considering the mean scalar distributions. The scalar mixing layer thickness θ is defined, in analogy with the energy layer thickness E, as the distance between the points with means scalar θ equal to 0.75 and 0.25. After an initial transient of about one eddy-turnover time, the time evolution of these interaction widths follows the one observed for the self-diffusion 1 Corrisponding autor: daniela.tordella@polito.it

Figure 1. Three-dimensional visualization of the kinetic energy and passive scalar fields. Above: isosurfaces of the turbulent kinetic energy, E = 0.1Emax (white) and E = 0.35Emax (green). Below: isosurfaces of the scalar field, θ = 0.25 (yellow) and θ = 0.75 (light blue). The high turbulent energy velocity field is on the right of each image; only a portion of the computational domain has been shown. The images are at t/τ = 4, where τ is the initial eddy turnover time of the high energy region.

of the velocity field with the same dimensionality, and a stage of evolution with a power law scaling of the mixing thickness is reached in both two and three dimensions. Scalar diffusion in two dimensions is asymptotically faster than in three dimensions, a result already observed in the kinetic energy mixing (see [7, 8]). In three dimensions there is a fair agreement with the wind tunnel experiments on the scalar diffusion from a linear source in a shearless mixing [6, 9]. In the three-dimensional flow, the scalar variance has already reached its maximum after less than one eddy turnover time and, since that instant, it slowly decreases. In the following 10 eddy turnover times, about 20% of the variance present at t/τ = 1 is lost. In two dimensions, the scalar flow is almost twice as large and the initial variance generation last longer: the maximum is attained later and is about 50% higher. The presence of the turbulent energy gradient is instead felt on the distribution of higher order moments as the scalar skewness shown in figures 2. In contrast with the velocity field, the spatial distribution of the scalar moments shows the presence of two intermittent fronts at the borders of the mixed region. These two fronts are at a distance from the mixing centre approximatively equal to 2 θ. Because of the gradual thickening of the mixing layer, the kinetic energy and the scalar gradients, which drive the scalar spreading, become milder and milder. As a consequence, not only the scalar flux is reduced, but also the intermittency levels of the two scalar fronts decrease in time. Because of the energy decay, this reduction is faster in three dimensions. Intermittency is not limited to large scale scalar fluctuations, but is quickly spread to small scale fluctuations, as can be inferred from the skewness and kurtosis of the scalar derivative in the inhomogeneous direction. Two peaks of large intermittency are present in correspondence of the two intermittent fronts since t/τ = 1. However, their time evolution is different from the one seen for large scale scalar fluctuations: the peaks tend to decay in time, but small scale intermittency tends to last longer in the three dimensional case. In absence of an energy gradient (i.e., E 1 /E 2 = 1) we observed the same mixing lenght temporal growth but the two scalar intermittent fronts at the borders of the mixing layer become symmetric [7]. In the central part of the mixing layer, between the two intermittent fronts, the scalar fluctuations tend, after few eddy turnover times, to a slow varying state. In two dimensions, the passive scalar exponent in the inertial range is about 1.4, which is roughly one half of the 3 exponent of the velocity field and is far from the κ 1 inertial range scaling of homogeneous and statistically stationary flows (see, e.g., [10]). In three dimensions, the difference between scalar and velocity exponents is very mild and both tend to approach the homogeneous turbulence scaling. However, the scalar spectrum seems to show a wider inertial range region, a feature which has been observed also in homogeneous flows at moderate Reynolds numbers, see [11]. Conclusion The diffusion length, θ, of a passive scalar through a shearless turbulent mixing closely follows the temporal evolution of the self-diffusion length of the velocity field E. In two dimensions the growth is faster: E θ t 0.68, while in three dimensions E θ t 0.46. The scalar flow is about two times larger in two dimensions than in three dimensions. Also the scalar variance in the centre of the mixed layer is 50% higher in two dimensional case. An important observation concerns the presence of two scalar intermittent fronts which appear at the sides of the mixed region. The intermittency levels of the fronts is high and gradually decay in time. In all simulations, the fronts move from the initial position of the interface. However, he front toward the high energy side of the mixing shows a deeper penetration and a higher intermittency. This asymmetry is milder in the three-dimensional case. The intermittency is not limited to the large scale, but involves also the small scale. In

8 6 4 2 t/τ 1 5 10 20 S θ 0-2 -4-6 energy f ow scalar f ow -8-3 -2-1 0 1 2 3 8 6 4 2 (x x c )/ θ t/τ 1 5 10 12.5 S θ 0-2 -4-6 energy f ow scalar f ow -8-3 -2-1 0 1 2 3 (x x c )/ θ Figure 2. Scalar skewness distribution in the simulations with initial energy ratio E 1 /E 2 = 6.7. The vertical dashed lines indicate the position of maximum intermittence at the end of the simulation. two dimensions, the small scale intermittency decay is very fast. In three dimensions, it is more intense and lasts longer. The large scale intermittency follows an opposite behaviour. In the centre of the mixing layer, in three dimensions, the passive scalar and velocity spectra both show an inertial range with an exponent close to -5/3. In two dimensions, the velocity shows a κ 3 inertial range, while the passive scalar shows a κ 1.4 inertial range. An exponent value which is closer to the three dimensional κ 5/3 scaling than to the two dimensional k 1 scaling, which is observed in the case of passive scalar transport in homogeneous condition. Method From a numerical point of view, the flow is assumed to be contained in a parallelepiped (or a rectangle in two dimensions) and periodic boundary conditions are applied to all the spatial directions. The initial condition is obtained by the linear match of two homogeneous and isotropic fields with the same integral scale but with a different turbulent kinetic energy. The imposed initial energy ratio is 6.7 in both two and three dimensions. The passive scalar is introduced in the low energy region of the flow at t = 0: the scalar concentration is initially

uniform in the two isotropic regions, θ = 0 in the high energy region and θ = 1 in the low energy region. The discontinuity is replaced by a sufficiently smooth transition to avoid the Gibbs phenomenon. No scalar fluctuation is introduced in the initial conditions. In the three-dimensional simulations, the highest turbulent field has a Taylor microscale Reynolds number equal to 250. The directions normal to the energy gradient in this flow configuration remain statistically homogeneous during the decay, so that all the statistics can be computed as plane averages in these directions. The mass, momentum and and the scalar transport equation are solved by using a dealiased pseudospectral Fourier-Galerkin spatial discretization coupled with a fourth order Runge-Kutta explicit time integration. About 22 initial eddy turnover times have been simulated in two dimensions and 12.5 initial eddy turnover times in three dimensions. Acknowledgements The authors would like to thank CINECA computing centre (http://www.cineca.it) for supporting this work. References [1] B. I. Shraiman and B. I. Siggia. Scalar turbulence. Nature, 405:639 646, 2000. [2] Z. Warhaft. Passive scalar in turbulent flows. Ann. Rev. Fluid Mech., 32:203 240, 2000. [3] D. Tordella and M. Iovieno. Numerical experiments on the intermediate asymptotics of the shear-free turbulent transport and diffusion. J. Fluid Mech., 549:429 441, 2006. [4] D. Tordella, M. Iovieno, and P. R. Bailey. Sufficient condition for gaussian departure in turbulence. Phys. Rev. E, 77:016309/1 11, 2008. [5] D. Tordella and M. Iovieno. Small scale anisotropy in turbulent shearless mixings. Phys. Rev. Lett, 107:194501, 2011. [6] S. Veeravalli and Z. Warhaft. The shearless turbulence mixing layer. J. Fluid Mech., 207:191 229, 1989. [7] M.Iovieno, L. Ducasse, and D. Tordella. Dimensionality influence on passive scalar transport. J.Physics: Conference Series, 318:052042 (ETC 13, Warsaw, Sept.2011), 2011. [8] D. Tordella, M. Iovieno, and L. Ducasse. Passive scalar diffusion through a turbulent energy gradient. Bad Reichenhall, September 2010. EFMC-10. [9] S. Veeravalli and Z. Warhaft. Thermal dispersion from a line source in the shearless turbulence mixing layer. J. Fluid Mech., 216:35 70, 1990. [10] W. J. Bos, B. Kadoch, K. Schneider, and J. P. Bertoglio. Inertial range scaling of the scalar flux spectrum in two-dimensional turbulence. Phys. Fluids, 21:115105/1 8, 2009. [11] L. Danaila and R. A. Antonia. Spectrum of a passive scalar in moderate Reynolds number homogeneous isotropic turbulence. Phys. Fluids., 21:111702/1 4, 2009.