(U c. t)/b (U t)/b

Similar documents
VISUALIZATION OF THE TOPOLOGY OF THE LARGE-SCALE STRUCTURE IN THE PLANAR TURBULENT JET.

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

General introduction to Hydrodynamic Instabilities

Reduced-order models for flow control: balanced models and Koopman modes

Turbulence Modeling I!

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

Lecture 2. Turbulent Flow

Turbulence: Basic Physics and Engineering Modeling

Improvement of Reduced Order Modeling based on Proper Orthogonal Decomposition

Modeling of turbulence in stirred vessels using large eddy simulation

Turbulence and its modelling. Outline. Department of Fluid Mechanics, Budapest University of Technology and Economics.

Abstract The coherent structure in the self-similar region of a turbulent planar jet at moderate Reynolds number was experimentally investigated by ap

Proper Orthogonal Decomposition

Isotropic homogeneous 3D turbulence

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

Turbulent Rankine Vortices

Turbulence Instability

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

Anisotropic grid-based formulas. for subgrid-scale models. By G.-H. Cottet 1 AND A. A. Wray

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson, lada

UNIVERSITY OF CALGARY. Base Region Topology of Turbulent Wake around Finite Wall-Mounted Cylinder with. Application of Low Order Flow Representation

On the relationship between the mean flow and subgrid stresses in large eddy simulation of turbulent shear flows

Energy Transfer Analysis of Turbulent Plane Couette Flow

nek5000 massively parallel spectral element simulations

Melbourne, Victoria, 3010, Australia. To link to this article:

Several forms of the equations of motion

SECONDARY MOTION IN TURBULENT FLOWS OVER SUPERHYDROPHOBIC SURFACES

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

On the (multi)scale nature of fluid turbulence

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION

Turbulence - Theory and Modelling GROUP-STUDIES:

Parabolized Stability Analysis of Jets Issuing from Serrated Nozzles

Generation of initial fields for channel flow investigation

Dynamics of Large Scale Motions in Bubble-Driven Turbulent Flow

1. Introduction, tensors, kinematics

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

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

EXACT SOLUTIONS TO THE NAVIER-STOKES EQUATION FOR AN INCOMPRESSIBLE FLOW FROM THE INTERPRETATION OF THE SCHRÖDINGER WAVE FUNCTION

Isotropic homogeneous 3D turbulence

Chapter 7. Basic Turbulence

Simulating Drag Crisis for a Sphere Using Skin Friction Boundary Conditions

Introduction to Turbulence and Turbulence Modeling

7. Basics of Turbulent Flow Figure 1.

Estimation of Turbulent Dissipation Rate Using 2D Data in Channel Flows

Reduced-Order Modeling of Channel Flow Using Traveling POD and Balanced POD

Locality of Energy Transfer

ASSESSMENT OF ANISOTROPY IN THE NEAR FIELD OF A RECTANGULAR TURBULENT JET

Homogeneous Turbulence Dynamics

Using Statistical State Dynamics to Study the Mechanism of Wall-Turbulence

Toward low order models of wall turbulence using resolvent analysis

The Role of Splatting Effect in High Schmidt Number Turbulent Mass Transfer Across an Air-Water Interface

arxiv: v1 [physics.flu-dyn] 15 Jun 2016

Multiscale Computation of Isotropic Homogeneous Turbulent Flow

LES ANALYSIS ON CYLINDER CASCADE FLOW BASED ON ENERGY RATIO COEFFICIENT

The JHU Turbulence Databases (JHTDB)

Project Topic. Simulation of turbulent flow laden with finite-size particles using LBM. Leila Jahanshaloo

Validation of an Entropy-Viscosity Model for Large Eddy Simulation

LARGE EDDY SIMULATION AND FLOW CONTROL OVER A 25 RAMP MODEL

Aerodynamic force analysis in high Reynolds number flows by Lamb vector integration

Transition to turbulence in plane Poiseuille flow

A dynamic global-coefficient subgrid-scale eddy-viscosity model for large-eddy simulation in complex geometries

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

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

Chapter 6 An introduction of turbulent boundary layer

+ = + t x x x x u. The standard Smagorinsky model has been used in the work to provide the closure for the subgridscale eddy viscosity in (2):

6 VORTICITY DYNAMICS 41

Due Tuesday, November 23 nd, 12:00 midnight

GFD 2012 Lecture 1: Dynamics of Coherent Structures and their Impact on Transport and Predictability

A Solution Method for the Reynolds-Averaged Navier-Stokes Equation

IMPLEMENTATION OF POD AND DMD METHODS IN APACHE SPARK FRAMEWORK FOR SIMULATION OF UNSTEADY TURBULENT FLOW IN THE MODEL COMBUSTOR

Turbulent energy density and its transport equation in scale space

Turbulent drag reduction by streamwise traveling waves

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

Database analysis of errors in large-eddy simulation

Optimal rotary control of the cylinder wake using POD Reduced Order Model. Abstract

Eddy viscosity of cellular flows by upscaling

Stochastic excitation of streaky boundary layers. Luca Brandt, Dan Henningson Department of Mechanics, KTH, Sweden

J. Szantyr Lecture No. 4 Principles of the Turbulent Flow Theory The phenomenon of two markedly different types of flow, namely laminar and

Table of Contents. Foreword... xiii. Preface... xv

Wall turbulence with arbitrary mean velocity profiles

Multiscale Computation of Isotropic Homogeneous Turbulent Flow

Stability of Shear Flow

Résonance et contrôle en cavité ouverte

Turbulent boundary layer

CHAPTER 11: REYNOLDS-STRESS AND RELATED MODELS. Turbulent Flows. Stephen B. Pope Cambridge University Press, 2000 c Stephen B. Pope y + < 1.

Structures in the turbulent wake of an Ahmed body: LES vs Experiments using POD and some ideas about control

AER1310: TURBULENCE MODELLING 1. Introduction to Turbulent Flows C. P. T. Groth c Oxford Dictionary: disturbance, commotion, varying irregularly

Applied Computational Fluid Dynamics

LARGE EDDY SIMULATION OF MASS TRANSFER ACROSS AN AIR-WATER INTERFACE AT HIGH SCHMIDT NUMBERS

The JHU Turbulence Databases (JHTDB)

Figure 1. Schematic of experimental setup.

Dynamics and low-dimensionality of a turbulent near wake

LES of turbulent shear flow and pressure driven flow on shallow continental shelves.

Secondary vortices in turbulent square duct flow

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Entropy generation and transport

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

arxiv: v1 [physics.flu-dyn] 11 Oct 2012

Application of a Helmholtz resonator excited by grazing flow for manipulation of a turbulent boundary layer

Modelling of turbulent flows: RANS and LES

Transcription:

DYNAMICAL MODELING OF THE LARGE-SCALE MOTION OF A PLANAR TURBULENT JET USING POD MODES. S. Gordeyev 1 and F. O. Thomas 1 University of Notre Dame, Notre Dame, USA University of Notre Dame, Notre Dame, USA The existence and importance of the large-scale structures, that is large regions of coherent fluid motion in fully turbulent flows has been experimentally proven and a substantial body of research has been performed in this area. Research has clearly shown that these structures have a significant effect on the macrocharacteristics of the flow, like mass, heat transfer and mixing. Thus knowledge of dynamics and possible ways of controlling these structures have a great practical importance. An objective and rigorous way of extracting large-scale structure from turbulent flow is the Proper Orthogonal Decomposition (POD), proposed by Lumley [4]. The idea is to decompose a given turbulent flowinto a minimum number of modes by maximizing the energy content in each mode. A few important results from POD theory which make it a powerful tool for investigating turbulent flows are worth mentioning here. First, the POD gives a complete set of orthogonal modes and the flow can be expanded in this series. Second, the series exhibits the most rapid convergence among all possible complete orthogonal sets. Third, the input parameter to calculate the POD modes is a two-point correlation tensor which is relatively straightforward to measure. The rapid convergence POD modes makes them useful for building loworder dynamical models of different turbulent flows. Example 'success stories' are a turbulent boundary layer model developed by Aubry et al. [1] and the dynamical model of a turbulent mixing layer [7]. Another candidate flow for investigating large-scale structure dynamics is the turbulent planar jet. Twopoint correlation measurements across the centerline of the jet suggested the presence of well-defined large-scale structures. The interested reader is referred to [6] for a comprehensive review on structure in jets. Detailed measurements of the two-point cross-correlation tensor in the selfsimilar region of the planar turbulent jet in [] have shown the self-similarity of the underlying large-scale structure, when conventionally scaled by the local mean velocity half-width b and local maximum velocity. The large-scale structure was reconstructed in physical space by a Fourier transform-based projection of measured u, v; and w-component POD eigenmodes onto instantaneous flow field realizations in [3]. Results indicate that a dominant planar component of the large-scale structure in the planar jet consists of two lines

of large-scale spanwise vortices arranged approximately asymmetrically with respect to the jet centerline, as shown in Figure 1. This planar component of 1 1 4 4 38 36 34 3 3 8 6 4 (U c t)/b 1 1 18 16 14 1 1 8 6 4 (U t)/b Figure 1: The velocity field (u; v) and instantaneous streamlines of the experimentally reconstructed planar structure u planar ff (y; t) from [3]. y is the lateral spatial coordinate and x = U c t is a speudo-spatial streamwise coordinate. the structure resembles the classic Karman vortex street. There is a strong interaction between structures on opposite sides of the jet in the form of lateral streaming motions that extend well across the flow. Modeling In was shown in [4], [5] that POD modes are resolved up to an arbitrary phase in homogeneous directions. An alternative way to reconstruct the largescale structure is to project the experimentally obtained POD mode onto the governing Navier-Stokes equations and thus obtain the phase information necessary to reconstruct the topology and dynamics of the large-scale structure. In what follows spatial coordinates in the flow will be denoted x ff (x; y; z) where x; y; z are the streamwise, lateral (cross-stream) and spanwise spatial coordinates respectively. In a similar manner, the corresponding instantaneous velocity components are designated eu ff (x;t) = (eu; ev; ew). The conventional Reynolds decomposition is used to express the velocity as the sum of a timemean and zero-mean fluctuating component, eu ff (x;t)=u ff (x)+u ff (x;t). Due to the basic geometry of the turbulent planar jet flow field certain simplifying assumptions can be made and these are next described. 1. The flow is statistically homogeneous in the spanwise z direction.

. The lateral mean velocity component will be neglected. 3. The flow is considered to be homogeneous in the streamwise direction. The last assumption allows one to treat the streamwise coordinate through the temporal evolution by using a Taylor's frozen field x = U c t with U c taken as a constant convective speed. This allows one to transform the POD eigenmodes from the frequency to the streamwise wavenumber domain. Physically the last assumption is equivalent to letting the spreading rate of the jet be zero (i.e. imposing wall-like boundary conditions which prevent the jet from spreading). One then expects that the new modified system is locally dynamically similar to the original system. Using this approach is clearly inappropriate, however, for addressing issues related to global jet dynamics. Galerkin Projection After the spatial orthogonal POD modes ' (n) ff (y; k x ; ) are determined from experiment [], and exploiting the rapid energy convergence of the POD modes, the fluctuating velocity can be decomposed in the discrete series as u(x;t)= NX X n k c (n) (t; k)' (n) (y;k) exp(ikx) + unresolved modes; (1) where k =(k x ; ) and N is the number of POD modes kept in the expansion. We will term these modes as "resolved". The number of resolved modes can be varied to investigate the effect of neglected modes on the jet dynamics. To get a system of ODE's for the temporal coefficients c (n) (t; k x ; ), the Galerkin projection procedure onto Fourier-transformed Navier-Stokes is performed. Details can be found in [7], for instance. The final system of finite ODE's for the temporal coefficients c (n) (t; k x ; ) can be symbolically written in the matrix form X =(C (n) +P (n) +eνd (n) )c (n) (k)+ dc (n) (k) dt k ;l;m T (n;l;m) k;k c (l) (k )c (m) (k k ) () where convection, production, dissipation and nonlinear transfer terms are denoted as C (n), P (n), D (n) and T (n;l;m), respectively. The modified viscosity eν will be defined below. Because of the nonlinear nature of the N-S equations, there exists inertial energy transfer from the large resolved modes to smaller unresolved scales. Consequently, the effect of the unresolved scales may be modeled as a dissipative influence on the large scales. This interaction can be parametrically described by a Heisenberg type of spectral model for isotropic turbulence. The algorithm, proposed in [1] provides a way to estimate the turbulent eddy

viscosity ν T from the POD modes. The model introduces an effective turbulent viscosity ν T into the system and the molecular viscosity shall be replaced by a modified viscosity eν = ν + ffν T. The parameter ff allows one to vary the eddy viscosity, thus allowing to fine-tune the model in order to match experimentally observed dynamics of the resolved large-scale structures. In the expansion (1) we will restrict ourselves to the first POD mode only N mode =1. Furthermore, due todominance of the planar part of the underlying large-scale structure [3], only two spanwise wavenumbers are kept in the expansion, b=ß = and :5, so N z =. Finally, the number of the resolved streamwise wavenumbers was chosen as N x = 5. Thus the total number of solved complex equations in () is 5 = 1, or real equations. Results The numerical investigation of the low-order system () has revealed several regimes. Above ff = 4 the system has a trivial solution only. At ff =34 the system undergoes a Hopf bifurcation. For ff = 4::34 the coefficients c have a simple periodic behavior with amplitudes reaching constant values. At ff = 34 a second Hopf bifurcation occurs. The region ff =3:: corresponds to a situation, where the amplitudes exhibit an apparent chaotic behavior. Below ff ' the system eventually reaches infinite values for mode amplitudes in finite timedue to insufficient dissipation..35.3 (1) λ (St, k ) experiment z b/π=. b/π=.5.7.6 (1) λ (St, k ) model z b/π=. b/π=.5.5.5..4.15.3.1..5.1.1..3 St.1..3 St Figure : Experimental and model predicted (for ff = 1) eigenspectrum (1) (St; b). One of the criteria to gauge ff by comparison of experimental and computational results is to match the eigenvalue spectrum (1) (St; )=jc (1) (St; )j. For large ff, the dissipation effectively damps all modes except only few modes around St = :1, where production prevails. At small ff dissipation is not strong enough and many modes via a nonlinear energy transfer term in ()

4 6 8 1 1 14 16 18 x/b 4 6 8 3 3 34 36 38 4 4 x/b Figure 3: The velocity field (u; v) and instantaneous streamlines of the numerically reconstructed planar structure u ff (y; t; = ) for ff = 1. are excited. Comparison with the experimental spectrum allows one to choose the value of ff ß 1 where the shape of the numerical spectrum is close to the experimental one at b = : and :5, as shown in Figure. Overall, the shape of numerically generated eigenspectrum looks quite similar to the experimental one, although the numerical eigenvalue spectrum is less energetic at high St. The maximum peak is located at St =:1, where the production term P (1) is essentially non-zero. This matches the characteristic flapping frequency of the planar jet [6]. As can be seen in Figure the model overpredicts the amplitude of the eigenspectrum by a factor of two. The analysis of u rms and v rms also shows the model overshoots the maximum values by a factor of :::5 in the range ff = 1::15. This may be due to the very crude nature of the model, since it uses only one POD mode and a minimal number of spanwise wavenumber components. Viewed from this point, the model works surprisingly well considering the small number of resolved modes. Additional numerical investigation indicates that the inclusion of more POD modes, especially additional -components will most likely lower the model prediction for velocity amplitudes. A sample reconstructed velocity field for ff = 1 is presented in Figure 3. It reveals a striking similarity to the experimentally reconstructed topology of the large-scale structure shown in Figure 1. In fact, many essential features of the structure's topology is correctly predicted by the numerical model, including the y-locations of the vortical structures, as well as the typical size and spacing between them. Thus, the model correctly reconstructs the essential topology

of the self-similar jet large-scale structure. Rigorously speaking, the dynamical model is based on experimental data obtained at one Reynolds number, namely Re D =8; (based on exit velocity and nozzle width). While the set of basis POD functions is still complete, it could become non-optimal for planar jets with different Re. But the viscosity effects appear mostly on small scales. Investigations of large-scale structures in different turbulent flows indicate that the shapes of these structures are macroscale-dependent and are primarily a function of the boundary and initial conditions. So, the coherent structure is mostly governed by inertia forces, rather than viscous ones and the viscosity will not likely change the spatial shape of the first few POD modes. Rather, it will primerily effect the value of the number of involved modes and the dissipation energy rate from the large-scale modes. Hence it is expected that the set of POD modes will still be close to the optimal set and the application of the model can be relatively easily extended for a range of planar turbulent jets with different Re-numbers. References [1] N. Aubry, P. Holmes, J. Lumley and E. Stone,'The Dynamics of Coherent Structures in the Wall Region of Turbulent Shear Layer', J. Fluid Mech., 1988, 19, 115-175. [] S. Gordeyev and F.O. Thomas, 'Coherent Structure in the Turbulent Planar Jet. Part 1. Extraction of Proper Orthogonal Decomposition Eigenmodes and Their Self-Similarity', J. Fluid Mech.,, 414, 145-194. [3] S. Gordeyev and F.O. Thomas, 'Coherent Structure in the Turbulent Planar Jet. Part. Structural Topology via POD Eigenmode Projection', submitted to J. Fluid Mech., 1. [4] J. Lumley, 'The Structure of Inhomogeneous Turbulent Flows', Proceedings of the International Colloquium on the Fine Scale Structure of the Atmosphere and its Influence on Radio Wave Propagation, A.M. Yaglam and V.I. Tatarsky (Ed.), Doklady Akademii Nauk SSSR, Moscow, Nauka, 1967, 166-178. [5] J. Lumley, 'Stochastic Tools in Turbulence', Academic Press, New York 197. [6] F.O. Thomas, 'Structure of Mixing Layers and Jets ', Appl. Mech. Rev., 1991, 44(3), 119-153. [7] L. Ukeiley, 'Dynamics of Large Scale Structures in a Plane Turbulent Mixing Layer', Ph.D dissertation, Clarkson University, 1995.