You are on page 1of 247

Fluid mechanics, turbulent ow and turbulence modeling

Lars Davidson Division of Fluid Dynamics Department of Applied Mechanics Chalmers University of Technology SE-412 96 G teborg, Sweden o http://www.tfd.chalmers.se/lada, lada@chalmers.se August 17, 2011
Abstract This course material is used in two courses in the International Masters programme Applied Mechanics at Chalmers. The two courses are TME225 Mechanics of uids, and MTF270 Turbulence Modeling. MSc students who follow these courses are supposed to have taken one basic course in uid mechanics. This document can be downloaded at http://www.tfd.chalmers.se/lada/MoF/lecture notes.html and http://www.tfd.chalmers.se/lada/comp turb model/lecture notes.html The Fluid courses in the MSc programme are presented at http://www.tfd.chalmers.se/lada/msc/msc-programme.html

The MSc programme is presented at http://www.chalmers.se/en/education/programmes/masters-info/Pages/Applied-Mechanics

Contents
1 Motion, ow 1.1 Eulerian, Lagrangian, material derivative . . . . . . . . . 1.2 Viscous stress, pressure . . . . . . . . . . . . . . . . . . 1.3 Strain rate tensor, vorticity . . . . . . . . . . . . . . . . . 1.4 Deformation, rotation . . . . . . . . . . . . . . . . . . . 1.5 Irrotational and rotational ow . . . . . . . . . . . . . . 1.5.1 Ideal vortex line . . . . . . . . . . . . . . . . . . 1.5.2 Shear ow . . . . . . . . . . . . . . . . . . . . . 1.6 Eigenvalues and and eigenvectors: physical interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 9 10 11 13 15 16 17 18 19 19 19 19 20 21 22 23 23 23 27 27 28 30 31 32 32 34 37 38 39 39 40 41 44 47 47 48 49 53 55 55 58

Governing ow equations 2.1 The Navier-Stokes equation . . . . . . . . . . . . . . . . . . . . . 2.1.1 The continuity equation . . . . . . . . . . . . . . . . . . . 2.1.2 The momentum equation . . . . . . . . . . . . . . . . . . 2.2 The energy equation . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Transformation of energy . . . . . . . . . . . . . . . . . . . . . . 2.4 Left side of the transport equations . . . . . . . . . . . . . . . . . 2.5 Material particle vs. control volume (Reynolds Transport Theorem) Exact solutions to the Navier-Stokes equation: two examples 3.1 The Rayleigh problem . . . . . . . . . . . . . . . . . . . 3.2 Flow between two plates . . . . . . . . . . . . . . . . . . 3.2.1 Curved plates . . . . . . . . . . . . . . . . . . . 3.2.2 Flat plates . . . . . . . . . . . . . . . . . . . . . 3.2.3 Force balance . . . . . . . . . . . . . . . . . . . 3.2.4 Balance equation for the kinetic energy . . . . . . Vorticity equation and potential ow 4.1 Vorticity and rotation . . . . . . . . . . . . . . . . 4.2 The vorticity transport equation in three dimensions 4.3 The vorticity transport equation in two dimensions . 4.3.1 Diffusion length from the Rayleigh problem Turbulence 5.1 Introduction . . . . . . . . . . . . . . . 5.2 Turbulent scales . . . . . . . . . . . . . 5.3 Energy spectrum . . . . . . . . . . . . . 5.4 The cascade process created by vorticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

Turbulent mean ow 6.1 Time averaged Navier-Stokes . . . . . . . . . . . 6.1.1 Boundary-layer approximation . . . . . . 6.2 Wall region in fully developed channel ow . . . 6.3 Reynolds stresses in fully developed channel ow 6.4 Boundary layer . . . . . . . . . . . . . . . . . . . Probability density functions Transport equations for kinetic energy 2

7 8

8.1 8.2 8.3 8.4 9

The Exact k Equation . . . . . . . . . Spatial vs. spectral energy transfer . . The overall effect of the transport terms The transport equation for vi vi /2 . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

58 61 62 63 65 70 71 71 73 76 76 76 77 77 78 80 80 81 82 82 83 84 84 87 93 95 96 96 97 98 99 99 101 103 104 105 106 107 110 113 114 114 114

Transport equations for Reynolds stresses 9.1 Reynolds shear stress vs. the velocity gradient . . . . . . . . . . . .

10 Correlations 10.1 Two-point correlations . . . . . . . . . . . . . . . . . . . . . . . . . 10.2 Auto correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 Reynolds stress models and two-equation models 11.1 Mean ow equations . . . . . . . . . . . . . . 11.1.1 Flow equations . . . . . . . . . . . . 11.1.2 Temperature equation . . . . . . . . . 11.2 The exact vi vj equation . . . . . . . . . . . . 11.3 The exact vi equation . . . . . . . . . . . . 11.4 The k equation . . . . . . . . . . . . . . . . . 11.5 The equation . . . . . . . . . . . . . . . . . 11.6 The Boussinesq assumption . . . . . . . . . . 11.7 Modelling assumptions . . . . . . . . . . . . 11.7.1 Production terms . . . . . . . . . . . 11.7.2 Diffusion terms . . . . . . . . . . . . 11.7.3 Dissipation term, ij . . . . . . . . . 11.7.4 Slow pressure-strain term . . . . . . . 11.7.5 Rapid pressure-strain term . . . . . . 11.7.6 Wall model of the pressure-strain term 11.8 The k model . . . . . . . . . . . . . . . . 11.9 The modelled vi vj equation with IP model . . 11.10 Algebraic Reynolds Stress Model (ASM) . . . 11.11 Explicit ASM (EASM or EARSM) . . . . . . 11.12 Boundary layer ow . . . . . . . . . . . . . . 12 Reynolds stress models vs. eddy-viscosity models 12.1 Stable and unstable stratication . . . . . . 12.2 Curvature effects . . . . . . . . . . . . . . . 12.3 Stagnation ow . . . . . . . . . . . . . . . 12.4 RSM/ASM versus k models . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

13 Realizability 13.1 Two-component limit . . . . . . . . . . . . . . . . . . . . . . . . . 14 Non-linear Eddy-viscosity Models 15 The V2F Model 15.1 Modied V2F model . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2 Realizable V2F model . . . . . . . . . . . . . . . . . . . . . . . . . 15.3 To ensure that v 2 2k/3 [1] . . . . . . . . . . . . . . . . . . . . . 16 The SST Model

17 Large Eddy Simulations 17.1 Time averaging and ltering . . . . . . . . . . . . . . . . . . . . . . 17.2 Differences between time-averaging (RANS) and space ltering (LES) 17.3 Resolved & SGS scales . . . . . . . . . . . . . . . . . . . . . . . . 17.4 The box-lter and the cut-off lter . . . . . . . . . . . . . . . . . . 17.5 Highest resolved wavenumbers . . . . . . . . . . . . . . . . . . . . 17.6 Subgrid model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.7 Smagorinsky model vs. mixing-length model . . . . . . . . . . . . . 17.8 Energy path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.9 SGS kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . 17.10 LES vs. RANS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.11 The dynamic model . . . . . . . . . . . . . . . . . . . . . . . . . . 17.12 The test lter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.13 Stresses on grid, test and intermediate level . . . . . . . . . . . . . . 17.14 Numerical dissipation . . . . . . . . . . . . . . . . . . . . . . . . . 17.15 Scale-similarity Models . . . . . . . . . . . . . . . . . . . . . . . . 17.16 The Bardina Model . . . . . . . . . . . . . . . . . . . . . . . . . . 17.17 Redened terms in the Bardina Model . . . . . . . . . . . . . . . . 17.18 A dissipative scale-similarity model. . . . . . . . . . . . . . . . . . 17.19 Forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.20 Numerical method . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.20.1 RANS vs. LES . . . . . . . . . . . . . . . . . . . . . . . . 17.21 One-equation ksgs model . . . . . . . . . . . . . . . . . . . . . . . 17.22 Smagorinsky model derived from the k 5/3 law . . . . . . . . . . . 17.23 A dynamic one-equation model . . . . . . . . . . . . . . . . . . . . 17.24 A Mixed Model Based on a One-Eq. Model . . . . . . . . . . . . . 17.25 Applied LES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.26 Resolution requirements . . . . . . . . . . . . . . . . . . . . . . . . 18 Unsteady RANS 18.1 Turbulence Modelling . . . . . . . . . . . . . . . . . . . . . . . . . 18.2 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 DES 19.1 DES based on two-equation models . . . . . . . . . . . . . . . . . . 19.2 DES based on the k SST model . . . . . . . . . . . . . . . . . 20 Hybrid LES-RANS 20.1 Momentum equations in hybrid LES-RANS . . . . . . . . . . . . . 20.2 The equation for turbulent kinetic energy in hybrid LES-RANS . . . 20.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 The SAS model 21.1 Resolved motions in unsteady . . . . . . . . . . . . . . . 21.2 The von K rm n length scale . . . . . . . . . . . . . . . . a a 21.3 The second derivative of the velocity . . . . . . . . . . . . 21.4 Evaluation of the von K rm n length scale in channel ow a a 22 The PANS Model . . . . . . . . . . . . . . . . . . . .

119 119 120 121 122 123 123 124 124 125 125 126 127 128 129 130 131 131 132 133 134 135 136 136 137 138 138 138 140 143 143 144 145 147 147 152 152 153 154 154 154 156 156 158

23 Inlet boundary conditions 23.1 Synthesized turbulence . . . . 23.2 Random angles . . . . . . . . . 23.3 Highest wave number . . . . . 23.4 Smallest wave number . . . . . 23.5 Divide the wave number range 23.6 von K rm n spectrum . . . . . a a 23.7 Computing the uctuations . . 23.8 Introducing time correlation . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

161 161 162 162 162 162 162 163 164 166 166 166 167 167 168

24 Best practice guidelines (BPG) 24.1 EU projects . . . . . . . . . . . . . . . . 24.2 Ercoftac workshops . . . . . . . . . . . . 24.3 Ercoftac Classical Database . . . . . . . . 24.4 ERCOFTAC QNET Knowledge Base Wiki A TME225: identity

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

B TME225 Assignment 1: laminar ow B.1 Fully developed region . . . . . . . . . . . . B.2 Wall shear stress . . . . . . . . . . . . . . . . B.3 Inlet region . . . . . . . . . . . . . . . . . . . B.4 Wall-normal velocity in the developing region B.5 Vorticity . . . . . . . . . . . . . . . . . . . . B.6 Deformation . . . . . . . . . . . . . . . . . . B.7 Dissipation . . . . . . . . . . . . . . . . . . . B.8 Eigenvalues . . . . . . . . . . . . . . . . . . B.9 Eigenvectors . . . . . . . . . . . . . . . . . . C TME225: Fourier series C.1 Orthogonal functions . . . . . . C.2 Trigonometric functions . . . . . C.3 Fourier series of a function . . . C.4 Derivation of Parsevals formula C.5 Complex Fourier series . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

169 169 170 170 170 171 171 171 171 172 173 173 174 176 176 178 179 179 179 181 183 184 184 184 185 185 186 186 186

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . . . . . .

D TME225: Compute energy spectra from LES/DNS data using Matlab D.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . D.2 An example of using FFT . . . . . . . . . . . . . . . . . . . . . . D.3 Energy spectrum from the two-point correlation . . . . . . . . . . D.4 Energy spectra from the autocorrelation . . . . . . . . . . . . . . . E TME225 Assignment 2: turbulent ow E.1 Time history . . . . . . . . . . . . . . . E.2 Time averaging . . . . . . . . . . . . . E.3 Mean ow . . . . . . . . . . . . . . . . E.4 The time-averaged momentum equation E.5 Wall shear stress . . . . . . . . . . . . . E.6 Resolved stresses . . . . . . . . . . . . E.7 Fluctuating wall shear stress . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

E.8 E.9 E.10 E.11

Production terms . . Pressure-strain terms Dissipation . . . . . Do something fun! .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

186 186 187 187 199 200 202 202 203 204 205 207 207 207 208 209 209 211 212 213 213 214 214 215 216 216 217 218 218 218 218 219 219 219 219 220 220 221 221 223 224 225

F MTF270: Some properties of the pressure-strain term G MTF270: Galilean invariance H MTF270: Computation of wavenumber vector and angles H.1 The wavenumber vector, n . . . . . . . . . . . . . . . . . . . . . . j n H.2 Unit vector i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . I J MTF270: 1D and 3D energy spectra I.1 Energy spectra from two-point correlations . . . . . . . . . . . . . . MTF270, Assignment 1: Reynolds averaged Navier-Stokes J.1 Two-dimensional ow . . . . . . . . . . . . . . . . . . J.2 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . J.2.1 The momentum equations . . . . . . . . . . . . J.2.2 The turbulent kinetic energy equation . . . . . J.2.3 The Reynolds stress equations . . . . . . . . . J.3 Compute derivatives on a curvi-linear mesh . . . . . . . J.3.1 Geometrical quantities . . . . . . . . . . . . . LES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

K MTF270, Assignment 2: K.1 Task 2.1 . . . . . K.2 Task 2.2 . . . . . K.3 Task 2.3 . . . . . K.4 Task 2.4 . . . . . K.5 Task 2.5 . . . . . K.6 Task 2.6 . . . . . K.7 Task 2.7 . . . . . K.8 Task 2.9 . . . . . K.9 Task 2.10 . . . . . K.10 Task 2.11 . . . . . K.11 Task 2.12 . . . . .

L MTF270, Assignment 4: Hybrid LES-RANS L.1 Time history . . . . . . . . . . . . . . . L.2 Mean velocity prole . . . . . . . . . . L.3 Resolved stresses . . . . . . . . . . . . L.4 Turbulent kinetic energy . . . . . . . . . L.5 The modelled turbulent shear stress . . . L.6 Turbulent length scales . . . . . . . . . L.7 SAS turbulent length scales . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

M MTF270: Transformation of a tensor M.1 Rotation to principal directions . . . . . . . . . . . . . . . . . . . . M.2 Transformation of a velocity gradient . . . . . . . . . . . . . . . . .

N MTF270: Greens formulas N.1 Greens rst formula . . . . . . . . . . . N.2 Greens second formula . . . . . . . . . N.3 Greens third formula . . . . . . . . . . N.4 Analytical solution to Poissons equation O MTF270: Learning outcomes P References

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

226 226 226 226 229 230 237

TME225 Mechanics of uids


L. Davidson
Division of Fluid Dynamics, Department of Applied Mechanics Chalmers University of Technology, G teborg, Sweden o http://www.tfd.chalmers.se/lada, lada@chalmers.se
This report can be downloaded at http://www.tfd.chalmers.se/lada/MoF/

1. Motion, ow
T (Xi , t1 ) Xi T (x1 , t1 ) i T (Xi , t2 )

T (x2 , t2 ) i Figure 1.1: The temperature of a uid particle described in Lagrangian, T (Xi , t), or Eulerian, T (xi , t), approach.

1 Motion, ow
1.1 Eulerian, Lagrangian, material derivative
See also [2], Chapt. 3.2. Assume a uid particle is moving along the line in Fig. 1.1. We can choose to study its motion in two ways: Lagrangian or Eulerian. In the Lagrangian approach we keep track of its original position (Xi ) and follow its path which is described by xi (Xi , t). For example, at time t1 the temperature of the particle is T (Xi , t1 ), and at time t2 its temperature is T (Xi , t2 ), see Fig. 1.1. This approach is not used for uids because it is very tricky to dene and follow a uid particle. It is however used when simulating movement of particles in uids (for example soot particles in gasoline-air mixtures in combustion applications). The speed of the particle, for example, is then expressed as a function of time and its position at time zero, i.e. vi = vi (Xi , t). In the Eulerian approach we pick a position, e.g. x1 , and watch the particle pass i by. This approach is used for uids. The temperature of the uid, T , for example, is expressed as a function of the position, i.e. T = T (xi ), see Fig. 1.1. It may be that the temperature at position xi , for example, varies in time, t, and then T = T (xi , t). Now we want to express how the temperature of a uid particle varies. In the Lagrangian approach we rst pick the particle (this gives its starting position, Xi ). If the starting position is xed, temperature varies only with time, i.e. T (t) and the temperature gradient can be written dT /dt. In the Eulerian approach it is a little bit more difcult. We are looking for the temperature gradient, dT /dt, but since we are looking at xed points in space we need to express the temperature as a function of both time and space. From classical mechanics, we know that the velocity of a uid particle is the time derivative of its space location, i.e. vi = dxi /dt. The chain-rule now gives dT T T T dxj T = = + + vj dt t dt xj t xj (1.1)

Note that we have to use partial derivative on T since it is a function of more than one (independent) variable. The rst term on the right side is the local rate of change; by this we mean that it describes the variation of T in time at position xi . The second term on the right side is called the convective rate of change, which means that it describes

local rate of change Conv. rate of change

1.2. Viscous stress, pressure

10

x2

1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0
x1

12 11 13

Figure 1.2: Denition of stress components on a surface. the variation of T in space when is passes the point xi . The left side in Eq. 1.1 is called the material derivative and is in this text denoted by dT /dt. Equation 1.1 can be illustrated as follows. Put your nger out in the blowing wind. The temperature gradient youre nger experiences is T /t. Imagine that youre a uid particle and that you ride on a bike. The temperature gradient you experience is the material derivative, dT /dx. Exercise 1 Write out Eq. 1.1, term-by-term.

Material derivative

1.2 Viscous stress, pressure


See also [2], Chapts. 6.3 and 8.1. We have in Part I [3] derived the balance equation for linear momentum which reads vi ji,j fi = 0 (1.2) Switch notation for the material derivative and derivatives so that dvi ji + fi = dt xj (1.3)

where the rst and the second term on the right side represents, respectively, the net force due to surface and volume forces (ij denotes the stress tensor). Stress is force per unit area. The rst term includes the viscous stress tensor, ij . As you have learnt earlier, the rst index relates to the surface at which the stress acts and the second index is related to the stress component. For example, on a surface whose normal is ni = (1, 0, 0) act the three stress components 11 , 12 and 13 , see Fig. 1.2. In the present notation we denote the velocity vector by v = vi = (v1 , v2 , v3 ) and the coordinate by x = xi = (x1 , x2 , x3 ). In the literature, you may nd other notations of the velocity vector such as ui = (u1 , u2 , u3 ). If no tensor notation is used the velocity vector is usually denoted as (u, v, w) and the coordinates as (x, y, z). The diagonal components of ij represent normal stresses and the off-diagonal components of ij represent the shear stresses. In Part I [3]you learnt that the pressure is dened as minus the sum of the normal stress, i.e. p = kk /3. The pressure, p, acts as a normal stress. In general, pressure is a thermodynamic property, pt , which can

1.3. Strain rate tensor, vorticity

11

be obtained for example from the ideal gas law. In that case the thermodynamics pressure, pt , and the mechanical pressure, p, may not be the same. The viscous stress tensor, ij , is obtained by subtracting the trace, kk /3 = p, from ij ; the stress tensor can then be written as ij = pij + ij (1.4) ij is the deviator of ij . The expression for the viscous stress tensor is found in Eq. 2.4 at p. 19. The minus-sign in front of p appears because the pressure acts into the surface. When theres no movement, the viscous stresses are zero and then of course the normal stresses are the same as the pressure. In general, however, the normal stresses are the sum of the pressure and the viscous stresses, i.e. 11 = p + 11 , 22 = p + 22 , 33 = p + 33 , (1.5)

Exercise 2 Consider Fig. 1.2. Show how 21 , 22 , 23 act on a surface with normal vector ni = (0, 1, 0). Show also how 31 , 32 , 33 act on a surface with normal vector ni = (0, 0, 1). Exercise 3 Write out Eq. 1.4 on matrix form.

1.3 Strain rate tensor, vorticity


See also [2], Chapt. 3.6. The velocity gradient tensor can be split into two parts as vi 1 vi vi vj vj + = + xj 2 xj xj xi xi
2vi /xj =0

(1.6) = Sij + ij

= where

1 2

vi vj + xj xi

1 2

vi vj xj xi

Sij is a symmetric tensor called the strain-rate tensor ij is a anti-symmetric tensor called the vorticity tensor The vorticity tensor is related to the familiar vorticity vector which is the curl of the velocity vector, i.e. = v, or in tensor notation i = ijk If we set, for example, i = 3 we get 3 = v2 /x1 v1 /x2 . (1.8) vk xj (1.7)

Strain-rate tensor vorticity tensor

The vorticity represents rotation of a uid particle. Inserting Eq. 1.6 into Eq. 1.7 gives i = ijk (Skj + kj ) = ijk kj (1.9)

1.3. Strain rate tensor, vorticity

12

since ijk Skj = 0 because the product of a symmetric tensor (Skj ) and a anti-symmetric tensor (ijk ) is zero. Let us show this for i = 1 by writing out the full equation. Recall that Sij = Sji (i.e. S12 = S21 , S13 = S31 , S23 = S32 ) and ijk = ikj = jki etc (i.e. 123 = 132 = 231 . . . , 113 = 221 = . . . 331 = 0) 1jk Skj = 111 S11 + 112 S21 + 113 S31 + 121 S12 + 122 S22 + 123 S32 + 131 S13 + 132 S23 + 133 S33 = 0 S11 + 0 S21 + 0 S31 (1.10)

+ 0 S12 + 0 S22 + 1 S32 + 0 S13 1 S23 + 0 S33 = S32 S23 = 0

Now les us invert Eq. 1.9. We start by multiplying it with im so that im i = im ijk kj The --identity gives (see Table A.1 at p. A.1) im ijk kj = (j mk k mj )kj = m m = 2m (1.12) (1.11)

This can easily be proved by writing all the components, see Table A.1 at p. A.1. Hence we get with Eq. 1.7 m = or, switching indices 1 1 1 im i = mi i = mi i 2 2 2 (1.13)

1 ij = ijk k (1.14) 2 A much easier way to go from Eq. 1.9 to Eq. 1.14 is to write out the components of Eq. 1.9. Here we do it for i = 1 1 = 123 32 + 132 23 = 32 23 = 223 and we get 1 23 = 1 2 which indeed is identical to Eq. 1.14. (1.15)

(1.16)

Exercise 4 Write out the second and third component of the vorticity vector given in Eq. 1.7 (i.e. 2 and 3 ). Exercise 5 Complete the proof of Eq. 1.10 for i = 2 and i = 3. Exercise 6 Write out Eq. 1.15 also for i = 2 and i = 3 and nd an expression for 12 and 13 (cf. Eq. 1.16). Show that you get the same result as in Eq. 1.14. Exercise 7 In Eq. 1.16 we proved the relation between ij and i for the off-diagonal components. What about the diagonal components of ij ? What do you get from Eq. 1.6?

1.4. Deformation, rotation


x1 v1 x2 t x2

13

x2 v2 x1 t x1

x2 x1

Figure 1.3: Rotation of a uid particle during time t. Here v1 /x2 = v2 /x1 so that 12 = 3 /2 = v2 /x1 > 0. Exercise 8 From you course in linear algebra, you should remember how to compute a vector product using Sarrus rule. Use it to compute the vector product 1 e 2 e 3 e = v = x1 x2 x3 v1 v2 v3

Verify that this agrees with the expression in tensor notation in Eq. 1.7.

1.4 Deformation, rotation


See also [2], Chapt. 3.3. The velocity gradient can, as shown above, be divided into two parts: Sij and ij . We have shown that the latter is connected to rotation of a uid particle. During rotation the uid particle is not deformed. This movement can be illustrated by Fig. 1.3. It is assumed that the uid particle is rotated the angle during the time t. The vorticity during this rotation is 3 = v2 /x1 v1 /x2 = 212 . The vorticity 3 should be interpreted as twice the average rotation of the horizontal edge (v2 /x1 ) and vertical edge (v1 /x2 ). Next let us have a look at the deformation caused by Sij . It can be divided into two parts, namely shear and elongation (also called extension or dilatation). The deformation due to shear is caused by the off-diagonal terms of Sij . In Fig. 1.4, a pure shear deformation by S12 = (v1 /x2 + v2 /x1 )/2 is shown. The deformation due to elongation is caused by the diagonal terms of Sij . Elongation caused by S11 = v1 /x1 is illustrated in Fig. 1.5.

rotation

1.4. Deformation, rotation

14

x1 v1 x2 t x2

x2 v2 x1 t x1

x2 x1

Figure 1.4: Deformation of a uid particle by shear during time t. Here v1 /x2 = v2 /x1 so that S12 = v1 /x2 > 0. x1 v1 x1 t x1

x2

x2 x1 Figure 1.5: Deformation of a uid particle by elongation during time t. In general, a uid particle experiences a combination of rotation, deformation and elongation as indeed is given by Eq. 1.6. Exercise 9 Consider Fig. 1.3. Show and formulate the rotation by 1 .

1.5. Irrotational and rotational ow


ti d

15

x2 S x1

Figure 1.6: The surface, S, is enclosing by the line . The vector, ti , denotes the unit tangential vector of the enclosing line, . Exercise 10 Consider Fig. 1.4. Show and formulate the deformation by S23 . Exercise 11 Consider Fig. 1.5. Show and formulate the elongation by S22 .

1.5 Irrotational and rotational ow


In the previous subsection we introduced different types of movement of a uid particle. One type of movement was rotation, see Fig. 1.3. Flows are often classied based on rotation: they are rotational (i = 0) or irrotational (i = 0); the latter type is also called inviscid ow or potential ow. Well talk more about that later on. In this subsection we will give examples of one irrotational and one rotational ow. In potential ow, there exists a potential, , from which the velocity components can be obtained as (1.17) vk = xk Consider a closed line on a surface in the x1 x2 plane, see Fig. 1.6. When the velocity is integrated along this line and projected onto the line we obtain the circulation = vm tm d (1.18)

Using Stokess theorem we can relate the circulation to the vorticity as =

vm tm d =
S

ijk

vk ni dS = xj

3 dS
S

(1.19)

where ni = (0, 0, 1) is the unit normal vector of the surface S. Equation 1.19 reads in vector notation =

v td =

( v) ndS =

3 dS
S

(1.20)

The circulation is useful in aeronautics and windpower engineering where the lift of an airfoil or a rotorblade is expressed in the circulation for a 2D section. The lift force is computed as L = V (1.21) where V is the velocity around the airfoil (for a rotorblade it is the relative velocity, since the rotorblade is rotating). In a recent MSc thesis project, an inviscid simulation method (based on the circulation and vorcitity sources) was used to compute the aerodynamic loads for windturbines [4].

1.5. Irrotational and rotational ow


v x2 r x1

16

Figure 1.7: Transformation of v into Cartesian components. Exercise 12 In potential ow i = ijk vk /xj = 0. Multiply Eq. 1.17 by ijk and derivate with respect to xk (i.e. take the curl of) and show that the right side becomes zero as it should, i.e. ijk 2 /(xk xj ) = 0. 1.5.1 Ideal vortex line The ideal vortex line is an irrotational (potential) ow where the uid moves along circular paths, see Fig. 1.8. The velocity eld in polar coordinates reads v = , 2r vr = 0 (1.22)

where is the circulation. Its potential reads = The velocity, v , is then obtained as v = 1 r 2 (1.23)

To transform Eq. 1.22 into Cartesian velocity components, consider Fig. 1.7. The Cartesian velocity vectors are expressed as v1 = v sin() = v x2 x2 = v 2 r (x1 + x2 )1/2 2 x1 x1 v2 = v cos() = v = v 2 r (x1 + x2 )1/2 2

(1.24)

Inserting Eq. 1.24 into Eq. 1.22 we get v1 = x2 , 2(x2 + x2 ) 1 2 v2 = x1 . 2(x2 + x2 ) 1 2 (1.25)

To verify that this ow is a potential ow, we need to show that the vorticity, i = ijk vk /xj is zero. Since it is a two-dimensional ow (v3 = /x3 = 0), 1 =

1.5. Irrotational and rotational ow


a

17

Figure 1.8: Ideal vortex. The uid particle (i.e. its diagonal, see Fig. 1.3) does not rotate. 2 = 0, we only need to compute 3 = v2 /x1 v1 /x2 . The velocity derivatives are obtained as x2 x2 v1 1 2 = , x2 2 (x2 + x2 )2 1 2 and we get 3 = 1 (x2 x2 + x2 x2 ) = 0 1 1 2 2 (x2 + x2 )2 2 1 2 (1.27) x2 x2 v2 2 1 = x1 2 (x2 + x2 )2 1 2 (1.26)

which shows that the ow is indeed a potential ow, i.e. irrotational (i 0). Note that the deformation is not zero, i.e. S12 = 1 2 v1 v2 + x2 x1 = x2 2 2 + x2 )2 2 (x1 2 (1.28)

Hence a uid particle in an ideal vortex does deform but it does not rotate (i.e. its diagonal does not rotate, see Fig. 1.8). It may be little confusing that the ow path forms a vortex but the ow itself has no vorticity. Thus one must be very careful when using the words vortex and vorticity. By vortex we usually mean a recirculation region of the mean ow. That the ow has no vorticity (i.e. no rotation) means that a uid particle moves as illustrated in Fig. 1.8. As a uid particle moves from position a to b on its counter-clockwise-rotating path the particle itself is not rotating. This is true for the whole ow eld, except at the center where the uid particle does rotate. This is a singular point as is seen from Eq. 1.22 for which 3 . Note that generally a vortex has vorticity, see Section 4.2. The ideal vortex is a very special ow case. 1.5.2 Shear ow Another example which is rotational is a shear ow in which v1 = cx2 , 2 v2 = 0 (1.29)

vortex vs. vorticity

with c, x2 > 0, see Fig. 1.9. The vorticity vector for this ow reads 1 = 2 = 0, 3 = v1 v2 = 2cx2 x1 x2 (1.30)

1.6. Eigenvalues and and eigenvectors: physical interpretation

18

v1 a

v1 c

x2 a x3 x1

Figure 1.9: A shear ow. The uid particle rotates. v1 = cx2 . 2 23 21 12 x2 x1 Figure 1.10: A two-dimensional uid element. Left: in original state; right: rotated to principal coordinate directions. 1 and 2 denote eigenvalues; v1 and v2 denote unit eigenvectors. When the uid particle is moving from position a, via b to position c it is indeed rotating. It is rotating in clockwise direction. Note that the positive rotating direction is dened as the counter-clockwise direction, indicated by in Fig. 1.9. This is why the vorticity, 3 , is negative (= 2cx2 ). 11 v2 2 1 x2 x1 v1

1.6 Eigenvalues and and eigenvectors: physical interpretation


See also [2], Chapt. 2.5.5. Consider a two-dimensional uid (or solid) element, see Fig. 1.10. In the left gure it is oriented along the x1 x2 coordinate system. On the surfaces act normal stresses (11 , 22 ) and shear stresses (12 , 21 ). The stresses form a tensor, ij . Any tensor has eigenvectors and eigenvalues (also called principal vectors and principal values). Since ij is symmetric, the eigenvalues are real (i.e. not imaginary). The eigenvalues are obtained from the characteristic equation, see [2], Chapt. 2.5.5 or Eq. 13.5 at p. 105. When the eigenvalues have been obtained, the eigenvectors can be computed. Given the eigenvectors, the uid element is rotated degrees so that its edges are aligned with the eigenvectors, v1 = x1 and v2 = x2 , see right part of Fig. 1.10. Note that the the sign of the eigenvectors is not dened, which means that the eigenvectors can equally well be chosen as 1 and/or 2 . In the principal coordinates (i.e. the v v uid element aligned with the eigenvectors), there are no shear stresses on surface of the uid element. There are only normal stresses. This is the very denition of

2. Governing ow equations

19

eigenvectors. Furthermore, the eigenvalues are the normal stresses in the principal coordinates, i.e. 1 = 1 1 and 2 = 2 2 .

2 Governing ow equations
See also [2], Chapts. 5 and 8.1.

2.1 The Navier-Stokes equation


2.1.1 The continuity equation The rst equation is the continuity equation (the balance equation for mass) which reads [3] + vi,i = 0 (2.1) Change of notation gives vi d =0 + dt xi For incompressible ow ( = const) we get vi =0 xi 2.1.2 The momentum equation The next equation is the momentum equation. We have formulated the constitutive law for Newtonian viscous uids [3] 2 ij = pij + 2Sij Skk ij 3 Inserting Eq. 2.4 into the balance equations, Eq. 1.3, we get ji p p dvi + + fi = + = dt xi xj xi xj 2 vk 2Sij ij 3 xk + fi (2.5) (2.4) (2.3) (2.2)

where denotes the dynamic viscosity. This is the Navier-Stokes equations (sometimes the continuity equation is also included in the name Navier-Stokes). It is also called the transport equation for momentum. If the viscosity, , is constant it can be moved outside the derivative. Furthermore, if the ow is incompressible the second term in the parenthesis on the right side is zero because of the continuity equation. If these two requirements are satised we can also re-write the rst term in the parenthesis as (2Sij ) = xj xj vj vi + xj xi = 2 vi xj xj (2.6)

because of the continuity equation. Equation 2.5 can now for constant and incompressible ow be written 2 vi p dvi + + fi = dt xi xj xj (2.7)

2.2. The energy equation

20

In inviscid (potential) ow, there are no viscous (friction) forces. In this case, the Navier-Stokes equation reduces to the Euler p dvi + fi = dt xi (2.8)

Euler

Exercise 13 Formulate the Navier-Stokes equation for incompressible ow but nonconstant viscosity.

2.2 The energy equation


See also [2], Chapts. 6.4 and 8.1. We have in Part I [3] derived the energy equation which reads u vi,j ji + qi,i = z (2.9)

where u denotes internal energy. qi denotes the conductive heat ux and z the net radiative heat source. The latter can also be seen as a vector, zi,rad ; for simplicity, we neglect the radiation from here on. Change of notation gives vi qi du = ji dt xj xi (2.10)

In Part I [3]we formulated the constitutive law for the heat ux vector (Fouriers law) T (2.11) qi = k xi Inserting the constitutive laws, Eqs. 2.4 and 2.11, into Eq. 2.10 gives du 2 vi + 2Sij Sij Skk Sii + = p dt xi 3 xi

T xi

(2.12)

where we have used Sij vi /xj = Sij (Sij + ij ) = Sij Sij because the product of a symmetric tensor, Sij , and an anti-symmetric tensor, ij , is zero. Two of the viscous terms (denoted by ) represent irreversible viscous heating (i.e. transformation of kinetic energy into thermal energy); these terms are important at high-speed ow1 (for example re-entry from outer space) and for highly viscous ows (lubricants). The rst term on the right side represents reversible heating and cooling due to compression and expansion of the uid. Equation 2.12 is the transport equation for (internal) energy, u. Now we assume that the ow is incompressible for which du = cp dT (2.13)

where cp is the heat capacity (see Part I) [3] so that Eq. 2.12 gives (cp is assumed to be constant) T dT k (2.14) =+ cp dt xi xi
1 High-speed ows relevant for aeronautics will be treated in detail in the course Compressible ow in the MSc programme.

2.3. Transformation of energy

21

The dissipation term is simplied to = 2Sij Sij because Sii = vi /xi = 0. If we furthermore assume that the heat conductivity coefcient is constant and that the uid is a gas or a common liquid (i.e. not an lubricant oil), we get dT 2T = dt xi xi (2.15) thermal diffusivity

where = k/(cp ) is the thermal diffusivity. From the thermal diffusivity, the Prandtl number (2.16) Pr = is dened where = / is the kinematic viscosity. The physical meaning of the Prandtl number is the ratio of how well the uid diffuses momentum to the how well it diffuses internal energy (i.e. temperature). Exercise 14 Write out and simplify the dissipation term, , in Eq. 2.12. The rst term is positive and the second term is negative; are you sure that > 0?

2.3 Transformation of energy


Now we will derive the equation for the kinetic energy, k = v 2 /2. Multiply Eq. 1.3 with vi ji dvi vi fi = 0 (2.17) vi vi dt xj The rst term on the left side can be re-written vi (vi vi /2 = v 2 /2 = k) so that dk ji + vi fi = vi dt xj (2.19) dvi 1 d(vi vi ) dk = = dt 2 dt dt (2.18)

Re-write the stress-velocity term so that vi vi ji dk ji + vi fi = dt xj xj (2.20)

This is the transport equation for kinetic energy, k. Adding Eq. 2.20 to Eq. 2.10 gives qi ji vi d(u + k) + vi fi = dt xj xi (2.21)

This is an equation for the sum of internal and kinetic energy, u + k. This is the transport equation for total energy, u + k. Let us take a closer look at Eqs. 2.10, 2.20 and 2.21. First we separate the term ji vi /xj in Eqs. 2.10 and 2.20 into work related to the pressure and viscous stresses respectively (see Eq. 1.4), i.e. ji vi vi vi + ji = p xj xi xj
a b=

(2.22)

The following things should be noted.

2.4. Left side of the transport equations

22

The physical meaning of the a-term in Eq. 2.22 which include the pressure, p is heating/cooling by compression/expansion. This is a reversible process, i.e. no loss of energy but only transformation of energy. The physical meaning of the b-term in Eq. 2.22 which include the viscous stress tensor, ij is a dissipation, which means that kinetic energy is transformed to thermal energy. It is denoted , see Eq. 2.12, and is called viscous dissipation. It is always positive and represents irreversible heating. The dissipation, , appears as a sink term in the equation for the kinetic energy, k (Eq. 2.20) and it appears a source term in the equation for the internal energy, u (Eq. 2.10). The transformation of kinetic energy into internal energy takes place through this source term. does not appear in the equation for the total energy u+k (Eq. 2.21); this makes sense since represents a energy transfer between u and k and does not affect their sum, u + k. Dissipation is very important in turbulence where transfer of energy takes place at several levels. First energy is transferred from the mean ow to the turbulent uctuations. The physical process is called production of turbulent kinetic energy. Then we have transformation of kinetic energy from turbulence kinetic energy to thermal energy; this is turbulence dissipation (or heating). At the same time we have the usual viscous dissipation from the mean ow to thermal energy, but this is much smaller than that from the turbulence kinetic energy. For more detail, see section 2.4 in [5]2 .

2.4 Left side of the transport equations


So far, the left side in transport equations have been formulated using the material derivative, d/dt. Let denote a transported quantity (i.e. = vi , u, T . . .); the left side of the equation for momentum, thermal energy, total energy, temperature etc reads d = + vj dt t xj (2.23) nonconservative

This is often called the non-conservative form. Using the continuity equation, Eq. 2.2, it can be re-written as d + = + vj dt t xj vj d + dt xj
=0

= (2.24)

+ vj + t xj

vj + vj + t xj xj

The two underlined terms will form a time derivative term, and the other three terms can be collected into a convective term, i.e. vj d = + dt t xj (2.25) conservative

This is called the conservative form. With = 1 we get the continuity equation. When
2 can

be downloaded from http://www.tfd.chalmers.se/lada

2.5. Material particle vs. control volume (Reynolds Transport Theorem)

23

solving the Navier-Stokes equations numerically using nite volume methods, the left side in the transport equation is always written in the form of Eq. 2.25; in this way we ensure that the transported quantity is conserved. The results may be inaccurate due to too coarse a numerical grid, but no mass, momentum, energy etc is lost (provided a transport equation for the quantity is solved): what comes in goes out.

2.5 Material particle vs. control volume (Reynolds Transport Theorem)


See also [2], Chapt. 5.2. In Part I [3]we initially derived all balance equations (mass, momentum and energy) for a collection of material particles. The conservation of mass, d/dt dV = 0, Newtons second law, d/dt vi = Fi etc were derived for a collection of particles in the volume Vpart , where Vpart is a volume that includes the same uid particles all the time. This means that the volume, Vpart , must be moving, otherwise particles would move across its boundaries. The equations we have looked at so far (the continuity equation 2.3, the Navier-Stokes equation 2.7, the energy equations 2.12 and 2.20) are all given for a xed control volume. How come? The answer is the Reynolds transport theorem, which converts the equations from being valid for a moving volume with a collection, Vpart , to being valid for a xed volume, V . The Reynolds transport theorem reads d dt =
V

dV =
Vpart V

vi d + dt xi vi + t xi

dV (2.26) dV

vi + + vi t xi xi

dV =
V

where V denotes a xed volume in space. In order to keep the uid particles in the volume during the time derivation the volume may change; this is accounted for by the dilatation (divergence) term vi /xi which appears on the right side of the rst line. in the equation above can be (mass), vi (momentum) or u (energy). This equation applies to any volume at every instant and the restriction to a collection of a material particles is no longer necessary. Hence, in uid mechanics the transport equations (Eqs. 2.2, 2.5, 2.10, . . . ) are valid both for a material particle as well as for a volume; the latter is usually xed (this is not necessary).

3 Exact solutions to the Navier-Stokes equation: two examples


3.1 The Rayleigh problem
Imagine the sudden motion of an innitely long at plate. For time greater than zero the plate is moving with the speed V0 , see Fig. 3.1. Because the plate is innitely long, there is no x1 dependency. Hence the ow depends only on x2 and t, i.e. v1 = v1 (x2 , t) and p = p(x2 , t). Furthermore, v1 /x1 = v3 /x3 = 0 so that the continuity equation gives v2 /x2 = 0. At the lower boundary (x2 = 0) and at the upper boundary (x2 ) the velocity component v2 = 0, which means that v2 = 0 in the entire domain. So, Eq. 2.7 gives (no

3.1. The Rayleigh problem


x2 x1 V0

24

Figure 3.1: The plate moves to the right with speed V0 for t > 0.

x2 t2
0.2

t3

t1
0

0.4

0.6

0.8

v1 /V0 Figure 3.2: The v1 velocity at three different times. t3 > t2 > t1 . body forces, i.e. f1 = 0) for the v1 velocity component 2 v1 v1 = 2 t x2 (3.1)

We will nd that the diffusion process depends on the kinematic viscosity, = /, rather than the dynamic one, . The boundary conditions for Eq. 3.1 are v1 (x2 , t = 0) = 0, v1 (x2 = 0, t) = V0 , v1 (x2 , t) = 0 (3.2)

The solution to Eq. 3.1 is shown in Fig. 3.2. For increasing time (t3 > t2 > t1 ), the moving plate affects the uid further and further away from the plate. It turns out that the solution to Eq. 3.1 is a similarity solution; this means that the number of independent variables is reduced by one, in this case from two (x2 and t) to one (). The similarity variable, , is related to x2 and t as x2 = 2 t (3.3)

similarity solution

If the solution of Eq. 3.1 depends only on , it means that the solution for a given uid will the same (similar) for many (innite) values of x2 and t as long as the ratio be x2 / t is constant. Now we need to transform the derivatives in Eq. 3.1 from /t and /x2 to d/d so that it becomes a function of only. We get dv1 x2 t3/2 dv1 1 dv1 v1 = = = t d t 4 d 2 t d v1 dv1 1 dv1 = = x2 d x2 2 t d 2 v1 1 dv1 1 v1 = = = x2 x2 x2 x2 2 t d 2 t x2 2

dv1 d

1 d2 v1 4t d 2 (3.4)

3.1. The Rayleigh problem

25

He introduce a non-dimensional velocity f= Inserting Eqs. 3.4 and 3.5 in Eq. 3.1 gives d2 f df + 2 =0 2 d d (3.6) v1 V0 (3.5)

We have now successfully transformed Eq. 3.1 and reduced the number of independent variables from two to one. Now let us nd out if the boundary conditions, Eq. 3.2, also can be transformed in a physically meaningful way; we get v1 (x2 , t = 0) = 0 f ( ) = 0 v1 (x2 = 0, t) = V0 f ( = 0) = 1

(3.7)

v1 (x2 , t) = 0 f ( ) = 0 Since we managed to transform both the equation (Eq. 3.1) and the boundary conditions (Eq. 3.7) we conclude that the transformation is suitable. Now let us solve Eq. 3.6. Integration once gives df = C1 exp( 2 ) d Integration a second time gives

(3.8)

f = C1
0

exp( 2 )d + C2

(3.9)

The integral above is the error function 2 erf()


0

exp( 2 )

(3.10)

At the limits, the error function takes the values 0 and 1, i.e. erf(0) = 0 and erf( ) = 1. Taking into account the boundary conditions, Eq. 3.7, the nal solution to Eq. 3.9 is (with C2 = 1 and C1 = 2/ ) f () = 1 erf() (3.11)

The solution is presented in Fig. 3.3. Compare this gure with Fig. 3.2 at p. 24; all graphs in that gure collapse into one graph in Fig. 3.3. To compute the velocity, v1 , we pick a time t and insert x2 and t in Eq. 3.3. Then f is obtained from Eq. 3.11 and the velocity, v1 , is computed from Eq. 3.5. This is how the graphs in Fig. 3.2 were obtained. From the velocity prole we can get the shear stress as 21 = V0 df V0 v1 = = exp 2 x2 2 t d t (3.12)

where we used = /. Figure 3.4 below presents the shear stress, 21 . The solid line is obtained from Eq. 3.12 and circles are obtained by evaluating the derivative, df /d, numerically using central differences (fj+1 fj1 )/(j+1 j1 ).

3.1. The Rayleigh problem

26

3 2.5 2

1.5
1 0.5 0 0

0.2

0.4

0.6

0.8

f Figure 3.3: The velocity, f = v1 /V0 , given by Eq. 3.11.


3 2.5 2

1.5
1 0.5 0 6

21 /(V0 ) Figure 3.4: The shear stress for water ( = 106 ) obtained from Eq. 3.12 at time t = 100 000. As can be seen from Fig. 3.4, the magnitude of the shear stress increases for de creasing and it is largest at the wall, w = V0 / t The vorticity, 3 , across the boundary layer is computed from its denition (Eq. 1.30) V0 df V0 v1 = = exp( 2 ) (3.13) x2 2 t d t From Fig. 3.2 at p. 24 it is seen that for large times, the moving plate is felt further and further out in the ow, i.e. the thickness of the boundary layer, , increases. Often the boundary layer thickness is dened by the position where the local velocity, v1 (x2 ), reaches 99% of the freestream velocity. In our case, this corresponds to the point where v1 = 0.01V0 . From Fig. 3.3 and Eq. 3.11 we nd that this occurs at 3 = = 1.8 = 2 t = 3.6 t (3.14)

It can be seen that the boundary layer thickness increases with t1/2 . Equation 3.14 can also be used to estimate the diffusion length. After, say, 10 minutes the diffusion length for air and water, respectively, are air = 10.8cm water = 2.8cm (3.15)

diffusion length

3.2. Flow between two plates


V P1 P2 P1 V x2 x1 Figure 3.5: Flow in a horizontal channel. The inlet part of the channel is shown. h

27

As mentioned in the beginning of this section, note that the diffusion length is determined by the kinematic viscosity, = / rather than by dynamic one, . Exercise 15 Consider the graphs in Fig. 3.3. Create this graph with Matlab. Exercise 16 Consider the graphs in Fig. 3.2. Note that no scale is used on the x2 axis and that no numbers are given for t1 , t2 and t3 . Create this graph with Matlab for both air and engine oil. Choose suitable values on t1 , t2 and t3 . Exercise 17 Repeat the exercise above for the shear stress, 21 , see Fig. 3.4.

3.2 Flow between two plates


Consider steady, incompressible ow in a two-dimensional channel, see Fig. 3.5, with constant physical properties (i.e. = const). 3.2.1 Curved plates Provided that the walls at the inlet are well curved, the velocity near the walls is larger than in the center, see Fig. 3.5. The reason is that the ow (with velocity V ) following the curved wall must change its direction. The physical agent which accomplish this is the pressure gradient which forces the ow to follow the wall as closely as possible (if the wall is not sufciently curved a separation will take place). Hence the pressure in the center of the channel, P2 , is higher than the pressure near the wall, P1 . It is thus easier (i.e. less opposing pressure) for the uid to enter the channel near the walls than in the center. This explains the high velocity near the walls. The same phenomenon occurs in a channel bend, see Fig. 3.6. The ow V approaches the bend and the ow feels that it is approaching a bend through an increased pressure. The pressure near the outer wall, P2 , must be higher than that near the inner wall, P1 , in order to force the ow to turn. Hence, it is easier for the ow to sneak along the inner wall where the opposing pressure is smaller than near the outer wall: the result is a higher velocity near the inner wall than near the outer wall. In a threedimensional duct or in a pipe, the pressure difference P2 P1 creates secondary ow downstream the bend (i.e. a swirling motion in the x2 x3 plane).

3.2. Flow between two plates

28

P2 P1 x2 x1

Figure 3.6: Flow in a channel bend. 3.2.2 Flat plates The ow in the inlet section (Fig. 3.5) is two dimensional. Near the inlet the velocity is largest near the wall and further downstream the velocity is retarded near the walls due to the large viscous shear stresses there. The ow is accelerated in the center because the mass ow at each x1 must be constant because of continuity. The acceleration and retardation of the ow in the inlet region is paid for by a pressure loss which is rather high in the inlet region; if a separation occurs because of sharp corners at the inlet, the pressure loss will be even higher. For large x1 the ow will be fully developed; the region until this occurs is called the entrance region, and the entrance length can, for moderately disturbed inow, be estimated as [6] x1,e V Dh = 0.016ReDh 0.016 Dh (3.16)

where V denotes the bulk (i.e. the mean) velocity, and Dh = 4A/Sp where Dh , A and Sp denote the hydraulic diameter, the cross-sectional area and the perimeter, respectively. For ow between two plates we get Dh = 2h. Let us nd the governing equations for the fully developed ow region; in this region the ow does not change with respect to the streamwise coordinate, x1 (i.e. v1 /x1 = v2 /x1 = 0). Since the ow is two-dimensional, it does not depend on the third coordinate direction, x3 (i.e. /x3 ), and the velocity in this direction is zero, i.e. v3 = 0. Taking these restrictions into account the continuity equation can be simplied as (see Eq. 2.3) v2 =0 (3.17) x2 Integration gives v2 = C1 and since v2 = 0 at the walls, it means that v2 = 0 (3.18)

across the entire channel (recall that we are dealing with the part of the channel where the ow is fully developed; in the inlet section v2 = 0, see Fig. 3.5). Now let us turn our attention to the momentum equation for v2 . This is the vertical direction (x2 is positive upwards, see Fig. 3.5). The gravity acts in the negative x2 direction, i.e. fi = (0, g, 0). The momentum equation can be written (see Eq. 2.7 at p. 19) v2 v2 p 2 v2 dv2 + v2 = + 2 g (3.19) v1 dt x1 x2 x2 x2

3.2. Flow between two plates


Since v2 = 0 we get

29

p = g x2 p = gx2 + C1 (x1 )

(3.20) (3.21)

Integration gives where the integration constant C1 may be a function of x1 but not of x2 . If we denote the pressure at the lower wall (i.e. at x2 = 0) as P we get p = gx2 + P (x1 ) (3.22)

Hence the pressure, p, decreases with vertical height. This agrees with our experience that the pressure decreases at high altitudes in the atmosphere and increases the deeper we dive into the sea. Usually the hydrostatic pressure, P , is used in incompressible ow. This pressure is zero when the ow is static, i.e. when the velocity eld is zero. However, when you want the physical pressure, the gx2 as well as the surrounding atmospheric pressure must be added. We can now formulate the momentum equation in the streamwise direction v1 v1 dP 2 v1 dv1 + v2 = + 2 v1 dt x1 x2 dx1 x2 (3.23)

hydrostatic pressure

where p was replaced by P using Eq. 3.22. Since v2 = v1 /x1 = 0 the left side is zero so 2 v1 dP 2 = (3.24) x2 dx1 Since the left side is a function of x2 and the right side is a function of x1 , we conclude that they both are equal to a constant. The velocity, v1 , is zero at the walls, i.e. v1 (0) = v1 (h) = 0 Integrating Eq. 3.24 twice and using Eq. 3.25 gives v1 = x2 h dP x2 1 2 dx1 h (3.26) (3.25)

The minus sign on the right side appears because the pressure gradient is decreasing for increasing x1 ; the pressure is driving the ow. The negative pressure gradient is constant and can be written as dP/dx1 = P/L. The velocity takes its maximum in the center, i.e. for x2 = h/2, and reads v1,max = h P h 2 L 2 1 1 2 = h2 P 8 L (3.27)

We often write Eq. 3.26 on the form v1 x2 4x2 1 = v1,max h h (3.28)

The mean velocity (often called the bulk velocity) is obtained by integrating Eq. 3.28 across the channel, i.e. v1,mean = v1,max h
h 0

4x2 1

2 x2 dx2 = v1,max h 3

(3.29)

3.2. Flow between two plates

30

0.8

0.6

x2 /h
0.4

0.2

0 0

0.2

0.4

0.6

0.8

v1 /v1,max Figure 3.7: The velocity prole in fully developed channel ow, Eq. 3.28. The velocity prole is shown in Fig. 3.7 Since we know the velocity prole, we can compute the wall shear stress. Equation 3.26 gives v1 h dP h P w = = = (3.30) x2 2 dx1 2 L Actually, this result could have been obtained by simply taking a force balance of a slice of the ow far downstream. 3.2.3 Force balance To formulate a force balance in the x1 direction, we start with Eq. 1.3 which reads for i=1 dv1 j1 (3.31) = dt xj The left hand side is zero since the ow is fully developed. Forces act on a volume and its bounding surface. Hence we integrate Eq. 3.31 over the volume of a slice (length L), see Fig. 3.8 j1 0= dV (3.32) V xj Recall that this is the form on which we originally derived the the momentum balance (Newtons second law) in Part I. [3] Now use Gauss divergence theorem 0=
V

j1 dV = xj

j1 nj dS
S

(3.33)

The bounding surface consists in our case of four surfaces (lower, upper, left and right) so that 0=
Slef t

j1 nj dS +
Sright

j1 nj dS +
Slower

j1 nj dS +
Supper

j1 nj dS (3.34)

The normal vector on the lower, upper, left and right are ni,lower = (0, 1, 0), ni,upper = (0, 1, 0), ni,lef t = (1, 0, 0), ni,right = (1, 0, 0). Inserting the normal vectors and us-

3.2. Flow between two plates


L w,U

31

P1 walls x2 w,L x1

P2 h

Figure 3.8: Force balance of the ow between two plates. ing Eq. 1.4 give 0= (p + 11 )dS +
Slef t Sright

(p + 11 )dS

21 dS +
Slower Supper

21 dS

(3.35) 11 = 0 because v1 /x1 = 0 (fully developed ow). The shear stress at the upper and lower surfaces have opposite sign because w = (v1 /x2 )lower = (v1 /x2 )upper . Using this and Eq. 3.22 give (the gravitation term on the left and right surface cancels and P and w are constants and can thus be taken out in front of the integration) 0 = P1 W h P2 W h 2w LW (3.36)

where W is the width (in x3 direction) of the two plates (for convenience we set W = 1). With P = P1 P2 we get Eq. 3.30. 3.2.4 Balance equation for the kinetic energy In this subsection we will use the equation for kinetic energy, Eq. 2.20. Let us integrate this equation in the same way as we did for the force balance. The left side of Eq. 2.20 is zero because we assume that the ow is fully developed; using Eq. 1.4 gives 0= vi vi ji ji + vi fi xj xj
=0

vj p vi ji vi vi = + + pij ji xj xj xj xj

(3.37)

On the rst line vi fi = v1 f1 + v2 f2 = 0 because v2 = f1 = 0. The third term on the second line pij vi /xj = pvi /xi = 0 because of continuity. The last term corresponds to the viscous dissipation term, (i.e. loss due to friction), see Eq. 2.22 (term b). Now we integrate the equation over a volume 0=
V

pvj ji vi + dV xj xj

(3.38)

4. Vorticity equation and potential ow

32

Gauss divergence theorem on the two rst terms gives 0=


S

(pvj + ji vi )nj dS

dV
V

(3.39)

where S is the surface bounding the volume. The unit normal vector is denoted by nj which points out from the volume. For example, on the right surface in Fig. 3.8 it is nj = (1, 0, 0) and on the lower surface it is nj = (0, 1, 0). Now we apply Eq. 3.39 to the uid enclosed by the at plates in Fig. 3.8. The second term is zero on all four surfaces and the rst term is zero on the lower and upper surfaces (see Exercises below). We replace the pressure p with P using Eq. 3.22 so that (P v1 + gx2 v1 )n1 dS = (P2 P1 ) v1 n1 dS
Slef t &Sright

Slef t &Sright

= P v1,mean W h because gx2 n1 v1 on the left and right surfaces cancels; P can be taken out of the integral as it does not depend on x2 . Finally we get P = 1 W hv1,mean dV
V

(3.40)

Exercise 18 For the fully developed ow, compute the vorticity, i , using the exact solution (Eq. 3.28). Exercise 19 Show that the rst and second terms in Eq. 3.39 are zero on the upper and the lower surfaces in Fig. 3.8. Exercise 20 Show that the second term in Eq. 3.39 is zero also on the left and right surfaces Fig. 3.8. Exercise 21 Using the exact solution, compute the dissipation, , for the fully developed ow. Exercise 22 From the dissipation, compute the pressure drop. Is it the same as that obtained from the force balance (if not, nd the error; it should be!).

4 Vorticity equation and potential ow


4.1 Vorticity and rotation
Vorticity, i , was introduced in Eq. 1.7 at p. 11. As shown in Fig. 1.3 at p. 13, vorticity is connected to rotation of a uid particle. Figure 4.1 shows the surface forces in the x1 momentum equation acting on a uid particle in a shear ow. Looking at Fig. 4.1 it is obvious that only the shear stresses are able to rotate the uid particle; the pressure acts through the center of the uid particle and is thus not able to affect rotation of the uid particle. Let us have a look at the momentum equations in order to show that the viscous terms indeed can be formulated with the vorticity vector, i . In incompressible ow the viscous terms read (see Eqs. 2.4, 2.5 and 2.6) 2 vi ji = xj xj xj (4.1)

4.1. Vorticity and rotation


21 (x + 0.5x2 ) 2 v1 = cx2 p(x 0.5x1 ) 1 (x , x ) 1 2 p(x + 0.5x1 ) 1

33

x2

21 (x 0.5x2 ) 2

x1 Figure 4.1: Surface forces in the x1 direction acting on a uid particle (assuming 11 = 0). The right side can be re-written using the tensor identity 2 vi 2 vj = xj xj xj xi Lets verify that 2 vi 2 vj xj xi xj xj = inm mjk 2 vk xj xn (4.3) 2 vj 2 vi xj xi xj xj = 2 vk 2 vj inm mjk xj xi xj xn (4.2)

Use the -identity (see Table A.1 at p. 38 inm mjk 2 vk 2 vk 2 vk 2 vi = (ij nk ik nj ) = xj xn xj xn xi xk xj xj (4.4)

The rst term on the right side is zero because of continuity and hence we nd that Eq. 4.2 can indeed be written as 2 vi 2 vk 2 vj = inm mjk xj xj xj xi xj xn At the right side we recognize the vorticity, m = mjk vk /xj , so that 2 vi m 2 vj = inm xj xj xj xi xn where the rst on the right side is zero because of continuity. so that m 2 vi = inm xj xj xn (4.7) (4.6) (4.5)

4.2. The vorticity transport equation in three dimensions

34

In vector notation the identity Eq. 4.6 reads 2 v = ( v) v = Using Eq. 4.7, Eq. 4.1 reads ji m = inm xj xn (4.9) (4.8)

Thus, there is a one-to-one relation between the viscous term and vorticity: no viscous terms means no vorticity and vice versa. An imbalance in shear stresses (left side of Eq. 4.9) causes a change in vorticity, i.e. creates vorticity (right side of Eq. 4.9). Hence, inviscid ow (i.e. friction-less ow) has no rotation. (The exception is when vorticity is transported into an inviscid region, but also in that case no vorticity is created or destroyed: it stays constant, unaffected.) Inviscid ow is often called irrotational ow (i.e. no rotation) or potential ow. The vorticity is always created at boundaries, see Section 4.3.1. The main points that we have learnt in this section are: 1. The viscous terms are responsible for creating vorticity; this means that the vorticity cant be created or destroyed in inviscid (friction-less) ow 2. The viscous terms in the momentum equations can be expressed in i ; considering Item 1 this was to be expected. Exercise 23 Prove the rst equality of Eq. 4.7 using the --identity. Exercise 24 Write out Eq. 4.9 for i = 1 and verify that it is satised.

potential

4.2 The vorticity transport equation in three dimensions


Up to now we have talked quite a lot about vorticity. We have learnt that physically it means rotation of a uid particle and that it is only the viscous terms that can cause rotation of a uid particle. The terms inviscid, irrotational and potential ow all denote frictionless ow which is equivalent to zero vorticity. There is a small difference between the three terms because there may be vorticity in inviscid ow that is convected into the ow at the inlet(s); but also in this case the vorticity is not affected once it has entered the inviscid ow region. However, mostly no distinction is made between the three terms. In this section we will derive the transport equation for vorticity in incompressible ow. As usual we start with the Navier-Stokes equation, Eq. 2.7 at p.19. First, we re-write the convective term of the incompressible momentum equation (Eq. 2.7) as vj vi 1 = vj (Sij + ij ) = vj Sij ijk k xj 2 (4.10)

frictionless

where Eq. 1.14 on p. 12 was used. Inserting Sij = (vi /xj + vj /xi )/2 and multiplying by two gives 2vj vi = vj xj vi vj + xj xi ijk vj k (4.11)

4.2. The vorticity transport equation in three dimensions

35

The second term on the right side can be written as vj 1 (vj vj ) vj = xi 2 xi (4.12)

Equation 4.11 can now be written as vj ( 1 v 2 ) vi ijk vj k = 2 xj xi


no rotation rotation

(4.13)

where v 2 = vj vj . The last term on the right side is the vector product of v and , i.e. v . The trick we have achieved is to split the convective term into one term without rotation (rst term on the right side of Eq. 4.13) and one term including rotation (second term on the right side). Inserting Eq. 4.13 into the incompressible momentum equation (Eq. 2.7) yields 1 v2 2 vi 1 p vi + + fi + 2 ijk vj k = t xi xi xj xj
no rotation rotation

(4.14)

The volume source is in most engineering ows represented by the gravity which is conservative meaning that it is uniquely determined by the position (in this case the vertical position). Hence it can be expressed as a potential, h; we write fi = gi = h xi (4.15)

The negative sign appears because height is dened positive upwards and the direction of gravity is downwards. Since the vorticity vector is dened by the cross product pqi vi /xq ( v in vector notation, see Exercise 8), we start by applying the operator pqi /xq to the Navier-Stokes equation (Eq. 4.14) so that 2 1 v2 vj k 2 vi 2 + pqi pqi ijk txq xi xq xq 2 3 1 p vi 2h = pqi + pqi pqi xi xq xj xj xq xq xi pqi

(4.16)

where the body force in Eq. 4.16 was re-written using Eq. 4.15. We know that ijk is anti-symmetric in all indices, and hence the second term on line 1 and the rst and the last term on line 2 are all zero (product of a symmetric and an anti-symmetric tensor). The last term on line 1 is re-written using the - identity (see Table A.1 at p. A.1) pqi ijk vj k vp k vq p vj k = (pj qk pk qj ) = xq xq xk xq (4.17)

Using the denition of i we nd that the divergence of i is zero i = xi xi ijk vk xj = ijk 2 vk =0 xj xi (4.18)

4.2. The vorticity transport equation in three dimensions


1 x2 x1 Figure 4.2: Vortex stretching. 1

36

(product of a symmetric and an anti-symmetric tensor). Using the continuity equation (vq /xq = 0) and Eq. 4.18, Eq. 4.17 can be written pqi ijk vj k vp p = k vk xq xk xk (4.19)

The second term on line 2 in Eq. 4.16 can be written as pqi 2 3 vi = xj xj xq xj xj pqi vi xq = 2 p xj xj (4.20)

Inserting Eqs. 4.19 and 4.20 into Eq. 4.16 gives nally vp p 2 p p dp = k + + vk dt t xk xk xj xj (4.21)

We recognize the usual unsteady term, the convective term and the diffusive term. Furthermore, we have got rid of the pressure gradient term. That makes sense, because as mentioned in connection to Fig. 4.1, the pressure cannot affect the rotation (i.e. the vorticity) of a uid particle since the pressure acts through the center of the uid particle. Equation 4.21 has a new term on the right-hand side which represents amplication and rotation/tilting of the vorticity lines. If we write it term-by-term it reads v1 + v1 + v1 1 2 3 x1 x2 x3 vp v2 v2 v2 (4.22) = k +2 + 3 1 x1 xk x2 x3 v3 v3 v3 1 +2 + 3 x1 x2 x3 The diagonal terms in this matrix represent vortex stretching. Imagine a slender, cylindrical uid particle with vorticity i and introduce a cylindrical coordinate system with the x1 -axis as the cylinder axis and r2 as the radial coordinate (see Fig. 4.2) so that i = (1 , 0, 0). We assume that a positive v1 /x1 is acting on the uid cylinder; it will stretch the cylinder. From the continuity equation in polar coordinates v1 1 (rvr ) + =0 x1 r r

Vortex stretching

(4.23)

we nd that since v1 /x1 > 0 the radial derivative of the radial velocity, vr , must be negative, i.e. the radius of the cylinder will decrease. Hence vortex stretching will either make a uid element longer and thinner (as in the example above) or shorter and thicker (when v1 /x1 < 0). The off-diagonal terms in Eq. 4.22 represent vortex tilting. Again, take a slender uid particle, but this time with its axis aligned with the x2 axis, see Fig. 4.3. The

Vortex tilting

4.3. The vorticity transport equation in two dimensions

37

v1 (x2 )

x2 2 x1

Figure 4.3: Vortex tilting. velocity gradient v1 /x2 will tilt the uid particle so that it rotates in clock-wise direction. The second term 2 v1 /x2 in line one in Eq. 4.22 gives a contribution to 1 . This means that vorticity in the x2 direction creates vorticity in the x1 direction.. Vortex stretching and tilting are physical phenomena which act in three dimensions: uid which initially is two dimensional becomes quickly three dimensional through these phenomena. Vorticity is useful when explaining why turbulence must be threedimensional. For more detail, see section 2.4 in [5]3 . You will learn more about this in other MSc courses on turbulent ow.

4.3 The vorticity transport equation in two dimensions


It is obvious that the vortex stretching/tilting has no inuence in two dimensions; in this case the vortex stretching/tilting term vanishes because the vorticity vector is orthogonal to the velocity vector (for a 2D ow the velocity vector reads vi = (v1 , v2 , 0) and the vorticity vector reads i = (0, 0, 3 ) so that the vector k vp /xk = 0). Thus in two dimensions the vorticity equation reads dq 2 3 = dt x x (4.24)

(Greek indices are used to indicate that they take values 1 or 2). This equation is exactly the same as the transport equation for temperature in incompressible ow, see Eq. 2.15. This means that vorticity diffuses in the same way as temperature does. In fully developed channel ow, for example, the vorticity and the temperature equations reduce to 0= 2 3 x2 2 2T 0=k 2 x2 (4.25a) (4.25b)

For the temperature equation the heat ux is given by c2 = T /x2 ; with a hot lower wall and a cold upper wall (constant wall temperatures) the heat ux is constant and goes from the lower wall to the upper wall. We have the same situation for the vorticity.
3 can

be downloaded from http://www.tfd.chalmers.se/lada

4.3. The vorticity transport equation in two dimensions


x2 x1 V0

38

111111111 000000000

111111 000000 1111111111111111111111111111111111111111 0000000000000000000000000000000000000000 1111111111111111111111111111111111111111 0000000000000000000000000000000000000000 1111111111111111111111111111111111111111 0000000000000000000000000000000000000000

L Figure 4.4: Boundary layer. The boundary layer thickness, , increases for increasing streamwise distance from leading edge (x1 = 0). Its gradient, i.e. the vorticity ux, 3 /x2 , is constant across the channel. You have plotted this quantity in Assignment 2. If wall-normal temperature derivative T /x2 = 0 at both walls (adiabatic walls), the heat ux is zero at the walls and the temperature will be equal to an arbitrary constant in the entire domain. It is only when the wall-normal temperature derivative at the walls are non-zero that a temperature eld is created in the domain. The same is true for q: if q/x2 = 0 at the walls, the ow will not include any vorticity. Hence, vorticity is in the same way as temperature created at the walls. 4.3.1 Diffusion length from the Rayleigh problem In Section 3.1 we studied the Rayleigh problem (unsteady diffusion). The diffusion time, t, or the diffusion length, , in Eq. 3.14 can now be used to estimate the thickness of the boundary layer. The vorticity is generated at the wall and it diffuses away from the wall. There is vorticity in a boundary layer and outside the boundary layer it is zero. Consider the boundary layer in Fig. 4.4. At the end of the plate the boundary thickness is (L). The time it takes for a uid particle to travel along the plate is L/V0 . During this time vorticity will diffuse the length according Eq. 3.14. If we assume that the uid is air with the speed V0 = 3m/s and that the length of the plate L = 2m we get from Eq. 3.14 that (L) = 1.2cm. Exercise 25 Note that the estimate above is not quite accurate because in the Rayleigh problem we assumed that the convective terms are zero, but in a developing boundary layer they are not (v2 = 0 and v1 /x1 = 0). The proper way to solve the problem is to use Blasius solution (you have probably learnt about this in your rst uid mechanics course; if not, you should go and nd out). Blasius solution gives 5 = 1/2 x Rex Compute what (L) you get from Eq. 4.26. Exercise 26 Assume that we have a developing ow in a pipe (radius R) or between two at plates (separation distance h). We want to nd out how long distance it takes for the the boundary layers to merge. Equation 3.14 can be used with = R or h. Make a comparison with this and Eq. 3.16. (4.26)

5. Turbulence

39

5 Turbulence
5.1 Introduction
Almost all uid ow which we encounter in daily life is turbulent. Typical examples are ow around (as well as in) cars, aeroplanes and buildings. The boundary layers and the wakes around and after bluff bodies such as cars, aeroplanes and buildings are turbulent. Also the ow and combustion in engines, both in piston engines and gas turbines and combustors, are highly turbulent. Air movements in rooms are turbulent, at least along the walls where wall-jets are formed. Hence, when we compute uid ow it will most likely be turbulent. In turbulent ow we usually divide the velocities in one time-averaged part vi , which is independent of time (when the mean ow is steady), and one uctuating part vi so that vi = vi + vi . There is no denition on turbulent ow, but it has a number of characteristic features (see Pope [7] and Tennekes & Lumley [8]) such as: I. Irregularity. Turbulent ow is irregular, random and chaotic. The ow consists of a spectrum of different scales (eddy sizes). We do not have any exact denition of an turbulent eddy, but we suppose that it exists in a certain region in space for a certain time and that it is subsequently destroyed (by the cascade process or by dissipation, see below). It has a characteristic velocity and length (called a velocity and length scale). The region covered by a large eddy may well enclose also smaller eddies. The largest eddies are of the order of the ow geometry (i.e. boundary layer thickness, jet width, etc). At the other end of the spectra we have the smallest eddies which are dissipated by viscous forces (stresses) into thermal energy resulting in a temperature increase. Even though turbulence is chaotic it is deterministic and is described by the Navier-Stokes equations. II. Diffusivity. In turbulent ow the diffusivity increases. This means that the spreading rate of boundary layers, jets, etc. increases as the ow becomes turbulent. The turbulence increases the exchange of momentum in e.g. boundary layers, and reduces or delays thereby separation at bluff bodies such as cylinders, airfoils and cars. The increased diffusivity also increases the resistance (wall friction) in internal ows such as in channels and pipes. III. Large Reynolds Numbers. Turbulent ow occurs at high Reynolds number. For example, the transition to turbulent ow in pipes occurs that ReD 2300, and in boundary layers at Rex 500 000. IV. Three-Dimensional. Turbulent ow is always three-dimensional and unsteady. However, when the equations are time averaged, we can treat the ow as two-dimensional (if the geometry is two-dimensional). V. Dissipation. Turbulent ow is dissipative, which means that kinetic energy in the small (dissipative) eddies are transformed into thermal energy. The small eddies receive the kinetic energy from slightly larger eddies. The slightly larger eddies receive their energy from even larger eddies and so on. The largest eddies extract their energy from the mean ow. This process of transferring energy from the largest turbulent scales (eddies) to the smallest is called the cascade process. VI. Continuum. Even though we have small turbulent scales in the ow they are much larger than the molecular scale and we can treat the ow as a continuum.

turbulent eddy

cascade process

5.2. Turbulent scales


ow of kinetic energy

40

....

large scales

intermediate scales

dissipative scales

Figure 5.1: Cascade process with a spectrum of eddies. The energy-containing eddies are denoted by v0 ; 1 and 2 denotes the size of the eddies in the inertial subrange such that 2 < 1 < 0 ; is the size of the dissipative eddies.

5.2 Turbulent scales


The largest scales are of the order of the ow geometry (the boundary layer thickness, for example), with length scale 0 and velocity scale v0 . These scales extract kinetic energy from the mean ow which has a time scale comparable to the large scales, i.e. 1 v = O(t1 ) = O(v0 /0 ) 0 x2 (5.1)

The kinetic energy of the large scales is lost to slightly smaller scales with which the large scales interact. Through the cascade process the kinetic energy is in this way transferred from the largest scale to the smallest scales. At the smallest scales the frictional forces (viscous stresses) become large and the kinetic energy is transformed (dissipated) into thermal energy. The dissipation is denoted by which is energy per unit time and unit mass ( = [m2 /s3 ]). The dissipation is proportional to the kinematic viscosity, , times the uctuating velocity gradient up to the power of two (see Section 8.1). The friction forces exist of course at all scales, but they are largest at the smallest eddies. Thus it is not quite true that eddies, which receive their kinetic energy from slightly larger scales, give away all of that to the slightly smaller scales. This is an idealized picture; in reality a small fraction is dissipated. However it is assumed that most of the energy (say 90%) that goes into the large scales is nally dissipated at the smallest (dissipative) scales. The smallest scales where dissipation occurs are called the Kolmogorov scales whose velocity scale is denoted by v , length scale by and time scale by . We assume that these scales are determined by viscosity, , and dissipation, . The argument is as follows.

5.3. Energy spectrum

41

viscosity: Since the kinetic energy is destroyed by viscous forces it is natural to assume that viscosity plays a part in determining these scales; the larger viscosity, the larger scales. dissipation: The amount of energy that is to be dissipated is . The more energy that is to be transformed from kinetic energy to thermal energy, the larger the velocity gradients must be. Having assumed that the dissipative scales are determined by viscosity and dissipation, we can express v , and in and using dimensional analysis. We write v = a b 2 2 3 [m/s] = [m /s] [m /s ] (5.2)

where below each variable its dimensions are given. The dimensions of the left and the right side must be the same. We get two equations, one for meters [m] 1 = 2a + 2b, and one for seconds [s] 1 = a 3b, (5.4) (5.3)

which give a = b = 1/4. In the same way we obtain the expressions for and so that v = ()
1/4

, =

1/4

, =

1/2

(5.5)

5.3 Energy spectrum


As mentioned above, the turbulence uctuations are composed of a wide range of scales. We can think of them as eddies, see Fig. 5.1. It turns out that it is often convenient to use Fourier series to analyze turbulence. In general, any periodic function, f , with a period of 2L (i.e. f (x) = f (x + 2L)), can be expressed as a Fourier series, i.e. f (x) = 1 a0 + (an cos(n x) + bn sin(n x)) 2 n=1

(5.6)

where x is a spatial coordinate and n = n/L. Variable n is called the wavenumber. The Fourier coefents are given by an = bn = Parsevals formula states that
L L

1 L 1 L

f (x) cos(n x)dx


L L

f (x) sin(n x)dx


L

f (x)2 dx =

L 2 a +L (a2 + b2 ) n n 2 0 n=1

(5.7)

5.3. Energy spectrum


Pk E() I
II= in e

42

rti al

ran

ge

Figure 5.2: Spectrum for turbulent kinetic energy, k. I: Range for the large, energy containing eddies. II: the inertial subrange. III: Range for small, isotropic scales. For readers not familiar to Fourier series, a brief introduc tion is given in Section C. Let now f be a uctuating velocity component, say v1 . The left side of Eq. 5.7 expresses 2 2 v1 in physical space (vs. x) and the right side v1 in wavenumber space (vs. n ). The reader who is not familiar to the term wavenumber, is probably more familiar to frequency. In that case, express f in Eq. 5.6 as a series in time rather than in space. 2 In this case the left side of Eq. 5.7 expresses v1 in the time domain and the right side 2 expresses v1 in the frequency domain. The turbulent scales are distributed over a range of scales which extends from the largest scales which interact with the mean ow to the smallest scales where dissipation occurs, see Fig. 5.1. Now let us think about how the kinetic energy of the eddies varies with eddy size. Intuitively we assume that large eddies have large uctuating velocities which implies large kinetic energy, vi vi /2. It is now convenient to study the kinetic energy of each eddy in wavenumber space. In wavenumber space the energy of eddies from to + d can be expressed as E()d (5.8)

where Eq. 5.8 expresses the contribution from the scales with wavenumber between and + d to the turbulent kinetic energy k. The energy, E(), corresponds to f 2 () in Eq. 5.7. The dimension of wavenumber is one over length; thus we can think of wavenumber as proportional to the inverse of an eddys diameter, i.e 1/d. The total turbulent kinetic energy is obtained by integrating over the whole wavenumber space i.e.

E ) (
3 5/

III=dissipating range

k=
0

E()d = L

f 2 (n )

(5.9)

Think of this equation as a way to compute the kinetic energy by rst sorting them by size (i.e. wavenumber), then computing the kinetic energy of each eddy size (i.e. E()d), and nally summing the kinetic energy of all eddy sizes (i.e. carrying out the integration).

5.3. Energy spectrum

43

The kinetic energy is the sum of the kinetic energy of the three uctuating velocity components, i.e. k= 1 2 1 2 2 v + v2 + v3 = vi vi 2 1 2 (5.10)

The spectrum of E is shown in Fig. 5.2. We nd region I, II and III which correspond to: I. In this region we have the large eddies which carry most of the energy. These eddies interact with the mean ow and extract energy from the mean ow. This energy transfer takes places via the production term, P k , in the transport equation for turbulent kinetic energy, see Eq. 8.14. The energy of the largest eddies is transferred to slightly smaller scales. The eddies velocity and length scales are v0 and 0 , respectively. III. Dissipation range. The eddies are small and isotropic and it is here that the dissipation occurs. The energy transfer from turbulent kinetic energy to thermal energy (increased temperature) is governed by in the transport equation for turbulent kinetic energy, see Eq. 8.14. The scales of the eddies are described by the Kolmogorov scales (see Eq. 5.5) II. Inertial subrange. The existence of this region requires that the Reynolds number is high (fully turbulent ow). The eddies in this region represent the mid-region. This region is a transport region (in wavenumber space i.e.) in the cascade process. The transport in wavenumber space is called spectral transfer. Energy per time unit, P k = , is coming from the large eddies at the lower part of this range and is given off to the dissipation range at the higher part (note that the relation P k = is given by the concept of the cascade process). Since the cascade concept assumes that all turbulent kinetic energy is transferred from large to small eddies, it means than the number of small eddies must be much larger than that of large eddies, i.e. N k is constant (N and k = v,i v,i /2 denote the number of eddies and kinetic energy eddies of size proportional to 1/, respectively). The eddies in this region are independent of both the large, energy-containing eddies and the eddies in the dissipation range. One can argue that the eddies in this region should be characterized by the spectral transfer of energy () and the size of the eddies 1/. Dimensional analysis gives E = [m /s2 ] =
3

spectral transfer

a b 2 3 [1/m] [m /s ]

(5.11)

We get two equations, one for meters [m] 3 = a + 2b, and one for seconds [s] so that b = 2/3 and a = 5/3. Inserted in Eq. 5.11 we get E() = const. 3 3 This is a very important law (Kolmogorov spectrum law or the 5/3 law) which states that, if the ow is fully turbulent (high Reynolds number), the energy spectra should exhibit a 5/3-decay in the inertial region.
2 5

2 = 3b,

5.4. The cascade process created by vorticity

44

Above we state that the small eddies are isotropic. This means that in average the eddies have no preferred direction, i.e. the uctuations in all directions are the 2 2 2 same so that v1 = v2 = v3 . Note that is not true instantaneously, i.e. in general v1 = v2 = v3 . Furthermore, isotropic turbulence implies that if a coordinate direction is switched, nothing should be changed. For example if the x1 coordinate direction is rotated 180o the v1 v2 should remain the same, i.e. v1 v2 = v1 v2 . This is possible v = 0. Hence, all shear stresses are zero in isotropic turbulence. only if v1 2 As discussed on p. 40, the concept of the cascade process assumes that the energy extracted by the large turbulent eddies is transferred by non-linear interactions through the inertial range to the dissipative range where the kinetic energy is transformed to thermal energy (increased temperature). The spectral transfer rate of kinetic energy from eddies of size to slightly smaller eddies can be estimated as follows. An eddy loses its kinetic energy during one revolution. The kinetic energy of the eddy is pro2 portional to v and the time for one revolution is proportional to /v . Hence, the energy spectral transfer rate, , for a an eddy of length scale can be estimated as (see Fig. 5.2) 3 2 v v =O (5.12) = O v The kinetic energy is transferred to smaller and smaller eddies until it is dissipated at the dissipative scales. In the inertial subrange the cascade process assumes that = . Applying Eq. 5.12 for the large energy-containing eddies gives 0 = O
2 v0 0 v0

isotropic turbulence

=O

3 v0 0

= =

(5.13)

Hence we nd that the dissipation at small scales (large wavenumbers) is determined by how much energy that enters the cascade process at the large scales (small wavenumbers). We can now estimate the ratio between the large eddies (with v0 and 0 ) to the Kolmogorov eddies (v and ). Equations 5.5 and 5.13 give v0 3 = ()1/4 v0 = v0 /0 v 0 = o = 3 0 3 v0
1/4 1/4

v0 = (v0 0 /) 0 = 3 3 v0 3 0 v0 0

1/4

= Re1/4 = Re3/4 (5.14)

0 =
1/2

3 0 3 v0
3 v0 0

1/4

1/4

1/2

0 =

0 = v0

1/2

= Re1/2

where Re = v0 0 /. We nd that the ratio of the velocity, length and time scales of the energy-containing eddies to the Kolmogorov eddies increases with increasing Reynolds number. This means that the eddy range (wavenumber range) of the intermediate region, (region II, the inertial region), increases with increasing Reynolds number. Hence, the larger the Reynolds number, the larger the wavenumber range of the intermediate range where the eddies are independent of both the large scales and the viscosity.

5.4 The cascade process created by vorticity


The interaction between vorticity and velocity gradients is an essential ingredient to create and maintain turbulence. Disturbances are amplied by interaction between

5.4. The cascade process created by vorticity


1

45

Figure 5.3: Family tree of turbulent eddies (see also Table 5.1). Five generations. The large original eddy, with axis aligned in the x1 direction, is 1st generation. Adapted from [9] the vorticity vector and the velocity gradients; the disturbances are turned into chaotic, three-dimensional, random uctuations, i.e. into turbulence. Two idealized phenomena in this interaction process can be identied: vortex stretching and vortex tilting. The equation for the instantaneous vorticity (i = qi + qi ) reads (see Eq. 4.21) vi i 2 i i = j + + vj t xj xj xj xj vk i = ijk xj

(5.15)

As we learnt in Section 4.2 this equation is not an ordinary convection-diffusion equation: it has an additional term on the right side which represents amplication and rotation/tilting of the vorticity lines (the rst term on right side). The i = j components of this term represent (see Eq. 4.22) vortex stretching. A positive v1 /x1 will stretch the cylinder, see Fig. 4.2 and from the continuity equation we nd that the radial derivative of the radial velocity v2 must be negative (see Eq. 4.23), i.e. the radius of the cylinder will decrease. We have neglected the viscosity since viscous diffusion at high Reynolds number is much smaller than the turbulent one and since viscous dissipation occurs at small scales (see p. 40). Thus we can assume that there are no viscous stresses acting on the cylindrical uid element surface which means that the rotation momentum r2 1 = const. (5.16)

Vortex stretching

remains constant as the radius of the uid element decreases. Note that also the cir-

5.4. The cascade process created by vorticity


generation 1st 2nd 3rd 4th 5th 6th 7th x1 1 0 2 2 6 10 22 x2 0 1 1 3 5 11 21 x3 0 1 1 3 5 11 21

46

Table 5.1: Number of eddies at each generation with their axis aligned in the x1 , x2 or x3 direction, see Fig. 5.3. culation, which is the integral of the tangential velocity round the perimeter is constant. Equation 5.16 shows that the vorticity increases if the radius decreases (and vice versa). As was mentioned above, the continuity equation shows that stretching results in a decrease of the radius of a slender uid element and an increase of the vorticity component (i.e. the tangential velocity component) aligned with the element. Hence an extension of a uid element in one direction (x1 direction) decreases the length scales and increases the velocity scales in the other two coordinate directions (x2 and x3 ). The increased v2 and v3 velocity components will in next stage stretch smaller uid elements aligned in these two directions and so on. At each stage, the length scale of the eddies whose velocity scale are increased decreases. Figure 5.3 illustrates how a large eddy whose axis is oriented in the x1 axis in a few generations creates through vortex stretching smaller and smaller eddies with larger and larger velocity gradients. Here a generation is related to a wavenumber in the energy spectrum (Fig. 5.2); young generations correspond to high wavenumbers. The smaller the eddies, the less the original orientation of the large eddy is recalled. In other words, the small eddies dont remember the characteristics of their original ancestor. The small eddies have no preferred direction. They are isotropic. The creation of multiple eddies by vortex stretching from one original eddies is illustrated in Fig. 5.3 and Table 5.1 The large original eddy (1st generation) is aligned in the x1 direction. It creates eddies in the x2 and x3 direction (2nd generation), which in turn each create new eddies in the x1 and x3 (3rd generation) and so on. For each generation the eddies become more and more isotropic as they get smaller. The i = j components in the rst term on the right side in Eq. 4.22 represent vortex tilting. Again, take a slender uid element, now with its axis aligned with the x2 axis, Fig. 4.2. The velocity gradient v1 /x2 will tilt the uid element so that it rotates in the clock-wise direction. As a result, the second term 2 v1 /x2 in line one in Eq. 4.22 gives a contribution to 1 . This shows how vorticity in one direction is transferred to the other two directions through vortex tilting. Vortex stretching and vortex tilting qualitatively explain how interaction between vorticity and velocity gradient create vorticity in all three coordinate directions from a disturbance which initially was well dened in one coordinate direction. Once this process has started it continues, because vorticity generated by vortex stretching and vortex tilting interacts with the velocity eld and creates further vorticity and so on. The vorticity and velocity eld becomes chaotic and random: turbulence has been created. The turbulence is also maintained by these processes. From the discussion above we can now understand why turbulence always must be

Vortex tilting

6. Turbulent mean ow

47

three-dimensional (Item IV on p. 39). If the instantaneous ow is two-dimensional (x1 x2 plane) we nd that the vortex-stretching/tilting term on the right side of Eq. 5.15 vanishes because the vorticity vector and the velocity vector are orthogonal. The only non-zero component of vorticity vector is 3 because v3 v2 0 x2 x3 v3 v1 0. x3 x1

1 2

= =

Since v3 = 0, we get j vi /xj = 0.

6 Turbulent mean ow
6.1 Time averaged Navier-Stokes
When the ow is turbulent it is preferable to decompose the instantaneous variables (for example the velocity components and the pressure) into a mean value and a uctuating value, i.e.
vi = vi + vi p = p + p

(6.1)

where the bar, , denotes the time averaged value. One reason why we decompose the variables is that when we measure ow quantities we are usually interested in their mean values rather than their time histories. Another reason is that when we want to solve the Navier-Stokes equation numerically it would require a very ne grid to resolve all turbulent scales and it would also require a ne resolution in time (turbulent ow is always unsteady). The continuity equation and the Navier-Stokes equation for incompressible ow with constant viscosity read vi xi vi vj vi + t xj = 0 = p 2 vi + xi xj xj (6.2) (6.3)

The gravitation term, gi , has been omitted which means that the p is the hydrostatic pressure (i.e. when vi 0, then p 0, see p. 29). Inserting Eq. 6.1 into the continuity equation (6.2) and the Navier-Stokes equation (6.3) we obtain the time averaged continuity equation and Navier-Stokes equation i v xi i vj v xj = = 0 p + xi xj i v vi vj xj (6.4) (6.5)

It is assumed that the mean ow is steady. This equation is the time-averaged Navier-Stokes equation and it is often called the Reynolds equation. A new term vi vj appears on the right side of Eq. 6.5 which is called the Reynolds stress tensor. The

Reynolds equations

6.1. Time averaged Navier-Stokes

48

x2 v1 (x2 ) xA 2 2

x1 x3

Figure 6.1: Flow between two innite parallel plates. The width (i.e. length in the x3 direction) of the plates, Zmax , is much larger that the separation between the plates, i.e. Zmax .
tensor is symmetric (for example v1 v2 = v2 v1 ). It represents correlations between uctuating velocities. It is an additional stress term due to turbulence (uctuating ve locities) and it is unknown. We need a model for vi vj to close the equation system in Eq. 6.5. This is called the closure problem: the number of unknowns (ten: three velocity components, pressure, six stresses) is larger than the number of equations (four: the continuity equation and three components of the Navier-Stokes equations). The continuity equation applied both for the instantaneous velocity, vi (Eq. 6.2), and for the time-averaged velocity, vi (Eq. 6.4); hence it applied also for the uctuating velocity, vi , i.e. vi =0 (6.6) xi

closure problem

6.1.1 Boundary-layer approximation For steady (/t = 0), two-dimensional (3 = /x3 = 0) boundary-layer type of v ow (i.e. boundary layers along a at plate, channel ow, pipe ow, jet and wake ow, etc.) where v2 v1 , 1 v 1 v , x1 x2 (6.7)

Eq. 6.5 reads (we write the left side on non-conservative form, see Section 2.4 at p. 22) 1 v 1 v 1 v + 2 v x1 x2 = 1 v p + v1 v2 x1 x2 x2
tot

(6.8)

x1 and x2 denote the streamwise and wall-normal coordinate, respectively, see Fig. 6.1. Note that the two terms on the left side are of the same order, because they both include the product of one large (1 or /x2 ) and one small (2 or /x1 ) part. v v In addition to the viscous shear stress, 1 /x2 , an additional turbulent one a v Reynolds shear stress appears on the right side of Eq. 6.8. The total shear stress is thus 1 v v1 v2 tot = (6.9) x2

shear stress

6.2. Wall region in fully developed channel ow

49

6.2 Wall region in fully developed channel ow


The region near the wall is very important. Here the velocity gradient is largest as the velocity drops down to zero at the wall over a very short distance. One important quantity is the wall shear stress which is dened as w = 1 v x2 (6.10)
w

From the wall shear stress, we can dene a wall friction velocity, u , as w = u2 u = w
1/2

(6.11)

wall friction velocity

In order to take a closer look at the near-wall region, let us, again, consider fully developed channel ow between two innite plates, see Fig. 6.1. In fully developed channel ow, the streamwise derivative of the streamwise velocity component is zero (this is the denition of fully developed ow), i.e. 1 /x1 = 0. The continuity v equation gives now v2 = 0, see Eq. 3.18 at p. 28. The rst term on the left side of Eq. 6.8 is zero because we have fully developed ow (1 /x1 = 0) and the last term v is zero because v2 0. The streamwise momentum equation, Eq. 6.8, can now be written p 1 v (6.12) 0= + v1 v2 x1 x2 x2 We know that the rst term is a function only of x1 and the two terms in parenthesis are functions of x2 only; hence they must be constant (see Eq. 3.24 and text related to this equation), i.e. p = constant x1 tot = = constant x2

x2

1 v v1 v2 x2

(6.13)

where the total stress, tot , is given by Eq. 6.9. Equation 6.13 shows that it varies linearly in the channel, i.e. tot = w x2 / (6.14) see Fig. 6.3a. The wall region can be divided into one outer and one inner region, see Fig. 6.2. The inner region includes the viscous region (dominated by the viscous diffusion) and the logarithmic region (dominated by turbulent diffusion); the logarithmic region is sometimes called the inertial region, because the turbulent stresses stem from the inertial (i.e. the non-linear convection) term. The buffer region acts as a transition region between these two regions where viscous diffusion of streamwise momentum is gradually replaced by turbulent diffusion. In the inner region, the total shear stress is approximately constant and equal to the wall shear stress w , see Fig. 6.3. Note that the total shear stress is constant only close to the wall (Fig. 6.3b); further away from the wall it decreases (in fully developed channel ow it decreases linearly with the distance from the wall, see Eq. 6.14 and Fig. 6.3a). The Reynolds shear stress vanishes at the wall because v1 = v2 = 0, and the viscous shear stress attains its wall-stress value w = u2 . As we go away from the wall the viscous stress decreases and turbulent one

6.2. Wall region in fully developed channel ow


104 103 102 101 outer region overlap region log-law region Wall 1 inner region buffer region viscous region Centerline 1

50

x2 /

x+ = y + 2 5 10 30 100 1000 10000

Figure 6.2: The wall region (adapted from Ch.7 in [7]) for Re = 10 000. denotes half width of the channel, see Fig. 6.1 and x+ = x2 u / denotes the normalized wall 2 distance. increases and at x+ 11 they are approximately equal. In the logarithmic layer the 2 viscous stress is negligible compared to the Reynolds stress. At the wall, the velocity gradient is directly related to the wall shear stress, i.e. (see Eq. 6.10 and 6.11) 1 1 v = u2 = u2 (6.15) x2 w Integration gives (recall that both and u2 are constant) v1 = 1 2 u x2 + C1

Since the velocity, v1 , is zero at the wall, the integration constant C1 = 0 so that v1 u x2 = u (6.16)

Equation 6.16 is expressed in inner scaling (or wall scaling) which means that v1 and x2 are normalized with quantities related to the wall, i.e. the friction velocity stemming from the wall shear stress and the viscosity (here we regard viscosity as a quantity related to the wall, since the ow is dominated by viscosity). Often the plus-sign ( + ) is used to denote inner scaling and equation Eq. 6.16 can then be written v1 = x+ + 2 (6.17)

Further away from the wall at 30 x+ 3000 (or 0.003 x2 / 0.3), we 2 encounter the log-law region, see Fig. 6.2. In this region the ow is assumed to be independent of viscosity. The Reynolds shear stress, v1 v2 , is in the region x+ 200 2 (i.e. x2 / 0.1) fairly constant and approximately equal to w , see Fig. 6.3b. Hence the friction velocity, u , is a suitable velocity scale in the inner logarithmic region; it is used in the entire region. What about the length scale? Near the wall, an eddy cannot be larger than the distance to the wall and it is the distance to the wall that sets an upper limit on the

6.2. Wall region in fully developed channel ow


2000 200

51

1500

150

x+ 2

1000

x+ 2

100

500

50

a)

0 0

0.2

0.4

0.6

0.8

b)

0.2

0.4

0.6

0.8

Figure 6.3: Reynolds shear stress. Re = 2000. a) lower half of the channel; b) zoom : v1 v2 /w ; : (1 /x2 )/w . v near the wall. DNS data [10, 11].
2000 24 20 16 12 8 500 4 0 0

1500

x+ 1000 2

a)

10

15

20

25

v1 /u

v1 /u

b)

10

100

1000

x+ 2 : DNS

Figure 6.4: Velocity proles in fully developed channel ow. Re = 2000. data [10, 11]; : v1 /u = (ln x+ )/0.41 + 5.2; : v1 /u = x+ . 2 2

eddy-size. Hence it seems reasonable to take the wall distance as the characteristic length scale; a constant is added so that = x2 The velocity gradient can be estimated as 1 v u = x2 x2 (6.19) (6.18)

based on the velocity scale, u , and the length scale x2 . Another way of deriving the expression in Eq. 6.19 is to use the Boussinesq assumption (see Eq. 11.26) in which a turbulent Reynolds stress is assumed to be equal to the product between the turbulent viscosity and the velocity gradient as
v1 v2 = t

1 v x2

(6.20)

The turbulent viscosity, t , represents the turbulence and has the same dimension as , i.e. [m2 /s]. Hence t can be expressed as a product of a turbulent velocity scale and a turbulent length scale, and in the log-law region that gives t = u x2 (6.21)

6.3. Reynolds stresses in fully developed channel ow


200 2000 1800 150 1600 1400

52

x+ 2

100

x+ 2

1200 1000 800

50

600 400

a)

0 150

100

50

50

100

150

b)

200 2

Figure 6.5: Fully developed channel ow. Re = 2000. Forces in the v1 equation, see Eq. 6.12. a) near the lower wall of the channel; b) lower half of the channel ex cluding the near-wall region. DNS data [10, 11]. : (v1 v2 /x2 )/w ; : : ( p/x1 )/w . ( 2 v1 /x2 )/w ; 2
so that Eq. 6.20 gives (inserting v1 v2 = u2 ) read

u2 = u x2

1 v u 1 v = x2 x2 x2

(6.22)

In non-dimensional form Eqs. 6.19 and 6.22 read 1 1 v+ + = x2 x+ 2 Integration gives now 1 ln x+ + B or 2 v1 1 x2 u +B = ln u v1 = + (6.23)

(6.24)

where B is an integration constant. From Eq. 6.24 we can dene the viscous length scale, , as (6.25) x+ = x2 / = 2 u Equation 6.24 is the logarithmic law due to von K rm n [12]. The constant, , is called a a the von K rm n constant. The constants in the log-law are usually set to = 0.41 and a a B = 5.2. 3000 (x2 / 0.3). As can be seen in Fig. 6.2 the log-law applies for x+ 2 Figure 6.4 where the Reynolds number is lower than in Fig. 6.2 show that the log-law t the DNS up to x+ 500 (x2 / 0.25). 2 In the outer region of the boundary layer, the relevant length scale is the boundary layer thickness. The resulting velocity law is the defect law v1,c v1 x2 = FD u where c denotes centerline. (6.26) log-law

6.3. Reynolds stresses in fully developed channel ow

53

32,tot x2

x3 Figure 6.6: Symmetry plane of channel ow.

6.3 Reynolds stresses in fully developed channel ow


The ow is two-dimensional (3 = 0 and /x3 = 0). Consider the x2 x3 plane, v see Fig. 6.6. Since nothing changes in the x3 direction, the viscous shear stress 32 = 3 v 2 v + x2 x3 =0 (6.27)

because v3 = 2 /x3 = 0. The turbulent part shear stress, v2 v3 , can be expressed v using the Boussinesq assumption (see Eq. 11.26) v2 v3 = t

2 v 3 v + x2 x3

=0

(6.28)

and it is also zero since v3 = 2 /x3 = 0. With the same argument, v1 v3 = 0. v 2 2 However note that v3 = v3 = 0. The reason is that although the time-averaged ow is two-dimensional (i.e. v3 = 0), the instantaneous turbulent ow is always three 2 dimensional and unsteady. Hence v3 = 0 and v3 = 0 so that v3 = 0. Consider, for example, the time series v3 = v3 = (0.25, 0.125, 0.125, 0.2, 0.2). This gives

v3 = (0.25 + 0.125 + 0.125 0.2 + 0.2)/5 = 0 but


2 2 v3 = v3 = (0.25)2 + 0.1252 + 0.1252 + (0.2)2 + 0.22 /5 = 0.03475 = 0.

Figure 6.3 presents the Reynolds and viscous shear stresses for fully developed ow. As can be seen, the viscous shear stress is negligible except very near the wall. It is equal to one near the wall and decreases rapidly for increasing wall distance. On the other hand, the Reynolds shear stress is zero at the wall (because the uctuating velocities are zero at the wall) and increases for increasing wall distance. The intersection of the two shear stresses takes place at x+ 11. 2 Looking at Eq. 6.12 we nd that it is not really the shear stress that is interesting, but its gradient. The gradient of the shear stress, (v1 v2 )/x2 and 2 v1 /x2 2 represent, together with the pressure gradient, p/x1 , the forces acting on the uid. Figure 6.5 presents the forces. Start by looking at Fig. 6.5b which shows the forces in the region away from the wall. The pressure gradient is constant and equal to one: this

6.4. Boundary layer


2000 100

54

1500

80

60

x+ 2

1000

x+ 2
40 20

500

a)

0 0

b)

0 0

Figure 6.7: Normal Reynolds stresses and turbulent kinetic energy. Re = 2000. DNS 2 2 2 : v1 /w ; : v2 /w ; : v3 /w ; : k/u2 . data [10, 11]. is the force driving the ow. This agrees fortunately with our intuition. We can imagine that the uid (air, for example) is driven by a fan. Another way to describe the behaviour of the pressure is to say that there is a pressure drop. The pressure must decrease in the streamwise direction so that the pressure gradient term, p/x1 , in Eq. 6.12 takes a positive value which pushes the ow in the x1 direction. The force that balances the pressure gradient is the gradient of the Reynolds shear stress. This is the force opposing the movement of the uid. This opposing force has its origin at the walls due to the viscous wall force (viscous shear stress multiplied by area). Now lets have a look at the forces in the wall region, see Fig. 6.5a. Here the forces are two orders of magnitude larger than in Fig. 6.5b but they act over a very thin region (x+ 40 or x2 / < 0.02). In this region the shear stress gradient term is driving 2 the ow and the opposing force is the viscous force. We can of course make a force balance for a section of the channel, as we did for laminar ow, see Eq. 3.36 at p. 31 and Fig. 3.8 at p. 31 which reads 0 = p1 Zmax 2 p2 Zmax 2 2w LZmax where L is the length of the section. We get p w p = = L x1 (6.30) (6.29)

As can be seen the pressure drop is directly related to the wall shear stress. In turbulent ow the velocity prole in the center region is much atter than in laminar ow (cf. Fig. 6.4 and Fig. 3.7 at p. 30). This makes the velocity gradient near the wall (and the wall shear stress, w ) much larger in turbulent ow than in laminar ow: Eq. 6.30 shows why the pressure drop is larger in the former case compared to the latter; or in other words why a larger fan is required to push the ow in turbulent ow than in laminar ow. 2 2 2 Figure 6.7 presents the normal Reynolds stresses, v1 , v2 and v3 . As can be seen, the streamwise stress is largest and the wall-normal stress is smallest. The former is largest because the mean ow is in this direction; the latter is smallest because the turbulent uctuations are dampened by the wall. The turbulent kinetic energy, 2 k = vi vi /2, is also included. Note that this is smaller than v1 .

6.4. Boundary layer


2000

55

24 20 16 12 8

1500

x+ 1000 2
500

v2 /u
5 10 15 20 25

4 0 0

a)

v2 /u

b)

10

100

1000

x+ 2 : DNS data [13];

Figure 6.8: Velocity proles in a boundary layer along a at plate. : v2 /u = (ln x+ )/0.41 + 5.2; : v2 /u = x+ . 2 2

6.4 Boundary layer


Up to now we have mainly discussed fully developed channel ow. What is the difference between that ow and a boundary layer ow? First, in a boundary layer ow the convective terms are not zero (or negligible), i.e. the left side of Eq. 6.8 is not zero. The ow in a boundary layer is continuously developing, i.e. its thickness, , increases continuously for increasing x1 . The ow in a boundary layer is described by Eq. 6.8. Second, in a boundary layer ow the wall shear stress is not determined by the pressure drop; the convective terms must also be taken into account. Third, the outer part of the boundary layer is highly intermittent, consisting of turbulent/non-turbulent motion. However, the inner region of a boundary layer (x2 / < 0.1) is principally the same as for the fully developed channel ow, see Fig. 6.8: the linear and the log-law regions are very similar for the two ows. However, in boundary layer ow the log-law is valid only up to approximately x2 / 0.1 (compared to approximately x2 / 0.3 in channel ow)

7 Probability density functions


Some statistical information is obtained by forming the mean and second moments, for 2 example v and v2 , as was done in Section 6. The root-mean-square (RMS) can be dened from the second moment as vrms = v 2
1/2

root-meansquare RMS standard deviation variance

(7.1)

The RMS is the same as the standard deviation which is equal to the square-root of the variance. In order to extract more information, probability density function is a useful statistical tool to analyze turbulence. From the velocity signals we can compute the probability densities (sometimes called histograms). With a probability density, fv , of the v velocity, the mean velocity is computed as

v=

vfv (v)dv

(7.2)

Normalize the probability functions, so that

fv (v)dv = 1

(7.3)

7. Probability density functions


20 15 10 1.5 1 0.5 0 5 0.5 0 1 5 10 1.5 10 4 6 10 2 6 4 2 0

56

12

14

(a) Point 1. vrms = 2.15.

16

18

20

12

14

(b) Point 2. vrms = 0.23.

16

18

20

12

14

(c) Point 3. vrms = 1.44.

16

18

20

Figure 7.1: Time history of v . Horizontal red lines show vrms .


3 0.3 0.25 0.2 2.5 2 1.5 1 0.5 0 5 0.25 0.2 0.15 0.1 0.05 0 4

fv 0.15
0.1 0.05 0 2

v /vrms

10

v /vrms

v /vrms

(a) Point 1. S = 1.51 and F = (b) Point 2. S = 0.27 and F = (c) Point 3. S = 0.09 and F = 7.68. 5.81. 2.77.

Figure 7.2: Probability density functions av time histories in Fig. 7.1. Vertical red lines show vrms . The skewness, S, and the atness, F , are given for the three time histories. Here we integrate over v. The mean velocity can of course also be computed by integrating over time, as we do when we dene a time average, (see Eq. 6.1 at p. 47), i.e. T 1 vdt (7.4) v= 2T T where T is sufciently large. Consider the probability density functions of the uctuations. The second moment corresponds to the variance of the uctuations (or the square of the RMS, see Eq. 7.1), i.e. v 2 =

v 2 fv (v )dv

As in Eq. 7.4, v 2 is usually computed by integrating in time, i.e. v 2 = 1 2T


T T

v 2 (t)dt

A probability density function is symmetric if positive values are as frequent and large as the negative values. Figure 7.1 presents the time history of the v history at three different points in a ow (note that v = 0). The red horizontal lines indicate the RMS value of v . The resulting probability densities functions are shown in Fig. 7.2. The red vertical lines show plus and minus RMS of v . Let us analyze the data at the three points.

7. Probability density functions

57

Point 1. The time history of the velocity uctuation (Fig. 7.1a) shows that there exists large positive values but no large negative values. The positive values are often larger than +vrms (the peak is actually close to 8vrms ) but the negative values are seldom smaller than vrms . This indicates that the distribution of v is skewed towards the positive side. This is conrmed in the PDF distribution, see Fig. 7.2a. Point 2. The uctuations at this point are much smaller and the positive values are as large the negative values; this means that the PDF should be symmetric which is conrmed in Fig. 7.2b. The extreme values of v are approximately 1.5vrms , see Figs. 7.1b and 7.2b. Point 3. At this point the time history (Fig. 7.1c) shows that the uctuations are clustered around zero and much values are within vrms . The time history shows that the positive and the negative values have the same magnitude. The PDF function in Fig. 7.2c conrms that there are many value around zero, that the extreme value are small and that positive and negative values are equally frequent (i.e. the PDF is symmetric). In Fig. 7.2 we can judge whether the PDF is symmetric, but instead of looking at the probability density functions, we should use a denition of the degree of symmetry, which is the skewness. It is dened as

skewness

v 3 =

v 3 fv (v )dv

3 and is commonly normalized by vrms , so that the skewness, Sv , of v is dened as

Sv =

1
3 vrms

v 3 fv (v )dv =

1
3 2vrms T

T T

v 3 (t)dt =

Note that f must be normalized (see Eq. 7.3). There is yet another statistical quantity which sometimes is used for describing turbulent uctuations, namely the atness. The variance (the square of RMS) tells us how large the uctuations are in average, but it does not tell us if the time history includes few very large uctuations or if all are rather close to vrms . The atness gives 4 this information, and it is dened computed from v 4 and normalized by vrms , i.e. F = 1
4 vrms

atness

v 4 fv (v)dv

The uctuations at Point 1 (see Fig. 7.1a) includes some samples which are very large and hence its atness is large (see caption in Fig. 7.2a), whereas the uctuation for Point 3 all mostly clustered within 2vrms giving a small atness, see Fig. 7.1c and the caption in Fig. 7.2c. For a Gaussian distribution f (v ) = for which F = 3. 1 vrms exp v vrms 2 2vrms

8. Transport equations for kinetic energy

58

8 Transport equations for kinetic energy


In this section and Section 9 we will derive various transport equations. There are two tricks which often will be used. Both tricks simply use the chain-rule for derivative backwards. Trick 1: Using the chain rule we get Ai Bj Bj Ai = Ai + Bj xk xk xk This expression can be re-written as Ai Ai Ai Bj Bj = Bj xk xk xk (8.2) (8.1)

and then we call it the chain-rule backwards. Trick 2: Using the chain rule we get 1 Ai Ai 1 = 2 xj 2 Ai Ai Ai + Ai xj xj = Ai Ai xj (8.3)

This trick is usually used backwards, i.e. Ai Ai 1 Ai Ai = xj 2 xj (8.4)

8.1 The Exact k Equation


The equation for turbulent kinetic energy, k = 1 vi vi , is derived from the Navier-Stokes 2 equation. Again, we assume incompressible ow (constant density) and constant viscosity (cf. Eq. 6.3). We subtract Eq. 6.5 from Eq. 6.3 and divide by density, multiply by vi and time average which gives vi

[vi vj vi vj ] = xj

vi vj 2 1 vi [p p] + vi [vi vi ] + v xi xj xj xj i Using vj = vj + vj , the left side can be rewritten as vi

(8.5)

v (i + vi )(j + vj ) vi vj = vi v vi vj + vi vj + vi vj . xj xj

(8.6)

Using the continuity equation vj /xj = 0 (see Eq. 6.6), the rst term is rewritten as vi

v i . vi vj = vi vj xj xj

(8.7)

v For the second term in Eq. 8.6 we start using j /xj = 0


vi

v (vi vj ) = vj vi i xj xj

(8.8)

8.1. The Exact k Equation

59

Next, we use Trick 2 vj


vi vi xj

= vj

xj

1 vv 2 i i

= vj

(k) = (j k) v xj xj

(8.9)

The third term in Eq. 8.6 can be written as (replace vj by vj and use the same technique as in Eq. 8.9)

1 v v v . 2 xj j i i as
1 p vi 1 p = vi xi xi

(8.10)

The rst term on the right side of Eq. 8.5 is re-written using the continuity equation

(8.11)

The second term on the right side of Eq. 8.5 can be written
vi 2 vi = xj xj xj vi v xj i

vi vi xj xj

(8.12)

applying Trick 1 (if we apply the chain-rule on the rst term on the right side of Eq. 8.12 we get the left side and the second term on the right side). For the rst term in Eq. 8.12 we use the same trick as in Eq. 8.9 so that xj
vi v xj i

xj

1 2

vi v vi + i vi xj xj

= (8.13)

xj

1 2

vi vi xj

1 2 vi vi 2k = = 2 xj xj xj xj

The last term on the right side of Eq. 8.5 is zero because it is time averaging of a uctuation, i.e. ab = ab = 0. Now we can assemble the transport equation for the turbulent kinetic energy. Equations 8.7, 8.9, 8.11, 8.12 and 8.13 give v j k v i = vi vj xj xj xj
I II v vi k 1 1 i vj p + vj vi vi 2 xj xj xj III IV

(8.14)

The terms in Eq. 8.14 have the following meaning. I. Convection. II. Production, P k . The large turbulent scales extract energy from the mean ow. This term (including the minus sign) is almost always positive. It is largest for the energy-containing eddies, i.e. for small wavenumbers, see Fig. 5.2. III. The two rst terms represent turbulent diffusion by pressure-velocity uctuations, and velocity uctuations, respectively. The last term is viscous diffusion. IV. Dissipation, . This term is responsible for transformation of kinetic energy at small scales to thermal energy. The term (excluding the minus sign) is always positive (it consists of velocity gradients squared). It is largest for large wavenumbers, see Fig. 5.2

8.1. The Exact k Equation

60

where C k , P k , Dk and correspond to terms I-IV in Eq. 8.14. Above, it is stated that the production takes place at the large energy-containing eddies, i.e. we assume that the large eddies contribute much more to the production term more than the small eddies. There are two arguments for this: 1. The Reynolds stresses (which appear in P k ) are larger for large eddies than for small eddies 2. In order to extract energy from the mean ow, the time scale of the eddy and the mean velocity gradient, i /xj , must be of the same magnitude. This rev quirement is best satised by the large scales. Actually, in the cascade process we argue that the smaller the eddies, the less they remember the characteristic of mean ow gradient (i.e. its magnitude, direction, time scale etc) In the cascade process (see Section 5.3) we assume that the dissipation takes places at the smallest scales. How do we know that the majority of the dissipation takes place at the smallest scales? First, note that the the dissipation, , is determined by the product of the viscosity and the squared gradient of the uctuating velocity vector (see term IV in Eq. 8.14). The dimension of a velocity gradient is the inverse of time, i.e. t1 . To nd out which eddies that contribute most to the dissipation term, we need to nd out how the time scale varies with eddy size. Consider the inertial subrange. The energy that is transferred in spectral space is equal to which is estimated by the large scales so that = u3 /0 , see Eq. 5.13. The energy transferred from eddy-to-eddy in 0 spectral space (i.e. the cascade process) can also be estimated using the eddys velocity 3 3 and length scales, i.e. = = u3 / = 2 / = 2 /0 , see Eq. 5.12. Since 0 and 0 0 are determined by the large eddies and are constant for a given spectrum, we nd 3 3 from 2 / = 2 /0 that for decreasing size of eddies (decreasing ), the time scale, 0 , also decreases, i.e. 2/3 0 (8.16) = 0
1 Hence we nd that the velocity gradient of an eddy, proportional to , increases for decreasing eddy size, i.e. the dissipation is indeed largest at high wavenumbers. In 2D boundary-layer ow, for which /x2 /x1 and v2 v1 , the exact k equation reads 2/3

The transport equation for k can also be written in a simplied easy-to-read symbolic form as C k = P k + Dk (8.15)

v 1 k 2 k v v 1 + = v1 v2 x1 x2 x2 v vi k 1 1 p v2 + v2 vi vi i x2 2 x2 xj xj

(8.17)

Note that the dissipation includes all derivatives. This is because the dissipation term is at its largest for small, isotropic scales for which all derivatives are of the same order and hence the usual boundary-layer approximation /x1 /x2 does not apply for these scales. Figure 8.1 presents the terms in Eq. 8.17 for fully developed channel ow. The left side is since the ow is fully developed zero. In the outer region (Fig. 8.1b) all

8.2. Spatial vs. spectral energy transfer


0.3 0.2 0.1 0 0.1 0.2 0.04 0.03 0.02 0.01 0 0.01 0.02 0.03 10 20 30 40 0.04

61

a)

x+ 2

b)

100

200

300

400

500

x+ 2

Figure 8.1: Channel ow at Re = 2000. Terms in the k equation scaled by u4 /. : P k; Re = 2000. a) Zoom near the wall; b) Outer region. DNS data [10, 11]. : ; : v p /x2 ; +: v2 vi vi /2/x2 ; : 2 k/x2 . 2
40 35 30 25

x+ 20 2
15 10 5 0 0 0.2 0.4 0.6 0.8 1

1 /x+ v+ 2 Figure 8.2: Channel ow at Re = 2000. DNS data [10, 11].Velocity gradient vs. x+ . 2 terms are negligible except the production term and the dissipation term which balance each other. Closer to the wall (Fig. 8.1a) the other terms do also play a role. Note that the production and the dissipation terms close to the wall are two orders of magnitude larger than in the logarithmic region (Fig. 8.1b). At the wall the turbulent uctuations are zero which means that the production term is zero. Since the region near the wall is dominated by viscosity the turbulent diffusion terms due to pressure and velocity are also small. The dissipation term and the viscous diffusion term attain their largest value at the wall and they much be equal to each other since all other terms are zero or negligible. The turbulence kinetic energy is produced by its main source term, the production v term, P k = v1 v2 1 /x2 . The velocity gradient is largest at the wall (see Fig. 8.2) where the shear stress is zero (see Fig. 6.3); the former decreases and the magnitude of the latter increases with wall distance and their product takes its maximum at x+ 11. 2 Since P k is largest here so is also k, see Fig. 6.7. k is transported in the x2 direction by viscous and turbulent diffusion and it is destroyed (i.e. dissipated) by .

8.2 Spatial vs. spectral energy transfer


In Section 5.3 we discussed spectral transfer of turbulent kinetic energy from large eddies to small eddies (which also applies to transport of the Reynolds stresses). In Section 8.1 we derived the equation for spatial transport of turbulent kinetic energy.

8.3. The overall effect of the transport terms

62

How are the spectral transfer and the spatial transport related? The reason that we in Section 5.3 only talked about spectral transfer was that we assumed homogeneous turbulence in which the spatial derivatives of time-averaged turbulent quantities are 2 zero, for example v1 /xi = 0, k/xi = 0 etc. (Note that the derivatives of the instantaneous turbulent uctuations are non-zero even in homogeneous turbulence, i.e. v1 /xi = 0; the instantaneous ow eld in turbulent ow is as we mentioned at the very beginning at p. 39 always three-dimensional and unsteady). In homogeneous turbulence the spatial transport terms (i.e. the convective term, term I, and the diffusion terms, term III in Eq. 8.14) are zero. Hence, in homogeneous turbulence there is no time-averaged spatial transport. However, there is spectral transfer of turbulent kinetic energy which takes place in wavenumber space, from large eddies to small eddies. The production term (term II in Eq. 8.14) corresponds to the process in which large energy-containing eddies extract energy from the mean ow. The dissipation term (term IV in Eq. 8.14) corresponds to transformation of the turbulent kinetic energy at the small eddies to thermal energy. However, real ows are hardly ever homogeneous. Some ows may have one or two homogeneous directions. Consider, for example, fully developed channel turbulent ow. If the channel walls are very long and wide compared to the distance between the walls, 2, then the turbulence (and the ow) is homogeneous in the streamwise 2 2 direction and the spanwise direction, i.e. 1 /x1 = 0, vi /x1 = 0 vi /x3 = 0 v etc. In non-homogeneous turbulence, the cascade process is not valid. Consider a large, turbulent eddy at a position xA (see Fig. 6.1) in fully developed channel ow. The 2 instantaneous turbulent kinetic energy, k = v,i v,i /2, of this eddy may either be transferred in wavenumber space or transported in physical (spatial) space, or both. It may rst be transported in physical space towards the center, and there lose its kinetic energy to smaller eddies. This should be kept in mind when thinking in terms of the cascade process. Large eddies which extract their energy from the mean ow may not give their energy to the slightly smaller eddies as assumed in Figs. 5.2 and 5.1, but k may rst be transported in physical space and then transferred in spectral space. In the inertial range (Region II), however, the cascade process is still a good approximation even in non-homogeneous turbulence. The reason is that the transfer of turbulent kinetic energy, k , from eddy-to-eddy occurs at a much faster rate than the spatial transport by convection and diffusion. In other words, the time scale of the cascade process is much smaller than that of convection and diffusion which have no time to transport k in space before it is passed on to a smaller eddy by the cascade process. We say that the turbulence at these scales are in local equilibrium. In summary, care should be taken in non-homogeneous turbulence, regarding the validity of the cascade process for the large scales (Region I).

homogeneous turbulence

local equilibrium

8.3 The overall effect of the transport terms


The overall effect (i.e. the net effect) of the production term is to increase k, i.e. if we integrate the production term over the entire domain, V , we get PK dV > 0
V

(8.18)

Similarly, the net effect of the dissipation term is a negative contribution, i.e.
V

dV < 0

(8.19)

8.4. The transport equation for vi vi /2

63

1 0.8 0.6 0.4 0.2 0 0

20

40

60

80

100

x+ 2 Figure 8.3: Channel ow at Re = 2000. Comparison of mean and uctuating : dissipation terms. Both terms are normalized by u4 /. DNS data [10, 11]. (1 /x2 )2 ; v : . What about the transport terms, i.e. convection and diffusion? Integration of the convection term over V gives, using Gauss divergence law
V

j k v dV = xj

vj knj dS
S

(8.20)

where S is the bounding surface of V . This shows that the net effect of the convection term occurs only at the boundaries. Inside the domain, the convection merely transports k with out adding or subtracting anything to k; the convection acts as a source term in part of the domain, but in the remaining part of the domain it acts as an equally large sink term. Similarly for the diffusion term, we get k 1 vj vk vk + p vj V xj xj V 1 k 1 v v v + p v = nj dS 2 j k k j xj S

(8.21)

The only net contribution occurs at the boundaries. Hence, Eqs. 8.20 and 8.21 show that the transport terms only as the word implies transports k without giving any net effect except at the boundaries. Mathematically these terms are called divergence terms, i.e. they can both be written as the divergence of a vector Aj , Aj xj where Aj for the convection and the diffusion term reads vj k convection term 1 k 1 Aj = v v v + pv diffusion term 2 j k k j xj (8.22)

divergence terms

(8.23)

8.4 The transport equation for vi vi /2

The equation for K = vi vi /2 is derived in the same way as that for vi vi /2. Multiply the time-averaged Navier-Stokes equations, Eq. 6.5, by vi so that

vi

vi vj 2 vi 1 p i vj v = vi + i v vi . xj xi xj xj xj

(8.24)

8.4. The transport equation for vi vi /2

64

Using the continuity equation and Trick 2 the term on the left side can be rewritten as vi i vj v i v 1 i vi v j K v = vj vi = vj = xj xj 2 xj xj (8.25)

Using the continuity equation the rst term on the right side of Eq. 8.24 can be written as i p v p = . (8.26) i v xi xi The viscous term in Eq. 8.24 is rewritten in the same way as the viscous term in Section 8.1, see Eqs. 8.12 and 8.13, i.e. i v K i i v v 2 vi = . xj xj xj xj xj xj (8.27)

Equations 8.25, 8.26 and 8.27 inserted in Eq. 8.24 gives


vi vj v j K v 2K 1 i p i i v v = vi . xj xj xj xi xj xj xj

(8.28)

The last term is rewritten using Trick 1 as i v Inserted in Eq. 8.28 gives j K v v 2K 1 i p i i v v = xj xj xj xi xj xj i vi vj v v i . + vi vj xj xj On the left side we have the usual convective term. On the right side we nd: viscous diffusion diffusion by pressure-velocity interaction viscous dissipation diffusion by velocity-stress interaction loss of energy to the uctuating velocity eld, i.e. to k Note that the last term in Eq. 8.30 is the same as the rst term in Eq. 8.14 but with opposite sign: here we clearly can see that the main source term in the k equation (the production term) appears as a sink term in the K equation. equation the dissipation term and the negative production term (representIn the K ing loss of kinetic energy to the k eld) read v i i v v i , + vi vj xj xj xj (8.31)
i vi vj v vi vj v i . = + vi vj xj xj xj

(8.29)

(8.30)

9. Transport equations for Reynolds stresses


and in the k equation the production and the dissipation terms read
vi vj v vi i v i xj xj xj

65

(8.32)

The gradient of the time-averaged velocity eld, vi , is much smoother than that of the uctuating velocity eld, vi . In fully turbulent ow, the dissipation by the uctuations, , is much larger than the dissipation by the mean ow (left side of Eq. 8.31). This is seen in Fig. 8.3. In the viscous-dominated region, the mean dissipation, (1 /x2 )2 , v is much larger than . At the wall, the mean dissipation takes the value = 1/2000 (normalized by u4 /).

9 Transport equations for Reynolds stresses


In Section 8 we derived transport equations for kinetic turbulent energy, k, which is the trace of the Reynolds stress tensor vi vj divided by two, i.e. k = vi vi /2. This means
2 that k is equal to twice the sum of the diagonal components of vi vj , i.e. k = 0.5(v1 + 2 2 v2 +v3 ). Here we will now derive the transport equation for the Reynolds stress tensor. This is an unknown in the time-averaged Navier-Stokes equations, Eq. 6.5, which must be known before Eq. 6.5 can be solved. The most accurate way to nd vi vj is, of course, to solve a transport equation for it. This is computationally expensive since we then need to solve six additional transport equations (recall that vi vj is symmetric, i.e. v1 v2 = v2 v1 etc.) Often, some simplications are introduced, in which vi vj is modelled by expressing it in a turbulent viscosity and a velocity gradient. This is the subject of Turbulence Modelling which you will learn about in other courses in the MSc programme. Now lets start to derive the equation for vi vj . This approach is very similar to that we used when deriving the k equation in Section 8.1. Steady, incompressible ow with constant density and viscosity is assumed. Subtract Eq. 6.5 from Eq. 6.3 and divide by density, multiply by vj and time average and we obtain

vj

[vi vk vi vk ] = xk

2 v v 1 vi p + vj + i k vj vj xi xk xk xk

(9.1)

Equation 6.5 is written with the index i as free index, i.e. i = 1, 2 or 3 so that the equation is an equation for v1 , v2 or v3 . Now write Eq. 6.5 as an equation for vj and multiply this equation by vi . We get
vi

[vj vk vj vk ] = xk

vj vk 2 vj 1 p + vi + v vi xj xk xk xk i

(9.2)

It may be noted that Eq. 9.2 is conveniently obtained from Eq. 9.1 by simply switching

9. Transport equations for Reynolds stresses


indices i and j. Adding Eqs. 9.1 and 9.2 together gives
vj

66

[v v v v ] = [vi vk vi vk ] + vi j k j k xk xk 1 p 1 p vi vj xj xi
+vi 2 2 vj vi + vj xk xk xk xk

(9.3)

vj vk vi vk vi + v xk xk j

Note that each line in the equation is symmetric: if you switch indices i and j in any of the terms nothing changes. This is important since the tensor vi vj is symmetric. Furthermore, you can check that the equation is correct according to the tensor notation rules. Indices i and j appear once in each term (not more and not less) and index k (the dummy index) appears exactly twice in each term (implying summation). Note that it is correct to use any other index than k in some terms (but you must not use i and j). You could, for example, replace k with m in the rst term and with q in the second term; it is permissible, but usually we use the same dummy index in every term. Using vi = vi + vi , the rst line can be rewritten as
vj

[i vk + vi vk + vi vk ] + vi v vj vk + vj vk + vj vk xk xk i v v j + vi vk xk xk

(9.4)

Using the continuity equation the rst terms in the two groups are rewritten as
vj vk

(9.5)

We merge the second terms in the two groups in Eq. 9.4.


vj vj vk vj vi vk v + vi = vk vj i + vk vi xk xk xk xk v v v vi j k vi j = = vk xk xk

(9.6)

The continuity equation was used twice (to get the right side on the rst line and to get the nal expression) and the the chain-rule was used backwards to get the second line. Re-writing also the third terms in the two groups in Eq. 9.4 in the same way, the second and the third terms in Eq. 9.4 can be written
vi vj vk k vi vj v + xk xk

(9.7)

The second line in Eq. 9.3 is also re-written using Trick 1


1 1 1 vi 1 vj vi p vj p + p + p xj xi xj xi

(9.8)

It will later turn out that it is convenient to express all derivatives as /xk . Therefore we re-write the derivative in the two rst terms as = jk and = ik (9.9) xj xk xi xk

9. Transport equations for Reynolds stresses

67

so that jk

1 1 vj 1 1 vi + p vi p ik vj p + p xk xk xj xi

(9.10)

The third line in Eq. 9.3 is also re-written using Trick 1


vj vi xk xk

v vj i xk xk

vi vj xk xk

Trick 1 is used again to merge the two rst terms so that the third line in Eq. 9.3 reads = xx
vi vj xk

xk

vi

vj v + vj i xk xk

vi vj xk xk

2 vi vj v vj v vj 2 i = 2 i xk xk xk xk xk xk

(9.11)

We can now put everything together. Put the rst term in Eq. 9.7 on the left side and the second term on the right side together with Eqs. 9.5, 9.10 and 9.11 so that v v i v v j (k vi vj ) = vj vk v i k xk xk xk
I II

xk

vi vj 1 1 vi vj vk + jk vi p + ik vj p xk III

(9.12)

1 + p

vj vi + xj xi V

vi vj xk xk IV

Note that the manipulation in Eq. 9.9 allows the diffusion (term III) to be written on a more compact form. After a derivation, it is always useful to check that the equation is correct according to the tensor notation rules. Every term or group of terms should include the free indices i and j (only once); Every term or group of terms should be symmetric in i and j; A dummy index (in this case index k) must appear exactly twice (=summation) in every term Equation 9.12 can also be written in a simplied easy-to-read symbolic form as Cij = Pij + Dij + ij ij where ij denotes the pressure-strain term ij = p
vj vi + xj xi

(9.13)

(9.14)

9. Transport equations for Reynolds stresses


0.5 0.1

68

0.05

0.05

a)

0.5 0

10

20

30

40

0.1

x+ 2

b)

100

200

300

400

500

x+ 2

2 Figure 9.1: Channel ow at Re = 2000. Terms in the v1 equation scaled by u4 /. a) Zoom near the wall; b) Outer region. DNS data [10, 11]. : P11 ; : 11 ; : v 2 )/x ; : 2 v 2 /x2 . 11 ; +: (v2 1 2 2 1 Equation 9.12 is the (exact) transport equation of the Reynolds stress, vi vj . It is called the the Reynolds stress equations. Since it is an equation for a second-order tensor, it consists of nine equations, but since it is symmetric we only need to consider six of them. Compare Eq. 9.12 with the equation for turbulent kinetic energy, Eq. 8.14. An alternative and maybe easier way to derive Eq. 8.14 is to rst derive Eq. 9.12 and then take the trace (setting i = j) and dividing by two. In both the k and the vi vj equations there is a convection term (I), a production term (II), a diffusion term (III) and a dissipation term (IV). In the vi vj equation there is a fth term (V), see Eq. 9.14, which is called the pressure strain term. The physical meaning of this term is to redistribute energy between the normal stress components (if we transform Eq. 9.12 to the principal coordinates of vi vj there are no shear stresses, only normal stresses). 2 The average of the normal stresses is vav = vi vi /3. For a normal stress that is larger 2 than vav , the pressure-strain term is negative and vice-versa. It is often called the Robin Hood term because it as Robin Hood takes from the rich and gives to the poor. Note that the trace of the pressure-strain term is zero, i.e.

Reynolds stress equations

pressure strain

Robin Hood

ii =

1 p

v vi + i xi xi

=0

(9.15)

because of the continuity equation and this is the reason why this term does not appear in the k equation. For 2D boundary layer ow, Eq. 9.12 reads v v i v v j (1 vi vj ) + v (2 vi vj ) = vj v2 v i 2 x1 x2 x2 x2 v vi j 1 1 vi vj v2 + j2 vi p + i2 vj p x2 x2 1 + p
vj vi + xj xi

(9.16)

vi vj xk xk

Now lets look at this equation for fully developed channel ow for which v2 = v3 = 0 () () = =0 x1 x3 (9.17)

9. Transport equations for Reynolds stresses

69

The second line shows that it is the streamwise and spanwise derivative that operate on time-averaged quantities that are zero, not those that operate on instantaneous quantities such as in ij and ij . The production term in Eq. 9.16 reads
Pij = vj v2

i v v j vi v2 x2 x2

(9.18)

2 2 2 For the v1 (i = j = 1), v2 (i = j = 2), v3 (i = j = 3) and v1 v2 (i = 1, j = 2) equations we get P11 = 2v1 v2

P22 P33 P12

1 v x2 v 2 = 0 = 2v2 v2 x2 v 3 = 0 = 2v3 v2 x2 1 v v v v v 2 = v 2 1 = v2 v2 1 2 2 x2 x2 x2

(9.19a) (9.19b) (9.19c) (9.19d)

using Eq. 9.17. 2 Figure 9.1 presents the terms in the v1 equation (Eq. 9.16 with i = j = 1). As we saw for the k equation, the production term, P11 , reaches its maximum at x2 11 2 where also v1 takes its maximum (Fig. 6.7). The pressure-strain term, 11 , and the dissipation term act as sink terms. In the outer region (Fig. 9.1b) the production term balances the pressure-strain term and the dissipation term. 2 The terms in the wall-normal stress equation, v2 , are shown in Fig. 9.2. Here we nd as expected that the pressure-strain term, 22 , acts as the main source term. 2 As mentioned previously, 22 the Robin Hood term takes from the rich v1 2 2 2 equation and gives to the poor v2 equation energy because v1 is large and v2 is small. Figure 9.3 presents the terms in the v1 v2 equation. The production term which should be a source term is here negative. Indeed it should be. Recall that v1 v2 is here negative and hence its source must be negative; or, rather, the other way around: 2 v v1 v2 is negative because its production term, P12 = v2 1 /x2 , is negative since 1 /x2 > 0. Note that in the upper half of the channel 1 /x2 < 0 and hence P12 v v and v1 v2 are positive. Furthermore, note that the dissipation, 12 , is zero. This is because dissipation takes place at the smallest scales and they are isotropic. That implies there is no correlation between two uctuating velocity components, e.g. v1 v2 = 0 (in v in isotropic turbulence are zero). Hence, also their general, for i = j, the stresses vi j gradients are zero so that v v 12 = 2 1 2 = 0 (9.20) xk xk However, very close to the wall, x+ 10, 12 = 0 because here the wall affects 2 the dissipative scales making them non-isotropic; 12 is positive since v1 v2 < 0, see Fig. 9.3. If you want to learn more how to derive transport equations of turbulent quantities, see [14] which can be downloaded here http://www.tfd.chalmers.se/lada/allpapers.html

9.1. Reynolds shear stress vs. the velocity gradient


0.05 0.02 0.015 0.01 0.005 0 0 0.005 0.01 0.015

70

a)

0.05 0

10

20

30

40

0.02

x+ 2

b)

100

200

300

400

500

x+ 2

2 Figure 9.2: Channel ow at Re = 2000. Terms in the v2 equation scaled by u4 /. a) Zoom near the wall; b) Outer region. DNS data [10, 11]. : P22 ; : 22 ; 2 2 : 22 ; +: (v2 v2 )/x2 ; : 2 v2 /x2 . : 2v2 p /x2 ; 2
0.15 0.1 0.05 0 0.05 0.1 0.05 0 0.05

a)

10

20

30

40

x+ 2

b)

100

200

300

400

500

x+ 2

Figure 9.3: Channel ow at Re = 2000. Terms in the v1 v2 equation scaled by u4 /. a) Zoom near the wall; b) Outer region. DNS data [10, 11]. : P12 ; : 12 ; 2 : 12 ; +: (v1 v2 )/x2 ; : 2 v1 v2 /x2 . : v2 p /x2 ; 2

9.1 Reynolds shear stress vs. the velocity gradient


In boundary-layer type of ow, the Reynolds shear stress and the velocity gradient 1 /x2 have nearly always opposite signs. For channel ow, for example, Eq. 9.19 v shows that P12 is negative (and hence also v1 v2 ) in the lower half because 1 /x2 > 0 v and it is positive in the upper half because 1 /x2 < 0. This can also be shown by v physical argumentation. Consider the ow in a boundary layer, see Fig. 9.4. A uid particle is moving downwards (particle drawn with solid line) from x2,B to x2,A with (the turbulent uctuating) velocity v2 . At its new location the v1 velocity is in average smaller than at its old, i.e. v1 (x2,A ) < v1 (x2,B ). This means that when the particle at x2,B (which has streamwise velocity v1 (x2,B )) comes down to x2,A (where the streamwise velocity is v1 (x2,A )) it has an excess of streamwise velocity compared to its new environment at x2.A . Thus the streamwise uctuation is positive, i.e. v1 > 0 v < 0). and the correlation between v1 and v2 is in average negative (v1 2 If we look at the other particle (dashed line in Fig. 9.4) we reach the same con clusion. The particle is moving upwards (v2 > 0), and it is bringing a decit in v1 so that v1 < 0. Thus, again, v1 v2 < 0. If we study this ow for a long time and average over time we get v1 v2 < 0. If we change the sign of the velocity gradient so that 1 /x2 < 0 we will nd that the sign of v1 v2 also changes. v

10. Correlations

71

x2,B
v2 < 0 v2 > 0

x2,A

v1 (x2 )

x2 x1
Figure 9.4: Sign of the Reynolds shear stress v1 v2 in a boundary layer.

Figure 10.1: Two-point correlation.

|xA xC | 1 1

In cases where the shear stress and the velocity gradient have the same sign (for example, in a wall jet) the reason is that the other terms (usually the transport terms) are more important than the production term.

10 Correlations
10.1 Two-point correlations
Two-point correlations are useful when describing some characteristics of the turbulence. Pick two points along the x1 axis, say xA and xC , and sample the uctuating 1 1 velocity, in, for example, the x1 direction. We can then form the correlation of v1 at these two points as (10.1) B11 (xA , xC ) = v1 (xA )v1 (xC ) 1 1 1 1 Often, it is expressed as
B11 (xA , x1 ) = v1 (xA )v1 (xA + x1 ) 1 1 1

(10.2)

where x1 = xC xA is the separation distance between point A and C. 1 1

10.1. Two-point correlations

72

xA 1

Lint

x1
(a) Small integral length scale

xA 1 Lint

x1
(b) Large integral length scale

Figure 10.2: Schematic relation between the two-point correlation, the largest eddies (thick lines) and the integral length scale, Lint .

10.2. Auto correlation

73

It is obvious that if we move point A and C closer to each other, B11 increases; when the two points are moved so close that they merge, then B11 = v 2 (xA ). If, 1 on the other hand, we move point C further and further away from point A, then B11 will go to zero. Furthermore, we expect that the two-point correlation function will be related to the largest eddies. It is convenient to normalize B11 so that it varies between 1 and +1. The normalized two-point correlation reads
norm B11 (xA , x1 ) = 1

1 v1,rms (xA )v1,rms (xA 1 1 + x1 )

v1 (xA )v1 (xA + x1 ) 1 1

(10.3)

where subscript rms denotes root-mean-square, which for v1 , for example, is dened as 2 v1,rms = v1 1/2

(10.4)

RMS is the same as standard deviation (Matlab command std) which is the squareroot of the variance (Matlab command var). Consider a ow where the largest eddies have an eddy scale (length scale) of Lint , see Fig. 10.2. We expect that the two point correlation, B11 , approaches zero for separation distance, |xA xC | > Lint because for separation distances larger than |xA xB | 1 1 1 1 there is no correlation between v1 (xA ) and v1 (xC ). Hence, ows with large eddies will 1 1 have a two-point correlation function which decreases slowly with separation distance. For ows with small eddies, the two-point correlation, B11 , decreases rapidly with x1 . If the ow is homogeneous (see p.61) in the x1 direction, the two-point correlation does not depend on the location of xA , i.e. it is only dependent on the separation of the 1 two points, x1 . From the two-point correlation, B11 , an integral length scale, Lint which is dened as the integral of B11 over the separation distance, i.e.

integral length scale

Lint (x1 ) =
0

B11 (x1 , x1 ) d1 x A C v1,rms v1,rms

(10.5)

If the ow is homogeneous in the x1 direction then Lint does not depend on x1 .

10.2 Auto correlation


Auto correlation is a two-point correlation in time, i.e. the correlation of a turbulent uctuation with a separation in time. If we again choose the v1 uctuation, the auto correlation reads (10.6) B11 (tA , t ) = v1 (tA )v1 (tA + t ) where t = tC tA , is the time separation distance between time A and C. If the mean ow is steady, the time direction is homogeneous and then B is independent on tA ; in this case the auto-correlation depends only on time separation, t, i.e.
B11 (t ) = v1 (t)v1 (t + t )

(10.7)

where any value t can be used. The normalized auto-correlation reads


norm B11 (t ) =

1 v (t)v1 (t 2 v1,rms 1

+ t)

(10.8) integral time scale

In analogy to the integral length scale, Lint , the integral time scale, Tint , is dened

10.2. Auto correlation

74

as (assuming steady ow)

Tint =
0

norm B11 (t)dt

(10.9)

10.2. Auto correlation

75

MTF270 Turbulence Modelling


L. Davidson
Division of Fluid Dynamics, Department of Applied Mechanics Chalmers University of Technology, G teborg, Sweden o http://www.tfd.chalmers.se/lada, lada@chalmers.se
This report can be downloaded at http://www.tfd.chalmers.se/lada/turbulent flow/

11. Reynolds stress models and two-equation models

76

11 Reynolds stress models and two-equation models


11.1 Mean ow equations
11.1.1 Flow equations For incompressible turbulent ow, all variables are divided into a mean part (time averaged) and uctuating part. For the velocity vector this means that vi is divided into a mean part vi and a uctuating part vi so that vi = vi + vi . Time average and we get (see Eq. 6.4 at. p. 47): (j ) = 0 v + t xj p 2 vi ij i v (i vj ) = v + ( 0 )gi + t xj xi xj xj xj (11.1) (11.2)

(note that denotes temperature) where fi = ( 0 )gi and the turbulent stress tensor (also called Reynolds stress tensor) is written as:
ij = vi vj

Reynolds stress tensor

The pressure, p, denotes the hydro-static pressure, see Eq. 3.22, which means that when the ow is still (i.e. vi 0), then the pressure is zero (i.e. p 0). The body force fi which was omitted for convenience in Eq. 6.4 has here been re-introduced. The body force in Eq. 11.2 is due to buoyancy, i.e. density differences. The basic form of the buoyancy force is fi = gi where gi denotes gravitational acceleration. Since the pressure, p, is dened as the hydro-static pressure we must rewrite the buoyancy source as fi = ( 0 )gi (11.3)

so that p 0 when vi 0 (note that the true pressure decreases upwards as gh where h denotes change in height). If we let density depend on pressure and temperature, differentiation gives d = d +
p

dp

(11.4)

Our ow is incompressible, which means that the density does not depend on pressure, i.e. /p = 0; it may, however, depend on temperature and mixture composition. Hence the last term in Eq. 11.4 is zero and we introduce the volumetric thermal expansion, , so that d = 0 d 0 = 0 ( 0 ) = 1 0
p

(11.5)

Now we can re-write the buoyancy source as fi = ( 0 )gi = 0 ( 0 )gi (11.6)

which is the last term in Eq. 11.2. Consider the case where x3 is vertically upwards. Then gi = (0, 0, g) and a large temperature in Eq. 11.6 results in a force vertically upwards, which agrees well with our intuition.

11.2. The exact vi vj equation

77

11.1.2 Temperature equation The instantaneous temperature, , is also decomposed into a mean and a uctuating component as = + . The transport equation for reads (see Eq. 2.15 where temperature was denoted by T ) 2 vi = + t xi xi xi Introducing = + we get i 2 v v = i + t xi xi xi xi (11.8) (11.7)

The last term on the right side is an additional term whose physical meaning is turbulent heat ux vector. This is similar to the Reynolds stress tensor on the right side of the time-averaged momentum equation, Eq. 11.2. The total heat ux vector viscous plus turbulent in Eq. 11.8 reads (cf. Eq. 2.11) ci,tot = ci + ci,turb =
11.2 The exact vi vj equation

vi xi

(11.9)

Now we want to solve the time-averaged continuity equation (Eq. 11.1) and the three momentum equations (Eq. 11.2). Unfortunately there are 10 unknowns; the four usual ones (i , p) plus six turbulent stresses, vi vj . We must close this equation system; it is v called the closure problem. We must nd some new equations for the turbulent stresses. We need a turbulence model. The most comprehensive turbulence model is to derive exact transport equations for the turbulent stresses. An exact equation for the Reynolds stresses can be derived from the Navies-Stokes equation. It is emphasized that this equation is exact; or, rather, as exact as the Navier-Stokes equations. The derivation follows the steps below.
Set up the momentum equation for the instantaneous velocity vi = vi + vi Eq. (A)

closure problem

Time average equation for vi , Eq. (B)


Subtract Eq. (B) from Eq. (A) equation for vi , Eq. (C) Do the same procedure for vj equation for vj , Eq. (D) Multiply Eq. (C) with vj and Eq. (D) with vi , time average and add them together v equation for vi j

In Section 9 at p. 65 these steps are given in some detail. More details can also be a found in [14] (set the SGS tensor to zero, i.e. ij = 0).

11.3. The exact vi equation

78

The nal vi vj -equation (Reynolds Stress equation) reads (see Eq. 9.12) vi vj vi vj v v j v v i + p = vi vk + vk j k t xk xk xk vj vi + xj xi

Cij

Pij
p vj

ij
2 vi vj

p vi vi vj vk + ik + jk + xk xk xk

(11.10)

Dij,t
gi vj gj vi 2 Gij vi vj xk xk ij

Dij,

where Dij,t and Dij, denote turbulent and viscous diffusion, respectively. The total diffusion reads Dij = Dij,t + Dij, . This is analogous to the momentum equation where we have gradients of viscous and turbulent stresses which correspond to viscous and turbulent diffusion. Equation 11.10 can symbolically be written Cij = Pij + ij + Dij + Gij ij where Cij Convection Pij Production ij Pressure-strain Dij Diffusion Gij Buoyancy production ij Dissipation
11.3 The exact vi equation

If temperature variations occurs we must solve for the mean temperature eld, see Eq. 11.8. To obtain the equation for the uctuating temperature, subtract Eq. 11.8 from Eq. 11.7 v 2 (vk + vk + vk ) = + k + t xk xk xk xk (11.11)

To get the equation for the uctuating velocity, vi , subtract the equation for the mean velocity vi (Eq. 11.2) from the equation for the instantaneous velocity, vi (Eq. 6.3) so that v v vi 1 p 2 vi (vk vi + vk vi + vk vi ) = + + i k gi (11.12) + t xk xi xk xk xk

11.3. The exact vi equation

79

Multiply Eq. 11.11 with vi and multiply Eq. 11.12 with , add them together and time average vi (v + v + v ) + ( v + v v + v v ) k vi k k i + vi i k k t xk k xk 2 2 vi p = + vi + gi xi xk xk xk xk

(11.13)

The Reynolds stress term in Eq. 11.12 multiplied by and time averaged is zero, i.e.
vi vj vi vj = 0 = xk xk

The rst term in the two parentheses on line 1 in Eq. 11.13 are combined into two production terms (using the continuity equation, /xk (vk ) = 0)
vi vk

v + vk xk xk

(11.14)

The second term in the two parenthesis on the rst line of Eq. 11.13 are re-written using the continuity equation
vi

k v k vi v + = vk xk xk

vi

v + i xk xk

(11.15)

Now the two terms can be merged (chain-rule backwards) vk


vi k vi v = xk xk

(11.16)

where we used the continuity equation to obtain the right side. The last two terms in Eq. 11.13 are re-cast into turbulent diffusion terms using the same procedure as in Eqs. 11.15 and 11.16 vi vk (11.17) xk The viscos diffusion terms on the right side are re-written using the chain-rule backwards (Trick 1, see p. 58)
vi

2 = vi xk xk xk
2 vi = xk xk xk

xk
vi xk

xk

vi

xk

vk xk xk

= xk

vi xk

vk xk xk

(11.18)

Inserting Eqs. 11.14, 11.16, 11.17 and 11.18 into Eq. 11.13 gives the transport equation for the heat ux vector vi
vi v v i p v v vk vi = vi vk + k t xk xk xk xi xk k i Pi i Di,t

+ xk

vi

xk

+ xk
Di,

v i

xk

( + )

vi

(11.19)

xk xk

gi

Gi

11.4. The k equation

80

where Pi , i and Di,t denote the production, scramble and turbulent diffusion term, respectively. The production term include one term with the mean velocity gradient and one term with the mean temperature gradient. On the last line, Di, , i and Gi denote viscous diffusion, dissipation and buoyancy term, respectively. The unknown terms i , Di , i , Gi have to be modelled as usual; this is out of the scope of the present course but the interested reader is referred to [15].

11.4 The k equation


The turbulent kinetic energy is the sum of all normal Reynolds stresses, i.e. k= 1 1 2 2 2 v + v2 + v3 vi vi 2 1 2

By taking the trace (setting indices i = j) of the equation for vi vj and dividing by two we get the equation for the turbulent kinetic energy: k v k i vi vi + vj = vi vj t xj xj xj xj

Ck xj

Pk p
k Dt

(11.20)

vj

1 + vi vi 2

k gi vi xj xj
k D Gk

k k where as in the vi vj equation Dt and D denotes turbulent and viscous diffusion, k k k respectively. The total diffusion reads D = Dt + D . Equation 11.20 can symbolically be written: C k = P k + Dk + Gk (11.21)

11.5 The equation


Two quantities are usually used in eddy-viscosity model to express the turbulent viscosity. In the k model, k and are used. The turbulent viscosity is then computed as k2 t = C where C = 0.09. An exact equation for the transport equation for the dissipation =
vi vi xj xj

can be derived (see, e.g., [16]), but it is very complicated and in the end many terms are found negligible. It is much easier to look at the k equation, Eq. 11.21, and to setup a similar equation for . The transport equation should include a convective term, C , a diffusion term, D , a production term, P , a production term due to buoyancy, G , and a destruction term, , i.e. C = P + D + G (11.22)

11.6. The Boussinesq assumption

81

The production and destruction terms in the k equation are used to formulate the corresponding terms in the equation. The terms in the k equation have the dimension k/t = [m2 /s3 ] whereas the terms in the equation have the dimension /t = [m2 /s4 ]. Hence, we must multiply P k and by a quantity which dimension [1/s]. One quantity with this dimension is the mean velocity gradient which might be relevant for the production term, but not for the destruction. A better choice should be /k = [1/s]. Hence, we get P + G = (c1 P k + c1 Gk c2 ) k (11.23)

The turbulent diffusion term is expressed in the same way as that in the k equation (see Eq. 11.32) but with its own turbulent Prandtl number, (see Eq. 11.29), i.e. D = xj + t xj (11.24)

The nal form of the transport equation reads = (c1 P k + c1 Gk c2 ) + + vj t xj k xj + t xj (11.25)

11.6 The Boussinesq assumption


In the Boussinesq assumption an eddy (i.e. a turbulent) viscosity is introduced to model the unknown Reynolds stresses in Eq. 11.2. The stresses are modelled as
vi vj = t

i v j v + xj xi

2 2 + ij k = 2t sij + ij k 3 3

(11.26)

The last term is added to make the equation valid also when it is contracted (i.e taking the trace); after contraction both left and right side are equal (as they must be) and equal to vi vi = 2k. When Eq. 11.26 is included in Eq. 11.2 we replace 6 turbulent stresses with one new unknown (the turbulent viscosity, t ). This is of course a drastic simplication. If the mean temperature equation is solved for we need an equation for the heat ux vector, vi . One option is to solve its transport equation, Eq. 11.19. If an eddyviscosity model (i.e. Eq. 11.26) is used for the Reynolds stresses, an eddy-viscosity model is commonly used also for the heat ux vector. The Boussinesq assumption reads (11.27) vi = t xi where t denotes the turbulent thermal diffusivity. It is usually obtained from the turbulent viscosity as t t = (11.28) where is the turbulent Prandtl number; it is an empirical constant which is usually set to 0.7 0.9. The physical meaning of the turbulent Prandtl number, , is analogous to the physical meaning of the usual Prandtl number, see Eq. 2.16; it denes how efcient the turbulence transports (by diffusion) momentum compared to how efcient it transports thermal energy, i.e. = t t (11.29)

11.7. Modelling assumptions

82

It is important to recognize that the viscosity (), the Prandtl number (P r), the thermal diffusivity () are physical parameters which depend on the uid (e.g. water or air) and its conditions (e.g. temperature). However, the turbulent viscosity (t ), the turbulent thermal diffusivity (t ) and the turbulent Prandtl number ( ) depend on the ow (e.g. mean ow gradients and turbulence).

11.7 Modelling assumptions


Now we will compare the modelling assumptions for the unknown terms in the vi vj , vi , k and equations and formulate modelling assumptions for the remaining terms in the Reynolds stress equation. This will give us the Reynolds Stress Model [RSM] (also called the Reynolds Stress underline Transport Model [RSTM]) where a (modelled) transport equation is solved for each stress. Later on, we will introduce a simplied algebraic model, which is called the Algebraic Stress Model [ASM] (this model is also called Algebraic Reynolds Stress Model, ARSM) Summary of physical meaning: Pij , Pi and P k are production terms of vi vj , vi and k Gij , Gi and Gk are production terms of vi vj , vi and k due to buoyancy k Dij,t , Di,t , Dt are the turbulent diffusion terms of vi vj , vi and k i is the pressure-scramble terms of vi

ij is the pressure-strain correlation term, which promotes isotropy of the turbulence


ij , i and are dissipation of vi vj , vi and k, respectively. The dissipation takes place at the small-scale turbulence.

11.7.1 Production terms In RSM and ASM the production terms are computed exactly
Pij = vi vk

Pi

j v v i , vj vk xk xk v v i = vi vk k xk xk

Pk =

1 v i Pii = vi vj 2 xj

(11.30)

The k is usually not solved for in RSM but a length-scale equation (i.e. or ) is always part of an RSM and that equation includes P k . In the k model, the Reynolds stresses in the production term are computed using the Boussinesq assumption, which gives
vi vj = t

P k = t sij = 1 2

j v i v + xj xi i v j v + xj xi i v j v + xj xi

2 ij k 3 i v = 2t sij sij xj

(11.31)

11.7. Modelling assumptions

83

11.7.2 Diffusion terms The diffusion terms in the k and -equations in the k model are modelled using the standard gradient hypothesis which reads Dk = xj D = xj t k t + + k xj xj

(11.32)

The gradient hypothesis simply assumes that turbulent diffusion acts as to even out k all inhomogeneities. In other words, it assumes that the turbulent diffusion term, Dt , transports k from regions where k is large to regions where k is small. The turbulent diffusion ux of k is expressed as dk = j,t t k 1 v vv = 2 j i i k xj (11.33)

Only the triple correlations are included since the pressure diffusion usually is negligible (see Fig. 8.1 at p. 61). Taking the divergence of Eq. 11.33 (including the minus sign in Eq. 11.20) gives the turbulent diffusion term in Eq. 11.32. Solving the equations for the Reynolds stresses, vi vj , opens possibilities for a more advanced model of the turbulent diffusion terms. Equation 11.33 assumes that if the gradient is zero in xi direction, then there is no diffusion ux in that direction. A more general gradient hypothesis can be formulated without this limitation, e.g.
dk j,t,G vj vk

k xk

(11.34)

which is called the general gradient diffusion hypothesis (GGDH). It was derived in [17] from the transport equation of the triple correlation vj vi vi . In GGDH the turbulent k ux d1,t,G , for example, is computed as
dk 1,t,G v1 v1

k k + v v k + v1 v2 1 3 x1 x2 x3

(11.35)

Hence, even if k/x1 = 0 the diffusion ux dk 1,t,G may be non-zero. A quantity of dimension [s] must be added to get the correct dimension, and as in Eq. 11.23 we take k/ so that k k (11.36) dk j,t,G = ck vj vk xk
k The diffusion term, Dt , in the k equation is obtained by taking the divergence of this equation dk k k j,t,G k = ck vj vk (11.37) Dt = xj xj xk

This diffusion model may be used when the k equation is solved in an RSM or an ASM. The corresponding diffusion terms for the and vi vj equations read
Dt =

xj xk

c vj vk

k xk
k vi vj xm

(11.38)

Dij,t =

ck vk vm

11.7. Modelling assumptions

84

Equation 11.38 often causes numerical problems. A more stable alternative is to model the diffusion terms as in 11.32 which for vi vj reads Dij,t = 11.7.3 Dissipation term, ij The dissipation term ij (see Eq. 11.10) is active for the small-scale turbulence. Because of the cascade process and vortex stretching (see Figs. 5.2 and 5.3) the smallscale turbulence is isotropic. This means that the velocity uctuations of the smallscale turbulence have no preferred direction, see p. 44. This gives:
2 2 2 1. v1 = v2 = v3 .

xm

t vi vj k xm

(11.39)

2. All shear stresses are zero because the uctuations in two different coordinate directions are not correlated. What applies for the small-scale uctuations (Items 1 and 2, above) must also apply to the gradients of the uctuations, i.e.
v1 v1 v2 v2 v3 v3 = = xk xk xk xk xk xk v vi j = 0 if i = j xk xk

(11.40)

The relations in Eq. 11.40 are conveniently expressed in tensor notation as ij = 2 ij 3 (11.41)

1 where the factor 2/3 is included so that = 2 ii is satised, see Eqs. 11.10 and 11.20.

11.7.4 Slow pressure-strain term


The pressure-strain term, ij , makes a large contribution to the vi vj equation. In Section 9 it was shown that for channel ow it is negative for the streamwise equation, 2 2 2 v1 , and positive for the wall-normal, v2 , and spanwise, v3 , equations. Furthermore, it acts as a sink term for the shear stress equation. In summary, it was shown that the term acts as to make the turbulence more isotropic, i.e. decreasing the large normal stresses and the shear stress and increasing the small normal stresses. The pressure-strain term is often called the Robin Hood terms, because it takes from the rich and gives to the poor. The role of the pressure strain can be described in physical terms as follows. As sume that two uid particles with uctuating velocities v1 bounce into each other at O so that v1 /x1 < 0, see Fig. 11.1. As a result the uctuating pressure p increases at O so that v p 1 < 0 x1

The uid in the x1 direction is performing work, moving uid particles against the pressure gradient. The kinetic energy lost in the x1 direction is transferred to the x2

11.7. Modelling assumptions

85

v1

v1

x2

x1 Figure 11.1: Physical illustration of the pressure-strain term. and x3 directions and we assume that the collision makes uid particles move in the other two directions, i.e. v2 v3 > 0, >0 (11.42) x2 x3
Indeed, if v1 /x1 < 0, the continuity equation gives v2 /x2 + v3 /x3 > 0. However, in Eq. 11.42 we assume that not only their sum is positive but also that they both are positive. If this is to happen the kinetic energy in the x1 direction must be 2 2 2 2 larger than that in the x2 and x3 direction, i.e. v1 > v2 and v1 > v3 . If this were not true, the uctuation v1 would not be able to create an acceleration of both v2 and v3 . The amount of kinetic energy transferred from the x1 direction to the x2 and x3 directions, should be proportional to the difference of their energy, i.e. 1 v1 1 p x1 2

2 2 2 2 v1 v2 + v1 v3

2 = v1

1 2 2 v + v3 2 2

3 2 1 2 2 2 v v + v2 + v3 2 1 2 1

3 2 v k 2 1

(11.43)

The expression in Eq. 11.43 applies only to the normal stresses, i.e. the principal axis of vi vj . By transforming to a coordinate system which is rotated /4 it is shown that the sign of p (vi /xj + vj /xi ) and vi vj are opposite. Assume that we express Eq. 11.43 in principal coordinates, (x1 , x2 ), and then transform the equation to (x1 , x2 ) by rotating it angle = /4, see Appendix M. Replacing u12 in Eq. M.6b by v1 v2 we get
2 2 v1 v2 = 0.5 v1 v2

(11.44)

since v1 v2 = v2 v1 . Now we have transformed the right side of Eq. 11.43. Next step is to transform the left side, i.e. the velocity gradients. We use Eqs. M.6b and M.6c: replacing u12 and u21 by v1 /x2 and v2 /x1 , respectively, and adding them gives v2 v v1 v + 1 = 2 (11.45) x1 x2 x1 x2

11.7. Modelling assumptions

86

v1

x2

x2

a)

x1

b)

x3

Figure 11.2: Decaying grid turbulence. The circles (a) and the thin rectangles (b) illustrates past of the grid which consists of a mesh of circular cylinders. the pressure-strain term in Eqs. 11.10 and 11.43 can be written p
v2 v + 1 x1 x2

= p

v1 v 2 x1 x2

(11.46)

Now we apply Eq. 11.43 using the right side of Eq. 11.46 p
v1 v 2 x1 x2 2 2 v1 v2

(11.47)

Inserting Eqs. 11.44 and 11.46 into Eq. 11.47 gives nally p
v v2 + 1 x1 x2 = v1 v2

(11.48)

This shows that the pressure-strain term acts as a sink term in the shear stress equation. Thus, Eqs. 11.43 and 11.48 lead as to write ij,1 p
vj vi + xj xi

= c1

2 vi vj ij k 3

(11.49)

where denotes the modelled pressure-strain term and subscript 1 means the slow part; the concept slow and rapid is discussed at p. 87. We have introduced the turbulent time scale k/. This pressure-strain model for the slow part was proposed by Rotta in 1951 [18]. Let us investigate how Eq. 11.49 behaves for decaying grid turbulence, see Fig. 11.2. Flow from left with velocity v1 passes through a grid. The grid creates velocity gradi ents behind the grid which generate turbulence. Further downstream the velocity gradients are smoothed out and the mean ow becomes constant. From this point and further downstream the ow represents homogeneous turbulence which is slowly approaching

11.7. Modelling assumptions

87

isotropic turbulence; furthermore the turbulence is slowly dying (i.e. decaying) due to dissipation. The exact vi vj equation for this ow reads (no production or diffusion because of homogeneity) v1
vi vj p = x1 vj vi + xj xi

ij

(11.50)

Rottas pressure-strain model is supposed to reduce anisotropy. Thus it should be interesting to re-write Eq. 11.50 expressed in the normalized anisotropy Reynolds stress tensor which is dened as vi vj 1 ij (11.51) bij = 2k 3 We introduce bij (Eq. 11.51), Rottas model (Eq. 11.49) and the model for the dissipation tensor (11.41) into Eq. 11.50 so that 21 v 1 k (kbij ) + ij x1 3 x1 2 = 2c1 bij ij 3 (11.52)

Analogously to Eq, 11.50, the k equation in decaying grid turbulence reads v1 k = x1 (11.53)

Inserting Eq. 11.53 in Eq. 11.52 and dividing by 2k we obtain v1 1 1 bij = c1 bij ij + bij + ij = bij (1 c1 ) x1 k 3 k k 3 k k (11.54)

Provided that c1 > 1 Rottas model does indeed reduce non-isotropy as it should. The model of the slow pressure-strain term in Eq. 11.49 can be extended by in cluding terms which are non-linear in vi vj . To make it general it is enough to include terms which are quadratic in vi vj , since according to the Cayley-Hamilton theorem, a second-order tensor satises its own characteristic equation (see Section 1.20 in [19]); this means that terms cubic in vi vj can be expressed in terms linear and quadratic in v . The most general form of vi j ij,1 can be formulated as [20] 1 ij,1 = c1 aij + c aik akj ij ak ak 1 3
vi vj 2 aij = ij k 3

(11.55)

aij is the anisotropy tensor whose trace is zero. In isotropic ow all its components are zero. Note that the right side is trace-less (i.e. the trace is zero). This should be so since the exact form of ij is trace-less, i.e. ii = 2p vi /xi = 0. 11.7.5 Rapid pressure-strain term Above a model for the slow part of the pressure-strain term was developed using physical arguments. Here we will carry out a mathematical derivation of a model for the rapid part of the pressure-strain term. The notation rapid comes from a classical problem in turbulence called the rapid distortion problem, where a very strong velocity gradient i /xj is imposed so that v

11.7. Modelling assumptions

88

y V

x2 , y2

x1 , y1 Figure 11.3: The exact solution to Eq. 11.58. The integral is carried out for all points, y, in volume V . initially the second term (the slow term) can be neglected, see Eq. 11.57. It is assumed that the effect of the mean gradients is much larger than the effect of the turbulence, i.e. i v (/k) (11.56) xj Thus in this case it is the rst term in Eq. 11.57 which gives the most rapid response in p . The second slow term becomes important rst at a later stage when turbulence has been generated. Now we want to derive an exact equation for the pressure-strain term, ij . Since it includes the uctuating pressure, p , we start by deriving an exact equation for p starting from Navier-Stokes equations. 1. Take the divergence of incompressible Navier-Stokes equation assuming con2 stant viscosity (see Eq. 6.2) i.e. x xi (vi vj ) = . . . Equation A. j 2. Take the divergence of incompressible time-averaged Navier-Stokes equation as2 suming constant viscosity (see Eq. 6.4) i.e. x xi (i vj ) = . . . Equation B. v j Subtracting of Equation B from Equation A gives a Poisson equation for the uctuating pressure p
i vj v 1 2 p 2 = 2 v v vi vj xj xj xj xi xi xj i j rapid term slow term

(11.57)

For a Poisson equation 2 =f (11.58) xj xj there exists an exact analytical solution given by Greens formula (it is derived from Gauss divergence law) (x) = 1 4 f (y)dy1 dy2 dy3 |y x| (11.59)

11.7. Modelling assumptions

89

where the integrals at the boundaries vanish because it is assumed that f 0 at the boundaries, see Fig. 11.3. Applying Eq. 11.59 on Eq. 11.57 gives p (x) = 4
V

where dy3 = dy1 dy2 dy3 . Now make two assumptions in Eq. 11.60:

(y) v (y) 2 vi j + v (y)vj (y) vi (y)vj (y) 2 yi yi yj i yj


rapid term slow term

dy3 |y x|

(11.60)

i) the turbulence is homogeneous (i.e. the spatial derivative of all time-averaged uctuating quantities is zero). This requirement is not as drastic as it may sound (very few turbulent ows are homogeneous). This term is indeed very small compared to the second derivative of the instantaneous uctuations, vi (y)vj (y. ii) the variation of i /xj in space is small. The same argument can be used as v above: the mean gradient i /xj varies indeed much more slowly than the v instantaneous velocity gradient, vj (y)/yi Assumption i) means that the last term in the integral in Eq. 11.60 is zero, i.e.
2 vi vj =0 yi yj

Assumption ii) means that the mean velocity gradient can be taken outside the integral. Now multiply Eq. 11.60 with vi /xj + vj /xi . Since this term is not a function of y it can be moved in under the integral. We obtain after time averaging 1 p (x) =
vi (x) vj (x) + xj xi vi (x) vj (x) + xj xi Mijk v (y) dy3 yk |y x|

k (x) 1 v x 2

(11.61)

1 4

vi (x) vj (x) + xj xi

dy3 2 (vk (y)v (y)) yk y |y x|


Aij

Note that the mean velocity gradient, /x , is taken at point x because it has been v moved out of the integral. In order to understand this better, consider the integral
L

f (x) =
0

g()d |x |

(11.62)

Note that x and are coordinates along the same axis (think of them as two different points along the x axis). If the two points, x and , are far from each other, then the denominator is large and the contribution to the integral is small. Hence, we only need to consider points which are close to x. If we assume that g() varies slowly with ,

11.7. Modelling assumptions

90

g() can be moved out of the integral and since x is close to , Eq. 11.62 can be written as L d f (x) = g(x) (11.63) |x | 0 Going from Eq. 11.62 to Eq. 11.63 corresponds to moving the mean velocity gradient out of the integral. Equation 11.61 can be written on shorter form as p
vj vi + xj xi

= Aij + Mijk

k v = ij,1 + ij,2 x

(11.64)

where the rst term represents the slow term, ij,1 (see Eq. 11.49), and second term the rapid term, ij,2 (index 2 denotes the rapid part). Now we will take a closer look at rapid part (i.e. the second term) of Mijk . The second term of Mijk in the integral in Eq. 11.61 can be rewritten as
vj (x) v (y) = xi yk yk v (y) vj (x) xi v (y) 2 vj (x) yk xi vj (x) v (y) xi

= =

2 v (y)vj (x) yk xi yk 2 v (y)vj (x) yk xi

(11.65)

2 vj (x)/yk xi on line 1 is zero because vj (x) is not a function of y. For the same reason the last term on line 2 is zero. Note that the terms above as well as in Eq. 11.61 are two-point correlations, the two points being x and y. Introduce the distance vector between the two points

ri = yi xi Differentiating Eq. 11.66 gives = ri yi xi Equation 11.66 is a coordinate transformation where we replace xi and yi with I. xi and ri , or II. yi and ri .

(11.66)

(11.67)

Assumption i) at p. 89 gives that /xi = 0 (Item I) or /yi = 0 (Item II). In other words, the two-point correlations are independent of where in space the two points are located; they are only dependent on the distance between the two points (i.e. ri ). Hence we can replace the spatial derivative by the distance derivative, i.e. = xi ri = yi ri

(11.68)

11.7. Modelling assumptions


We can now write Mijk in Eq. 11.61, using Eqs. 11.65 and 11.68, as Mijk = 1 2 2 2 v vj + v v rk ri rk rj i dr3 |r|

91

(11.69)

= aijk + ajik

It can be shown that aijk is symmetric with respect to index j and (recall that v and vj are not at the same point but separated by ri ), i.e.

aijk = aikj

(11.70)

see Appendix F on p. 199. Furthermore, Eq. 11.69 is independent of in which order the two derivatives are taken, so that aijk is symmetric with respect to i and k, i.e. aijk = akji (11.71)

Now let us formulate a general expression of aijk which is linear in vi vj and symmetric in (j, ) and (i, k). We get aijk = c1 ik vj v + c2 j vi vk + c3 (ij vk v + kj vi v + i vk vj + k vi vj )

(11.72)

+ c4 j ik k + c5 (ij k + jk i )k Each line is symmetric in (j, ) and (i, k). For example, on line 3, term 1 & term 3 and term 2 & term 4 are symmetric with respect to j and and term 1 & term 2 and term 3 & term 4 are symmetric with respect to i and k. Consider Eq. 11.61. Here it is seen that if i = j then Mijk = 0 due to the continuity equation; looking at Eq. 11.69 we get aiik = 0 Applying this condition to Eq. 11.72 gives
0 = c1 ik vi v + c2 i vi vk + c3 (3vk v + ki vi v + i vk vi + k vi vi ) + c4 i ik k + c5 (3k + ik i )k = c1 vk v + c2 v vk + c3 (3vk v + vk v + vk v + 2k k) + c4 k k + c5 (3k + k )k = vk v (c1 + c2 + 5c3 ) + kk (c4 + 2c3 + 4c5 )

(11.73)

(11.74)

Greens third formula reads (see Appendix F on p. 199)


aiji = 2vj v

(11.75)

Using Eq. 11.75 in Eq. 11.72 gives


2vj v = 3c1 vj v + c2 j vi vi + c3 (ij vi v + ij vi v + i vi vj + i vi vj )

+ (3c4 j + c5 (ij i + ji i ))k


= 3c1 vj v + 2c2 j k + 4c3 vj v + (3c4 + 2c5 )j )k = vj v (3c1 + 4c3 ) + j k(2c2 + 3c4 + 2c5 )

(11.76)

11.7. Modelling assumptions

92

Equations 11.74 and 11.76 give four equations c1 + c2 + 5c3 = 0, 3c1 + 4c3 2 = 0, 4c2 + 10 , 11 3c2 + 2 , 11 c4 + 2c3 + 4c5 = 0 2c2 + 3c4 + 2c5 = 0 (11.77)

for the ve unknown constants. Let us express all constants in c2 which gives c1 = c3 = c4 = 50c2 + 4 , 55 c5 = 20c2 + 6 55 (11.78)

Inserting Eq. 11.78 into Eq. 11.72 and 11.64 gives k v k v = (aijk + ajik ) x x v v v v i + v v j k + v v k = c1 vj v + c2 vi vk i j k x x xj xi v v v v v k + v v j + v v i + v v k + v v k 2ij vk v i j k j k i x x x xi xj j v i v j v i v + c5 k + + +c4 k xj xi xi xj ij,2 = Mijk

(11.79)

+c3

We nd that the c1 term and the second and third part of the c3 term can be merged. Furthermore, the c2 term and the third and fourth part of the c3 term can be merged as well as the c4 and c5 terms; using Eq. 11.77 we get ij,2 = c2 + 8 8c2 2 6c2 + 4 k 4 60c2 Pij Dij + P + kij s 11 11 11 55 v v k v v k Dij = vi vk j k xj xi

(11.80)

Finally we re-write this equation so that it is expressed in trace-less tensors ij,2 = 2 c2 + 8 Pij ij P k 11 3 8c2 2 2 60c2 4 Dij ij P k kij s 11 3 55

(11.81)

where c2 = 0.4. Note that ii = 0 as we required in Eq. 11.73. This pressure-strain model is called the LRR model and it was proposed in [21]. All three terms in Eq. 11.81 satisfy continuity and symmetry conditions. It might be possible to use a simpler pressure-strain model using one or any two terms. Since the rst term is the most important one, a simpler model has been proposed [21, 22] 2 ij,2 = c2 Pij ij P k 3 (11.82)

It can be noted that there is a close similarity between the Rotta model and Eq. 11.82: both models represent return-to-isotropy, the rst expressed in vi vj and the second in Pij . The model in Eq. 11.82 is commonly called the IP model (IP=Isotropization by Production) . Since two terms are omitted we should expect that the best value of should be different than (c2 + 8)/11; a value of = 0.6 was (c2 = 1.4) found to give good agreement with experimental data. Since Eq. 11.82 is a truncated form of

11.7. Modelling assumptions


LRR model 1.5 0.4 LRR-IP model 1.5 0.6

93

c1 (Eq. 11.49 c2 (Eq. 11.81) c2 (Eq. 11.82)

Table 11.1: Constants in the LRR and LRR-IP pressure-strain models. x2

x1

Figure 11.4: Modelling of wall correction in pressure-strain terms. Eq. 11.81 it does not satisfy all requirements that Eq. 11.81 do. Equation 11.82 does satisfy symmetry condition and continuity but it does not satisfy the integral condition in Eq. 11.75. Although Eq. 11.82 is a simpler, truncated version of Eq. 11.81, it is often found to give more accurate results [23]. Since the IP model is both simpler and seems to be more accurate than Eq. 11.81, it is one of the most popular models of the rapid pressure-strain term. The coefcients for the slow and rapid terms in the LRR and LRR-IP models are summarized in Table 11.1 11.7.6 Wall model of the pressure-strain term When we derived the rapid pressure-strain model using Greens function in Eq. 11.60 we neglected the inuence of any boundaries. In wall-bounded domains it turns out that the effect of the walls must be taken into account. Both the rapid term in the LRR model and the IP model must be modied to include wall modelling. The effect of the wall is to dampen turbulence. There are two main effects whose underlying physics are entirely different. 1. Viscosity. Close to the wall the viscous processes (viscous diffusion and dissipation) dominate over the turbulent ones (production and turbulent diffusion). 2. Pressure. When a uid particle approaches a wall, the presence of the wall is felt by the uid particle over a long distance. This is true for a uid particle carried by the wind approaching a building as well as for a uid particle carried by a uctuating velocity approaching the wall in a turbulent boundary layer. In both cases it is the pressure that informs the uid particle of the presence of the wall. Since the pressure-strain term includes the uctuating pressure, it is obviously the second of these two processes that we want to include in the wall model. Up to now we have introduced two terms for modelling the pressure-strain term, the slow and the

11.7. Modelling assumptions

94

fast term. It is suitable to include a slow and a fast wall model term, i.e. ij = ij,1 + ij,2 + ij,1w + ij,2w (11.83)

where subscript w denotes wall modelling. Consider a wall, see Fig. 11.4. The pressure uctuations dampens the wall-normal uctuations. Furthermore, the damping effect of the wall should decrease for increasing wall distance. We need to scale the wall-normal distance with a relevant quantity and the turbulent length scale, k 3/2 /, seems to be a good candidate. For the wall-normal uctuations, the IP wall model reads [24] 2 22,1w = 2c1w v2 f k 3 k2 f= 2.55ni,w (|xi xi,w |)

(11.84)

where xi xi,w and ni,w denotes the distance vector to the wall and the unit wallnormal vector, respectively. As explained above, this damping is inviscid (due to pressure) and affects the turbulent uctuations well into the log-region. It has nothing to do with viscous damping. Away from the wall, in the fully turbulent region, the damping function goes to zero since the distance to the wall (|xi xi,w |) increases faster than the turbulence length scale, k 3/2 /. In the viscous region the wall model term, 22,1w , is not relevant since it should account only for inviscid damping. Function f should not exceed one. The IP wall model for the wall-parallel uctuations read 2 11,1w = 33,1w = c1w v2 f k (11.85)

The requirement that the sum of the pressure strain term should be zero. i.e. ii,1w = 0, is now satised since 11,1w + 22,1w + 33,1w = 0. The wall model for the shear stress is set as 3 12,1w = c1w v1 v2 f 2 k (11.86)

The factor 3/2 is needed to ensure that ii,1w = 0 is satised when the coordinate system is rotated. You can prove this by rotating the matrix [11,1w , 12,1w ; 21,1w , 22,1w ] and taking the trace of in the principal coordinates system (i.e. taking the sum of the eigenvalues). The general formula for a wall that is not aligned with a Cartesian coordinate axis reads [24] 3 3 vk vm nk,w nm,w ij vk vi nk,w nj,w vk vj ni,w nk,w f 2 2 (11.87) An analogous wall model is used for the rapid part which reads ij,1w = c1w 3 3 ij,2w = c2w km,2 nk,w nm,w ij ki,2 nk,w nj,w kj,2 ni,w nk,w f 2 2 (11.88) k

11.8. The k model

95

11.8 The k model

The exact k equation is given by Eq. 11.20. By inserting the model assumptions for the turbulent diffusion (Eq. 11.32), the production (Eq. 11.31) and the buoyancy term (Eqs. 11.27 and 11.28) we get the modelled equation for k k i v k j i v t v = t + + gi + vj t xj xj xi xj xi t k + + xj k xj In the same way, the modelled equation is obtained from Eq. 11.25 = c1 t + vj t xj k i v j i v v + xj xi xj 2 t c2 + + c1 gi k xi k xj

(11.89)

xj

(11.90)

The turbulent viscosity is computed as t = c The standard values for the coefcients read (c , c1 , c2 , k , ) = (0.09, 1.44, 1.92, 1, 1.3) (11.92) k2 (11.91)

11.9. The modelled vi vj equation with IP model

96

11.9 The modelled vi vj equation with IP model

With the models for diffusion, pressure-strain and dissipation we get


vi vj = xk v v j v v i vi vk j k xk xk 2 c1 vi vj ij k k 3 2 c2 Pij ij P k 3 3 +c1w [ vk vm nk nm ij vi vk nk nj k 2 3 vj vk nk ni ]f 2 3 +c2w [ km,2 nk nm ij ik,2 nk nj 2 3 jk,2 nk ni ]f 2 2 vi vj + xk xk vi vj k ck vk vm + xk xm

vk

(convection) (production) (slow part) (rapid part)

(wall, slow part)

(11.93)

(wall, rapid part) (viscous diusion) (turbulent diusion)

gi vj gj vi (buoyancy production) 2 ij (dissipation) 3

11.10 Algebraic Reynolds Stress Model (ASM)


The Algebraic Reynolds Stress Model is a simplied Reynolds Stress Model. The RSM and k models are written in symbolic form (see p. 78 & 80) as: RSM : Cij Dij = Pij + ij ij k : C k Dk = P k (11.94)

In ASM we assume that the transport (convective and diffusive) of vi vj is related to that of k, i.e. vi vj Cij Dij = C k Dk k Inserting Eq. 11.94 into the equation above gives vi vj Pk k

Pij + ij ij =

(11.95)

Thus the transport equation (PDE) for vi vj has been transformed into an algebraic equation based on the assumption in Eq. 11.94.

11.11. Explicit ASM (EASM or EARSM)

97

Now we want to re-write this equation as an equation for vi vj . Insert the IP models for ij,1 (Eq. 11.49) and ij,2 (Eq. 11.82) and the isotropic model for ij (Eq. 11.41) in Eq. 11.95 and multiply by k/ so that

k k 2 Pij c1 vi vj ij k c2 3

2 Pij ij P k 3

2 ij k 3

vi vj k + (ij,1w + ij,2w ) = Pk Collect all vi vj terms so that vi vj

k 2 Pij c2 Pij ij P k 3 = 2 k 2 Pij ij P k c2 Pij ij P k 3 3

+ ij,1w + ij,2w

Pk 1 + c1 = 2 + ij k(1 + c1 ) 3

+ ij,1w + ij,2w +

2 ij k(P k / 1 + c1 ) 3

where (2/3)ij P k k/ was added and subtracted at the last line (shown in boxes). Dividing both sides by P k / 1 + c1 gives nally
vi vj = 2 2 k (1 c2 ) Pij 3 ij P k + ij,1w + ij,2w ij k + 3 c1 + P k / 1

(11.96)

In boundary layer ow Eq. 11.96 reads


v1 v2 =

v c1 1 + c2 P k / k 2 2 (1 c2 ) 3 (c1 1 + P k /) y c

As can be seen, this model can be seen as an extension of an eddy-viscosity model where the c constant is made a function of the ratio P k /.

11.11 Explicit ASM (EASM or EARSM)


Equation 11.96 is an implicit equation for vi vj , i.e. the Reynolds stresses appear both on the left and the right side of the equation. It would of course be advantageous to be able to get an explicit expression for the Reynolds stresses. Pope [25] managed to derive an explicit expression for ASM in two dimensions. He assumed that the Reynolds stress tensor can expressed in the strain-rate tensor, sij , and the vorticity tensor, ij . Furthermore, he showed that the coefcients, G(n) , in that expression can be a function of not more than the following ve invariants

(k 2 /2 )ij sji , s

s (k 2 /2 )ij ji , (k 3 /3 )ij sjk ski 4 4 3 3 (k / )ij jk ski , (k / )ij jk skm smi

(11.97)

In two dimension the expression reads


vi vj =

2 k2 k3 s kij + G(1) sij + +G(2) 2 (ik kj ik skj ) 3

(11.98)

11.12. Boundary layer ow

98

x2

n
v1 (x2 )

Xx x1 s Xx

Figure 11.5: Boundary layer ow. In general three-dimensional ow, the Reynolds stress tensor depends on 10 tenn sors, Tij [25], i.e.
10 vi vj kij = 1 Tij n G(n) Tij n=1

= sij ,

2 Tij

= sik kj sjk ki ,

1 4 Tij = ik kj ij ik ki , 3 2 6 Tij = im mk skj + sik km mj ij pm mk skp 3 8 7 Tij = im smk kn nj im mk skn nj , Tij = sim mk skn snj sim smk kn snj 2 9 Tij = im mk skn snj sim smk kn nj ij pm mk skn snp 3 T 10 = im smk skn np pj im mk skn snp pj
ij

1 3 Tij = sik skj ij sik ski 3 5 Tij = ik skm smj sim smk kj

(11.99)

n where Tij may depend on the ve invariants in Eq. 11.97. Equation 11.99 is a general form of a non-linear eddy-viscosity model. Any ASM may be written on the form of Eq. 11.99. It may be noted that Eq. 11.99 includes only linear and quadratic terms of sij and ij . That is because of Cayley-Hamilton theorem which states that a second-order tensor satises its own characteristic equation (see Section 1.20 in [19]); hence cubic terms or higher can recursively be expressed in linear ( sij ) and quadratic tensors (ik skj ). Furthermore, note that all terms in Eq. 11.99 are symmetric and traceless as s required by the left side, vi vj 2ij k/3.

11.12 Boundary layer ow


Let us study boundary layer ow (Fig. 11.5) where v2 = 0, v1 = v1 (x2 ). In general the production Pij has the form:
Pij = vi vk

j v v i vj vk xk xk

12. Reynolds stress models vs. eddy-viscosity models

99

In this special case we get:


P11 = 2v1 v2

1 v x2

2 P12 = v2

1 v x2

P22 = 0
2 Is v2 zero because its production term P22 is zero? No! The sympathetic term ij 2 2 which takes from the rich (i.e. v1 ) and gives to the poor (i.e. v2 ) saves the unfair situation! The IP model for ij,1 and ij,2 gives

22,1 = c1 22,2

2 2 k v2 > 0 k 3 2 1 1 v >0 = c2 P11 = c2 v1 v2 3 3 x2

2 Note also that the dissipation term for the v1 v2 is zero, but it takes the value 3 for the 2 2 v1 and v2 equations (see p. 84). Since the modelled v1 v2 does not have any dissipation term, the question arises: what is the main sink term in the v1 v2 equation? The answer is, again, the pressure strain term ij,1 and ij,2 .

12 Reynolds stress models vs. eddy-viscosity models


In this section we present three fundamental physical processes which Reynolds stress models are able to handle whereas eddy-viscosity models fail. The reason for the superiority of the former model is in all cases that the production term is treated exactly, whereas it in eddy-viscosity models is modelled.

12.1 Stable and unstable stratication


In ows where buoyancy is dominating, the temperature has a large effect on the turbulence through the buoyancy term Gij , see Eq. 11.10. If the temperature increases upwards (i.e. /x3 > 0), then the ow is stably stratied. This is illustrated in Fig. 12.1. Consider /x3 > 0. This means that the density decreases with increasing vertical height, i.e. /x3 < 0. If a uid particle is displaced from its equilibrium level 0 up to level 2, see Fig. 12.1, it is heavier then the surrounding at this new level (0 > 2 ). Hence, the buoyancy forces the particle back to its original position 0. In this way the vertical turbulent uctuations are dampened. Similarly if a particle originating at level 0, is moved down to level 1. Here it is lighter than its new environment, and hence buoyancy makes it to move make to its original level 0. For the case of unstable stratication, the situation is reversed. Cold uid is located on top of hot uid, i.e. /x3 < 0 and /x3 > 0. In Fig. 12.1 we would then have 2 > 0 . If a uid particle at level 0 is displaced upwards to level 2, it is at this location lighter than its new environment; hence it continues to move upwards. If it is moved down to level 1 it is heavier than its new environments and it will then

12.1. Stable and unstable stratication


2 > 0 2 < 0 F

100

/x3 > 0 /x3 < 0 x3

F 1 < 0 1 > 0 1

Figure 12.1: Stable stratication due to positive temperature gradient /x3 > 0. continue downwards. Hence, turbulent uctuations are enhanced. This ow situation is called unstable stratication. The production term due to buoyancy reads (see Eq. 11.10)
G33 = 2gv3

(12.1)

since gi = (0, 0, g). From the equation for the turbulent heat ux, v3 (i.e. Eq. 11.19 with i = 3), we nd the production term for v3 P3 = v3 vk

3 v vk xk xk

(12.2)

In the case illustrated in Fig. 12.1, the production term due to temperature gradient 2 reads P3 = v3 /x3 < 0 (recall that we assume that buoyancy dominates so that the rst term in Eq. 12.2 is much larger than the second one). Since the main source term in the v3 equation, P3 , is negative, it makes v3 < 0 so that G33 < 0 (see Eq. 12.1). Thus, for the case illustrated in Fig. 12.1, we nd that the production term, G33 , due to buoyancy yields a damping of the vertical uctuations as it should. Note that the horizontal turbulent uctuations are not affected by the buoyancy term, Gij , since G11 = G22 = 0 because the gravity is in the x3 direction (i.e. g1 = g2 = 0). If the situation in Fig. 12.1 is reversed so that /x3 < 0 the vertical uctuations are instead augmented. This is called unstably stratied conditions. When eddy-viscosity models are used, transport equations are usually not solved for vi . Instead the heat ux tensor is modelled with an eddy-viscosity assumption using the Boussinesq assumption, see Eq. 11.27. The buoyancy term, Gk , in the k equation reads, see Eq. 11.10 (take the trace of Gij and divide by two)
Gk = 0.5Gii = gi vi For gi = (0, 0, g), it reads Gk = gv3 which with Eq. 11.27 gives

(12.3)

Gk = g

t x3

(12.4)

12.2. Curvature effects

101

A 0 B r B 0

V (r)

Figure 12.2: Flow in a polar coordinate system illustrating streamline curvature. The streamline is aligned with the axis.

r v1 x2 x1 streamline

Figure 12.3: Streamline curvature occurring when the ow approaches, for example, a separation region or an obstacle. Hence it is seen that in stably stratied conditions, Gk < 0 as required. The difference between an eddy-viscosity model and a Reynolds stress model, is that the former reduces k whereas the latter reduces only the vertical uctuations.

12.2 Curvature effects


When the streamlines in boundary layer ow have a convex curvature, the turbulence is stabilized. This dampens the turbulence [26, 27], especially the shear stress and the Reynolds stress normal to the wall. Concave curvature destabilizes the turbulence. The ratio of boundary layer thickness to curvature radius R is a common parameter for quantifying the curvature effects on the turbulence. The work reviewed by Bradshaw [26] demonstrates that even such small amounts of convex curvature as /R = 0.01 can have a signicant effect on the turbulence. In [28] they carried out an experimental investigation on a conguration simulating the ow near a trailing edge of an airfoil, where they measured /R 0.03. They reported a 50 percent decrease 2 of v2 (Reynolds stress in the normal direction to the wall) owing to curvature. The 2 reduction of v1 and v1 v2 was also substantial. In addition they reported signicant damping of the turbulence in the shear layer in the outer part of the separation region. An illustrative model case is curved boundary layer ow, see Fig. 12.2. A polar

12.2. Curvature effects


V /r > 0 convex curvature concave curvature stabilizing destabilizing V /r < 0 destabilizing stabilizing

102

Table 12.1: Effect of streamline curvature on turbulence. coordinate system r with locally aligned with the streamline is introduced. As v = v (r) (with v /r > 0 and vr = 0), the radial inviscid momentum equation degenerates to
2 v p =0 r r

(12.5)

Here the variables are instantaneous or laminar. The centrifugal force exerts a force in the normal direction (outward) on a uid following the streamline, which is balanced by the pressure gradient. Since we have assumed that v /r > 0, Eq. 12.5 shows that the pressure gradient increases with r. If the uid is displaced by some disturbance (e.g. turbulent uctuation) outwards to level A, it encounters a pressure gradient larger than that to which it was accustomed at r = r0 , as (v )A > (v )0 , which from Eq. 12.5 gives (p/r)A > (p/r)0 . Hence the uid is forced back to r = r0 . Similarly, if the uid is displaced inwards to level B, the pressure gradient is smaller here than at r = r0 and cannot keep the uid at level B. Instead the centrifugal force drives it back to its original level. It is clear from the model problem above that convex curvature, when v /r > 0, has a stabilizing effect on (turbulent) uctuations, at least in the radial direction. It is discussed below how the Reynolds stress model responds to streamline curvature. Assume that there is a at-plate boundary layer ow, see Fig. 12.3. The ratio of 2 2 the normal stresses v1 to v2 is typically 5. At one x1 station, the ow is deected upwards. How will this affect turbulence? Let us study the effect of concave streamline curvature. The production terms Pij owing to rotational strains (1 /x2 , 2 /x1 ) v v can be written as (see Eq. 11.10):
2 RSM, v1 eq. : P11 = 2v1 v2 2 RSM, v1 v2 eq. : P12 = v1

1 v x2

(12.6a) (12.6b) (12.6c)


2

2 v v 2 1 v2 x1 x2 2 v x1

2 RSM, v2 eq. : P22 = 2v1 v2

P k = t

2 v 1 v + x2 x1

(12.6d)

The terms in boxes appear because of the streamline curvature. As long as the streamlines are parallel to the wall, all production is a result of 1 /x2 . However as soon as the streamlines are deected, there are more terms v resulting from 2 /x1 . Even if 2 /x1 is much smaller than 1 /x2 it will still v v v 2 2 contribute non-negligibly to P12 as v1 is much larger than v2 . Thus the magnitude of P12 will increase (P12 is negative) as 2 /x1 > 0. An increase in the magnitude of v

12.3. Stagnation ow
v1 (x2 )

103

x2 x1

Figure 12.4: The velocity prole for a wall jet.


2 P12 will increase v1 v2 , which in turn will increase P11 and P22 . This means that v1 2 and v2 will be larger and the magnitude of P12 will be further increased, and so on. It is seen that there is a positive feedback, which continuously increases the Reynolds stresses. The turbulence is destabilized owing to concave curvature of the streamlines. Note than eddy-viscosity models such as k and k models cannot account for streamline curvature since the two rotational strains, 1 /x2 and 2 /x1 , in the v v production term are multiplied by the same coefcient (the turbulent viscosity). If the ow (concave curvature) is a wall jet ow where 1 /x2 < 0 in the outer v part (see Fig. 12.4) the situation will be reversed: the turbulence will be stabilized. If the streamline (and the wall) is deected downwards, the situation will be as follows: the turbulence is stabilizing when 1 /x2 > 0, and destabilizing for 1 /x2 < 0. v v The stabilizing or destabilizing effect of streamline curvature is thus dependent on the type of curvature (convex or concave), and whether there is an increase or decrease in momentum in the tangential direction with radial distance from its origin (i.e. the sign of V /r). For convenience, these cases are summarised in Table 12.1. It should be noted that concave or convex depends on from which the streamline is viewed. The streamline in Fig. 12.3, for example, is concave when viewed from the wall but convex when viewed from the orig of the circle with radius r. It should be mentioned that one part of the effect of curved streamlines in Eq. 12.6 is due to the transformation of the advective term of the vi vj -equation (cf. polar coordinates where additional terms appear both in the momentum equations and the transport equation for vi vj ). In [29] they proposed a correction term to take this effect into account.

12.3 Stagnation ow
The k model does not model the normal stresses properly, whereas ASM/RSM do. The production term in the k equations for RSM/ASM and k model in stagnation v v ow (see Fig. 12.5) due to 1 /x1 and 2 /x2 is:
2 RSM : 0.5 (P11 + P22 ) = v1

1 v v v 2 2 = 1 (v 2 v 2 ) v2 2 x1 x2 x1 1 1 v x1
2

(12.7) (12.8)

k : P k = 2t

2 v x2

where continuity 1 /x1 = 2 /x2 has been employed. In RSM, the two terms v v are added with sign. In the k model, however, the production will be large because

12.4. RSM/ASM versus k models

104

x2
x2

x1

x1

x2 x1

Figure 12.5: The ow pattern for stagnation ow. the difference in sign of the two terms is not taken into account.

12.4 RSM/ASM versus k models

Advantages with k models (or eddy viscosity models): i) simple due to the use of an isotropic eddy (turbulent) viscosity ii) stable via stability-promoting second-order gradients in the mean-ow equations iii) work reasonably well for a large number of engineering ows

Disadvantages:
2 2 2 i) isotropic, and thus not good in predicting normal stresses (v1 , v2 , v3 )

ii) as a consequence of i) it is unable to account for curvature effects iii) as a consequence of i) it is unable to account for irrotational strains (stagnation ow) iv) in boundary layers approaching separation, the production due to normal stresses is of the same magnitude as that due to shear stresses [30]. Advantages with ASM/RSM: i) the production terms do not need to be modelled ii) thanks to i) it can selectively augment or damp the stresses due to curvature effects (RSM is better than ASM because the convective terms are accounted for), boundary layers approaching separation, buoyancy etc. Disadvantages with ASM/RSM: i) RSM is complex and difcult to implement, especially implicit ASM ii) numerically unstable because small stabilizing second-order derivatives in the momentum equations (only laminar diffusion) iii) CPU time consuming

13. Realizability

105

13 Realizability
There are a number of realizability constraints. The usual two ones are that all normal stresses should stay positive and that the correlation coefcient for the shear stress should not exceed one, i.e.
2 vi 0 for all i 1/2

vi vj 2 vi 2 vj

1 no summation over i and j, i = j

(13.1)

These criteria are seldom used in RSMs. However, satisfying the rst criteria is actually of importance for eddy-viscosity models in stagnation ow [31]. Assume that the ow is in the x1 direction and that it approaches the wall (see Fig. 12.5). The Boussinesq 2 assumption for the normal stress v1 reads (cf. Eq. 12.7)
2 v1 =

2 1 v 2 = k 2t s11 k 2t 3 x1 3

(13.2)

2 It is seen that if s11 gets too large then v1 < 0 which is unphysical, i.e. non-realizable. Lets now briey repeat the concept invariants. This means something that is independent of the coordinate system. Here we mean independent of rotation of the coordinate system. If a tensor is symmetric, then we know that it has real eigenvalues which means that we can rotate the coordinate system so that the off-diagonal components vanish (see, e.g., [19]). For the strain tensor this means that the off-diagonal components of sij vanish and this is the coordinate system where the diagonal com ponents become largest (e.g. s11 in Eq. 13.2). Thus this is the coordinate system in 2 which the danger of negative v1 from Eq. 13.2 is largest. The equation for nding the eigenvalues of a tensor Cij is (see e.g. [19] or [32])

|Cij ij | = 0 which gives in 2D C11 C21 The resulting equation is


2D 2D 2 I1 + I2 = 0

(13.3)

C12 C22

=0

(13.4)

2D I1 = Cii 1 2D I2 = (Cii Cjj Cij Cij ) = det(Cij ) 2

(13.5)

Since the above equation is the same irrespectively of how the coordinate system is 2D 2D rotated, it follows that its coefcients I1 and I2 are invariants. In 3D Eq. 13.3 gives C11 C12 C21 C22 C31 C32 C13 C23 C33 =0 (13.6)

13.1. Two-component limit

106

which gives
3D 3D 3D 3 I1 2 + I2 I3 = 0

3D I1 = Cii 1 3D I2 = (Cii Cjj Cij Cij ) 2 1 3D I3 = (2Cij Cjk Cki 3Cij Cji Ckk + Cii Cjj Ckk ) = det(Cij ) 6

(13.7)

3D 3D 3D The invariants are I1 , I2 and I2 . Lets go back to Eq. 13.2 and assume incompressible 2D ow. The rst invariant reads (cf. Eq. 13.5) 2D I1 = sii = s11 + s22 = 1 + 2 = 0

(13.8)

It is zero due to the continuity equation. The second invariant of sij reads
2D I2 = ij sij /2, s

(13.9)

(see Eq. 13.5) which is the same in all coordinate systems (hence the name invariant). The solution to Eq. 13.5, using Eq. 13.8, is
2D 1,2 = I2 1/2

sij sij 2

1/2

(13.10)

The eigenvalues of sij correspond to the strains in the principal axis. As discussed above, we apply Eq. 13.2 in principal coordinate directions of sij . Hence, s11 in Eq. 13.2 is replaced by the largest eigenvalue so that
2 v1 =

2 k 2t 1 3

(13.11)

2 The requirement v1 0 gives now together with Eq. 13.11

k k = 3|1 | 3

2 sij sij

1/2

(13.12)

In 3D, Eq. 13.7 instead of Eq. 13.5 is used, and Eq. 13.10 is replaced by [31] |k | = k 2ij sij s 3
1/2

(13.13)

13.1 Two-component limit


2 Another realizability constraint is to require that when vi approaches zero near walls, 2 it should do so smoothly. One way to ensure this is to require that the derivative of vi 2 should go to zero as vi goes to zero, i.e. 2 vi 0 2 Dvi 0 Dt

(13.14)

where D/Dt denotes the material derivative (think of Eq. 13.14 in Lagrangian coordinates, i.e. we follow a uid particle as it approaches the wall). Eq. 13.14 requires that

14. Non-linear Eddy-viscosity Models

107

2 when vi approaches zero, the left side (and thus also the right side) of the transport 2 equation of vi should also do so. Since we are here concerned about the pressure2 strain term, well take a look at how it behaves near walls when vi 0. This is of some relevance in near-wall turbulence where the wall-normal stress goes to zero faster than the wall-parallel ones: this state of turbulence is called the two-component limit [33]. Neither the form of ij,2 in Eq. 11.82 nor Eq. 11.81 satisfy the requirement 2 that 22,2 = 0 when v2 = 0 [20]. In Eq. 11.82, for example,

2 22,2 ij P k 3

(13.15)

Very complex forms of ij,2 have been proposed [34] [CL96] which include terms cu bic in vi vj . The CL96 model does satisfy the two-component limit. Another advantage of the CL96 model is that it does not need any wall distances, which is valuable in complex geometries. The models of the slow pressure-strain in Eq. 11.49 (linear model) and Eq. 11.55 (non-linear model) do also not satisfy the two-component limit. The Rotta model, for example, gives 2 22,1 c1 (13.16) 3 The only way to ensure this is to make c1 0 when the wall is approached. A convenient parameter proposed by [33] is A which is an expression of A2 and A3 (the second and third invariant of aij , respectively), i.e. 9 A2 = aij aji , A3 = aij ajk aki , A = 1 (A2 A3 ) 8 (13.17)

The parameter A = 0 in the two-component limit and A = 1 in isotropic turbulence. Thus A is a suitable parameter to use when damping the constant c1 as the wall is approached.

14 Non-linear Eddy-viscosity Models


In traditional eddy-viscosity models the turbulent stress vi vj is formulated from the Boussinesq assumption, i.e.

aij = 2t sij = 1 2

sij k i v j v + xj xi

(14.1)

where the anisotropy tensor is dened as aij


vi vj 2 ij k 3

(14.2)

The relation between the stress vi vj and the velocity gradient in Eq. 14.1 is, as can be seen, linear. One way to make eddy-viscosity models more general, is to include

14. Non-linear Eddy-viscosity Models

108

non-linear terms [25]. A subset of the most general form reads [35] aij = 2c sij 1 + c1 2 sik skj sk sk ij + c2 2 ik skj sik kj 3 1 + c3 2 ik jk k k ij + c4 3 sik sk j i sk skj 3 2 + c5 3 i m smj + si m mj mn n sm ij 3 + c6 3 sk sk sij + c7 3 k k sij 1 ij = 2 i v j v xj xi

(14.3)

where is a turbulent time scale; for a non-linear k model = k/, and for a nonlinear k model = 1/. The tensor groups correspond to a subset of Eq. 11.99:
1 Line 1: Tij , 3 2 Line 2: Tij and Tij 4 5 Line 3: Tij and Tij 6 Line 4: Tij 1 Line 5: Tij multiplied by the invariants sk sk and k k

v The expression in Eq. 14.3 is cubic in i /xj . However, note that it is only quadratic in sij and ij . This is due to Cayley-Hamilton theorem which states that a tensor is only linearly independent up to quadratic terms, see p. 87; this means that, for example, s3 = sik sk sj can be expressed as a linear combination of s2 = sik skj and ij ij sij . aij is symmetric and its trace is zero; it is easily veried that the right side of Eq. 14.3 also has these properties. Examples of non-linear models (sometimes also called explicit algebraic Reynolds stress models, EARSM) in the literature are the models presented in [36, 37, 35, 38]. EARSMs are very popular especially the model in [38] in the aeronautical community where explicit time-marching solver are used. They are computationally cheap, more accurate than linear eddy-viscosity models and they dont give rise to any numerical instabilities as in implicit solvers (like SIMPLE). In implicit solvers a large turbulent viscosity in the diffusion term of the momentum equations is needed to stabilize the solution procedure.

14. Non-linear Eddy-viscosity Models

109

Lets take a closer look on Eq. 14.3 in fully developed channel ow (2 = v3 = v /x1 = /x3 0); we obtain a11 = a22 = a33 1 2 12 1 2 12 1 v x2 1 v x2 1 v x2
2

(c1 + 6c2 + c3 )
2

(c1 6c2 + c3 )
2

1 = 2 6

(14.4)

(c1 + c3 ) 1 v x2
3

a12 = c

1 1 v + 3 x2 4

(c5 + c6 + c7 )

Using values on the constants as in [35], i.e c1 = 0.05, c2 = 0.11, c3 = 0.21, c4 = 0.8 c5 = 0, c6 = 0.5 and c7 = 0.5 we get a11 = a22 = a33 a12 0.82 2 12 0.5 2 12 1 v x2 u1 x2
2 2 v1 = 2 2 v2 = 2

2 0.82 2 k+ k 3 12 2 0.5 2 k k 3 12

1 v x2 1 v x2

v 0.16 2 1 = 12 x2 k 1 v = c x2

2 0.16 2 2 v3 = k k 3 12

1 v x2

(14.5)

We nd that indeed the non-linear model gives anisotropic normal Reynolds stresses. v In Eqs. 14.4 and 14.5 we have assumed that the only strain is 1 /x2 . When we discussed streamline curvature effects at p. 103 we found that it is important to investigate the effect of secondary strains such as 2 /x1 . Lets write down Eq. 14.3 v for the strain 2 /x1 v a11 = a22 = a33 1 2 12 1 2 12 2 v x1 2 v x1 2 v x1 2 v x1
2

(c1 6c2 + c3 )
2

(c1 + 6c2 + c3 )
2

1 = 2 6

(14.6) (c1 + c3 )

1 a12 = 3 4

(c5 + c6 + c7 )

Inserting with values on the constants from [35] (see above) we obtain a11 = a22 = 0.5 2 12 2 v x1 2 v x1 2 v x1
2

0.82 2 12 0.16 2 12

(14.7)
2

a33 =

, a12 = 0

15. The V2F Model

110

As can be seen the coefcient for a22 is larger than that in Eq. 14.5, and hence the model is slightly more sensitive to the secondary strain 2 /x1 than to the primary v one 1 /x2 . Thus, the non-linear models are able to account for streamline curvature, v but due to the choice of constants so that c5 + c6 + c7 = 0 this effect is weak.

15 The V2F Model


In the V2F model of [39, 40, 31] two additional equations, apart from the k and 2 equations, are solved: the wall-normal stress v2 and a function f . This is a model which is aimed at improving modelling of wall effects on the turbulence. 2 Walls affect the uctuations in the wall-normal direction, v2 , in two ways. The wall 2 is felt by the turbulence fairly far from the wall (x+ 200) through the damping of v2 2 pressure eld whereas the viscous damping takes place within the viscous and buffer 10). In usual eddy-viscosity models both these effects are accounted for layer (x+ 2 2 through damping functions. The damping of v2 is in the RSM accounted for through the modelled pressure-strain terms 22,1w and 22,2w (see Eqs. 11.87 and Eq. 11.88). They go to zero far away from the wall. 2 In the V2F model the problem of accounting for the wall damping of v2 is simply 2 equation in boundary-layer form resolved by solving its transport equation. The v2 reads (see Eq. 9.16 at p. 68) v 2 2 v2 v 2 1 v2 v 2 ( + t ) 2 2v2 p /x2 22 + = x1 x2 x2 x2 (15.1)

in which the diffusion term has been modelled with an eddy-viscosity assumption, see Eq. 11.39 at p. 84. Note that the production term P22 = 0 because in boundary-layer approximation v2 v1 and /x1 /x2 . The model for the dissipation 22 is taken as v 2 model = 2 22 k Add and subtract model on the right side of Eq. 15.1 yields 22 2 v2 v 2 1 v2 v 2 + = x1 x2 v 2 v 2 v 2 ( + t ) 2 2v2 p /x2 22 + 2 2 x2 x2 k k In the V2F model P is now dened as 2 v 2 P = v2 p /x2 22 + 2 k so that Eq. 15.2 can be written as 1 v2 v 2 v 2 2 v2 v 2 v 2 ( + t ) 2 + P 2 + = x1 x2 x2 x2 k (15.4) (15.3)

(15.2)

15. The V2F Model

111

2 P is the source term in the v2 -equation above, and it includes the velocity-pressure gradient term and the difference between the exact and the modelled dissipation. Note that this term is commonly split into the pressure-strain term and a diffusion term as v2 p /x2 = v2 p p v2 /x2 x2

Physically, the main agent for generating wall-normal stress is indeed the pressurestrain term via re-distribution, see example in Section 11.12. A new variable f = P/k is dened and a relaxation equation is formulated for f as L2 2f 22 1 f = x2 k T 2 T = max C1 22 = k T k , CT
2 2 v2 3 k 2 2 v2 k 3

1/2

t + C2 k 3

1 v x2
1/4

(15.5)

L = CL max

k 3/2 , C

where 22 is the IP model of the pressure-strain term, see Eqs. 11.49 and 11.82, the rst term being the slow term, and the second the rapid term. The constants are given the following values: c = 0.23, CT = 6, ce1 = 1.44, c2 = 1.9, k = 0.9, = 1.3, C1 = 1.3, C2 = 0.3, CL = 0.2, C = 90. 2 2 The boundary condition for f is obtained from v2 equation. Near the wall, the v2 equation reads 2 2 v2 2 v2 0= (15.6) 2 + fk k x2 Near the wall, Taylor analysis gives = 2k/x2 [5]; using this expression to replace 2 k in Eq. 15.6 gives 2 2 2 v2 f x2 2v2 2 0= (15.7) 2 + 2 2 x2 x2 2 Assuming that f and are constant very close to the wall, this equation turns into an ordinary second-order differential equation with the solution
2 v2 = Ax2 + 2

x4 B f 2 2 x2 20

2 Since v2 = O(x4 ) as x2 0, both constants must be zero, i.e. A = B = 0, so we get 2

f =

2 20 2 v2 x4 2

(15.8)

For more details, see [41]. 2 Above we have derived the v2 equation in boundary layer form assuming that x2 is the wall-normal coordinate. In general, three-dimensional ow it reads v 2 v2 j v 2 v ( + t ) + f k = xj xj xj k (15.9)

15. The V2F Model

112

1.5

0.5

L=0.04,S=1 L=0.08,S=1 L=0.12,S=1 L=0.2,S=1 L=0.2, S=2 L=0.2, S=0.5


0.2 0.4 0.6 0.8 1

0 0

x2

Figure 15.1: Illustration of Eq. 15.12 In the V2F model a transport equation for the normal stress normal to walls is solved. 2 If the wall lies in the x1 x3 plane, then v 2 = v2 . However, if a wall lies in the 2 x2 x3 plane, for example, this means that the transport equation for v2 is turned 2 2 into an equation for v1 , i.e. v 2 = v1 . This is done automatically since in the general formulation in Eq. 15.9, 1 /x2 in the expression for 22 is replaced by P k . If the v wall lies in the x2 x3 plane the largest velocity gradient will be 2 /x1 or 3 /x1 . v v Why does the right side of Eq. 15.5 has the form it has? Far from the wall, 2 the source term in the v2 -equation simplies to 22 plus isotropic dissipation (see Eq. 15.1). This is what happens, because far from the wall when 2 f /x2 0, and 2 Eq. 15.5 yields (T = k/)
2 kf P 22 + (v2 /k 2/3)

(15.10)

When this expression is inserted in Eq. 15.4 we get 1 v2 v 2 v 2 2 2 v2 v 2 ( + t ) 2 + 22 + = x1 x2 x2 x2 3 (15.11)

2 which is the usual form of the modelled v2 -equation with isotropic dissipation. Thus the f equation acts so as to let f go from the value of its source term to its (negative) wall value (see Eq. 15.8) over lengthscale L. This is how the reduction of the source term P in Eq. 15.4 is achieved as the wall is approached. The behavior of the equation for f (Eq. 15.5) for different right sides is illustrated in the Fig. 15.1 where the equation

L2

2f f +S =0 x2 2

(15.12)

has been solved with f = 0 at the wall and with different L and S. As can be seen, f is, as required, reduced as the wall is approached. Furthermore, f approaches the value of the source term as x2 > L. The inuence of the lengthscale L is nicely illustrated: the larger L, the further away from the wall does f go to its far-eld value.

15.1. Modied V2F model

113

In the V2F model the turbulent viscosity is computed from


2 t = C v2 T

(15.13)

The k and -equations are also solved (without damping functions). For convenience, the boundary conditions are given again
2 k = 0, v2 = 0

= 2k/x2 2
2 20 2 v2 f = 4 x2

(15.14)

The boundary condition for f makes the equation system numerically unstable. 2 One way to get around that problem is to solve both the k, and v2 , f equations coupled [41]. An alternative is to use the f model [42] which is more stable. In this model 2 2 they solve for the ratio v2 /k instead of for v2 which gives a simpler wall boundary condition for f , namely f = 0.

15.1 Modied V2F model


In [43] they proposed a modication of the V2F model allowing the simple explicit boundary condition f = 0 at walls. They introduced a new variable f = f 5v 2 /k 2 and they neglected the term 5L2 2 xj xj v 2 k2

2 The resulting v2 and f -equation read [43]

j v 2 v v 2 v2 ( + t ) + kf 6 = xj xj xj k L2 v2 1 Pk 2f 2 (C1 6) (C1 1) + C2 + f = xj xj T k 3 k j i v i v v + P k = t xj xi xj k 1/2 T = max ,6 L = CL max k 3/2 , C 3


1/4

(15.15)

(15.16)

Boundary conditions at the walls are k = 0, v 2 = 0 = 2k/x2 2 f = 0 This modied model is numerically much more stable. Note that the modied model is identical to the original model far from the wall.

15.2. Realizable V2F model

114

15.2 Realizable V2F model


The realizable condition for stagnation ow (see p. 105) is used also for the V2F model, and they read [43] T = min k 0.6k , 2 ( s )1/2 6C v sij ij

k 3/2 k 3/2 L = min , 1/2 6C v 2 (2ij sij ) s

(15.17)

These realizable conditions have been further investigated by Sveningsson [41, 44, 45, 46, 47], and it was found that the limitation on T is indeed important, whereas that for L is not. Furthermore, it was found that it is important to impose the limitation on T in a consistent manner. For instance, if the limit is used in the f equation, it must for consistency also be used for /k in Eq. 15.15.

In the V2F model, v 2 denotes the generic wall-normal stress. Thus it should be the smallest one. This is not ensured in the V2F models presented above. Below the simple modication proposed by [1] is presented. The source term kf in the v 2 -equation (Eq. 15.15) includes the modelled velocitypressure gradient term which is dampened near walls as f goes to zero. Since v 2 represents the wall-normal normal stress, it should be the smallest normal stress, i.e. 2 2 2 2 2 2 v2 v1 and v2 v3 , and thus v2 should be smaller than or equal to 3 k. In the homogeneous region far away from the wall, the Laplace term is assumed to be negligible i.e. 2 f /xj xj 0. Then Eq. 15.16 reduces to f = right side. It turns out that in the region far away from the wall, the Laplace term is not negli2 gible, and as a consequence v 2 gets too large so that v 2 > 3 k. A simple modication is to use the right side of Eq. 15.16 as an upper bound on the source term kf in the v 2 -equation, i.e.
2 vsource = min kf,

15.3 To ensure that v 2 2k/3 [1]

2k (C1 6)v 2 (C1 1) + C2 P k k 3

(15.18)

This modication ensures that v 2 2k/3. For more details, see [1].

16 The SST Model


The SST (Shear Stress Transport) model of [48] is an eddy-viscosity model which includes two main novelties: 1. It is combination of a k model (in the inner boundary layer) and k model (in the outer region of the boundary layer as well as outside of it); 2. A limitation of the shear stress in adverse pressure gradient regions. The k model has two main weaknesses: it over-predicts the shear stress in adverse pressure gradient ows because of too large length scale (due to too low dissipation) and it requires near-wall modication (i.e. low-Re number damping functions/terms)

16. The SST Model

115

Figure 16.1: Flow around an airfoil. Pressure contours. Red: high pressure; blue: low pressure One example of adverse pressure gradient is the ow along the surface of an airfoil, see Fig. 16.1. Consider the upper surface (suction side). Starting from the leading edge, the pressure decreases because the velocity increases. At the crest (at x/c 0.15) the pressure reaches its minimum and increases further downstream as the velocity decreases. This region is called the adverse pressure gradient (APG) region. The k model is better than the k model at predicting adverse pressure gradient ow and the standard model of [49] does not use any damping functions. However, the disadvantage of the standard k model is that it is dependent on the free-stream value of [50]. In order to improve both the k and the k model, it was suggested in [48] to combine the two models. Before doing this, it is convenient to transform the k model into a k model using the relation = /( k), where = c . The lefthand side of the equation will consist of the convection term, d/dt, which denotes the material derivative assuming steady ow, see Eq. 2.23. Let us express the lefthand side of the equation as a combination of the left-hand sides of the and the k equations by using the chain rule, i.e. d/( k) 1 d d(1/k) d = = + dt dt k dt dt dk 1 d dk 1 d 2 = = k dt k dt k dt k dt

(16.1)

Now we have transformed the left side of the equation. The right side should be transformed in the same manner. For example, the production of the equation will consist of two terms, one term from the equation 1 k P (the equation is the rst term at the RHS in Eq. 16.1) (16.2)

and one from the k equation Pk k (the k equation is the second term at the RHS in Eq. 16.1) (16.3)

16. The SST Model

116

(cf. with Eq. 16.1). In the same way we transform the entire right side inserting the modelled equations for k and so that D 1 1 = P P k k + Dt k k k k
Production, P Destruction,

1 k

T D

T D k k

2 2 k k x2 k x2 j j
Viscous diffusion, D

(16.4)

T Turbulent diffusion, D

Production term P = 1 P P k = C1 2 P k P k k k k k k = (C1 1) P k (16.5)

Destruction term = 2 k = C2 k k k k = (C2 1) 2 1 (16.6)

Viscous diffusion term


D =

2 k 2 k 2 k 2 2 = k x2 k x2 k x2 k xj j j j k xj k +k xj xj 2k k x2 j

= k =

k 2k k 2 2k + 2 + +k 2 xj xj xj xj xj xj k x2 j xj

(16.7)

2 k + k xj xj xj

The turbulent diffusion term is obtained as (the derivation is found in [51] which can be downloaded from www.tfd.chalmers.se/lada)
T D =

2t k + k xj xj k 1 1 k

t t k

2k + x2 j t xj

+ k

t k + xj xj xj

(16.8)

In the standard k model we have k = 1 and = 1.3. If we assume that k = in the second and third term of the right-hand side, we can considerably simplify the turbulence diffusion so that
T D =

2t k + k xj xj xj

t xj

(16.9)

16. The SST Model


We can now nally write the equation formulated as an equation for + (j ) = v xj xj t 2 + + k t + P k 2 xj k k xi xi

117

(16.10)

= C1 1 = 0.44, = (C2 1) = 0.0828 Since the k model will be used for the outer part of the boundary layer, the viscous part of the cross-diffusion term (second line) is usually neglected when the equation is used as an equation. In the SST model the coefcients are smoothly switched from k values in the inner region of the boundary layer to k values in the outer region. Functions of the form k 500 42 k 4 , (16.11) F1 = tanh( ), = min max , y y 2 CD y 2 are used. F1 = 1 in the near-wall region and F1 = 0 in the outer region. The coefcient, for example, is computed as SST = F1 k + (1 F1 )k (16.12)

where k = 0.075 and k = 0.0828. Since the standard k model does not include any cross-diffusion term, the last term in the equation (second line in Eq. 16.10) should only be active in the k region; hence it is multiplied by (1 F1 ). At p. 114 it was mentioned that the k model is better than the k model in predicting adverse pressure-gradient ows because it predicts a smaller shear stress. Still, the predicted shear stress is too large. This brings us to the second modication (see p. 114). When introducing this second modication, the author in [48] noted that a model (the Johnson - King model [JK]) which is based on transport of the main shear stress v1 v2 , predicts adverse pressure gradient ows much better than the k model. In the JK model, the v1 v2 transport equation is built on Bradshaws assumption [52]
v1 v2 = a1 k

(16.13)

where a1 = c written as
v1 v2

1/2

= 1/2 . In boundary layer ow, the Boussinesq assumption can be


2 1/2 1/2

c k 2 v v c k 2 1 k 1 = = c1/2 k = x2 x2 2

1 v x2

= c1/2 k

Pk

(16.14) It is found from experiments that in boundary layers of adverse pressure gradient ows 1/2 the production is much larger than the dissipation (P k ) and v1 v2 c k, which explains why Eq. 16.14 over-predicts the shear stress and works poorly in this type of ow. To reduce |v1 v2 | in Eq. 16.14 in adverse pressure gradient ow, [48] proposed to re-dene the turbulent eddy viscosity including the expression in Eq. 16.13. We have

16. The SST Model

118

two expressions for the turbulent viscosity t =


v1 v2 c k = 1/2 c k k t = = 1/2 c 1/2

(16.15a) (16.15b)

where is the absolute vorticity (in boundary layer ow = 1 /x2 ); in (a) the v Boussinesq assumption together with Eq. 16.13 were used and (b) is taken from the k model. We want (a) to apply only in the boundary layer and hence we multiply it with a function F2 (similar to F1 ) which is 1 near walls and zero elsewhere. Then we take the minimum of (a) and (b) so that t = c k
1/2 max(c , F2 ) 1/2

(16.16)

When the production is large (i.e. when is large), Eq. 16.16 reduces t according to the Johnson - King model, i.e. Eq. 16.15a. It is important to ensure that this limitation is not active in usual boundary layer ows where P k . It can be seen that t is 1/2 reduced only in regions where P k > , because if P k < then < c since 1 2 = t 2 = P k < = c 2 t k k (16.17)

Hence, in regions where P k < , Eq. 16.16 returns to t = k/ as it should. Today, the SST model has been slightly further developed. Two modications have been introduced [53]. The rst modication is that the absolute vorticity in Eq. 16.16 1/2 has been replaced by || = (2ij sij ) s s which comes from the production term using the Boussinesq assumption (see Eq. 11.31), i.e. i v j i v v + = 2ij (ij + ij ) = 2ij sij s s s xj xi xj v j v 1 i ij = 2 xj xi || = s

(16.18)

where sij ij = 0 because sij is symmetric and ij is anti-symmetric. Equation 16.16 with || limits t in stagnation regions similar to Eq. 13.12. The second modication s in the SST model is that the production term in the new SST model is limited by 10, i.e. Pk,new = min P k , 10 (16.19) The nal form of the SST model is given in Eq. 19.5 at p. 147.

17. Large Eddy Simulations


0.5 0.45 0.4

119

v1 , v1

0.35 0.3 0.25 0.2 0.15 0.1 2 2.5 3 no filter one filter

x1

3.5

Figure 17.1: Filtering the velocity.

17 Large Eddy Simulations


17.1 Time averaging and ltering
In CFD we time average our equations to get the equations in steady form. This is called Reynolds time averaging: = 1 2T
T T

(t)dt, = +

(17.1)

(note that we use the notation . for time averaging). In LES we lter (volume average) the equations. In 1D we get (see Fig. 17.1)
x+0.5x 1 (, t)d (x, t) = x x0.5x = +

Since in LES we do not average in time, the ltered variables are functions of space and time. The equations for the ltered variables have the same form as Navier-Stokes, i.e. 1 p 2 vi ij i v (i vj ) = v + + t xj xi xj xj xj vi =0 xi where the subgrid stresses are given by ij = vi vj vi vj
Contrary to Reynolds time averaging where vi = 0, we have here vi = 0

(17.2)

(17.3)

v i = vi

17.2. Differences between time-averaging (RANS) and space ltering (LES)

120

This is true for box lters. Note that for the spectral cut-off lter vi = v i , see p. 122. However, in nite volume methods, box lters are always used. In this course we use box lters, if not otherwise stated. Lets look at the ltering of Eq. 17.2 in more detail. The pressure gradient term, for example, reads 1 p p = dV xi V V xi Now we want to move the derivative out of the integral. When is that allowed? The answer is if the integration region is not a function of xi , i.e. if V is constant. In nite volume methods, the ltering volume, V , is (almost always) identical to the control volume. In general, the size of the control volume varies in space. Fortunately, it can be shown that if V is a function of xi , the error we do when moving the derivative out of the integral is proportional to V 2 , i.e. it is an error of second order. Since this is the order of accuracy of our nite volume method anyway, we can accept this error. Now lets move the derivative out of the integral, i.e. p = xi xi 1 V pdV
V

+O V2 =

p +O V2 xi

All linear terms are treated in the same way. Now we take a look at the non-linear term in Eq. 17.2, i.e. the convective term. First we lter the term and move the derivative out of the integral, i.e. vi vj = xj xj 1 V vi vj dV
V

+O V2 =

(vi vj ) + O V 2 xj

There is still a problem with the formulation of this term: it includes an integral of a product, i.e. vi vj ; we want it to appear like a product of integrals, i.e. vi vj . To achieve this we simple add the term we want (i vj ) and subtract the one we dont want ( vi vj ) v on both the right and left side. This is how we end up with the convective term and the SGS term in Eq. 17.2.

17.2 Differences between time-averaging (RANS) and space ltering (LES)


In RANS, if a variable is time averaged twice ( v ), it is the same as time averaging once ( v ). This is because v is not dependent on time. From Eq. 17.1 we get v = 1 2T
T

v dt =
T

1 v 2T = v 2T

This is obvious if the ow is steady, i.e. v /t = 0. If the ow is unsteady, we must assume a separation in time scales so that the variation of v during the time interval T is negligible, i.e. /t 1/T . In practice this requirement is rarely satised. In LES, v = v (and since v = v + v we get v = 0). Lets lter vI once more (lter size x, see Fig. 17.2. For simplicity we do it in 1D. (Note that subscript I denotes node number.) vI = = 1 x 1 x
x/2

v ()d =
x/2

1 x

x/2

v ()d +
x/2 0

v ()d

x x vA + vB . 2 2

17.3. Resolved & SGS scales


x

121

x I 1 A I

x B I +1

x Figure 17.2: Box lter illustrated for a control volume. E() I cut-off

II

resolved scales

III
SGS Figure 17.3: Spectrum of velocity. The trapezoidal rule, which is second-order accurate, was used to estimate the integrals. v at locations A and B (see Fig. 17.2) is estimated by linear interpolation, which gives vI = 1 1 3 3 1 vI1 + vI + vI + vI+1 2 4 4 4 4 1 v v = (I1 + 6I + vI+1 ) = vI 8

(17.4)

17.3 Resolved & SGS scales


The basic idea in LES is to resolve (large) grid scales (GS), and to model (small) subgrid-scales (SGS). The limit (cut-off) between GS and SGS is supposed to take place in the inertial subrange (II), see Fig. 17.3.

17.4. The box-lter and the cut-off lter


I: II: III: large, energy-containing scales inertial subrange (Kolmogorov 5/3-range) dissipation subrange

122

17.4 The box-lter and the cut-off lter


The ltering is formally dened as (1D)

v (x) =

GB (r)v(x r)dr 1/, if r /2 0, if r >

GB (r) =

(17.5)

GB (r)dr = 1

It is often convenient to study the ltering process in the spectral space. The lter in spectral space is particular simple: we simply set the contribution from wavenumbers larger than cut-off to zero. Hence the cut-off lter lters out all scales with wavenumber larger than the cut-off wavenumber c = /. It is dened as GC () = 1/ if c 0 otherwise (17.6)

The Fourier transform is dened as (see Section C) v () = and its inverse v(r) =
0

1 2

v(r) exp(r)dr
0

(17.7)

v () exp(r)d

(17.8)

where denotes the wavenumber and = 1. Note that it is physically meaningful to use Fourier transforms only in a homogeneous coordinate direction; in nonhomogeneous directions the Fourier coefcients which are not a function of space have no meaning. Using the convolution theorem (saying that the integrated product of two functions is equal to the product of their Fourier transforms) the ltering in Eq. 17.5 is conveniently written

v () = v () =
0 0

v () exp()d

=
0 0 0

exp()GC ()v( )dd

(17.9)

=
0

exp() exp(( ))GC ()v( )dd

=
0

exp() exp()GC ()v()dd = GC ()() v

If we lter twice with the cut-off lter we get (see Eq. 17.9) v v = GC G = GC v = v (17.10)

17.5. Highest resolved wavenumbers

123

0.4
v1 0.2

0.4
v1 0.2

0 0.2 0

node 1
0.5

2
1

3
1.5

0 0.2 0

2
0.5

3 x1 /L

4
1

5
1.5

(a) One sinus period covering two cells

x1 /L

(b) One sinus period covering four cells

Figure 17.4: Physical and wavenumber space. Sinus curves with different wavenumbers illustrated in physical space. using Eqs. 17.9 and 17.6. Thus, contrary to the box-lter (see Eq. 17.4), nothing happens when we lter twice in spectral space. The box lter is sharp in physical space but not in wavenumber space; for the cut-off lter it is vice versa. In nite volume methods box ltering is always used. Furthermore implicit ltering is employed. This means that the ltering is the same as the discretization (=integration over the control volume which is equal to the lter volume, see Eq. 17.14).

17.5 Highest resolved wavenumbers


Any function can be expressed as a Fourier series as Eq. 17.8 provided that the coordinate direction is homogeneous. Lets choose the uctuating velocity in the x1 direction, i.e. v1 , and let it be a function of x1 . We require it to be homogeneous, i.e. its RMS, v1,rms , does not vary with x1 . Now we ask the question: on a given grid, what is the highest wavenumber that is resolved? Or, in other words, what is the cut-off wavenumber? In Fig. 17.4a, the highest wave
v1 = 0.25 [1 + 0.8 sin(1 x1 )] ,

1 = 2/L

(17.11)

and it covers two cells (x1 /L = 0.5). If we dene this as the cut-off wavenumber we get 1,c L = 1,c 2x1 = 2 so that 1,c = 2/(2x1 ) = /x1 (17.12)

It is of course questionable if v1 in Fig. 17.4a really is resolved since the sinus wave covers only two cells. However this is the usual denition of the cut-off wavenumber. If we require that the highest resolved wavenumber should be covered by four cells (x1 /L = 0.25), as in Fig. 17.4b, then the cut-off wavenumber is given by 1,c = 2/(4x1 ) = /(2x1 ).

17.6 Subgrid model


We need a subgrid model to model the turbulent scales which cannot be resolved by the grid and the discretization scheme. The simplest model is the Smagorinsky model [54]: 1 ij ij kk = sgs 3 sgs = (CS )
2

i v j v + xj xi

= 2sgs sij
2

(17.13)

2ij sij (CS ) || s s

17.7. Smagorinsky model vs. mixing-length model

124

and the lter-width is taken as the local grid size = (VIJK )


1/3

(17.14)

The scalar || is the norm (i.e. the length) of i /xj + j /xi in the Boussinesq s v v assumption, see Eq. 16.18. Near the wall, the SGS viscosity becomes quite large since the velocity gradient is very large at the wall. However, because the SGS turbulent uctuations near a wall go to zero, so must the SGS viscosity. A damping function f is added to ensure this f = 1 exp(x+ /26) 2 (17.15)

A more convenient way to dampen the SGS viscosity near the wall is simply to use the RANS length scale as an upper limit, i.e. = min (VIJK )
1/3

, n

(17.16)

where n is the distance to the nearest wall. Disadvantage of Smagorinsky model: the constant CS is not constant, but it is ow-dependent. It is found to vary in the range from CS = 0.065 [55] to CS = 0.25 [56].

17.7 Smagorinsky model vs. mixing-length model


The eddy viscosity according to the mixing length theory reads in boundary-layer ow [57, 58] 1 v . t = 2 x2 Generalized to three dimensions, we have t = 2 i v j v + xj xi i v xj
1/2

= 2 (2ij sij ) s

1/2

2 ||. s

In the Smagorinsky model the SGS turbulent length scale corresponds to = CS so that sgs = (CS )2 || s which is the same as Eq. 17.13

17.8 Energy path


The path of kinetic energy is illustrated in Fig. 17.5. At cut-off, SGS kinetic energy is dissipated sgs = ij sij = 2sgs sij sij (17.17) from the resolved turbulence. This energy is transferred to the SGS scales and act as production term (Pksgs ) in the ksgs equation. The SGS kinetic energy is then transferred to higher wave-numbers via the cascade effect and the kinetic energy is nally dissipated (=physical dissipation) in the dissipation range. It should be mentioned that this process is an idealized one. We assume that ALL dissipation takes place in the dissipation range. This is a good approximation, but in reality dissipation (i.e. transfer of energy from kinetic energy to internal energy, i.e. increase in temperature) takes place at all wave numbers, and the dissipation increases for increasing wave number.

17.9. SGS kinetic energy

125

E()

PK

in e

rti al
gs

ran P

ge
gs

E()

5/3

ks

dissipating range c Figure 17.5: Energy spectrum.

17.9 SGS kinetic energy


The SGS kinetic energy ksgs can be estimated from the Kolmogorov 5/3 law. The total turbulent kinetic energy is obtained from the energy spectrum as

k=
0

E()d

Changing the lower integration limit to wavenumbers larger than cut-off (i.e. c ) gives the SGS kinetic energy

ksgs =
c

E()d

(17.18)

The Kolmogorov 5/3 law now gives

ksgs =
c

C5/3 2/3 d

(Note that for these high wavenumbers, the Kolmogorov spectrum ought to be replaced by the Kolmogorov-Pau spectrum in which an exponential decaying function is added for high wavenumbers [57, Chapter 3]). Carrying out the integration and replacing c with / we get 2/3 3 (17.19) ksgs = C 2 In the same way as ksgs can be computed from Eq. 17.18, the resolved turbulent kinetic energy, kres , is obtained from
c

kres =
0

E()d

17.10 LES vs. RANS


LES can handle many ows which RANS (Reynolds Averaged Navier Stokes) cannot; the reason is that in LES large, turbulent scales are resolved. Examples are:

17.11. The dynamic model


test lter E() SGS grid scales I SGS test lter scales
II

126 cut-off, grid lter

resolved on test lter resolved on grid lter

III

= / c = /

Figure 17.6: Energy spectrum with grid and test lter. o Flows with large separation o Bluff-body ows (e.g. ow around a car); the wake often includes large, unsteady, turbulent structures o Transition In RANS all turbulent scales are modelled inaccurate In LES only small, isotropic turbulent scales are modelled accurate LES is very much more expensive than RANS. If we apply the second lter to the grid-ltered equations (Eq. 17.2) we obtain

17.11 The dynamic model


In this model of [59] the constant C is not arbitrarily chosen (or optimized), but it is computed. If we apply two lters to Navier-Stokes [grid lter and a second, coarser lter (test lter, denoted by . )] where = 2 we get vi + t xj vivj = 1 p 2 v i Tij + xi xj xj xj (17.20)

where the subgrid stresses on the test level now are given by Tij = vi vj v i v j (17.21)

17.12. The test lter


test lter grid lter

127

Figure 17.7: Control volume for grid and test lter.

vi + t xj

vivj

1 p 2 v i ij + xi xj xj xj vi vj v i v j xj =

(17.22)

Identication of Eqs. 17.20 and 17.22 gives vi vj v i v j + ij = Tij The dynamic Leonard stresses are now dened as Lij vi vj v i v j = Tij ij The trace of this relation reads Lii Tii ii With this expression we can re-formulate Eq. 17.24 as 1 1 Lij ij Lkk = Tij ij Tkk 3 3 1 ij ij kk 3 (17.25) (17.24) (17.23)

In the energy spectrum, the test lter is located at lower wave number than the grid lter, see Fig. 17.6.

17.12 The test lter


The test lter is twice the size of the grid lter, i.e. = 2. The test-ltered variables are computed by integration over the test lter. For ex ample, the 1D example in Fig. 17.7 v is computed as (x = 2x) v = = 1 2x
E

v dx =
W

1 2x

v dx +
W P

v dx (17.26)

1 1 (w x + ve x) = v 2x 2 1 = (W + 2P + vE ) v v 4

vW + vP vP + vE + 2 2

17.13. Stresses on grid, test and intermediate level


I, J + 1, K

128

I 1, J, K

I, J, K

I + 1, J, K

x2 x1 I, J 1, K

Figure 17.8: A 2D test lter control volume. For 3D, ltering at the test level is carried out in the same way by integrating over the test cell assuming linear variation of the variables [60], i.e. (see Fig. 17.8) v I,J,K = 1 (I1/2,J1/2,K1/2 + vI+1/2,J1/2,K1/2 v 8 +I1/2,J+1/2,K1/2 + vI+1/2,J+1/2,K1/2 v +I1/2,J1/2,K+1/2 + vI+1/2,J1/2,K+1/2 v +I1/2,J+1/2,K+1/2 + vI+1/2,J+1/2,K+1/2 ) v

(17.27)

17.13 Stresses on grid, test and intermediate level


The stresses on the grid level, test level and intermediate level (dynamic Leonard stresses) have the form ij = vi vj vi vj stresses with < Tij = vi vj v i v j stresses with < Lij = Tij ij stresses with < < Thus the dynamic Leonard stresses represent the stresses with lengthscale, , in the range between and . Assume now that the same functional form for the subgrid stresses that is used at the grid level (ij ) also can be used at the test lter level (Tij ). If we use the Smagorinsky model we get 1 ss (17.28) ij ij kk = 2C2 ||ij 3
2 1 Tij ij Tkk = 2C | s | s ij 3

(17.29)
1/2

where s ij

1 v i vj = + , | s |= 2 xj xi

2 s ij s ij

17.14. Numerical dissipation

129

Note that C in Eq. 17.28 is not squared (cf. the Smagorinsky model, Eq. 17.13 at 2 p.123). Hence, C should be compared with CS . Applying the test lter to Eq. 17.28 (assuming that C varies slowly), substituting this equation and Eq. 17.29 into Eq. 17.25 gives 2 1 (17.30) ss Lij ij Lkk = 2C | s | s ij 2 ||ij 3 Note that the constant C really is a function of both space and time, i.e. C = C(xi , t). Equation 17.30 is a tensor equation, and we have ve (ij is symmetric and traces less) equations for C. [61] suggested to satisfy Eq. 17.30 in a least-square sense. Let us dene the error as the difference between the left-hand side and the right-hand side of Eq. 17.30 raised to the power of two, i.e. Q= Mij = 1 Lij ij Lkk + 2CMij 3
2 2

(17.31a) (17.31b)

ss | s | s ij 2 ||ij

The error, Q, has a minimum (or maximum) when Q/C = 0. Carrying out the derivation of 17.31a gives Q 1 = 4Mij Lij ij Lkk + 2CMij C 3 =0 (17.32)

Since 2 Q/C 2 = 8Mij Mij > 0 it is a minimum. Equation 17.31 is re-written so that Lij Mij (17.33) C= 2Mij Mij It turns out that the dynamic coefcient C uctuates wildly both in space and time. This causes numerical problems, and it has been found necessary to average C in homogeneous direction(s). Furthermore, C must be clipped to ensure that the total viscosity stays positive ( + sgs 0). In real 3D ows, there is no homogeneous direction. Usually local averaging and clipping (i.e. requiring that C stays within pre-dened limits) of the dynamic coefcient is used. Use of one-equation models solve these numerical problems (see p. 137).

17.14 Numerical dissipation


The main function of an SGS model is to dissipate (i.e. to dampen) resolved turbulent uctuations. The SGS model is hopefully designed to give a proper amount of dissipation. This is the reason why in LES we should use a central differencing scheme, because this class of schemes does not give any numerical dissipation. All upwind schemes give numerical dissipation in addition to the modelled SGS dissipation. Indeed, there are LES-methods in which upwind schemes are used to create dissipation and where no SGS model is used at all (e.g. MILES [62]). However, here we focus on ensuring proper dissipation through an SGS model rather than via upwind differencing. It can be shown using Neumann stability analysis that all upwind schemes are dissipative (see Further reading at

17.15. Scale-similarity Models


vI I 1 I I +1

130

Figure 17.9: Numerical dissipation. http://www.tfd.chalmers.se/lada/comp turb model/). Below it is shown that rst-order upwind schemes are dissipative. The (exact) rst-derivative in the convective term is estimated by rst-order upwind differencing as (nite difference, see Fig. 17.9) vI v x = vI
I,exact

vI vI1 + O (x) x

(17.34)

where we have assumed vI > 0. Taylor expansion gives vI1 = vI x so that vI vI1 = x Insert this into Eq. 17.34 v v x = vI
I,exact

v x

1 + (x)2 2 I,num

2v x2

I,num

+ O (x)3

v x

1 x 2 I,num

2v x2

I,num

+ O (x)2

v x

1 x 2 I,num

2v x2

I,num

+ O (x)2

When this expression is inserted into the LES momentum equations, the second term on the right-hand side will act as an additional (numerical) diffusion term. The total diffusion term will have the form diusion term = x ( + sgs + num ) v x (17.35)

where the additional numerical viscosity, num 0.5|I |x. This means that the total v dissipation due to SGS viscosity and numerical viscosity is (cf. Eq. 17.17) sgs+num = 2(sgs + num )ij sij s For more details on derivation of equations transport equations of turbulent kinetic energies, see [14].

17.15 Scale-similarity Models


In the models presented in the previous sections (the Smagorinsky and the dynamic models) the total SGS stress ij = vi vj vi vj was modelled with an eddy-viscosity

17.16. The Bardina Model

131

hypothesis. In scale-similarity models the total stress is split up as


v ij = vi vj vi vj = (i + vi )(j + vj ) vi vj v = vi vj + vi vj + vj vi + vi vj vi vj

= (i vj vi vj ) + vi vj + vj vi + vi vj v

where the term in brackets is denoted the Leonard stresses, the term in square brackets is denoted cross terms, and the last term is denoted the Reynolds SGS stress. Thus ij = Lij + Cij + Rij Cij = vi vj + vj vi
Rij = vi vj .

Lij = vi vj vi vj

(17.36)

Note that the Leonard stresses Lij are computable, i.e. they are exact and dont need to be modelled. In scale-similarity models the main idea is that the turbulent scales just above cutoff wavenumber, c , (smaller than ) are similar to the ones just below c (larger than ); hence the word scale-similar. Looking at Eq. 17.36 it seems natural to assume that the cross term is responsible for the interaction between resolved scales (i ) and v modelled scales (vi ), since Cij includes both scales.

17.16 The Bardina Model


In the Bardina model the Leonard stresses Lij are computed explicitly, and the sum of the cross term Cij and the Reynolds term is modelled as [63, 64]
M Cij = cr (i vj v i v j ) v

(17.37)

M and Rij = 0 (superscript M denotes Modelled). It was found that this model was not sufciently dissipative, and thus a Smagorinsky model was added M Cij = cr (i vj v i v j ) v M 2 Rij = 2CS 2 ||ij ss

(17.38)

17.17 Redened terms in the Bardina Model


The stresses in the Bardina model can be redened to make them Galilean invariant for any value cr (see Appendix G). A modied Leonard stress tensor Lm is dened as [65] ij
m m m ij = ij = Cij + Lm + Rij ij

m Cij = 0

v Lm = cr (i vj v i v j ) ij

(17.39)

m Rij = Rij = vi vj

Note that the modied Leonard stresses is the same as the unmodied one plus the modelled cross term Cij in the Bardina model with cr = 1 (right-hand side of Eq. 17.37), i.e. M Lm = Lij + Cij ij

17.18. A dissipative scale-similarity model.


E()
10

132

+ SGS SGS

10

20

30 0 50 100 150

x+ 2

Figure 17.10: Dissipation terms and production term from DNS data. 963 mesh data : + ; : ; +: SGS . ltered onto a 483 mesh. Re = 500. SGS SGS In order to make the model sufciently dissipative a Smagorinsky model is added, and the total SGS stress ij is modelled as ij = vi vj v i v j 2(CS )2 ||ij ss Below we verify that the modied Leonard stress is Galilean invariant. 1 m v v v v L = vi vj v i v j = (i + Vi )(j + Vj ) (i + Vi ) (j + Vj ) cr ij = vi vj + vi Vj + vj Vi v i v j v i Vj Vi v j 1 = vi v j v i v j = Lm cr ij (17.40)

(17.41)

17.18 A dissipative scale-similarity model.


Above it was mentioned that when the rst scale-similarity model was proposed it was found that it is not sufciently dissipative [63]. An eddy-viscosity model has to be added to make the model sufciently dissipative; these models are called mixed models. [66] (can be downloaded from www.tfd.chalmers.se/lada) presents and evaluates a dissipative scale-similarity model. The ltered Navier-Stokes read 2 vi ik 1 p di v = + dt xi xk xk xk (17.42)

where d/dt and ik denote the material derivative and the SGS stress tensor, respectively. In the scale-similarity model, the SGS stress tensor is given by ik = vi vk vi vk (17.43)

This model is not sufciently dissipative. Let us take a closer look at the equation for

17.19. Forcing
the resolved, turbulent kinetic energy, K = vi vi /2, which reads

133

dK p vi 1 vk vi vi vi 2 vi + + = v + vk vi dt xk xi 2 xk xk xk i 2 vi ik ik ik vi = v v = xk xk xk xk i xk i 2K i i v v ik v xk xk xk xk xk i
SGS

(17.44)

The rst term on the last line is the viscous diffusion term and the second term, , is the viscous dissipation term which is always positive. The last term, SGS , is a source term arising from the SGS stress tensor, which can be positive or negative. When it is positive, forward scattering takes place (i.e. it acts as a dissipation term); when it is negative, back scattering occurs. Figure 17.10 presents SGS dissipation, SGS in Eq. 17.44, computed from ltered DNS data. The forward scatter, + , and back scatter, , SGS dissipation are SGS SGS dened as the sum of all instants when SGS is positive and negative, respectively. As can be seen, the scale-similarity model is slightly dissipative (i.e. SGS > 0) , but the forward and back scatter dissipation are both much larger than SGS . One way to make the SGS stress tensor strictly dissipative is to set the back scatter to zero, i.e. max(SGS , 0). This could be achieved by setting ik /xk = 0 when its sign is different from that of vi (see the last term in Eq. 17.44). This would work if we were solving for K. Usually we do not, and the equations that we do solve (the ltered Navier-Stokes equations) are not directly affected by the dissipation term, SGS . Instead we have to modify the SGS stress tensor as it appears in the ltered NavierStokes equations, Eq. 17.42. The second derivative on the right side is usually called a diffusion term because it acts like a diffusion transport term. When analyzing the stability properties of discretized equations to an imposed disturbance, v , using Neumann analysis (see, for example, Chapter 8 in [67]), this term is referred to as a dissipation term. In stability analysis the concern is to dampen numerical oscillations; in connection with SGS models, the aim is to dampen turbulent resolved uctuations. It is shown in Neumann analysis that the diffusion term in the Navier-Stokes equations is dissipative, i.e. it dampens numerical oscillations. However, since it is the resolved turbulent uctuations, i.e. K in Eq. 17.44, that we want to dissipate, we must consider the ltered Navier-Stokes equations for the uctuating velocity, vi . It is the diffusion term in this equation which appears in the rst term on the right side (rst line) in Eq. 17.44. To ensure that SGS > 0, we set ik /xk to zero when its sign is different from that of the viscous diffusion term (cf. the two last terms on the second line in Eq. 17.44). This is achieved by dening a sign function; for details, see [66].

17.19 Forcing
An alternative way to modify the scale-similarity model is to omit the forward scatter, i.e. to include instants when the subgrid stresses act as counter-gradient diffusion. In hybrid LES-RANS, the stresses can then be used as forcing at the interface between URANS and LES. This new approach is the focus of [68].

17.20. Numerical method

134

17.20 Numerical method


A numerical method based on an implicit, nite volume method with collocated grid arrangement, central differencing in space, and Crank-Nicolson ( = 0.5) in time is briey described below. The discretized momentum equations read vi
n+1/2

= vi + tH v n , vi n

n+1/2

pn+1/2 pn (1 )t xi xi

(17.45)

where H includes convective, viscous and SGS terms. In SIMPLE notation this equation reads pn+1/2 n+1/2 V = aP vi anb v n+1/2 + SU t xi
nb

where SU includes all source terms except the implicit pressure. The face velocities n+1/2 n+1/2 n+1/2 + vi,j1 ) (note that j denotes node number and i is a tensor vf,i = 0.5(i,j v index) do not satisfy continuity. Create an intermediate velocity eld by subtracting the implicit pressure gradient from Eq. 17.45, i.e. vi = vi + tH v n , vi n vi = vi
n+1/2 n+1/2

(1 )t

pn xi

(17.46a) (17.46b)

+ t

pn+1/2 xi
n+1/2

Take the divergence of Eq. 17.46b and require that f,i v v 1 f,i 2 pn+1 = xi xi t xi

/xi = 0 so that (17.47)

The Poisson equation for pn+1 is solved with an efcient multigrid method [69]. In the 3D MG we use a plane-by-plane 2D MG. The face velocities are corrected as vf,i = vf,i t n+1 pn+1 xi (17.48)

A few iterations (typically two) solving the momentum equations and the Poisson pressure equation are required each time step to obtain convergence. More details can be found [70] 1. Solve the discretized ltered Navier-Stokes equation for v1 , v2 and v3 . 2. Create an intermediate velocity eld vi from Eq. 17.46. 3. The Poisson equation (Eq. 17.47) is solved with an efcient multigrid method [69]. 4. Compute the face velocities (which satisfy continuity) from the pressure and the intermediate face velocity from Eq. 17.48 5. Step 1 to 4 is performed till convergence (normally two or three iterations) is reached. 6. The turbulent viscosity is computed.

17.20. Numerical method


RANS 2D or 3D steady or unsteady 2nd order upwind 1st order more than two-equations

135 LES always 3D always unsteady central differencing 2nd order (e.g. C-N) zero- or one-equation

Domain Time domain Space discretization Time discretization Turbulence model

Table 17.1: Differences between a nite volume RANS and LES code. v1

t t1 : start Figure 17.11: Time averaging in LES. 7. Next time step. Since the Poisson solver in [69] is a nested MG solver, it is difcult to parallelize with MPI (Message Passing Interface) on large Linux clusters. Hence, when we do large simulations (> 20M cells) we use a traditional SIMPLE method. 17.20.1 RANS vs. LES Above a numerical procedure suitable for LES was described. However, in general, any numerical procedure used for RANS can also be used for LES; for example pressurecorrection methods such as SIMPLE [71, 72] are often used for LES. What are the specic requirements to carry out LES with a nite volume code? If you have a RANS nite volume code, it is very simple to transform that into an LES code. An LES code is actually simpler than a RANS code. Both the discretization scheme and and the turbulence model are simpler in LES and RANS, see Table 17.1. It is important to use a non-dissipative discretization scheme which does not introduce any additional numerical dissipation, see Section 17.14; hence a second-order (or higher) central differencing scheme should be employed. The time discretization should also be non-dissipative. The Crank-Nicolson scheme is suitable. As mentioned above, turbulence models in LES are simple. There are two reasons: rst, only the small-scale turbulence is modelled and, second, no equation for the turbulent length scale is required since the turbulent length scale can be taken as the lter width, . In LES we are doing unsteady simulations. The question then arises, when can we start to time average and for how long? This is exactly the same question we must t2 : end

17.21. One-equation ksgs model

136

E()

I
II

III
Figure 17.12: Spectrum for k. I: Range for the large, energy containing eddies; II: the inertial subrange for isotropic scales, independent of the large scales () and the dissipative scales (); III: Range for small, isotropic, dissipative scales. ask our self whenever doing an experiment in, for example, a windtunnel. We start the windtunnel: when has the ow (and turbulence) reached fully developed conditions so that we can start the measure the ow. Next question: for how long should we carry out the measurements. Both in LES and the windtunnel, the recorded time history of the v1 velocity at a point may look like in Fig. 17.11. Time averaging can start at time t1 when the ow seems to have reached fully developed conditions. It is difcult to judge for how long one should carry out time averaging. Usually it is a good idea to form a nondimensional time scale from a velocity, V (free-stream or bulk velocity), and a length scale, L (width of a wake, width or length of a recirculation region), and use this to estimate the required averaging time; 100 time units, i.e. 100L/V , may be a suitable averaging time.

17.21 One-equation ksgs model


A one-equation model can be used to model the SGS turbulent kinetic energy. The equation can be written on the same form as the RANS k-equation, i.e. ksgs ksgs ( + sgs ) + Pksgs (j ksgs ) = v + t xj xj xj sgs =
1/2 ck ksgs ,

Pksgs = 2sgs sij sij ,

ksgs = C

3/2

(17.49)

Note that the production term, Pksgs , is equivalent to the SGS dissipation in the equation for the resolved turbulent kinetic energy (look at the ow of kinetic energy discussed at the end of [73]).

17.22 Smagorinsky model derived from the k 5/3 law


We can use the one-equation model to derive the Smagorinsky model, Eq. 17.13. The subgrid model is supposed to model the turbulence with scales II , where II is proportional to the inverse of the wave number (II 1/II ) in the inertial subrange, see the Fig. 17.12. In the inertial subrange the turbulence is at large Reynolds number independent of the large scale as well as of the small scales. This range is characterized by the

17.23. A dynamic one-equation model

137

dissipation and the wavenumber II 1/. Thus we can write the SGS viscosity as sgs = a (CS )b (17.50) Dimensional analysis yields a = 1/3, b = 4/3 so that sgs = (CS )4/3 1/3 . (17.51) The scales in the inertial subrange are isotropic, and thus they are in local equilibrium, i.e. in the ksgs equation we have that production balances dissipation (see Eq. 17.49) 2sgs sij sij = . Eq. 17.52 substituted into Eq. 17.51 gives
3 sgs = (CS )4 = (CS )4 sgs (2ij sij ) s

(17.52)

sgs = (CS )2 || s || = (2ij sij ) s s

(17.53)

1/2

17.23 A dynamic one-equation model


One of the drawbacks of the dynamic model of [59] (see p. 126) is the numerical instability associated with the negative values and large variation of the C coefcient. Usually this problem is xed by averaging the coefcient in some homogeneous ow direction. In real applications ad-hoc local smoothing and clipping is used. Below a dynamic one-equation model is presented. The main object when developing this model was that it should be applicable to real industrial ows. Furthermore, being a dynamic model, it has the great advantage that the coefcients are computed rather than being prescribed. The equation for the subgrid kinetic energy reads [74, 75] (see also [76, 77]) ksgs (j ksgs ) = Pksgs + v + t xj xj ef f ksgs xj
1 2

ksgs

3/2

(17.54)

a a Pksgs = ij vi,j , ij = 2Cksgs sij


1 2 with ef f = + 2Chom ksgs . The C in the production term Pksgs is computed dynamically (cf. Eq. 17.33). To ensure numerical stability, a constant value (in space) of C (Chom ) is used in the diffusion term in Eq. 17.54 and in the momentum equations. Chom is computed by requiring that Chom should yield the same total production of ksgs as C, i.e. 1 2 2Cksgs sij sij 1 2 = 2Chom ksgs sij sij

xyz

xyz

The dissipation term ksgs is estimated as: ksgs . (17.55) Now we want to nd a dynamic equation for C . The equations for ksgs and K read in symbolic form ksgs Tf (vi,j , vi,j ) = C T (ksgs ) Cksgs Dksgs = Pksgs C T (K) CK DK = P K C ksgs K 3/2
3/2 3/2

(17.56)

17.24. A Mixed Model Based on a One-Eq. Model

138

Since the turbulence on both the grid level and the test level should be in local equilibrium (in the inertial 5/3 region), the left-hand side of the two equations in Eq. 17.56 should be close to zero. An even better approximation should be to assume T (ksgs ) = T (K), i.e. 3/2 K 3/2 1 = P K C , P ksgs C ksgs so that
n+1 C =

P K P ksgs +

1 n 3/2 C k sgs

K2
3

(17.57)

The idea is to put the local dynamic coefcients in the source terms, i.e. in the production and the dissipation terms of the ksgs equation (Eq. 17.54). In this way the dynamic coefcients C and C dont need to be clipped or averaged in any way. This is a big advantage compared to the standard dynamic model of Germano (see discussion on p. 129).

17.24 A Mixed Model Based on a One-Eq. Model


Recently a new dynamic scale-similarity model was presented by [78]. In this model a dynamic one-equation SGS model is solved, and the scale-similarity part is estimated in a similar way as in Eq. 17.40.

17.25 Applied LES


At the Department we used LES for applied ows such as ow around a cube [79, 80], the ow and heat transfer in a square rotating duct [81, 82], the ow around a simplied bus [83, 80], a simplied car [84, 85, 86] and the ow around an airfoil [87, 88], detailed SUV [89], trains and buses subjected to sidewinds and wind gusts [90, 91, 92]. We have also done some work on buoyancy-affected ows [93, 94, 95, 96, 97, 98, 99].

17.26 Resolution requirements


The near-wall grid spacing should be about one wall unit in the wall-normal direction. This is similar to the requirement in RANS (Reynolds-Averaged Navier-Stokes) using low-Re number models. The resolution requirements in wall-parallel planes for a wellresolved LES in the near-wall region expressed in wall units are approximately 100 (streamwise) and 30 (spanwise). This enables resolution of the near-wall turbulent structures in the viscous sub-layer and the buffer layer consisting of high-speed inrushes and low-speed ejections [100], often called the streak process. At low to medium Reynolds numbers the streak process is responsible for the major part of the turbulence production. These structures must be resolved in an LES in order to achieve accurate results. Then the spectra of the resolved turbulence will exhibit 5/3 range, see gure on p. 42. In applied LES, this kind of resolution can hardly ever be afforded. In outer scaling (i.e. comparing the resolution to the boundary layer thickness, ), we can afford /x1 and /x3 in the region of 5 20 and 10 20, respectively. In this case, the spectra in the boundary layer will look something like that shown in Fig. 17.13 [101]. Energy spectra are actually not very reliable to judge if a LES simulation is well resolved or not. In [101, 102] different ways to estimate the resolution of an LES were investigated. The

17.26. Resolution requirements


Thick dashed line 5/3 slope

139

10

E33 (k3 )

10

10

10

10

10

10

3 = 2k3 /x3,max Figure 17.13: Energy spectra in fully developed channel ow [101]. denotes half channel width. Number of cells expressed as (/x1 , /x3 ). : (10, 20); : (20, 20); : (10, 40); : (5, 20); +: (10, 10). L1

V1,in y h x L2 L Figure 17.14: Onera bump. Computational domain (not to scale). suggestion in these works was that two-point correlations is the best way to estimate if an LES is sufciently resolved or not. Even if the turbulence in boundary layer seldom can be resolved, the ow in recirculation regions and shear layer can. In [103] the ow (Re 106 ) over a bump was computed. The geometry is shown in Fig. 17.14. The turbulence in the boundary layer on the bump was very poorly resolved: x1 /in = 0.33, x3 /in = 0.44, x+ = 1300 and x+ = 1800. Nevertheless, the turbulence in the recirculation re1 3 gion and in the shear layer downstream the bump turned out to be well resolved, see Fig. 17.15. Thus, for wall-bounded ows at high Reynolds numbers of engineering interest, the computational resource requirement of accurate LES is prohibitively large. Indeed, the requirement of near-wall grid resolution is the main reason why LES is too expensive for engineering ows, which was one of the lessons learned in the LESFOIL project [104, 105]. H

18. Unsteady RANS

140

10

E33 (3 )

10

10

10

100

200

3 = 2k3 /x3,max Figure 17.15: Energy spectra E33 (3 ) in the recirculation region and the shear layer downstream the bump (x1 /H = 1.2). Thick dashed line shows 5/3 slope. : : x2 /H = 0.13; : x2 /H = 0.34 (in the shear x2 /H = 0.0035 (near the wall); layer).

18 Unsteady RANS
To perform an accurate LES, a very ne mesh must be used. This causes problems, for example, near walls. LES is very good for wake ow, where the ow is governed by large, turbulent structures, which can be captured by a fairly coarse mesh. However, if attached boundary layers are important, LES will probably give poor predictions in these regions, unless ne grids are used. An alternative to LES for industrial ows can then be unsteady RANS (ReynoldsAveraged Navier-Stokes), often denoted URANS (Unsteady RANS) or TRANS (Transient RANS). In URANS the usual Reynolds decomposition is employed, i.e. v (t) = 1 2T
t+T tT

v(t)dt, v = v + v

(18.1)

The URANS equations are the usual RANS equations, but with the transient (unsteady) term retained, i.e. (on incompressible form)
vi vj 1 p 2 vi i v (i vj ) = v + + t xj xi xj xj xj vi =0 xi

(18.2)

Note that the dependent variables are now not only function of the space coordinates, but also function of time, i.e. vi = vi (x1 , x2 , x3 , t), p = p(x1 , x2 , x3 , t) and vi vj = v (x , x , x , t). vi j 1 2 3 Even if the results from URANS are unsteady, one is often interested only in the time-averaged ow. We denote here the time-averaged velocity as v , which means that we can decompose the results from an URANS as a time-averaged part, v , a resolved uctuation, v , and the modelled, turbulent uctuation, v , i.e. v = v + v = v + v + v (18.3)

18. Unsteady RANS

141

see Fig. 18.1. The modelled turbulent uctuation, v is not shown in the gure; if this is added to v + v we obtain v. What type of turbulence model should be used in URANS? That depends on type of ow. If the ow has strong vortex shedding, the standard high-Re number k model can be used, i.e. k j k v = + t xj xj j v = + t xj xj t + t k k + P k xj (18.4)

+ c1 P k c2 xj k

(18.5)

k2 With an eddy-viscosity, the URANS equations read t = c i v 1 p i vk v i v ( + t ) = + + t xk xi xk xk

(18.6)

(18.7)

So we are doing unsteady simulations, but still we time average the equations. How is this possible? The theoretical answer is that the time, T , in Eq. 18.1 should be much smaller than the resolved time scale, i.e. the modelled turbulent uctuations, v , should have a much smaller time scale than the resolved ones, v . This is called scale separa tion. In practice this requirement is often not satised [70]. On the other hand, how do the momentum equation, Eq. 18.7, know how they were time averaged? Or if they were volume ltered? The answer is that they dont. The URANS momentum equation and the LES momentum equation are exactly the same, except that we denote the turbulent viscosity in the former case by t and in the latter case by sgs . In URANS, much more of the turbulence is modelled than in LES, and, hence, the turbulent viscosity, t , is much larger than the SGS viscosity, sgs . The common denition of URANS is that the turbulent length scale is not determined by the grid, whereas in LES it is. In URANS we do usually not care about scale v, v v

v t Figure 18.1: Decomposition of velocities in URANS.

18. Unsteady RANS

142

Figure 18.2: Conguration of the ow past a triangular ameholder. Flow from left to right

Figure 18.3: 2D URANS k simulations [106]. One cycle of the v2 velocity in a cell near the upper-right corner of the ameholder. separation. What we care about is that the turbulence model and the discretization scheme should not be too dissipative, i.e. they should not kill the resolved uctuations, v . The standard k model (Eq. 18.4 and 18.5) was used by [106] for URANS simulations computing the ow around a triangular ame-holder in a channel, see Fig. 18.2. This ow has a very regular vortex shedding. and the ow actually has a scale separation. In the gure below the v2 velocity in a point above the ame-holder is shown. Figure 18.3 shows that the velocity varies with time in a sinusoidal manner. When were doing URANS, the question arises how the results should be time averaged, i.e. when should we start to average and for how long. This issue is the same when doing LES, and this was discussed in the LES-lecture.

18.1. Turbulence Modelling

143

Figure 18.4: 2D URANS k simulations compared with experiment [106]. Solid lines: total turbulent kinetic energy; dashed lines: resolved turbulent kinetic energy: : experimental data. Left gure: x = 0.43H; right gure: x = 1.1H (x = 0 at the downstream vertical plane of the ame-holder).

18.1 Turbulence Modelling


In URANS, part of the turbulence is modelled (v ) and part of the turbulence is resolved ( ). If we want to compare computed turbulence with experimental turbulence, v we must add these two parts together. Proles downstream the ameholder are shown in Fig. 18.4. It can be seen that here the resolved and the modelled turbulence are of the same magnitude. If the turbulence model in URANS generates too much eddy viscosity, the ow may not become unsteady at all, because the unsteadiness is dampened out; the reason for this is that the turbulence model is too dissipative. [107, 108] found when using URANS for the ow around a surface-mounted cube and around a car, that the standard k model was too dissipative. Non-linear models like that of [109] was found to be less dissipative, and was successfully applied in URANS-simulations for these two ows.

18.2 Discretization
In LES it is well-known that non-dissipative discretization schemes should be used. The reason is that we dont want to dampen out resolved, turbulent uctuations. The same is to some extent true also for URANS. In the predictions on the ame-holder presented above, the hybrid discretization scheme for the convective terms was used together with fully implicit rst-order discretization in time; this gives rst-order accuracy in both space and time. The turbulence model that was used was the standard k model. Thus, both the discretization and the turbulence model have high dissipation. The reason why the unsteadiness in these computations was not dampened out is that the vortex shedding in this ow is very strong. In general a discretization scheme which has little numerical dissipation should be

19. DES

144

Figure 18.5: URANS simulations of the ow around a surface-mounted cube. used. How dissipative a scheme needs to be in order to be stable is ow dependent; for some simple ows, it may work with no dissipation at all (i.e. central differencing), whereas for industrially complex ows maybe a bounded second-order scheme must be used. For time discretization, the second-order accurate Crank-Nicolson works in most cases. In [107] LES and URANS simulations were carried out of the ow around a surfacemounted cube (Fig. 18.5) with a coarse mesh using wall-functions. Two different discretization schemes were used: the central scheme and the Mars scheme (a blend between central differencing and a bounded upwind scheme of secondorder accuracy). In Fig. 18.6 the time-averaged velocity prole upstream of the cube (x1 = 0.6H) using URANS and LES with central differencing are shown together with URANS and Mars scheme. It is seen that with LES and central differencing unphysical oscillations are present (this was also found by [79]). However, LES with the Mars scheme (in which some numerical dissipation is present) and URANS with the central scheme (where the modelling dissipation is larger than in LES) no such unphysical oscillations are present. The main reason to the unphysical oscillations is that the predicted ow in this region does not have any resolved uctuations. If turbulent unsteady inlet uctuations are used, the unphysical oscillations do usually not appear, even if a central differencing scheme is used. In this case the turbulent, resolved uctuations dominate over any numerical oscillations.

19 DES
DES (Detached Eddy Simulation) is a mix of LES and URANS. The aim is to treat the boundary layer with RANS and capture the outer detached eddies with LES. The model was originally developed for wings at very high angles of attack. The RANS model that was originally used was the one-equation model by [110]. It can be written [110, 104, Sect. 4.6] t j t v = + t xj xj t = t f1 + t t t xj + Cb2 t t +P t xj xj

(19.1)

19.1. DES based on two-equation models

145

Figure 18.6: URANS simulations of the ow around a surface-mounted cube. Velocity proles upstream the cube [107]. The production term P and the destruction term have the form P = Cb1 s + s = (2ij sij ) s t f 2 d2 2 , t t d
2

(19.2)

1/2

= Cw1 fw

d in the RANS SA model is equal to the distance to the nearest wall. In [111] the DES model was proposed in which d is taken as the minimum of the RANS turbulent length scale d and the cell length = max(x , x , x ), i.e. d = min(d, Cdes ) (19.3)

x , x and x denote the cell length in the three grid directions , and . The constant Cdes is usually set to 0.65. In the boundary layer d < Cdes and thus the model operates in RANS mode. Outside the turbulent boundary layer d > Cdes so that the model operates in LES mode. The modelled length scale is reduced and the consequence is that the destruction term Y increases, which gives a reduction in the turbulent viscosity t . A reduced t gives a smaller production term P so that the turbulent viscosity is further reduced. At rst sight it may seem that as the model switches from RANS mode to LES mode thus reducing d, this would give rise to an increased production term P through the second term (see Eq. 19.2). However, this second term is a viscous term and is active only close to the wall. This term is sometimes neglected [112]

19.1 DES based on two-equation models


The model described above is a one-equation model. In RANS mode it takes its length scale from the wall distance, which in many situations is not a relevant turbulent length

19.1. DES based on two-equation models

146

scale. Recently, DES models based on two-equation models have been formulated by [113, 114, 115]. In these models the turbulent length scale is either obtained from the two turbulent quantities (e.g. k 3/2 / or k 1/2 /) or the lter width . A model based on the k model can read k (j k) = v + t xj (j ) = v + t xj xj xj t k t + + k + P k T xj + (C1 P k C2 ) xj k

P k = 2t sij sij ,

t = k 1/2 t

The turbulent length scale t and the turbulent dissipation, T , are computed as [115, 116] t = min C k 3/2 , Ck k 3/2

T = max , C

In other models [113, 53] only the dissipation term, T is modied. When the grid is sufciently ne, the length scale is taken as . The result is that the dissipation in the k equation increases, k decreases which gives a reduced t . A third alternative is to modify only the turbulent length scale appearing in the turbulent viscosity [116]. A rather new approach is to reduce the destruction term in the equation as in PANS [117, 118, 119] (Partially Averaged Navier-Stokes) and PITM [120] (Partially Integrated Transport Modelling). In these models increases because of its reduced destruction term which decreases both k and t . The values of the constants can be (C , , Ck , C , C1 , C2 ) = (0.09, 1.31, 0.07, 1.09, 1.44, 1.92) (19.4)

Note that a low-Re k model should be used. The AKN model [121] could be a suitable one. In regions where the turbulent length scales are taken from (LES mode) the equation is still solved, but is not used. However, it is needed as soon as the model switches to RANS model again. In the RANS mode the major part of the turbulence is modelled. When the model switches to LES mode, the turbulence is supposed to be represented by resolved turbulence. This poses a major problem with this type of models. If the switch occurs at location x1 , say, it will take some distance L before the momentum equations start to resolve any turbulence. This is exactly what happens at an inlet in an LES simulation if no real turbulence is given as inlet boundary conditions. One way to get around this is to impose turbulence uctuations as forcing conditions [122, 123, 73, 124, 125, ] at the location where the model switches from RANS mode to LES mode. The forcing is added in the form of a source term (per unit volume) in the momentum equations.

19.2. DES based on the k SST model

147

19.2 DES based on the k SST model


The standard k model SST reads [48, 53] k (j k) = v + t xj (j ) = v + t xj xj xj +

t k + Pk k k xj t Pk + + 2 xj t 1 k + 2(1 F1 )2 xi xi k 500 4 F1 = tanh( ), = min max , d d2 t = a1 k max(a1 , ||F2 ) s = max

42 k CD d2

(19.5)

F2 = tanh( 2 ),

2k 1/2 500 , d d2

where d is the distance to the closest wall node. The SST model behaves as a k model near the wall where F1 = 1 and a k model far from walls (F1 = 0). All coefcients are blended between the k and the k model using the function F1 . In DES the dissipation term in the k equation is modied as [53] Lt ,1 CDES k 1/2

k kFDES ,

FDES = max Lt =

= max {x1 , x2 , x3 } ,

Again, the DES modication is meant to switch the turbulent length scale from a RANS length scale ( k 1/2 /) to a LES length scale ( ) when the grid is sufciently ne. When FDES is larger than one, the dissipation term in the k equation increases which in turn decreases k and thereby also the turbulent viscosity. With a smaller turbulent viscosity in the momentum equations, the modelled dissipation (i.e the damping) is reduced and the ow is induced to go unsteady. The result is, hopefully, that a large part of the turbulence is resolved rather than being modelled. In some ows it may occur that the FDES term switches to DES in the boundary layer because z is too small (smaller than the boundary layer thickness, ). Different proposals have been made [126, 127] to protect the boundary layer from the LES mode FDES = max Lt (1 FS ), 1 CDES

where FS is taken as F1 or F2 (see Eq. 19.5) of the SST model.

20 Hybrid LES-RANS
When simulating bluff body ows, LES (Large Eddy Simulation) is the ideal method. Bluff body ows are dominated by large turbulent scales that can be resolved by LES

20. Hybrid LES-RANS

148

without too ne a resolution and accurate results can thus be obtained at an affordable cost. On the other hand, it is a challenging task to make accurate predictions of wallbounded ows with LES. The near-wall grid spacing should be about one wall unit in the wall-normal direction. This is similar to the requirement in RANS using low-Re number models. The resolution requirements in wall-parallel planes for a well-resolved LES in the near-wall region expressed in wall units are approximately 100 (streamwise) and 30 (spanwise). This enables resolution of the near-wall turbulent structures in the viscous sub-layer and the buffer layer consisting of high-speed in-rushes and low-speed ejections [100], often called the streak process. An event of a high-speed in-rush is illustrated in Fig. 20.1. In the lower part of the gure the spanwise vortex line is shown. Initially it is a straight line, but due to a disturbance e.g. a turbulent uctuation the mid-part of the vortex line is somewhat lifted up away from the wall. The mid-part of the vortex line experiences now a higher v1 velocity (denoted by U in the gure) than the remaining part of the vortex line. As a result the mid-part is lifted up even more and a tip of a hairpin vortex is formed. The vorticity of the legs lift each other through self-induction which helps lifting the tip even more. In the x1 x2 plane (upper part of Fig.. 20.1) the instantaneous and mean velocity proles (denoted by U and U in the gure, respectively) are shown as the hairpin vortex is created. It can be seen that an inexion point is created in the instantaneous velocity prole, U , and the momentum decit in the inner layer increases for increasing x1 . Eventually the momentum decit becomes too large and the highspeed uid rushes in compensating for the momentum decit. The in-rush event is also called a sweep. There are also events which occurs in the other direction, i.e. lowspeed uid is ejected away from the wall. These events are called bursts or ejections. The spanwise separation between sweeps and bursts is very small (approximately 100 viscous units, see Fig. 20.1). This is the main reason why the grid must be very ne in the spanwise direction. The streamwise distance between the events is related to the boundary layer thickness (4, see Fig. 20.1). The process by which the events are formed is similar to the later stage in the transition process from laminar to turbulent ow. Figure 20.2 presents the instantaneous eld of the streamwise velocity uctuation, v1 in the viscous wall region. As can be seen, the turbulent structures very elongated in the streamwise direction. At low to medium Reynolds numbers the streak process is responsible for the major part of the turbulence production. These structures must be resolved in an LES in order to achieve accurate results. Thus, for wall-bounded ows at high Reynolds numbers of engineering interest, the computational resource requirement of accurate LES is prohibitively large. Indeed, the requirement of near-wall grid resolution is the main reason why LES is too expensive for engineering ows, which was one of the lessons learned in the LESFOIL project [104, 105]. The object of hybrid LES-RANS (and of DES) is to eliminate the requirement of high near-wall resolution in wall-parallel planes. In the near-wall region (the URANS region), a low-Re number RANS turbulence model (usually an eddy-viscosity model) is used. In the outer region (the LES region), the usual LES is used, see Fig. 20.3. The idea is that the effect of the near-wall turbulent structures should be modelled by the RANS turbulence model rather than being resolved. In the LES region, coarser grid spacing in wall-parallel planes can be used. The grid resolution in this region is presumably dictated by the requirement of resolving the largest turbulent scales in the ow (which are related to the outer length scales, e.g. the boundary layer thickness) rather than the near-wall turbulent processes. The unsteady momentum equations are solved throughout the computational domain. The turbulent RANS viscosity is used in

20. Hybrid LES-RANS

149

Figure 20.1: Illustration of near-wall turbulence (taken from [57]).

x3

x1 Figure 20.2: Fluctuating streamwise velocity in a wall-parallel plane at x+ = 5. DNS 2 of channel ow [73]. wall URANS region

LES region x2 URANS region wall x1 Figure 20.3: The LES and URANS region. x+ 2,ml

20. Hybrid LES-RANS


1 30

150

20

2 B11 (1 )/v1,rms x
1 2 3

25

0.8

0.6

v1 15 +
10 5 0

0.4

0.2

10

x+ 2

10

10

0 0

0.5

x1

1.5

(a) Velocity proles

(b) Two-point correlation

Figure 20.4: Comparison of standard hybrid LES-RANS in channel ow on a very : hybrid coarse mesh (x+ = 2x+ = 785. /x1 2.5, /x3 5.) [73]. 1 3 LES-RANS; : 0.4 ln(y + ) + 5.2. Markers in right gure indicate resolution. interface
v1,f , v2,f , v3,f

LES region

URANS region x2 wall x1 Figure 20.5: Using forcing at the interface between the LES and URANS region. the URANS region, and the turbulent SGS viscosity is used in the LES region. Much work on hybrid LES-RANS has been carried out. In [128, 70, 129] twoequation models were used in the URANS region and a one-equation SGS model was employed in the LES region. One-equation models were used in both regions in [130, 131]. The locations of the matching planes were determined in different ways. In some work [70, 129] it was chosen along a pre-selected grid plane. In [130] it was determined by comparing the URANS and the LES turbulent length scales or was computed from turbulence/physics requirements. In [128] they used a two-equation model in the URANS region and blended it into a one-equation model in the LES region. Different partial differential equations for automatically nding the matching plane were investigated in [131]. A one-equation model was used in both regions in [132], and the c coefcient at the interface was computed dynamically to yield a smoother transition between the URANS and LES regions. In [133] they proposed a k turbulence model, later also used by [134], in which the c2 is made into a function of the ratio of the RANS and LES length scales. On a ne mesh the model switches smoothly to LES x+ 2,ml

20. Hybrid LES-RANS

151

j + 1/2
v2,f v1,f

Control volume j

LES region Interface j 1/2 URANS region An

Figure 20.6: Added uctuations, vf,1 , vf,2 , vf,3 , in a control volume (j = jml + 1) in the LES region adjacent to the interface. The uctuations are either synthesized (subscript f = S) or taken from channel DNS (subscript f = DN S).

and in the limit c1 = c2 so that a pure DNS solution is obtained. Hybrid LES-RANS is similar to DES (Detached Eddy Simulations) [111, 135, 127]. The main difference is that the original DES aims at covering the whole attached boundary layer with URANS, whereas hybrid LES-RANS aims at covering only the inner part of the boundary layer with URANS. In later work DES has been used as a wall model [136, 123], and, in this form, DES is similar hybrid LES-RANS. Figure 20.4a presents comparison of LES and hybrid LES-RANS in channel ow at Re = 2000 on a very coarse mesh. The momentum equations are solved in the entire domain and the turbulent viscosity is in both regions obtained from a one-equations ksgs equation and an algebraic length scale (see Sections 20.1 and 20.2). The resolution in the wall-parallel plane is comparable to what can be afforded for boundary layer in real, industrial ows, at least in terms of viscous units (x+ and x+ ). The LES 1 3 cannot resolve the ow at all. Hybrid LES-RANS gives much improved results, still not very good however. The normalized streamwise two-point correlation is shown in Fig. 20.4b. As can be seen, the streamwise lengthscale predicted with hybrid LESRANS is extremely large. It should be mentioned that standard hybrid LES-RANS does of course give better results on ner grids [101], but these ner grids are rarely affordable in industrial ows. Although the results obtained with hybrid LES-RANS are better than those obtained with LES, it has been found that the treatment of the interface between the URANS region and the LES region is crucial for the success of the method. The resolved turbulence supplied by the URANS region to the LES region has no reasonable turbulent characteristics and is not appropriate for triggering the LES equations to resolve turbulence. This results in too poorly resolved stresses in the interface region and thereby gives a ramp also referred to as a shift in the velocity prole approximately at the location of the matching plane [70, 136, 124, 129, 137, 130, 123]. The overly small resolved stresses in the LES region are translated into too small a wall shear stress. Several modications have been proposed to remove this deciency. In [137, 132], they suggested dampening the modelled stresses in the URANS region to reduce the total (i.e. resolved plus modelled) shear stress in the URANS region and thereby reduce the jump in shear stress across the matching plane. Numerical smoothing was used at the interface in [130]. [129] proposed a modication of the discretized streamwise equation at the interface in order to avoid ltering out any resolved uctu-

20.1. Momentum equations in hybrid LES-RANS


URANS region T C
3/4 c n[1 exp(0.2k 1/2 n/)] 1/4 c k 1/2 n[1 exp(0.014k 1/2n/)]

152 LES region = 0.07k 1/2 1.05

1.0

Table 20.1: Turbulent viscosity and turbulent length scales in the URANS and LES regions. n and denote the distance to the nearest wall and von K rm n constant a a (= 0.41), respectively. = (V )1/3 . ations at the interface. In [123] backscatter was introduced in the interface region with the object of generating resolved uctuations. One way to improve hybrid LES-RANS is to add uctuations to the momentum equations at the interface [124, 73], see Figs. 20.5 and 20.6. The object is to trigger the equations to resolve turbulence. Adding uctuations in order to trigger the equations to resolve turbulence is actually very similar to prescribing uctuating turbulent inlet boundary conditions for DNS or LES (or hybrid LES-RANS). If no triggering inlet boundary conditions are prescribed in DNS or LES, the resolved turbulence near the inlet will be too small and a large streamwise distance is required before the equations trigger themselves into describing turbulent ow. This is also the case in hybrid LESRANS: if no triggering (forcing) is applied at the interface between the LES region and the URANS region, the resolved turbulence in the LES region near the URANS region will be too small.

20.1 Momentum equations in hybrid LES-RANS


The incompressible Navier-Stokes equations with an added turbulent/SGS viscosity read i v i v 1 p ( + T ) (20.1) (i vj ) = v + + t xj xi xj xj where T = t (t denotes the turbulent RANS viscosity) for x2 x2,ml (see Fig. 20.3) and, for x2 > x2,ml , T = sgs . The turbulent viscosity, T , is computed from an algebraic turbulent length scale (see Table 20.1) and kT ; the latter is obtained by solving its transport equation, see Eq. 20.2.

20.2 The equation for turbulent kinetic energy in hybrid LES-RANS


A one-equation model is employed in both the URANS region and the LES region, which reads kT kT ( + T ) (j kT ) = v + t xj xj xj PkT = ij sij , + PkT C kT
3/2

(20.2)

ij = 2T sij

In the inner region (x2 x2,ml ) kT corresponds to the RANS turbulent kinetic energy, k; in the outer region (x2 > x2,ml ) it corresponds to the subgrid-scale kinetic turbulent energy (ksgs ). No special treatment is used in the equations at the matching plane except that the form of the turbulent viscosity and the turbulent length scale are different in the two regions, see Table 20.1. At the walls, kT = 0.

20.3. Results

153

30 25

30 25 20 15 10 5
0 2

v /u,w

20 15 10 5 0

10

x+ 2

10

10

x+ 2

10

(a) Solid lines: isotropic forcing, MS = 0.25; dashdotted lines: isotropic forcing, MS = 1; dashed lines: no forcing; +: present 963 DNS.

(b) DNS forcing. Solid line: MDNS = 0.25; dashed line: MDNS = 0.5.

Figure 20.7: Streamwise velocities [73]. v proles. : 2.5 ln(x+ ) + 5.2. 2 + 12 /u2 ,w kT /u2 ,w
0.5vi vi /u2 , ,w
1

7 6 5 4 3 2 1 0 0 0.2 0.4 0.6 0.8 1

0.8

0.6

v1 v2 /u2 , ,w

0.4

0.2

0 0

0.2

0.4

x2

0.6

0.8

x2

(a) Shear stresses. denotes viscous stress.

(b) Turbulent kinetic energy.

Figure 20.8: Shear stress and turbulent kinetic energy [73]. Solid lines: no forcing; dashed lines: forcing with isotropic uctuations with MS = 0.25; : present 963 DNS. Thick lines: resolved; thin lines: modelled.

20.3 Results
Fully developed channel ow at Re = u / = 2000 ( denotes the channel half width) is used as a test case to evaluate the effect of different forcing conditions. This ow may seem to be an easy test case, but it is not. In attempts to improve the performance of LES in wall-bounded ows, the Achilles heel is the near-wall ow region. The bulk velocity in fully developed channel ow with periodic boundary conditions (see Eq. 20.1) is entirely determined by the wall shear stress; consequently the ow is extremely sensitive to the turbulence in the near-wall region. The streamwise velocity proles obtained with and without forcing are compared in Fig. 20.7 with the present DNS and the log-law. It can be seen that the centerline velocity is strongly over-predicted when no forcing is used, whereas forcing with MS = MDN S = 0.25 gives excellent agreement with the log-law (MS and MDN S

21. The SAS model

154

denote forcing with synthetic and DNS uctuations, respectively). The reason for the overly large velocities without forcing is that the resolved shear is too small. It can be seen in Fig. 20.8a that it is the resolved shear stress that increases when forcing is introduced, indicating that the resolved shear stress without forcing is too small. This was also observed by [123]: when forcing is introduced, the resolved shear stress increases, which reduces the bulk and centerline velocity. Recently a novel way for generating uctuations to be used as forcing at the interface was presented [101]. In this work backscatter obtained from a scale-similarity model was used.

21 The SAS model


21.1 Resolved motions in unsteady
When doing URANS or DES, the momentum equations are triggered through instabilities to go unsteady in regions where the grid is ne enough. In URANS or in DES operating in RANS mode, high turbulent viscosity often dampens out these instabilities. In many cases this is an undesired feature, because if the ow wants to go unsteady, it is usually a bad idea to force the equations to stay steady. One reason is that there may not be any steady solution. Hence, the equations will not converge. Another reason is that if the numerical solution wants to go unsteady, the large turbulent scales i.e. part of the turbulent spectrum will be resolved instead of being modelled. This leads to a more accurate prediction of the ow. One way to improve a RANS models ability to resolve large-scale motions is to use the SAS (Scale- Adaptive Simulation) model

21.2 The von K rm n length scale a a


The von K rm n length scale a a LvK,1D = v /x2 2 v /x2 2 (21.1)

which includes the second velocity gradient is a suitable length scale for detecting unsteadiness. The von K rm n length scale is smaller for an instantaneous velocity a a prole than for a time averaged velocity, see Fig. 21.1. This is interesting because, as noted in [138], the von K rm n length scale decreases when the momentum equations a a resolve (part of) the turbulence spectrum. The rst and second derivatives in Eq. 21.1 are given in boundary layer form. We want to extend this expression to a general one, applicable in three dimensions. In the same way as in, for example, the Smagorinsky model, we take the rst derivative as || = (2ij sij )1/2 . The second derivative can be generalized in a number of ways. In s s the SAS model it is taken as U = 2 vi 2 vi xj xj xk xk
0.5

(21.2)

Hence, the general three-dimensional expression for the von K rm n length scale reads a a LvK,3D = || s |U | (21.3)

21.2. The von K rm n length scale a a

155

In [139] they derived a one-equation t turbulence model where the von K rm n a a length scale was used. The model was called the SAS model. Later, based on the k k 1/2 L model of Rotta [140], Menter & Egorov [138] derived a new k kL model using the von K rm n length scale. Finally, in [141] they modied the k -SST a a model to include the SAS features; they called this model the SST-SAS model. This model is described in more detail below.

The SST-SAS model


The k SST model is given in Eq. 19.5 at p. 147 (see also the section starting at p. 114) Now, Menter & Egorov [141] introduced a SAS-term in the equation. The object of this term is to decrease the turbulent viscosity when unsteadiness is detected, i.e. when the von K rm n length scale becomes small. The production term in the a a equation in the k-SST model reads P = P k /t ||2 . To decrease the turbulent s viscosity we should increase . Thus it seems reasonable to add a new production term proportional to P Lt /Lvk,3D where Lt denotes a RANS length scale. The additional term reads k 1/2 Lt s , Lt = 1/4 (21.4) 2 ||2 LvK,3D c When unsteadiness occurs i.e. when the momentum equations attempt to resolve part of the turbulence spectrum , this term reacts as follows: Local unsteadiness will create velocity gradients which decrease the turbulent length scale, see Fig. 21.1 This results in a decrease in the von K rm n length scale, LvK,3D a a As a consequence the additional source, Eq. 21.4, in the equation increases This gives an increase in and hence a decrease in t The decreased turbulent viscosity will allow the unsteadiness to stay alive and, perhaps, grow. The last item in the list above is the main object of the SAS model. The reaction to local unsteadiness in a eddy-viscosity model without the SAS feature is as follows: the increased local velocity gradients will create additional production of turbulent kinetic energy and give an increased turbulent viscosity which will dampen/kill the local unsteadiness. As mentioned in the introduction to this chapter, this is an undesirable feature. When incorporating the additional production term (Eq. 21.4) in the k -SST model, the last term in the equation is replaced by (for further details, see [141]) PSAS = FSAS max (T1 T2 , 0) L T1 = 2 S 2 LvK,3D 2k 1 1 k k T2 = max , 2 xj xj k 2 xj xj L= k 1/2 c
1/4

(21.5)

Note that the term T1 is the real additional SAS term; T2 is included to make sure that the model in steady ow works as a k SST model.

21.3. The second derivative of the velocity

156

0.8

x2

0.6

Lvk,1D
0.4 0.2

Lvk,3D

0 0

10

v1

15

20

Figure 21.1: Velocity proles from a DNS of channel ow. Solid line: time-averaged velocity with length scale Lvk,1D , Eq. 21.1 ; dashed line: instantaneous velocity with length scale Lvk,3D , Eq. 21.3.

21.3 The second derivative of the velocity


To compute U in Eq. 21.2, we need to compute the second velocity gradients. In nite volume methods there are two main options for computing second derivatives. Option I: compute the rst derivatives at the faces v x2 and then =
j+1/2

vj+1 vj , x2 =
j

v x2

=
j1/2

vj vj1 x2

2v x2 2

(x2 )2 4 v vj+1 2vj + vj1 + (x2 )2 12 x4 2

Option II: compute the rst derivatives at the center v x2 and then =
j+1

vj+2 vj , 2x2 =
j

v x2

=
j1

vj vj2 2x2

2v x2 2

vj+2 2vj + vj2 (x2 )2 4 v + 2 4(x2 ) 3 x4 2

In [142], Option I was used unless otherwise stated.

21.4 Evaluation of the von K rm n length scale in channel ow a a


In Fig. 21.2 the turbulent length scale, LvK,3D , is evaluated using DNS data of fully developed channel ow. When using DNS data only viscous dissipation of resolved turbulence affects the equations. This implies that the smallest scales that can be resolved are related to the grid scale. The von K rm n length scale based on instantaneous vea a locities, LvK,3D , is presented in Fig. 21.2. For x2 > 0.2, its magnitude is close to x2 which conrms that the von K rm n length scale is related to the smallest resolva a able scales. Closer to the wall, LvK,3D increases slightly whereas x2 continues to decrease.

21.4. Evaluation of the von K rm n length scale in channel ow a a


x+ 2 x+ 2

157

0 0.3

100

200

300

400

500

0 0.1

10

15

20

25

0.08 0.25 0.2 0.15 0.1 0.02 0.05 0 0 0.2 0.4 0.6 0.8 1 0 0 0.01 0.02 0.03 0.04 0.05 0.06

0.04

x2 /

x2 /

Figure 21.2: Turbulent length scales in fully developed channel ow. Left: global view; right: zoom. DNS. 963 mesh. Re = 500. x1 / = 0.065, x3 / = 0.016, x2 -stretching of 9%. : LvK,3D ; : LvK,1D ; : (x1 x2 x3 )1/3 ; : x2 .
0 500

x+ 2

1500

2000

0 0.1

20

40

x+ 2

60

80

100

0.25 0.2

0.08

0.06 0.15 0.04 0.1 0.05 0 0 0.02

0.2

0.4

0.6

0.8

0 0

0.01

0.02

0.03

0.04

0.05

x2 /

x2 /

Figure 21.3: Turbulent length scales in fully developed channel ow. Hybrid LESRANS. Left: global view; right: zoom. 32 64 32 mesh. Re = 2000. x1 / = 0.39, x3 / = 0.19, x2 -stretching of 17%. : LvK,3D ; : LvK,1D ; : 1/4 (x1 x2 x3 )1/3 ; : x2 ; +: k = k 0.5 /(c ). The von K rm n length scale, LvK,1D , based on the averaged velocity prole a a v1 = v1 (x2 ) is also included in Fig. 21.2, and as can be seen it is much larger than LvK,3D . Near the wall LvK,1D increases because the time-average second derivative, 2 v1 /x2 , goes to zero as the wall is approached. No such behavior is seen for the 2 three-dimensional formulation, LvK,3D . In Fig. 21.3, data from hybrid LES-RANS are used (taken from [73]). When using hybrid LES-RANS, part of the turbulence is resolved and part of the turbulence is modelled. The resolved turbulence is dissipated by a modelled dissipation, 2 T sij sij (T denotes SGS or RANS turbulent viscosity), and T . As a result, the length scale of the smallest resolved turbulence is larger in hybrid LES-RANS than in DNS. Close to the wall in the URANS region (x2 < 0.031), the resolved turbulence is

22. The PANS Model

158

dampened by the high turbulent viscosity, and as a results LvK,3D follows closely LvK,1D . The RANS turbulent length scale, k , from a 1D RANS simulation at Re = 2000 with the k SST model is also included in Fig. 21.3. In the inner region (x2 < 0.5), its behavior is close to that of the von K rm n length scale, LvK,1D . In the a a center region the RANS turbulent length scale continues to increase which is physically correct. However, the von K rm n length scale, LvK,1D , goes to zero because the a a velocity derivative goes to zero. Two lter scales are included in Figs. 21.2 and 21.3. In the DNS-simulations, x2 < (x1 x2 x3 )1/3 near the wall, whereas far from the wall x2 > (x1 x2 x3 )1/3 because of the stretching in the x2 direction and because of small x1 and x3 . In the hybrid simulations, it can be noted that the three-dimensional lter width is more that twice as large as the three-dimensional formulation of the von K rm n length scale, a a i.e. (x1 x3 x3 )1/3 > 2 LvK,3D . In [142], the SST-SAS model has been evaluated in channel ow, ow in an asymmetric diffuser and ow over an axi-symmetric hill.

22 The PANS Model


The PANS method uses the so-called partial averaging concept, which corresponds to a ltering operation for a portion of the uctuating scales [143]. For an instantaneous ow variable, F , we use f to denote the partially-averaged part, namely f = P(F ), where P denotes the partial-averaging operator. We consider incompressible ows. Applying the partial averaging to the governing equations gives i v =0 xi 1 p (i vj ) v i v = + + t xj xi xj i v ij xj (22.1)

(22.2)

where ij is the central second moment resulting from the partial averaging for the nonlinear terms, that is ij = (P(vi vj ) vi vj ), where vi indicates instantaneous ve locity components. This term is similar to the Reynolds stress tensor resulting from the Reynolds averaging in RANS or to the subgrid-scale (SGS) stress tensor after the spatial ltering in LES. For simplicity, we also use the terminology of Reynolds stresses for the term ij in Eq. 22.2. To close the system of the partially-averaged Navier-Stokes equations, as in RANS and LES, a model is needed for ij . In [143] they proposed using the conventional eddy viscosity concept so that ij = 2u sij , where sij is the strain-rate tensor of the computed ow and u is the PANS eddy viscosity. In order to formulate the PANS eddy viscosity, they dened in [143] another two quantities, the partially-averaged turbulent kinetic energy, k u and its dissipation rate u , 2 so that u = C k u /u . In the derivation of the transport equations for k u and u , two parameters, fk and f , have been introduced, relating the unresolved to the resolved uctuating scales. Parameter fk denes the ratio of unresolved (partially-averaged) turbulent kinetic energy (k u ) to the total kinetic energy (k), and f is the ratio between the unresolved (u ) and the total () dissipation rates. These give k= u ku and = fk f (22.3)

22. The PANS Model

159

The extent of the resolved part is now determined by fk and f . In [144, 143] they employed the standard k model as the base model. The k u equation is derived by multiplying the RANS k equation (Eq. 11.89) in the k model by fk , i.e. (for simplicity we omit the buoyancy term) fk k k + Vj t xj = fk P k + xj + t k k xj (22.4)

where Vi denotes the RANS velocity. The left side can be re-written fk k k + Vj t xj =
u u u u k u k = k + vj k + (Vj vj ) k + Vj t xj t xj xj

(22.5)

The convective term must be expressed in vj (the PANS averaged velocity) rather than in Vj (the RANS averaged velocity), because it is vj that transports k u because vj represents the PANS resolved part of vj . The last term on the right side in Eq. 22.5 is usually neglected. The diffusion term is re-written using Eq. 22.3 fk xj + t k k xj xj = xj =
2 fk f

t k u + ku +

k u xj k u xj

(22.6)

where ku = k

(22.7)

The sum of the source terms in Eq. 22.4 must be equal to the sum of the source terms of the k u equation, i.e. (22.8) fk P k = P u u This relation implies Pk = u 1 (P u u ) + fk f (22.9)

Using Eqs. 22.5, 22.6 and 22.8 the nal transport equation for ku can now be written as k u (k u vj ) = + t xj xj + u ku k u + Pu u xj (22.10)

where the production term, Pu , is expressed in terms of the PANS eddy viscosity, u , and the strain rate of PANS-resolved ow eld, i.e. P u = u i v j v + xj xi
2

i v xj

ku u = c where

(22.11)

ku u = c

(22.12)

22. The PANS Model


The u equation is derived by multiplying the RANS equation by f , i.e. u (Vj ) (u vj ) = f + + t xj t xj t = f + xj

160

2 + C1 Pk C2 xj k k

(22.13)

The diffusion term is re-written using Eq. 22.3 f xj + t xj xj = xj =


2 fk f

t u + u +

u xj u xj

(22.14)

where ku = k

(22.15)

In the same way, the production and destruction terms are re-formulated as (using Eqs. 22.3 and 22.9) f C1 Pk 2 C2 k k
2

= + =

C1

u fk ku
2

1 u (Pu u ) + fk f
2

C2

u fk f k u (22.16)

u u = C1 u Pu C1 u k k

u fk u fk C1 u C2 u u k f f k C1
u u P C2 u u u k k
2

where
C2 = C1 +

fk (C2 C1 ) f

(22.17)

The u equation in the PANS model now takes the following form u (u vj ) = + t xj xj + u u
u u u + C1 Pu u C2 u xj k k
2

(22.18)

v As in the k u equation, the the additional term (Vj j )u /xj has been neglected. The PANS equation for k u , Eq. 22.10, was derived by multiplying the RANS equation for k by fk which was assumed to be constant in space and in time. By referring to Eqs. 22.6, 22.12 and 22.7, the turbulent diffusion term was obtained as fk xj t k k xj t k u xj k xj u k u = xj ku xj = (22.19a) (22.19b)

The expression on the right-hand side of Eq. 22.19(a) suggests that the turbulent transport for the PANS-modelled turbulent kinetic energy, k u , is actually formulated in terms of the RANS turbulent viscosity from the base model. This is different from the

23. Inlet boundary conditions

161

turbulent diffusion in subgrid scale (SGS) modelling of LES with a one-equation ksgs model, which reads sgs ksgs (22.20) xj k xj In Eq. 22.20 the SGS turbulent viscosity is invoked for the transport of ksgs , whereas on the right-hand side of Eq. 22.19(a) the total (i.e. the RANS) turbulent viscosity has been used for ku . Equation 22.19(a) suggests that, when used as an SGS model, the 2 modelled turbulent diffusion in the PANS formulation is a factor of k /ku = f /fk larger than in Eq. 22.20, see Eqs. 22.10 and 22.19(b). With f = 1 and fk = 0.4, for example, this factor is larger than six. The modication of the diffusion coefcient, ku , is a unique property of the PANS model. In other models, such as DES [145], X-LES [115] and PITM [120], the sink term in the k, or equation is modied, but not the diffusion term. A Low Reynolds number PANS model was presented in [118].

23 Inlet boundary conditions


In RANS it is sufcient to supply proles of the mean quantities such as velocity and temperature plus the turbulent quantities (e.g. k and ). However, in unsteady simulations (LES, URANS, DES . . . ) the time history of the velocity and temperature need to be prescribed; the time history corresponds to turbulent, resolved uctuations. In some ows it is critical to prescribe reasonable turbulent uctuations, but in many ows it seems to be sufcient to prescribe constant (in time) proles [103, 146]. There are different ways to create turbulent inlet boundary conditions. One way is to use a pre-cursor DNS or well resolved LES of channel ow. This method is limited to fairly low Reynolds numbers and it is difcult (or impossible) to re-scale the DNS uctuations to higher Reynolds numbers. Another method based partly on synthesized uctuations is the vortex method [147]. It is based on a superposition of coherent eddies where each eddy is described by a shape function that is localized in space. The eddies are generated randomly in the inow plane and then convected through it. The method is able to reproduce rst and second-order statistics as well as two-point correlations. A third method is to take resolved uctuations at a plane downstream of the inlet plane, re-scale them and use them as inlet uctuations. Below we present a method of generating synthesized inlet uctuations.

23.1 Synthesized turbulence


The method described below was developed by [148, 149, 73] for creating turbulence for generating noise. It was later further developed for inlet boundary conditions by [150, 151, 152]. A turbulent uctuating velocity uctuating eld (whose average is zero) can be expressed using a Fourier series, see Eq. C.17. Let us re-write this formula as an cos(nx) + bn sin(nx) = cn cos(n ) cos(nx) + cn sin(n ) sin(nx) = cn cos(nx n ) (23.1)

where an = cn cos() , bn = cn sin(n ). The new coefcient, cn , and the phase angle,

23.2. Random angles


n , are related to an and bn as cn = a 2 + b 2 n n
1/2

162

n = arctan

bn an

(23.2)

A general form for a turbulent velocity eld is given by N random Fourier modes as v (x) = 2
n=1 N

un cos(n x + n ) n

(23.3)

n where un , n and i are the amplitude, phase and direction of Fourier mode n. The synthesized turbulence at one time step is generated as follows.

23.2 Random angles


The angles n and n determine the direction of the wavenumber vector , see Eq. 23.3 and Eq. 23.1; n denotes the direction of the velocity vector, v . For more details, see Appendix H.

23.3 Highest wave number


Dene the highest wave number based on mesh resolution max = 2/(2), where is the grid spacing. The uctuations are generated on a grid with equidistant spacing (or on a weakly stretched mesh), = x2,max /N2 , x3 = x3,max /N3 , where denotes the wall-normal direction and N2 and N3 denote the number of cells in the x2 and x3 direction, respectively. The uctuations are set to zero at the wall and are then interpolated to the inlet plane of the CFD grid (the x2 x3 plane).

23.4 Smallest wave number


Dene the smallest wave number from 1 = e /p, where e = 9/(55Lt), = 1.453. The turbulent length scale, Lt , may be estimated in the same way as in RANS simulations, i.e. Lt where denotes the inlet boundary layer thickness. In [150, 151, 152] it was found that Lt 0.1in is suitable. Factor p should be larger than one to make the largest scales larger than those corresponding to e . A value p = 2 is suitable.

23.5 Divide the wave number range


Divide the wavenumber space, max 1 , into N modes, equally large, of size .

23.6 von K rm n spectrum a a


A modied von K rm n spectrum is chosen, see Eq. 23.4 and Fig. 23.2. The amplitude a a un of each mode in Eq. 23.3 is then obtained from un = (E())1/2 E() = cE
2 (/e )4 u2 rms e[2(/ ) ] e [1 + (/e )2 ]17/6

(23.4)

= (i i )1/2 ,

= 1/4 3/4

23.7. Computing the uctuations

163

n 2

n i

x3
n 3 n 1

n
n 3 n 1

n n i
0

x2

x1

n Figure 23.1: The wave-number vector, n , and the velocity unit vector, i , are orthogi onal (in physical space) for each wave number n.

The coefcient cE is obtained by integrating the energy spectrum over all wavenumbers to get the turbulent kinetic energy, i.e.

k=
0

E()d

(23.5)

which gives [57] 4 (17/6) cE = 1.453 (1/3) where (z) =


0

(23.6)

ez xz1 dz

(23.7)

23.7 Computing the uctuations


n Having un , n , i and n , allows the expression in Eq. 23.3 to be computed, i.e. j N v1 = 2 n=1 N v2 = 2 n=1 N v3 = 2 n=1 n n n n = k1 x1 + k2 x2 + k3 x3 + n

un cos( n )1 un cos( n )2 un cos( n )3

(23.8)

23.8. Introducing time correlation

164

E() 1 E(n )

e E() 5/3 n

n
Figure 23.2: Modied von K rm n spectrum a a

log()

where un is computed from Eq. 23.4. In this way inlet uctuating velocity elds (v1 , v2 , v3 ) are created at the inlet x2 x3 plane. The code for generating the isotropic uctuations can be downloaded here http://www.tfd.chalmers.se/lada/projects/inlet-boundary-conditions/proright.ht

23.8 Introducing time correlation


A uctuating velocity eld is generated each time step as described above. They are independent of each other and their time correlation will thus be zero. This is unphysical. To create correlation in time, new uctuating velocity elds, V1 , V2 , V3 , are computed based on an asymmetric time lter
(V1 )m = a(V1 )m1 + b(v1 )m (V2 )m = a(V2 )m1 + b(v2 )m (V3 )m

(23.9)

a(V3 )m1

b(v3 )m

where m denotes the time step number and a = exp(t/T ). 2 2 The second coefcient is taken as b = (1 a2 )0.5 which ensures that V1 = v1 ( denotes averaging). The time correlation of will be equal to exp(t/T ) (23.10)

and thus Eq. 23.9 is a convenient way to prescribe the turbulent time scale of the uctuations. The inlet boundary conditions are prescribed as (we assume that the inlet is located at x1 = 0 and that the mean velocity is constant in the spanwise direction, x3 ) v1 (0, x2 , x3 , t) = V1,in (x2 ) + u (x2 , x3 , t) 1,in
v2 (0, x2 , x3 , t) = V2,in (x2 ) + v2,in (x2 , x3 , t)

(23.11)

v3 (0, x2 , x3 , t) = V3,in (x2 ) +

v3,in (x2 , x3 , t)

23.8. Introducing time correlation

165

1 0.8 0.6

B( )

0.4 0.2 0 0 0.2 0.4

0.6

0.8

Figure 23.3: Auto correlation, B( ) = v1 (t)v1 (t )t (averaged over time, t). Eq. 23.10; : computed from synthetic data, (V1 )m , see Eq. 23.9.

where = 0.4 and B = 5.2. The method to prescribed uctuating inlet boundary conditions have been used for channel ow [152], for diffuser ow [146] as well as for the ow over a bump and an axisymmetric hill [154].

where v1,in = (V1 )m , v2,in = (V2 )m and v3,in = (V3 )m (see Eq. 23.9). The mean inlet proles, V1,in , V2,in , V3,in , are either taken from experimental data, a RANS solution or from the law of the wall; for example, if V2,in = V3,in = 0 we can estimate V1,in as [153] + x+ 5 x2 2 + + (23.12) V1,in = 3.05 + 5 ln(x2 ) 5 < x+ < 30 2 1 ln(x+ ) + B x+ 30 2 2

24. Best practice guidelines (BPG)

166

24 Best practice guidelines (BPG)


In the early days of CFD, different CFD codes used to give different results. Even if the same grid and the same turbulence model were used, there could be substantial differences between the results. The reasons to these differences could be that the turbulence model was not implemented in exactly the same way in the two codes, or that the discretization scheme in one code was more diffusive than in the other. There could be small differences in the implementation of the boundary conditions in the two codes. Today the situation is much improved. Two different CFD codes usually give the same results on the same grid. The main reason for this improved situation is because of workshops and EU projects where academics, engineers from industry and CFD software vendors regularly meet and discuss different aspects of CFD. Test cases with mandatory grids, boundary conditions, turbulence models etc are dened and the participants in the workshops and EU projects carry out CFD simulations for these test cases. Then they compare and discuss their results.

24.1 EU projects
Four EU projects in which the author has taken part can be mentioned LESFOIL: Large Eddy Simulation of Flow Around Airfoils http://www.tfd.chalmers.se/lada/projects/lesfoil/proright.html FLOMANIA: Flow Physics Modelling: An Integrated Approach http://cfd.mace.manchester.ac.uk/omania/ DESIDER: Detached Eddy Simulation for Industrial Aerodynamics http://cfd.mace.manchester.ac.uk/desider ATAAC: Advanced Turbulence Simulation for Aerodynamic Application Challenges http://cfd.mace.manchester.ac.uk/ATAAC/WebHome

24.2 Ercoftac workshops


Workshops are organized by Ercoftac (European Research Community On Flow, Turbulence And Combustion). The Special Interest Group Sig15 is focused on evaluating turbulence models. The outcome from all workshop are presented here http://www.ercoftac.org/leadmin/user upload/bigles/sig15/database/index.html Ercoftac also organizes workshops and courses on Best Practice Guidelines. The publication Industrial Computational Fluid Dynamics of Single-Phase Flows can be ordered on Ercoftac www page http://www.ercoftac.org/publications/ercoftac best practice guidelines/single-phase ows spf/

24.3. Ercoftac Classical Database

167

24.3 Ercoftac Classical Database


A Classical Database, which includes some 100 experimental investigations, can be found at Ercoftacs www page http://www.ercoftac.org/products and services/classic collection database

24.4 ERCOFTAC QNET Knowledge Base Wiki


The QNET is also the responsibility of Ercoftac. Here you nd descriptions of how CFD simulations of more than 60 different ows were carried out. The ows are divided into Application Areas. These are sector disciplines such as Built Environment, Chemical and Process Engineering, External Aerodynamics, Turbomachinery, Combustion and Heat Transfer etc. Each Application Area is comprised of Application Challenges. These are realistic industrial test cases which can be used to judge the competency and limitations of CFD for a given Application Area. Underlying Flow Regimes. These are generic, well-studied test cases capturing important elements of the key ow physics encountered across the Application Areas. For more information, visit ERCOFTAC QNET Knowledge Base Wiki http://www.ercoftac.org/products and services/wiki/

A. TME225: identity

168

TME225: identity
inm mjk = min mjk = nmi mjk = ij nk ik nj In Table A.1 the components of the identity are given.

The identity reads

i 1 2 1 1 3 1 2 3 2

n 2 1 2 3 1 3 3 2 3

j 1 1 2 1 1 3 2 2 3

k 2 2 1 3 3 1 3 3 2

inm mjk 12m m12 = 123 312 = 1 1 = 1 21m m12 = 213 312 = 1 1 = 1 12m m21 = 123 321 = 1 1 = 1 13m m13 = 132 213 = 1 1 = 1 31m m13 = 312 213 = 1 1 = 1 13m m31 = 132 231 = 1 1 = 1 23m m23 = 231 123 = 1 1 = 1 32m m23 = 321 123 = 1 1 = 1 23m m32 = 231 132 = 1 1 = 1

ij nk ik nj 10=1 0 1 = 1 0 1 = 1 10=1 0 1 = 1 0 1 = 1 10=1 0 1 = 1 0 1 = 1

Table A.1: The components of the identity which are non-zero.

B. TME225 Assignment 1: laminar ow

169

V x2 x1 L Figure B.1: Flow between two plates (not to scale). h

B TME225 Assignment 1: laminar ow


You will get results of a developing two-dimensional channel ow (i.e. ow between two parallel plates), see Fig. B.1. The ow is steady and incompressible. The simulations have been done with Calc-BFC [155]. The inlet boundary condition (left boundary) is v1 = Vin = 0.7. The height of the channel is h = 0.011m and L = 0.6385m; the uid is air of 20o C. You will use Matlab to analyze the data. First, nd out and write down the governing equations (N.B:. you cannot assume that the ow is fully developed). From the course www page http://www.tfd.chalmers.se/lada/MoF/, download the data le channel flow data.dat and the m-le channel flow.m which reads the data and plot some results. Open Matlab and execute channel flow. Open channel flow.m in an editor and make sure that you understand it. There are three eld variables, v1 , v2 and p; the corresponding Matlab arrays are v1 2d, v2 2d and p 2d. The grid is 199 22, i.e. ni = 199 grid points in the x1 direction and nj = 22 grid points in the x2 direction. The eld variables are stored at these grid points. We denote the rst index as i and the second index as j, i.e. v1 2d(i,j). Hence in v1 2d(:,1) are the v1 values at the lower wall; v1 2d(:,nj) are the v1 values at the upper wall; v1 2d(1,:) are the v1 values at the inlet; v1 2d(ni,:) are the v1 values at the outlet; The work should be carried out in groups of two (you may also do it on your own, but we dont recommend it). At the end of this Assignment the group should write and submit a report (in English). Divide the report into sections corresponding to the sections B.1 B.9. In some sections you need to make derivations; these should clearly be described and presented. Present the results in each section with a gure (or a numerical value). The results should also be discussed and as far as you can explained.

B.1

Fully developed region

Fully developed conditions mean that the ow does not change in the streamwise direction, i.e. v1 /x1 = 0. If we dene fully developed as the location where the

B.2. Wall shear stress

170

velocity gradient in the center becomes smaller than 0.01, i.e. |v1 /x1 | < 0.01, how long distance from the inlet does the ow become fully developed? Another way to dene fully developed conditions can be the x1 position where the centerline velocity has reached, for example, 99% of its nal value. What x1 value do you get? In Section 3.2.2, a distance taken from the literature is given. How well does this agree with your values? In the fully developed region, compare the velocity prole with the analytical prole (see Section 3.2.2). Look at the vertical velocity component, v2 . What value should it take in the fully developed region (see Section 3.2.2)? What value does it take (at x2 = h/4, for example)?

B.2

Wall shear stress

On the lower wall, the wall shear stress, w,L (index L denotes Lower), is computed as v1 (B.1) x2 L Recall that 12 = (v1 /x2 + v2 /x1 ) (see Eqs. 2.4 and 1.4) but at the wall v2 /x1 = 0; Skk = 0 because of the continuity equation, Eq. 2.3. Plot w,L versus x1 . Why does it behave as it does? Now we will compute the wall shear stress at the upper wall, w,U . If you use Eq. B.1, you get the incorrect sign. Instead, use Cauchys formula (see [2], Chapt. 4.2) w,L 21,w,L = ti
( ) n

= ji nj

(B.2)

which is a general way to compute the stress vector on a surface whose (outward point ing) normal vector is n = nj . The expression for ij can be found in Eqs. 1.4 and 2.4; recall that the ow in incompressible. On the top wall, the normal vector points out from the surface (i.e. nj = (0, 1, 0)). Use Eq. B.2 to compute the wall shear stress at the upper wall. Plot the two wall shear stresses in the same gure. How do they compare? In the fully developed region, compare with the analytical value (see Eq. 3.30).

B.3

Inlet region

In the inlet region the ow is developing from its inlet prole (v1 = V = 0.7) to the fully developed prole somewhere downstream. The v1 velocity is decelerated in the near-wall regions, and hence the v1 velocity in the center must increase due to continuity. Plot v1 in the center and near the wall as a function of x1 . Plot also v1 /x1 . If you, for a xed x1 , integrate v1 , i.e.
h

(x1 ) =
0

v1 (x1 , x2 )dx2

what do you get? How does (x1 ) vary in the x1 direction? How should it vary?

B.4

Wall-normal velocity in the developing region

In Section B.3 we found that, in the developing region, v1 near the walls decreases for increasing x1 . What about v2 ? How do you explain the behaviour of v2 ?

B.5. Vorticity

171

B.5

Vorticity

Do you expect the ow to be irrotational anywhere? Lets nd out by computing the vorticity. Plot it in the fully developed region as 3 vs. x2 . Where is it largest? Plot the vorticity also in the inlet and developing regions; what happens with the vorticity in the inlet region? Now, is the ow rotational anywhere? Why? Why not?

B.6

Deformation

In Section 1.4, we divided the velocity gradient into a strain-rate tensor, Sij , and a vorticity tensor, ij . Since the ow is two-dimensional, we have only two off-diagonal terms (which ones?). Plot and compare one of the off-diagonal term of Sij and ij . Where are they largest? Why? What is the physical meaning of Sij and ij , respectively? Compare ij with the vorticity you plotted in Section B.5. Are they similar? Any comment?

B.7

Dissipation

Compute and plot the dissipation, = ji vi /xj . What is the physical meaning of the dissipation? Where do you expect it to be largest? Where is it largest? Any difference it its behaviour in the inlet region compared to in the fully developed region? The dissipation appears as a source term in the equation for internal energy, see Eq. 2.9. This means that dissipation increases the internal energy, i.e. the temperature. This is discussed in some detail at p. 21. Use Eq. 2.14 to compute the temperature increase that is created by the ow (i.e. by dissipation). Start by integrating the dissipation over the entire computational domain. Next, re-write the left side on conservative form and then apply the Gauss divergence theorem. Assume that the upper and lower walls are adiabatic; furthermore we can neglect the heat ux by conduction, q1 , (see Eq. 2.11) at the inlet and outlet. Now you can compute the increase in bulk temperature, Tb , from inlet to outlet. The bulk temperature is dened at h v1 T dx2 Tb = 0 h 0 v1 dx2

B.8

Eigenvalues

Compute and plot the eigenvalues of the viscous stress tensor, ij . Use the Matlab command eig. If you have computed the four elements of the ij matrix you can use the following commands: tau=[tau_11 tau 12; tau_21 tau_22]; [n,lambda]=eig(tau); where n and lambda denote eigenvalues and eigenvectors, respectively. What is the physical meaning of the eigenvalues (see Chapter 1.6)? Pick an x1 location where the ow is fully developed. Plot one eigenvalue as a x y graph (eigenvalue versus x2 ). Plot also the four stress components, ij , versus x2 . Is (Are) anyone(s) negligible? How does the largest component of ij compare with the largest eigenvalue? Any thoughts? And again: what is the physical meaning of the eigenvalues?

B.9. Eigenvectors

172

B.9

Eigenvectors

Compute and plot the eigenvectors of ij . Recall that at each point you will get two eigenvectors, perpendicular to each other. It is enough to plot one of them. An eigenvector is, of course, a vector. Use the Matlab command quiver to plot the eld of the eigenvectors. Recall that the sign of the eigenvector is not dened (for example, both v1 and 1 in Fig. 1.10 at p. 18 are eigenvectors). Try to analyze why the eigenvectors v behave as they do.

C. TME225: Fourier series


V2

173

(V2 |T)

(V1 |T) Figure C.1: Scalar product.

V1

TME225: Fourier series

Here a brief introduction to Fourier series extracted from [156] is given.

C.1 Orthogonal functions


Consider three vectors, V1 , V2 , V3 , in physical space which form an orthogonal base in R3 (i.e. their scalar products are zero). Let us call them basis functions. Any vector, T, in R3 can now be expressed in these three vectors, i.e. T = c1 V1 + c2 V2 + c3 V3 (C.1)

see Fig. C.1. Now dene the scalar product of two vectors, a and b, as a b = (a|b). The coordinates, ci , can be determined by making a scalar product of Eq. C.1 and Vi which gives (T|Vi ) = (c1 V1 |Vi ) + (c2 V2 |Vi ) + (c3 V3 |Vi ) = (c1 V1 |V1 ) + (c2 V2 |V2 ) + (c3 V3 |V3 ) = c1 |V1 | + c2 |V2 | + c3 |V3 | = ci |Vi |
2 2 2

(C.2)

where |Vi | denotes the length of Vi ; the second line follows because of the orthogonality of Vi . Hence the coordinates, ci , are determined by ci = (T|Vi )/|Vi |2 (C.3)

C.2. Trigonometric functions

174

Now let us dene an innite (-dimensional) functional space, B, with orthogonal basis functions {g} . The scalar product of two functions, f and gn , is dened as 1
b

(f |gn ) =

f (x)gn (x)dx
a

(C.4)

Then, in a similar way to Eq. C.1, any function can, over the interval [a, b], be expressed as

f=
n=1

cn g n

(C.5)

As above, we must now nd the coordinates (cf. the coordinates, ci , in Eq. C.1). Multiply, as in Eq. C.2, f with the basis functions, gi , i.e.

(f |gi ) =

n=1

cn (gn |gi )

(C.6)

Since we know that all gn are orthogonal, Eq. C.6 is non-zero only if i = n, i.e. (f |gi ) = (c1 g1 |gi ) + (c2 g2 |gi ) . . . ci (gi |gi ) . . . ci+1 (gi+1 |gi ) . . . = = ci (gi |gi ) = ci ||gi ||2 (C.7)

Similar to Eq. C.3, the coordinates can be found from (switch from index i to n) cn = (f |gn )/||gn ||2 (C.8)

The coordinates, cn , are called the Fourier coefcients to f in system {g} and 1 ||gn || is the length of gn (cf. |Vi | which is the length of Vi in Eq. C.3), i.e. ||gn || = (gn |gn )
1/2 b 1/2

=
a

gn (x)gn (x)dx

(C.9)

Let us now summarize and compare the basis functions in physical space and the basis functions in functional space 1. Any vector in R3 can be expressed in the orthogonal basis vectors Vi 2. The length of the basis vector, Vi , is |Vi | 3. The coordinates of Vi are computed as ci = (T|Vi )/|Vi |2 1. Any function in [a, b] can be expressed in the orthogonal basis functions gn 2. The length of the basis function, gn , is ||gn || 3. The coordinates of gn are computed as cn = (f |gn )/||gn ||2

C.2 Trigonometric functions


Here we choose gn as trigonometric functions which are periodic in [, ]. The question is now how to choose the orthogonal function system {g} on the interval 1 [, ]. In mathematics, we usually start by doing an intelligent guess, and then we prove that it is correct. So let us guess that the trigonometric series [1, sin x, cos x, sin(2x), . . . , sin(nx), cos(nx), . . .] (C.10)

C.2. Trigonometric functions

175

is an orthogonal system. The function system in Eq. C.10 can be dened as gn (x) = k (x), k (x), for n = 2k = 2, 4, . . . for n = 2k + 1 = 1, 3, . . . (C.11)

where k (x) = sin(kx) (k = 1, 2, . . .) and k (x) = cos(kx) (k = 0, 1, . . .). Now we need to show that they are orthogonal, i.e. that the integral of the product of any two functions k and k is zero on B[, ] and we need to compute their length (i.e. their norm). Orthogonality of n and k

(n |k ) = =

cos(nx) cos(kx)dx =

1 2

[cos((n + k)x) + cos((n k)x)] dx

1 1 1 sin((n + k)x) + sin((n k)x) 2 n+k nk

= 0 for k = n

(C.12)

Length of k (k |k ) = ||k ||2 = (0 |0 ) = ||0 ||2 =


cos2 (kx)dx = 1 dx = 2

x 1 + sin(2x) 2 4

= for k > 0

(C.13) Orthogonality of n and k

(n |k ) =

sin(nx) cos(kx)dx =

1 2

[sin((n + k)x) + sin((n k)x)] dx =

1 1 1 cos((n + k)x) + cos((n k)x) 2 n+k nk

=0

(C.14)

Orthogonality of n and k

(n |k ) = =

sin(nx) sin(kx)dx =

1 2

[cos((n k)x) cos((n + k)x)] dx

1 1 1 sin((n k)x) sin((n + k)x) 2 nk n+k

= 0 for k = n

(C.15)

Length of k (k |k ) = ||k ||2 =


sin2 (kx)dx =

x 1 sin(2x) 2 4

= for k 1 (C.16)

C.3. Fourier series of a function

176

C.3 Fourier series of a function


Now that we have proved that {g} in Eq. C.11 forms an orthogonal system of func1 tions, we know that we can express any periodic function, f (with a period of 2) in {g} as 1

f (x) = c +

(an cos(nx) + bn sin(nx))


n=1

(C.17)

where x is a spatial coordinate. The Fourier coefents are given by 1 1 an = (f |n )/||n ||2 = 1 c = (f |0 )/||0 ||2 = 2 bn = (f |n )/||n ||2 =

f (x) sin(nx)dx

(C.18a) (C.18b) (C.18c)

f (x) cos(nx)dx, n > 0


f (x)dx

If we set c = a0 /2, then a0 is obtained from Eq. C.18b, i.e. f (x) = a0 (an cos(nx) + bn sin(nx)) + 2 n=1

(C.19a) (C.19b) (C.19c)

1 1 an = (f |n )/||n ||2 = bn = (f |n )/||n ||2 =

f (x) sin(nx)dx

f (x) cos(nx)dx

Note that a0 corresponds to the average of f . Taking the average of f (i.e. integrating f from to ) gives (see Eq. C.19a) f= Hence, if f = 0 then a0 = 0.

f (x)dx = a0

(C.20)

C.4 Derivation of Parsevals formula


Parsevals formula reads

(f (x))2 dx =

2 a + (a2 + b2 ) n n 2 0 n=1

We will try to prove this formula. Assume that we want to approximate the function f as well as possible with an orthogonal series

an g n
n=1

(C.21)

Now we want to prove that the Fourier coefcients are the best choice to minimize the difference
N

||f

n=1

an gn ||

(C.22)

C.4. Derivation of Parsevals formula

177

Later we will let N . Using the denition of the norm and the laws of scalar product we can write
N N N

||f = (f |f ) = (f |f ) 2

n=1

an gn ||2 =

f
N

n=1

an g n f
N

ak g k
k=1

N n=1 N n=1

an (f |gn ) an (f |gn ) +

k=1 N n=1

ak (f |gk ) + a2 (gn |gn ) n

n=1 k=1

an ak (gn |gk ) =

(C.23)

because of the orthogonality of the function system, {g}N . Express f in the second 1 term using the Fourier coefcients cn (see Eqs. C.5 and C.8) gives
N N

(f |f ) 2

n=1

an cn (gn |gn ) +
N

n=1

a2 (gn |gn ) n (C.24)

= ||f ||2 +
N

n=1

||gn ||2 a2 2an cn n


N 2

= ||f ||2 +

n=1

||gn ||2 (an cn )

n=1

||gn ||2 c2 n

The left side of Eq. C.23 is thus minimized if the coefcients an are chosen as the Fourier coefcients, cn so that
N N

||f

n=1

an gn ||2 = ||f ||2

n=1

||gn ||2 c2 n

(C.25)

The left side must always be positive and hence


N n=1

||gn ||2 c2 ||f ||2 = n

(f (x))2 dx

for all N

(C.26)

As N is made larger, the magnitude of the left side increases, and its magnitude gets closer and closer to that of the right side, but it will always stay smaller than ||f ||2 . This means the series on the left side is convergent. Using the Fourier coefcients in Eq. C.19 and letting N it can be shown that we get equality of the left and right side, which gives Parsevals formula,

||f ||2

(f (x))2 dx =

2 a + (a2 + b2 ) n n 2 0 n=1

Note that /2 and on the right side correspond to the length of ||gn ||, i.e. ||0 ||, ||n || and ||n ||, respectively. Appendix D describes in detail how to create energy spectra from two-point correlations.

C.5. Complex Fourier series

178

C.5 Complex Fourier series


Equation C.19 gives the Fourier series of a real function. It is more convenient to express a Fourier series in complex variables even if the function f itself is real. On complex form it reads

f (x) = cn = 1 2

cn exp(nx))
n=

(C.27a) (C.27b)

f (x) exp(nx)dx

where the Fourier coefcients, cn , are complex. Below we verify that if f is real, then Eq. C.27 is equivalent to Eq. C.19. The Fourier coefcients, cn , read assuming that f is real according to Eq. C.27 cn = 1 2

f (x)(cos(nx) sin(nx))dx =

1 (an bn ), 2

n>0

(C.28)

where an and bn are given by Eq. C.19. For negative n in Eq. C.27 we get cn = c = n 1 2

f (x)(cos(nx) + sin(nx))dx =

1 (an + bn ), 2

n > 0 (C.29)

where c denotes the complex conjugate. For n = 0, Eq. C.27 reads n c0 = 1 2

f (x)dx =

1 a0 2

(C.30)

see Eq. C.19. Inserting Eqs. C.28, C.29 and C.30 into Eq. C.27 gives f (x) = = = 1 1 a0 + (an bn ) exp(nx) + (an + bn ) exp(nx) 2 2 n=1 1 1 a0 + (an bn )(cos(nx) + sin(nx)) + (an + bn )(cos(nx) sin(nx)) 2 2 n=1 1 1 a0 + an cos(nx) 2 bn sin(nx) = a0 + an cos(nx) + bn sin(nx) 2 2 n=1 n=1 (C.31) which veries that the complex Fourier series for a real function f is indeed identical to the usual formulation in Eq. C.19 although the Fourier coefcients, cn , are complex. One advantage of Eq. C.27 over the formulation in Eq. C.19 is that we dont need any special denition for the rst Fourier coefcient, a0 . The trick in the formulation in Eq. C.27 is that the imaginary coefcients for negative and positive n cancel whereas the real coefcients add. This means that the real coefcients are multiplied by a factor two except the rst coefcient, a0 , which makes up for the factor 1 in front of a0 in 2 Eq. C.19.

D. TME225: Compute energy spectra from LES/DNS data using Matlab

179

D TME225: Compute energy spectra from LES/DNS data using Matlab


D.1 Introduction
When analyzing DNS or LES data, we are interested to look at the energy spectra. From these we can nd out in which turbulence scales (i.e. at which wave numbers) the uctuating kinetic turbulent energy reside. By taking the Fourier transform of the time signal (a uctuating turbulent velocity) and then taking the square of the Fourier coefcients we obtain the energy spectrum versus frequency. If we want to have the energy spectrum versus wavenumber, we Fourier transform N instantaneous signals in space and then time average the N Fourier transforms. An alternative way is to Fourier transform of a (time-averaged) two-point correlation, B33 (3 ), which is dened as (see Eq. 10.2) x
B(x3 , x3 ) = v3 (x3 x3 )v3 (x3 )

(D.1)

where x3 is the separation between the two points. Here we assume that x3 is an homogeneous direction so that B33 is independent of x3 , i.e. B33 = B33 (3 ). The x two-point correlation for an innite channel ow is shown in Fig. D.1. On discrete form the expression for B33 reads B33 (kz) = 1 M
M m=1 v3 (x3 kz)v3 (x3 )

(D.2)

where m denotes summation in homogeneous directions (i.e. time plus spatial homogeneous directions). In the following section we give a simple example how to use Matlab to Fourier transform a signal where we know the answer. Then we show how to derive the energy spectrum from a spatial two-point correlation. Finally, some comments are given on how to create an energy spectrum versus frequency from an autocorrelation (i.e. from a two-point correlation in time).

D.2 An example of using FFT


Here we will present a simple example. Consider the function u = 1 + cos(2x/L) = 1 + cos(2(n 1)/N ) (D.3)

where L is the length of the domain and N = 16 is the number of discrete points, see Fig. D.2. Lets use this function as input vector for the discrete Fourier transform (DFT) using Matlab. The function u is symmetric, so we expect the Fourier coefcients to be real. In Matlab the DFT of u is dened as (type help fft at the Matlab prompt)
N

U (k) =
n=1

un exp

2(k 1)(n 1) N

(D.4)

where k is the non-dimensional wavenumber and = 1. The ratio (n 1)/N corresponds to the physical coordinate, x, in the the continuous FFT U c () = 1 L
L

1kN

u(x) exp(x)dx,
L

= 2/L

(D.5)

D.2. An example of using FFT

180

Note that the discrete Fourier U (k) coefcients in Eq. D.4 must be divided by N , i.e. U (k)/N , in order to correspond to the Fourier coefcients U c (N corresponds to L in Eq. D.5). Furthermore, it can be noted that in Eq. D.4 the period [0, 2] is used whereas the formulation in Eq. D.5 is based on the interval [, ]. In Matlab, we generate the function u in Eq. D.3 using the commands N=16; n=1:1:N; u=1+cos(2*pi*(n-1)/N); The u function is shown in Fig. D.2. 16 nodes are used; node 1 is located at x = 0 and node 16 is located at 15L/16.

1.4 1.2 1

B33 (3 ) x

0.8 0.6 0.4 0.2 0 0.2 0 0.1 0.2 0.3 0.4 0.5 0.6

x3
Figure D.1: Two-point correlation, B(3 ) = v3 (x3 x3 )v3 (x3 ) , of DNS data in x channel ow taken from [73].

2 2
1.5

16

0.5

9
0 0 0.5 1 1.5

2.5

Figure D.2: The u function.

D.3. Energy spectrum from the two-point correlation


Now we take the discrete Fourier transform of u. Type U=fft(u);

181

Instead of using the built-in fft command in Matlab we can program Eq. D.4 directly in Matlab as U=zeros(1,N); for k=1:N for n=1:N arg1=2*pi*(k-1)*(n-1)/N; U(k)=U(k)+u(n)*cos(-arg1); end end Note that since u is symmetric, we have only used cos(x) = cos(x) (the symmetric part of exp(x)). The resulting Fourier coefcients are shown in Fig. D.4. Since the function u includes only one cosine function and a mean (which is equal to one) only three Fourier coefcient are non-zero. Two of them, U (2)/N = 0.5, U (16)/N = 0.5, correspond to the cosine functions (there must be two since U is symmetric) cos((N 1)2(n 1)/N ) = cos(2(n 1)/N ) = cos(2(n 1)/N ) cos(2(n 1)/N ) (D.6)

which corresponds to cos(2x/L) in Eq. D.3. It can be noted that the interval [k = N/2 + 1, N = 9, 16] corresponds to the negative, symmetric part of the wavenumbers in the physical formulation (cf. Eqs. D.4 and D.5). The rst Fourier coefcient corresponds as always to the mean of u, i.e. U (1)/N = u . This is easily veried from Eq. D.4 by inserting k = 1. The remaining coefcients are zero. In Fig. D.3, U/N is plotted versus non-dimensional wavenumber, k, and versus wavenumber = 2(n 1)/L. The energy, u2 , of the signal in Fig. D.2 can be computed as u2 = 1 L
L N

u2 (x)dx =
0 n=1

u2 /N = 1.5 n

(D.7)

In wavenumber space the energy is according to Parsevals formula, see Eq. C.4 equal to the integral of the square of the Fourier coefcients, i.e. u2 = 1 L
0

U 2 ()d =

1 N

N 2 Un /N = 1.5 n=1

(D.8)

D.3 Energy spectrum from the two-point correlation


Now that we have learnt how to use the FFT command in Matlab, lets use it on our two-point correlation in Eq. D.1 and Fig. D.1. Equation D.4 reads
N

B33 (k) =
n=1

B33 (n) exp

2(k 1)(n 1) N

(D.9)

D.3. Energy spectrum from the two-point correlation

182

1.2 1

0.8 0.8

U/N

0.6 0.4 0.2

0.6

U/N
0.4 0.2

0 0.2 0 5 10 15 20 0 0 5 10 15 20

k
(a) Versus non-dimensional wavenumber, k.

(b) Versus wavenumber, .

Figure D.3: The U/N Fourier coefcients.


1.4 1.2 1

96

B33 (3 ) x

0.8 0.6 0.4 0.2 0 0.2 0 0.5 1 1.5

x3
Figure D.4: Periodic two-point correlation, B33 (3 ) = v3 (x3 )v3 (x3 + x3 ) , of DNS x data in channel ow taken from [73].

The simulations have been carried out with periodic boundary conditions in x3 direction (and x1 ), and hence B33 (3 ) is symmetric, see Fig. D.4. Thus, it is sufcient to x use the cosine part of Eq. D.9, i.e.
N

B33 (k) =
n=1

B33 (n) cos

2(k 1)(n 1) N

(D.10)

In Fig. D.5a the Fourier coefcients B33 3 are presented versus wavenumber 3 = 2(n 1)/x3,max , where x3,max 1.55, see Fig. D.4. Figure D.5b shows the same energy spectra in log-log scale (only half of the spectrum is included), which is the common way to present energy spectra. The dashed line shows the 5/3 slope which indicates that the energy spectra from the DNS follows the Kolmogorov 5/3 decay. As usual, the Fourier coefcient for the rst non-dimensional wavenumber, i.e. B33 (1)/N is equal to the mean of B33 , i.e. B33 = 1 N
N n=1

B33 (n)

1 B33 (1) N

(D.11)

compare with Eq. D.10. Note that this is almost the same expression as that for the

D.4. Energy spectra from the autocorrelation


0.25 10
1

183

0.2

B33 (k)/N

0.15

10 0.1 10 0.05 10

0 0

100

200

300

400

10

(a) Linear-linear scale

(b) Log-log scale

10

2 Figure D.5: The energy spectrum of v3 versus wavenumber, 3 . Dashed line in b) show 5/3 slope.

integral length scale which reads (see Eq. 10.5) Lint (3 ) = x 1


2 v3 0

B33 (x3 , x3 )d3 = x

B33
2 v3

(D.12)

Hence the integral length scale is related to the rst Fourier mode as Lint = B33 (1)
2 v3

(D.13)

2 2 The two-point correlation for zero separation is equal to v3 , i.e. B33 (0) = v3 = 2 is to integrate the energy spectrum in Fig. D.5, i.e. 1.51. Another way to obtain v3

2 v3 =

1 x3,max
0

1 B33 (3 )d3 = N

B33 (n) = 1.52


n=1

(D.14)

D.4 Energy spectra from the autocorrelation


When computing the energy spectra of the v3 velocity, say, versus frequency, the time series of v3 (t) is commonly Fourier transformed and the energy spectrum is obtained by plotting the square of the Fourier coefcients versus frequency, f . We can also split the time signal into a number subsets, Fourier transform each subset and then average. In Matlab, the command pwelch is a convenient command which does all this. In the previous section we computed the energy spectrum versus wavenumber by Fourier transforming the two-point correlation. We can use the same approach in time. First we create the autocorrelation B33 ( ) = v3 (t)v3 (t + ) (this can be seen as a two-point correlation in time). Then B33 ( ) is Fourier transformed to get B33 (f ) in 33 ( ) is a function of the same way as in Section D.3. The only difference is that B frequency whereas B33 (3 ) is a function of wavenumber.

E. TME225 Assignment 2: turbulent ow

184

E TME225 Assignment 2: turbulent ow


In this exercise you will use data from a Direct Numerical Simulation (DNS) for fully developed channel ow. In DNS the unsteady, three-dimensional Navier-Stokes equations are solved numerically. The Re number based on the friction velocity and the half channel width is Re = u h/ = 500 (h = = u = 1 so that = 1/Re ). A 96 96 96 mesh has been used. The streamwise, wall-normal and spanwise directions are denoted by x (x1 ), y (x2 ) and z (x3 ) respectively. The cell size in x and z directions are x = 0.0654 and z = 0.0164. Periodic boundary conditions were applied in the x and z direction (homogeneous directions). All data have been made non-dimensional by u and . You can do the assignment on your own or in a group of two. You should write a report where you analyze the results following the heading H1H13. It is recommended A A (but the not required) that you use LTEX(an example of how to write in LTEXis available on the course www page). It is available on Linux. On Windows you can use, for example, Lyx (www.lyx.org) or MikTex (www.miktex.org) which are both free to download.

E.1

Time history

At the course home page http://www.tfd.chalmers.se/lada/turbulent flow/ you nd a le u v time 4nodes.dat with the time history of v1 and v2 . The le has eight columns of v1 and v2 at four nodes: x2 / = 0.0039, x2 / = 0.0176, x2 / = 0.107 and x2 / = 0.47. With u = 1 and = 1/Re = 1/500 this correspond to x+ = 1.95, x+ = 8.8, x+ = 53.5 and x+ = 235. The sampling time step 2 2 2 2 is t = 0.0033 (every second time step). The four points are located in the viscous sublayer, the buffer layer and in the logarithmic layer, see Fig. 6.2 at p. 50. Use the Matlab program pl time.m which loads and plots the time history of v1 . Start Matlab and run the program pl time. Recall that the velocities have been scaled with the friction velocity u , and thus what you see is really v1 /u . The time history of v1 at x2 / = 0.0176 and x2 / = 0.107 are shown. Study the time history of the blue line (x2 / = 0.0176) more in detail. Make a zoom between, for example, t = 10 and t = 11 and v1,min = 3 and v1,min = 21. This is conveniently done with the command axis([10 11 3 21]) In order to see the value at each sampling time step, change the plot command to plot(t,u2,b-,t,u2,bo) Use this technique to zoom, to look at the details of the time history. Alternatively, you can use the zoom buttons above the gure. Plot v1 for all four nodes. How does the time variation of v1 vary for different positions? Plot also v2 at the four different positions. What is the differences between v1 and v2 ?

E.2

Time averaging

Compute the average of the v1 velocity at node 2. Add the following code (before the plotting section)

E.3. Mean ow

185

umean=mean(u2) Here the number of samples is n = 5000 (the entire u2 array). Find out how many samples must be used to get a correct mean value. Start by trying with 100 samples as umean_100=mean(u2(1:100)) What is the maximum and minimum value of v1 ? Compare those to the mean. Do the same exercise for the other three nodes. Compute and plot also the instantaneous uctuations; v1 at node 1, for example, is computed as u1_mean=mean(u1); u1_fluct=u1-u1_mean;

E.3

Mean ow

All data in the data les below have been stored every 10th time step. Download the le uvw inst small.dat, y.dat and the Matlab le pl vel.m which reads the data les. The data le includes v1 , v2 and v3 from the same DNS as above, but now you are given the time history of all x2 nodes at one chosen x1 and x3 node. There are nj = 98 nodes in the x2 direction; node 1 and nj are located at the lower and upper wall, respectively. Your data are instantaneous. Compute the mean velocity. Plot it both as linearlinear plot and a log-linear plot (cf. Fig. 6.4). + In the log-linear plot, use x+ for the wall distance. Include the linear law, v1 = x+ , 2 2 + + 1 and the log law, v1 = ln x2 + B ( = 0.41 is the von K rm n constant and a a B = 5.2). How far out from the wall does the velocity prole follow the linear law? At what x+ does it start to follow the log-law? 2 Compute the bulk velocity V1,b = 1 2h
2h

v1 dx2
0

(E.1)

(recall that h denote half the channel width) What is the Reynolds number based on V1,b and centerline velocity, V1,c , respectively?

E.4

The time-averaged momentum equation

Let us time average the streamwise momentum equation. Since the ow is fully developed and two dimensional we get 0=
2 v1 v1 v2 1 p + 2 x x1 x2 2

(E.2)

This equation is very similar to fully developed laminar ow which you studied in Assignment 1, see Eq. 3.24; the difference is that we now have an additional Reynolds stress term. Recall that all terms in the equation above represent forces (per unit volume). Let us investigate how these forces (the pressure gradient, the viscous term and the Reynolds stress term) affect uid particles located at different x2 locations. Com pute and plot the three terms. (the le uvw inst small.dat does not include p; set p/x = 1.)

E.5. Wall shear stress

186

If a term is positive it means that it pushes the uid particle in the positive x1 direction. What about the viscous term? Is it always negative? Where is it largest? At that point? which term balances it? How large is the third term? The pressure term should be a driving force. Where is the Reynolds shear stress positive and where is it negative?

E.5

Wall shear stress

Compute the wall shear stress at both walls. They should be exactly equal. Are they?

E.6

Resolved stresses

In Section E.3 you computed the mean velocities. From the instantaneous and the mean velocity, you can compute the uctuations as
vi = vi vi

(E.3)

Now you can easily compute all stresses vi vj . Plot the normal stresses in one gure and the shear stresses in one gure (plot the stresses over the entire channel, i.e. from x2 = 0 to x2 = 2h). Which shear stresses are zero?

E.7

Fluctuating wall shear stress

In the same way as the velocity, the wall shear stress can be decomposed into a mean value and a uctuation. In general, any uctuating variable, , can be decomposed into a mean and uctuation as = + . The root-mean-square (RMS) is then dened as rms = 2
1/2

(E.4)

Compute the RMS of the wall shear stress. This is a measure of the uctuating tangential force on the wall due to turbulence. If heat transfer is involved, the uctuating temperature at the wall inducing uctuating heat transfer may be damaging to the material of the walls causing material fatigue. This is probably the most common form of uid-solid interaction.

E.8

Production terms

In order to understand why a stress is large, it is useful to look at its transport equation, see Eq. 9.12. Usually, a stress is large when its production term, Pij , is large (there may be exceptions when other terms, such as the diffusion term, are largest). Plot the production terms for all non-zero stresses across the entire channel. Which ones are zero (or close to)? Does any production term change sign at the centerline? If so, what about the sign of the corresponding shear stress plotted in Section E.6?

E.9

Pressure-strain terms
vj vi + xj xi

The pressure-strain term reads (see Eq. 9.14) ij = p (E.5)

E.10. Dissipation

187

Our data are obtained from incompressible simulations, in which the pressure may vary unphysically in time (p/t does not appear in the equations). Hence, we prefer to compute the velocity-pressure gradient term p = ij
vj p vi p , xj xi

(E.6)

2 see the second line in Eq. 9.3. The pressure diffusion term in the v2 equation which is the difference between Eqs. E.5 and E.6 (the two rst terms in Eq. 9.8) is small except very close to the wall (see Figs. 9.2 and 9.3). Hence, the difference between p ij and ij is small. Download the data le p inst small.mat and the Matlab le pl press strain.m which reads the data le. The time histories of the pressure along ve x2 lines [(x1 , x2 , x3 ), (x1 x1 , x2 , x3 ) and (x1 , x2 , x3 x3 )] are stored in this le. This allows you to compute all the three spatial derivatives of p . Using the velocities stored in uvw inst small.dat (see Section E.3), you can compute all the terms in Eq. E.6. Plot the pressure strain, p , for the three normal stresses and the shear stress across ij the channel. For which stresses is it negative and positive? Why? Which term p is the largest source and sink term, respectively? ij

E.10

Dissipation

The physical meaning of dissipation, , is transformation of turbulent kinetic energy into internal energy, i.e. increased temperature. Download the le diss inst.mat and the Matlab le pl diss.m which reads it. The data le includes the time history of the velocities along ve x2 lines [(x1 , x2 , x3 ), (x1 x1 , x2 , x3 ) and (x1 , x2 , x3 x3 )] so that you can compute all spatial derivatives. The data cover only the lower half of the channel. Compute and plot =
vi vi xk xk

(E.7)

see Eq. 8.14. Where is it largest? In which equation does this quantity appear? Let us now consider the equations for the mean kinetic energy, K = vi vi /2 (Eq. 8.30) and turbulent kinetic energy, k = vi vi /2 (Eq. 8.14). The dissipation in the K equation reads i i v v mean = (E.8) xk xk The ow of kinetic energy between K, k and T is illustrated in Fig. E.1. Compute and plot also mean and Pk . Which is large and which is small? How is the major part of the kinetic energy transformed from K to T ? Is it transformed via k or directly from K to T ?

E.11

Do something fun!

You have been provided with a lot of data which you have analyzed in many ways. Now think of some other way to analyze the data. There are many interesting things yet to be analyzed!

E.11. Do something fun!

188

n ea

Pk

Figure E.1: Transfer of energy between mean kinetic energy (K), turbulent kinetic energy (k) and internal energy (denoted as an increase in temperature, T ). K = 1 1 2 vi vi and k = 2 vi vi . The dissipations, and mean , are dened in Eqs. E.7 and E.8, respectively.

E.11. Do something fun!

189

TME225 Old Learning outcomes: week 1& 2


Describe and explain 1. the difference between Lagrangian and Eulerian description 2. Watch the on-line lecture Eularian and Lagrangian Description, part 1 3 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. Part 1 described the differecne between Lagrangian and Eularian points and velocities. T T is nicely explained in Part 2 + vi ii. The formula t xi 3. which viscous stress components that act on a Cartesian surface 4. the physical meaning of diagonal and off-diagonal components of Sij 5. the physical meaning of ij 6. denition of irrotational ow 7. the physical meaning of irrotational ow 8. why in uid dynamics we can formulate our equations for a volume although the balance equations are derived for a given mass (see Section 2.5) 9. the ow physics at the entrance (smooth curved walls) to a plane channel 10. the ow physics in a channel bend 11. vortex stretching and vortex tilting 12. why the vortex stretching/tilting term is zero in two-dimensional ow 13. the similarities between the vorticity and temperature transport equations in fully developed ow between two parallel plates 14. the physical meaning of the eigenvectors and the eigenvalues of the stress tensor 15. where in a ow the dissipation is largest 16. the physical meaning of the terms in the transport equation for inner energy, u 17. the physical meaning of the terms in the transport equation for kinetic energy, vi vi /2 18. the energy transfer between kinetic energy and inner energy 19. the physical meaning of dissipation, 20. Watch the on-line lecture Boundary layers part 1 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. They talk vorticity and inviscid ow ii. They talk about boundary layer growth iii. How do they dene the boundary layer thickness?

E.11. Do something fun!

190

iv. The velocity gradient in the boundary layer is large upstream but decreases further downstream. Why? u (u is velocity in their notation). v. They dene shear stress, , as = y Does increase or decrease when going downstream? vi. By using the concept of circulation, show that the vorticity in the boundary layer stays constant for any x. vii. Use the diffusion of vorticity to show that for any x. (see also Eq. 3.14). U

21. Watch the on-line lecture Boundary layers parts 2 & 3 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. Why is the vorticity levels higher and the boundary layer thinner after the contraction? ii. Is the wall shear stress lower or higher after the contraction? iii. At high angle of a diffuser, the ow separates. Why? How large is the shear stress at this point? iv. Describe the process of transition from laminar to turbulent ow v. The ow is tripped into turbulence. How? vi. When the ow along the lower wall of the diffuser is tripped into turbulent ow, the separation region is suppressed. Try to explain why. vii. Two boundary layers one on each side of the plate are shown. The upper one is turbulent and the lower one is laminar. What is the difference in their two velocity proles? Explain the differences. viii. Why is the turbulent wall shear larger for the turbulent boundary layer? What about the circulation? ix. Consider the airfoil: when the boundary layer on the upper (suction) side is turbulent, stall occurs at a higher angle of incidence compared when the boundary layer is laminar. Why? x. Vortex generator are place on the suction side in order prevent or delay separation. Explain why this happens. 22. Watch the on-line lecture Pressure eld and acceleration part 1 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. The water ow goes through the contraction. What happens with the velocity and pressure. Why? ii. Fluid particles become thinner and elongated in the contraction. Explain why. iii. They show that the acceleration along s (V dV /dS) is related to the pressure gradient dp/ds. Compare this relation with the three-dimensional form of Navier-Stokes equations for incompressible ow, Eq. 2.7 23. Watch the on-line lecture Pressure eld and acceleration part 2 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html

E.11. Do something fun!

191

i. Water ow in a manifold (a pipe with many outlets) is presented. The pressure decreases slowly downstream. Why? ii. The bleeders (outlets) are opened. The pressure now increases in the downstream direction. Why? iii. What is the stagnation pressure? How large is the velocity at a stagnation point? iv. What is the static pressure? How can it be measured? What is the difference between the stagnation and the static pressures? v. A venturi meter is a pipe that consists of a contraction and an expansion (i.e. a diffuser). The bulk velocities at the inlet and outlet are equal, but still the pressure at the outlet is lower than that at the inlet. There is a pressure drop. Why? vi. What happens with the pressure drop is there is a separation in the diffuser? vii. They increase the speed in the venturi meter. The pressure difference in the contraction region and the outlet increases. Since there is atmospheric pressure at the outlet, this means that the pressure in the contraction region must decrease as we increase the velocity of the water. Finally the water starts to boil, although the water temperature may be around 10o C. This is called cavitation (this causes large damages in water turbines). viii. Explain how suction can be created by blowing in a pipe. 24. Watch the on-line lecture Pressure eld and acceleration part 3 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. What is the Coanda effect? ii. The water from the tap hitting horizontal pipe follows attaches to the surface of the pipe because of the Coanda effect. How large is the pressure at the surface of the pipe relative to the surrounding pressure. iii. Explain the relation between streamline curvature and pressure (cf. Chapt. 3.2.1). iv. At the end of the contraction, there is an adverse pressure gradient (p/x > 0). Explain why. Derive 1. the relation between the vorticity vector and the vorticity tensor 2. the Navier-Stokes equation, Eq. 2.5 3. the incompressible Navier-Stokes equation, Eq. 2.7 4. the transport equation for the inner energy, u, Eq. 2.12 5. the transport equation for temperature in incompressible ow, Eq. 2.15 6. the transport equation for kinetic energy, vi vi /2, Eq. 2.20 7. the transport equation for total energy, u + vi vi /2, Eq. 2.21 8. the left side of the transport equations on conservative and non-conservative form

E.11. Do something fun!

192

9. the ow equations for fully developed ow between two parallel plates (Eqs. 3.18, 3.22 and 3.26) 10. the ow equations governing the Rayleigh problem expressed in f and 11. how the boundary layer thickness can be estimated from the Rayleigh problem using f and 12. the exact solution (v1 ) for the Rayleigh problem 13. the exact solution (v1 , v2 and p) for the fully developed ow between two parallel plates 14. the pressure drop from the force balance for the ow between two parallel plates 15. the pressure drop from the kinetic energy equation for the ow between two parallel plates 16. that the viscous terms at the right side of the incompressible Navier-Stokes can be expressed in the vorticity vector 17. that the divergence of the vorticity vector is zero 18. that the convective term at left side of the incompressible Navier-Stokes can be re-written so that it includes one non-rotational term and one rotational term 19. the transport equation (3D) for the vorticity vector, Eq. 4.21 20. 2D equation transport equation for the vorticity vector from the 3D transport equation, Eq. 4.24

E.11. Do something fun!

193

TME225 Old Learning outcomes: week 3


1. What characterizes turbulence? Explain the characteristics. What is a turbulent eddy? 2. Explain the cascade process. How large are the largest scales? What is dissipation? What dimensions does it have? Which eddies extract energy from the mean ow? Why are these these eddies best at extracting energy from the mean ow? 3. What are the Kolmogorov scales? Use dimensional analysis to derive the expression for the velocity scale, v , the length scale, and the time scale, . 4. Make a gure of the energy spectrum. Show the relation between the Fourier coefents of a uctuating velocity and the energy spectrum. The energy spectrum consists of thee subregions: which? describe their characteristics. Show the ow of turbulent kinetic energy in the energy spectrum. Given the energy spectrum, E(), how is the turbulent kinetic energy, k, computed? Use dimensional analysis to derive the 5/3 Kolmogorov law. 5. What does isotropic turbulence mean? 6. How is the energy transfer from eddy-to-eddy, , estimated? Show how the ratio of the large eddies to the dissipative eddies depend on the Reynolds number. 7. Describe the cascade process created by vorticity. Write the vortex stretching/tilting term in tensor notation. What is its physical meaning? Describe the physical process of vortex stretching which creates smaller and smaller eddies. Show and discuss the family tree of turbulence eddies and their vorticity. Show that in 2D ow the vortex stretching/tilting term vanishes. 8. Watch the on-line lecture Turbulence part 1 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. Why does the irregular motion of wave on the sea not qualify as turbulence? ii. How is the turbulence syndrome dened? iii. The movie laminar shows ow in a pipe. The viscosity is decreased, and the pressure drop (i.e. the resistance) decreases. Why? The viscosity is further decreased, and the pressure drop increases. Why? How does the waterow coming out of the pipe change due to the second decrease of viscosity? iv. It is usually said that the ow in a pipe gets turbulent at a Reynolds number of 2300. In the movie they show that the ow can remain laminar up to 8 000. How? 9. Watch the on-line lecture Turbulence part 2 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. Draw a laminar and turbulent velocity prole for pipe ow. What is the 1 v main difference? In which ow is the wall shear stress w = largest, x2 laminar ot turbulent? ii. In turbulent ow, the velocity near the wall is larger than in laminar ow. Why?

E.11. Do something fun!

194

iii. Discuss the connection between mixing and the cross-stream (i.e. v2 ) uctuations.

iv. Two turbulent jet ows are shown, one at low Reynolds number and one at high Reynolds number. They look very similar in one way and very different in another way. v. The two turbulent jet ows have the same energy input and hence the same dissipation. Use this fact to explain why the smallest scales in the high Rweynolds number jet must be smaller that those in the low Reynolds number jet. vi. Explain the analogy of a water wall and the cascade process.

E.11. Do something fun!

195

TME225 Old Learning outcomes: week 4

August 17, 2011

1. Derive the exact transport equation for turbulent kinetic energy, k. Discuss the physical meaning of the different terms in the k equation. Which is the main source and sink term, respectively? Which terms do only transport k? 2. In the cascade process, we assume that the dissipation is largest at the smallest
1 0 , see Eq. 8.16 at p. 60. Show this. For which scales, i.e. = 0 eddies is the production largest? Why? 2/3

3. Watch the on-line lecture Turbulence part 3 at http://www.tfd.chalmers.se/lada/MoF/flow viz.html i. The lm says that there is a similarity of the small scales in a channel ow and in a jet ow. What do they mean? ii. What happens with the small scales when the Reynolds number is increased? What happens with the large scales? Hence, how does the ratio of the large scales to the small scales change when the Reynolds number increases (see Eq. 5.14) iii. Even though the Reynolds number may be large, there are a couple of physical phenomena which may inbit turbulence and keep the ow laminar: mention three. iv. Consider ow in a channel where the uid on the top (red) and the bottom (yellow) are separated by a horizontal partition. Study how the two uids mix downstream of the partition. In the next exampe, the uid on the top is hot (yellow) and the one at the bottom (dark blue) is cold: how do the uids mix downstream of the partition, better or worse than in the previous example? This ow situation is called stable stratication. In the last example, the situation is reverser: hot uid (dark blue) is moving on top of light uid (yellow). How is the mixing affected? This ow situation is called unstable stratication.

E.11. Do something fun!

196

TME225 Old Learning outcomes: week 5

August 17, 2011

1. Use the decomposition vi = vi + vi to derive the time-averaged Navier-Stokes equation. A new terms appears: what is it called? Simplify the time-averaged Navier-Stokes equation for boundary layer. What is the total shear stress? How is the friction velocity, u , dened? Dene x+ and v + . 2

2. The wall region is divided into an inner and outer region. The inner region is furthermore divided into a viscous sublayer, buffer layer and log-layer. Make a gure and show where these regions are valid (Fig. 6.2) 3. Consider fully developed channel ow. The total shear stress consists of a viscous and turbulent shear stress: show how they vary across the channel (show also a zoom near the wall). In which region (viscous sublayer, buffer layer or log-layer) does the viscous stress dominate? In which region is the turbulent shear stress large? Integrate the boundary layer equations and show that the total shear stress varies as x2 / 1. 4. What are the relevant velocity and length scales in the viscous-dominated region? Derive the linear velocity law in this region (Eq. 6.16). What are the suitable velocity and length scales in the inertial region? Derive the log-law. 5. In fully developed channel ow, the time-averaged Navier-Stokes consists only of three terms. Make a gure and show how the velocity and shear stress vary across the channel. After that, show how the three terms (i.e. their gradients plus the pressure gradient) vary across the channel. Which two terms balance each other in the outer region? Which terms drives (pushes) the ow in the x1 direction? Which two terms are large in the inner region? Which term drives the ow? 6. How is the mean velocity computed with a PDF function? How should the function be normalized? Show how the second moment of a uctuating velocity is computed with a PDF function. How are the skewness and the atness of a uctuating velocity dened? What is the physical meaning of a small and large skewness? What is the physical meaning of a small and large atness?

E.11. Do something fun!

197

TME225 Old Learning outcomes: week 6

August 17, 2011

1. Given the exact k equation, give the equation for boundary-layer ow. All spatial derivatives are kept in the dissipation term: why? In the turbulent region of the boundary layer, the k equation is dominated by two terms. Which ones? Which terms are non-zero at the wall? Why? Where is the viscous and turbulent diffusion terms large? Why? Where is the production term largest? 2. Discuss the difference of spatial transport of k and spectral transfer of k. Give an example of how they are combined in non-homogeneous turbulence. What is homogeneous turbulence? 3. Derive the exact transport equation for mean kinetic energy, K. Discuss the physical meaning of the different terms. One term appears in both the k and the K equations: which one? Consider the dissipation terms in the k and the equations: which is largest? Why? Show where they appear in the energy K spectrum.
4. Derive the exact transport equation for turbulent Reynolds stress, vi vj . Take the trace of the vi vj equation to obtain the k equation.

5. Show that the role of the convection and diffusion terms is purely to transport and not give any net effect except at the boundaries (use the Gauss divergence theorem)

E.11. Do something fun!

198

TME225 Old Learning outcomes: week 7

August 17, 2011

1. Discuss the physical meaning of the different terms. The mean normal stress can 2 be dened as vrms,av = vi vi /3; what sign will the pressure-strain term have for 2 normal stresses, respectively, larger and smaller than vrms,av ? What role does 12 has? What sign? Why do we call the pressure-strain term the Robin Hood term? 2. Consider the dissipation term, 12 , for the shear stress: how large is it? 3. Consider fully developed channel ow: how are the expressions for the production terms simplied? Which production terms are zero and non-zero, respec tively? Consider the production term for v1 v2 : which sign does it have in the lower and upper part of the channel, respectively? Why is there no pressurestrain term in the k equation? 4. Consider the fully turbulent region in fully developed channel ow: which are 2 2 2 the main source and sink terms in the v1 , v2 , v3 and v1 v2 equations? Which terms are the largest terms at the wall? Which terms are zero at the wall?
5. Consider channel ow and use physical reasoning to show that v1 v2 must be negative and positive in the lower and upper half of the channel, respectively. Is this consistent with the sign of P12 ?

6. Dene the two-point correlation. How is it normalized? What is the physical meaning of the two-point correlation? How is it related to the largest eddies? How is the integral length scale dened? 7. Dene the auto correlation. How is it normalized? What physical meaning does it have? The integral time scale is dened in analogy to the integral length scale: show how it is dened.

F. MTF270: Some properties of the pressure-strain term

199

MTF270: Some properties of the pressure-strain term

In this Appendix we will investigate some properties of aijk in Eq. 11.69 at p. 91. Introduce the two-point correlation function
Bj (r) = vj (x)v (x + r)

Dene the point x = x + r so that


Bj (r) = vj (x r)v (x ) = v (x )vj (x r) = Bj (r)

We get Bj (r) 2 Bj (r) 2 Bj (r) Bj (r) = = ri ri rk ri rk ri (F.1)

Since Eq. F.1 in the denition of aijk in Eq. 11.69 is integrated over r3 covering both r and r (recall that v and vj are separated by r), aijk is symmetric with respect to index j and , i.e. aijk = aikj (F.2) Greens third formula (it is derived from Gauss divergence law) reads (x) = 1 4 2 dy3 |y x| (F.3)

where the boundary integrals have been omitted. Setting = v vj in Eq. F.3 gives

vj v =

1 4

2 v vj dy3 1 = aiji xi xi |y x| 2

(F.4)

where the last equality is given by Equation 11.69.

G. MTF270: Galilean invariance

200

MTF270: Galilean invariance

In [64] he found that the Leonard term Lij and the cross term Cij are not Galilean invariant by themselves, but only the sum Lij + Cij is. As a consequence, if the cross term is neglected, the Leonard stresses must not be computed explicitly, because then the modelled momentum equations do not satisfy Galilean invariance. Below we repeat some of the details of the derivation given in [64]. Galilean invariance means that the equations do not change if the coordinate system is moving with a constant speed Vk . Lets denote the moving coordinate system by , i.e. x = xk + Vk t, t = t, vk = vk + Vk k By differentiating a variable = (t , x ) we get i x t (xi , t) j = + = xk xk xj xk t x k x t (xi , t) k = + t t = Vk x + t . t t xk k (G.1)

(G.2)

From Eq. G.2 is it easy to show that the Navier-Stokes (both with and without lter) is Galilean invariant [64, 158]. Transforming the material derivative from the (t, xi )coordinate system to the (t , x )-coordinate system gives i = + Vk + (vk Vk ) + vk t xk t xk xk = + vk , t xk It shows that the left hand side does not depend on whether the coordinate system moves or not, i.e. it is Galilean invariant. Now, lets look at the Leonard term and the cross term. Since the ltering operation is Galilean invariant [64], we have vk = vk + Vk and consequently also vk = vk . For the Leonard and the cross term we get (note that since Vi is constant Vi = Vi = Vi ) v v v v L = vi vj vi vj = (i + Vi )(j + Vj ) (i + Vi )(j + Vj ) ij = vi vj + vi Vj + vj Vi vi vj vi Vj Vi vj v v = vi vj vi vj + Vj (i vi ) + Vi (j vj )
= Lij Vj vi Vi vj

(G.3)

v v Cij = vi vj + vj vi = (i + Vi )vj + (j + Vj )vi = = vi vj + vj Vi + vj vi + vi Vj = Cij + vj Vi + vi Vj

From Eq. G.3 we nd that the Leonard term and the cross term are different in the two coordinate systems, and thus the terms are not Galilean invariant. However, note that the sum is, i.e. L + Cij = Lij + Cij . (G.4) ij The requirement for the Bardina model to be Galilean invariant is that the constant must be one, cr = 1 (see Eq. 17.38). This is shown by transforming both the exact

G. MTF270: Galilean invariance

201

M Cij (Eq. 17.36) and the modelled one, Cij (i.e. Eq. 17.37). The exact form of Cij transforms as in Eq. G.3. The Bardina term transforms as M Cij = cr (i vj v i v j ) v

v v = cr (i + Vi )(j + Vj ) (i + Vi )(j + Vj ) v v
M = Cij + cr v i Vj + v j Vi .

(G.5)

v v = cr [i vj vi vj (i vi )Vj (j vj )Vi ] v

M M As is seen, Cij = Cij , but here this does not matter, because provided cr = 1 the M modelled stress, Cij , transforms in the same way as the exact one, Cij . Thus, as for the M M exact stress, Cij (see Eq. G.4), we have Cij + L = Cij + Lij . Note that in order ij to make the Bardina model Galilean invariant the Leonard stress must be computed explicitly.

H. MTF270: Computation of wavenumber vector and angles

202

MTF270: Computation of wavenumber vector and angles

For each mode n, create random angles n , n and n (see Figs. H.1 and 23.1) and random phase n . The probability distributions are given in Table H.1. They are chosen so as to give a uniform distribution over a spherical shell of the direction of the wavenumber vector, see Fig. H.1.

H.1 The wavenumber vector, n j

x3

dAi n x1
Figure H.1: The probability of a randomly selected direction of a wave in wave-space is the same for all dAi on the shell of a sphere. Compute the wavenumber vector, n , using the angles in Section H according to j Fig. H.1, i.e. n = sin(n ) cos(n ) 1 n = sin(n ) sin(n ) 2 n = cos(n ) 3 p(n ) = 1/(2) p( n ) = 1/(2) p(n ) = 1/2 sin() p(n ) = 1/(2) 0 n 2 0 n 2 0 n 0 n 2 (H.1)

n i
n

x2

Table H.1: Probability distributions of the random variables.

n H.2. Unit vector i

203 n i (1, 0, 0) (1, 0, 0) (0, 1, 0) (0, 1, 0) (0, 0, 1) (0, 0, 1)


n i (0, 0, 1) (0, 1, 0)

n 0 90 0 90 0 90

(0, 0, 1) (1, 0, 0) (0, 1, 0) (1, 0, 0)

n Table H.2: Examples of value of n , i and n from Eqs. H.1 and H.3. i

n H.2 Unit vector i


n Continuity requires that the unit vector, i , and n are orthogonal. This can be seen j by taking the divergence of Eq. 23.3 which gives N

v = 2

n=1

un cos(n x + n ) n n

(H.2)

n n i.e. i n = 0 (superscript n denotes Fourier mode n). Hence, i will lie in a plane i n normal to the vector i , see Fig. 23.1. This gives n 1 = cos(n ) cos(n ) cos(n ) sin(n ) sin(n ) n 2 = sin(n ) cos(n ) cos(n ) + cos(n ) sin(n ) n 3

(H.3)

= sin( ) cos( )

n n n The direction of i in this plane (the 1 2 plane) is randomly chosen through n . Table H.2 gives the direction of the two vectors in the case that i is along one coordinate direction and = 0 and = 90o .

I. MTF270: 1D and 3D energy spectra

204

MTF270: 1D and 3D energy spectra

The general two-point correlation Bij of vi and vj (see Eq. 10.2) can be expressed by the energy spectrum tensor as [57, Chapter 3] (cf. Eq. 17.8) +

Bij (x1 , x2 , x3 )) =

ij () exp(m xm )d1 d2 d3

(I.1)

where xm and m are the separation vector the two points and the wavenumber vector, respectively. The complex Fourier transform exp(m xm ) is dened in Appendix C. The two-point correlation, Bij , and the energy spectrum tensor, ij , form a Fouriertransform pair ij () = 1 (2)3
+

Bij () exp(m rm )d1 d2 d3 x x x x

(I.2)

The separation between the two points is described by a general three-dimensional vector, xm . Both in experiments and in LES it is usually sufcient to study the two point correlation and the energy spectra along a line. Hence, one-dimensional energy spectra, Eij (), which are a function of scalar wavenumber, (1 , 2 or 3 ), are often used. They are formed by integrating over a wavenumber plane; the energy spectrum for the wavenumber 1 , for example, reads Eij (1 ) = 1 2
+

ij ()d2 d3

(I.3)

A factor of two is included because E ii /2 is used to dene a energy spec trum for the turbulent kinetic energy k = vi vi /2, see Eqs. I.8 and I.10. Note that the maximum magnitude of the wavenumber vector contributing to Eij (1 ) is very large since it includes all 2 and 3 , i.e. < 2 < and < 3 < +. The one-dimensional two-point correlation, Bij (1 ), for example, and the one-dimensional x spectrum, Eij (1 ), form a Fourier-transform pair, i.e. Bij (1 ) = x Eij (1 ) = 1 2
+

Eij (1 ) exp(1 x1 )d1


+

(I.4) (I.5)

2 2

Bij (1 ) exp(1 x1 )d1 x x

where Eij is twice the Fourier transform of Bij because of the factor two in Eq. I.3. The diagonal components of the two-point correlation tensor are real and symmetric and hence the antisymmetric part of exp(1 x1 ) i.e. the sinus part is zero and Eqs. I.4 and I.5 are simplied as Bij (1 ) = x 1 2
+

Eij (1 ) cos(1 x1 )d1 =


+ 0

Eij (1 ) cos(1 x1 )d1


+

1 Eij (1 ) =

2 Bij (1 ) cos(1 x1 )d1 = x x

(I.6) Bij (1 ) cos(1 x1 )d1 x x

2 The Reynolds stress v1 , for example, is equal to the two-point correlation tensor 2 Bij with with zero separation distance. The v1 can be computed both from the three-

I.1. Energy spectra from two-point correlations

205

dimensional spectrum (Eq. I.1 ) and one-dimensional spectrum (Eq. I.6)


+ 2 v1 = B11 (x1 , 0, 0) = 2 v1

ii ()d1 d2 d3

(I.7)

= B11 (0) =
0

E11 (1 )d1

Hence the turbulent kinetic energy, k = vi vi /2, an be written as

k= k= 1 2

1 2

ii ()d1 d2 d3

(I.8)

E11 (1 )1 +
0

1 2

E22 (2 )2 +
0

1 2

E33 (3 )d3
0

(I.9)

The integral in Eq. I.8 has no directional dependence: it results in a scalar, k. Instead of integrating over d1 d2 d3 we can integrate over a shell with radius and letting the radius go from zero to innity, i.e. k= 1 2
0

42 ii d

(I.10)

where 42 is the surface area of the shell. We now dene an energy spectrum, E() = 42 ii so that

k=
0

E()d

(I.11)

where E() = 22 ii (). The energy spectra E11 (1 ) and E(), for example, correspond to the square of the Fourier coefcient of the velocity uctuation (see Parsevals formula, Eq. C.4), i.e. E11 (1 ) = v1 (1 ) 2 1 2 E() = v () + v3 () + v3 () 2 2 2 1 Below the properties of the three energy spectra are summarized. The three-dimensional spectrum tensor, ij (), is a tensor which is a function of the wavenumber vector. The one-dimensional spectrum, Eij (1 ), is a tensor which is a function of a scalar (one component of m ). The energy spectrum, E(), is a scalar which is a function of the length of the wavenumber vector, || .

(I.12)

I.1 Energy spectra from two-point correlations


In connection to Eqs. I.4, I.5 and I.6 we stated that the one-dimensional energy spectra and the two-point correlations form Fourier-transform pairs. The proof is given in this section. The energy spectrum is given by the square of the Fourier coefcients, see Parsevals formula, Eq. C.4. Let u be the Fourier coefcient of the velocity uctuation

I.1. Energy spectra from two-point correlations

206

u in the x direction which is periodic with period L. Take the covariance of the Fourier coefcients, u( ) and u() where and denote two different wavenumbers and x and x denote two points separated in the x directions so that u()( ) = u 1 L
L

u(x) exp(x)dx
L

1 L

L L

u(x ) exp( x )dx (I.13)


1 = 2 L

L L

L L

u(x)u(x ) exp((x + x )dxdx

where denotes averaging over time; this equation corresponds to Eq. I.4 except the factor of two. Since we are performing a Fourier transform in x we must assume that this direction is homogeneous, i.e. all averaged turbulence quantities are independent of x and the two-point correlation is not dependent on x (or x ) but only on the separation distance x x , see discussion in connection to Eq. 10.5. Hence we replace x by y + x so that u()( ) = u = 1 L2 1 L
L L L L Lx Lx

u(x)u(x + x ) exp((x + (x + x ))dx 1 L


Lx Lx

dx dx

exp(( + )x)

B11 (x ) exp( x ))dx

(I.14) The second integral (in parenthesis) is the Fourier transform of the two-point correlation B11 , i.e. u()( ) = u 1 B11 (x ) L
L L

exp(( + )x))dx

(I.15)

where B11 denotes the Fourier transform of B11 (cf. I.12) and since it does not depends on the spatial coordinate it has been moved out of the integral. Furthermore, B11 is real and symmetric since B11 is real and symmetric. The remaining integral includes trigonometric function with wavelengths and . They are orthogonal functions, see Appendix C, and the integral of these functions is zero unless = . This integral in Eq. I.15 for = is evaluated as (see length of of k in Appendix C, Eq. C.13, and use 1 = cos(2x/L))
L

(1 |1 ) = ||1 ||2 =

cos2
L

2x L
L

dx (I.16) =L

x L = + sin 2 8 Equation I.15 can now be written

4x L

u()() = B11 (x) u

(I.17)

Hence, it is seen that the Fourier transform of a two-point correlation (in this example B11 (x1 ) ) indeed gives the corresponding one-dimensional energy spectrum (in this example E11 (1 ) = (())2 ). u

J. MTF270, Assignment 1: Reynolds averaged Navier-Stokes

207

J MTF270, Assignment 1: Reynolds averaged NavierStokes


J.1 Two-dimensional ow
You can do the assignment on your own or in a group of two. It is recommended (but A A the not required) that you use LTEX(an example of how to write in LTEXis available on the course www page). It is available on Linux. On Windows you can use, for example, Lyx (www.lyx.org) or MikTex (www.miktex.org) which are both free to download. Youll use data from a coarse DNS. Although some of the data are probably not fully accurate, in this exercise we consider the data to be exact. You will use Matlab to 2 2 2 read data les of the mean ow (1 , v2 , p) and turbulent quantities. (v1 , v2 , v3 , v1 v2 , v and ). You will analyze one of the following ows: Case 1: Flow over a 2D hill. Re = 10 595 ( = 9.44 105 , = 1) based on the bulk velocity in the channel and the hill height. Case 2: Flow over two small hills. Re = 10 595 ( = 9.44 105 , = 1) based on the bulk velocity bulk velocity in the channel and the height of the hill at the lower wall. Case 3: Flow in a diverging/converging section Re = 18 000 ( = 5.56 105 , = 1) based on the bulk velocity in the channel and the width of the the channel. Periodic boundary conditions are imposed in streamwise (x1 ) and spanwise (x3 ) directions in all ows. The work should be carried out in groups of two (if you want to work on you own that is also possible) . Contact the teacher to get a Case No. Download the data from http://www.tfd.chalmers.se/lada/comp turb model. At the wwwpage you can download a M-le (pl vect.m) which reads the data and plots the vector eld and the pressure contours. You must also download the function dphidx dy.m which computes the gradients. Make sure you put this function in the directory where you execute pl vect.m. The report, along with the Matlab les(s), should be submitted electronically at the Student Portal www.student.portal.se; the deadline can be found at the Student Portal.

J.2 Analysis
Study the ow. In which regions do you expect the turbulence to be important? Now lets nd out. The two-dimensional time-averaged Navier-Stokes for the x1 momentum reads (the density is set to one, i.e. = 1) 1 v1 v 1 v2 v p 2 v1 v 2 2 v1 v v + = + 1 + 1 2 x1 x2 x1 x2 x1 x2 x2 1 2 (J.1)

Recall that all the terms on the right-hand side represent x components of forces per unit volume.

J.2. Analysis

208

J.2.1 The momentum equations


2 The le pl vect.m loads the data le and plots the proles of v1 at some x stations, the velocity vector eld and a contour plot of velocity gradient 1 /x2 . Compute v all terms in Eq. J.1. You will need to compute the derivatives of e.g. v1 and p. In pl vect.m the function dphidx dy.m is used to compute 1 /x1 and 1 /x2 . v v Use this function to compute all derivatives that you need. Find two (or more) x1 2 locations (vertical grid lines) where the v1 stress is large and small, respectively. One way to nd these locations is to use the Matlab surf command.

Assignment 1.1. Plot the stresses along vertical grid lines at these two locations using the Matlab command plot(x,y). Please make sure that in your report the numbering on the axis and the text in the legend is large enough; you can use the command h1=gca; set(h1,fontsize,[20]) %the number 20 gives the fontsize The size of the labels and the title is similarly controlled by xlabel(x/H,fontsize,[20]) ylabel(y/H,fontsize,[20]) title(velocity,fontsize,[20]) Assignment 1.2. Plot also all terms in Eq. J.1. To enhance readability you may omit the small terms or use two plots per vertical grid line. Make also a zoom near the walls. For example, for a x y plot plot(u,y,linew,2) % linewidth=2 you may want to zoom in on y=[0 0.01] and u=[-0.1 0.4]; this is achieved by axis([-0.1 0.4 0 0.01]) The axis command can be used together with any plot, e.g. with surf and quiver. Which terms are negligible? Can you explain why they are negligible? What about the viscous terms: where do they play an important role? Which terms are non-zero at the wall? (you can show that on paper). So far we have looked at the v1 -momentum equation. The database corresponds to a two-dimensional ow. Now lets think of the forces as vectors. The normal stresses in the x1 x2 plane represent the vector N =
2 v 2 v1 , 2 x1 x2

(J.2)

and the corresponding vector due to the shear stresses reads S =


v1 v2 v v , 1 2 x2 x1

(J.3)

This is the force per unit volume ([N/m3 ]) by which the normal stresses affect the ow.

J.2. Analysis

209

Assignment 1.3. Plot the vector eld N to nd out some features. Zoom-in on interesting regions. Assignment 1.4. Plot also vector elds of the shear stress, S (see Eq. J.3), the pressure gradient and the viscous terms. Zoom up in interesting regions. Anything interesting? 2 When v2 reaches a maximum or a minimum along a grid line normal to the wall, what happens with the vector eld? Zoom-in on interesting regions. J.2.2 The turbulent kinetic energy equation The exact transport equation for for the turbulent kinetic energy, k, reads 2k (j k) = v + Pk + Dk xj xj xj v i Pk = vi vj xj The diffusion term, Dk needs to be modelled. Assignment 1.5. Plot the production term along the two grid lines. Explain why it is large at some locations and small at others. The production term consists of the sum of four terms, two of which involve the shear stress while the other include the normal stresses. Compare the contributions due the shear stress and the normal stresses. Assignment 1.6. Plot the dissipation and compare it with the production. Do you have local equilibrium (i.e. P k ) anywhere? J.2.3 The Reynolds stress equations The modelled transport equation for the Reynolds stresses can be written as
2 vi vj vk vi vj = + Pij + ij + Dij ij xk xk xk v v j v v i Pij = vi vk j k xk xk

(J.4)

(J.5)

The pressure-strain term, ij , and the diffusion term, Dij , need to be modelled. Here we use the models in Eqs. 11.83, 11.49, 11.82, 11.87 and 11.88. 1. In the damping function, f (see Eq. 11.84), xn denotes the distance to the nearest wall. If, for example, the lower wall is the closest wall to node (I, J), then xn = (x(I, J) x(I, 1))2 + (y(I, J) y(I, 1))2
1/2

(J.6)

2. ni,w denotes the unit normal vector of the wall to which the distance xn is computed. If we assume, again, that the lower wall is the closest wall to cell (I, J) and that the lower wall is horizontal, then ni,w = (0, 1). The compute ni,w for the general case, compute rst the vector which is parallel to the wall, si,w , and compute then ni,w from si,w (see Eq. J.11)

J.2. Analysis

210

3. The diffusion terms Dij and D can be modelled using the Generalized Gradient Diffusion Hypothesis GGDH of [159] Dij = xm cuk um
k vi vj xk

(J.7)

This diffusion model can cause numerical problems, and the GGDH is then replaced by a simple eddy viscosity model Dij = xm
t vi vj k xm

, t = C k 2 /

(J.8)

The following constants should be used: (c , c1 , c2 , c1w , c2w , k ) = (0.09, 1.5, 0.6, 0.5, 0.3, 1) Assignment 1.7. Choose two stresses. Plot the different terms in the equations for one vertical grid line fairly close to the inlet (not too close!). Use the simple eddy viscosity model for the turbulent diffusion term. If the gure becomes too crowdy, use two plots per vertical grid line or simply omit terms that are negligible. Try to explain why some terms are large and vice versa. Usually, a stress is large in locations where its production (or pressure-strain) term is large. Is that the case for you?
Assignment 1.8. Compute the stresses using the Boussinesq assumption, i.e vi vj = 2t sij + 2 (2k/3)ij where t = c k /. Compare the eddy-viscosity stresses with two of the Reynolds stresses from the database. Make also a zoom-in near walls.

When using the Boussinesq assumption the production of turbulent kinetic energy P k = 2t sij sij (J.9)

is always positive. The exact production of turbulent kinetic energy (see Eq. J.4) is usually positive. It can however become negative. Assignment 1.9. Compute the exact production in Eq. J.4 in the entire domain to investigate if the production is negative anywhere. If so, explain why. The reason why the eddy-viscosity production in Eq. J.9 must be positive is of course that neither t nor sij sij can go negative. Another way to explain this fact is v that the modelled Reynolds stress, vi vj , and the strain rate tensor, i /xj are parallel. To nd out to what degree the exact Reynolds stress and the strain rate are parallel, one can compute the eigenvectors. Assignment 1.10. Compute the eigenvalues and eigenvectors of the strain tensor, sij . The eigen values correspond to the normal strain in the direction of the eigenvectors (see Section 13). If the shear strains (i.e. the off-diagonal components) dominate, you will get eigenvectors in the direction /4 /2 and if the normal strains (i.e. the diagonal components) dominate the direction of the eigenvectors will be along the x1 and x2 axes (explain why!). Plot the eigenvectors as a vector eld. Our ow is 2D; thus we get two eigenvectors and two eigenvalues. Since the two eigenvectors are perpendicular to each other it is sufcient to plot one of them ( for example, the eigenvectors (/4, /4), (/4, /4), (/4, /4) and (/4, /4), all represent the same principal coordinate system). Zoom in on interesting regions.

J.3. Compute derivatives on a curvi-linear mesh


N ne nw se = (s1e , s2e ) P W sw se S ne = (n1e , n2e ) E

211

Figure J.1: Control volume. The velocity v1 is stored at the corners (ne, nw, . . . ). Coordinates x1 , x2 are given at the corners (ne, nw, ...).
Assignment 1.11. Compute the eigenvalues and eigenvectors of the Reynolds stresses, vi vj . The eigenvalues correspond to the normal stresses in the direction of the eigenvectors. Zoom in on interesting regions. In which regions are the eigenvectors of the Reynolds stress tensor and those of the strain tensor not parallel? This should indicate regions in which an eddy-viscosity model would perform poorly. Zoom in on interesting regions.

J.3 Compute derivatives on a curvi-linear mesh


In this appendix we describe how the derivatives on a curvi-linear grid are computed in the provided Matlab function dphidx dy.m. On a Cartesian grid it is more convenient to use the built-in Matlab function gradient, but the approach used below works for all meshes, including Cartesian ones. The data you have been given, x1 and x2 and all variables are stored at the grid points, i.e. at (x1,sw , x2,sw ), (x1,se , x2,se ), (x1,nw , x2,nw ) and (x1,ne , x2,ne ). When you need a variable, say v1 , at the center of the cell, compute it as v1,P = 1 (v1,sw + v1,se + v1,nw + v1,ne ) 4 (J.10)

Lets compute v1 /x1 . In order to do that we use Gauss law over a control volume centered at face e (dashed control volume in Fig. J.1). To compute v1 /x1 we use the x1 -component of the normal vector n = (n1 , n2 ), i.e. v1 dV = x1 v1 n1 dA
A

Assuming that v1 /x1 is constant in the volume V we obtain v1 1 = x1 V v1 n1 dA


A

J.3. Compute derivatives on a curvi-linear mesh

212

In discrete form we can write (see Fig. J.1) v1 x1 = 1 V (v1 n1 A)i =


i=e,n,w,w

1 {(v1 An1 )e + (v1 An1 )n + (v1 An1 )w + (v1 An1 )s } V J.3.1 Geometrical quantities It is useful to rst compute the unit vectors s along the control volume. For the east face, for example, we get s1e = x1,ne x1,se de x2,ne x2,se de (x1,ne x1,se )2 + (x2,ne x2,se )2

s2e =

de =

(note that the area of the east face Ae is equal to de since z = 1). The relation between the normal vector n, s and the unit vector in the z-direction sn =0

s z = n, gives us the normal vector for the east face as n1e = s2e n2e = s1e . (J.11)

K. MTF270, Assignment 2: LES

213

MTF270, Assignment 2: LES

You can do the assignment on your own or in a group of two. You will receive data from a DNS of fully developed ow in a channel. It is recommended (but the not A A required) that you use LTEX(an example of how to write in LTEXis available on the course www page). It is available on Linux. On Windows you can use, for example, Lyx (www.lyx.org) or MikTex (www.miktex.org) which are both free to download. The equations that have been solved are vi =0 xi vi p 1 2 vi (vi vj ) = i1 + + t xj xi Re xj xj

(K.1)

The Re number based on the friction velocity and the half channel width is Re = u h/ = 500 (h = = u = 1 so that = 1/Re ). A 96 96 96 mesh has been used. The streamwise, wall-normal and spanwise directions are denoted by x (x1 ), y (x2 ) and z (x3 ) respectively. The cell size in x and z directions are x = 0.0654 and z = 0.0164. Periodic boundary conditions were applied in the x and z direction (homogeneous directions). At the www-page (http://www.tfd.chalmers.se/lada/comp turb model) you nd data les with three instantaneous ow elds (statistically independent). The data les include the instantaneous variables u (v1 ), v (v2 ), w (v3 )and p (made nondimensional by u and ). Use Matlab to analyze the data. You nd a Matlab program at the www-page which reads the data and computes the mean velocity. The data les are Matlab binary les. Since the data les are rather large, it is recommended that you do all tasks using only data les 1. When everything works, then use also data les 2 and 3 averaging by use of the three les.

K.1 Task 2.1


We decompose the instantaneous variables in time-averaged and uctuating quantities as vi = vi + vi , v = p + p The symbol . denotes averaging in the homogeneous directions x and z. Note that in reality . always denote time averaging. It is only in this special academic test case where we have three homogeneous directions (x, z, t) where we can in addition to time averaging also can use x and z averaging. Compute the six stresses of the stress tensor, vi vj . Use the denition to compute the stresses, for example
2 v1 = (v1 v1 ) 2 = v1 2 v1 2 2 2 = v1 2 v1 v1 2

v1

+ v1

2 = v1 v1 2 .

Wait with analysis of the results till you have done next part.

K.2. Task 2.2

214

K.2 Task 2.2


Compute the production term and the pressure-strain terms
Pk = v1 v2

P11 P12 11 12 22

v1 y v1 = 2 v1 v2 y v1 = v2 v2 y v = 2 p 1 x v v = p 1 + p 2 y x v = 2 p 2 y

Do the production terms and the pressure-strain term appear as you had expected? (see the previous course MTF256) Now analyze the uctuations in the previous subsection. Which stresses do you think are symmetric with respect to the centerline? or anti-symmetric? Whats the reason? When averaging, we use only three time steps (three les). If we would use many more time steps or, in general, if we let T when time averaging, e.g. = lim 1 T 2T
+T

dt
T

then some of the stresses would be zero: which ones? Why?

K.3 Task 2.3


Plot v1 and v2 along x1 at two different x2 values at x3 = x3,max /2. 1. Filter v1 and v2 to get v and v2 using a 1D box-lter (in the x1 direction) with lter width = 2x1 (this corresponds to a test lter, see Eq. 17.26. Compare v1 and v2 with v1 and v2 . 2. Do the same thing again but with a lter width of = 4x1 (now you must derive the expression on your own!). Discuss the differences between no lter, = 2x1 and = 4x1 . In LES we almost always assume that the lter width is equal to the control volume (i.e. we use an implicit lter). Above, in Item 1 and 2 you have just carried out explicit ltering. Repeat Item 1, but now for a 2D lter (x1 and x3 direction); the formula for a 3D lter is given in Eq. 17.27. Compare the results along the same lines as in Item 1 and 2.

K.4. Task 2.4


cut-off

215

E() I

II
G

III
S

Figure K.1: Spectrum with cut-off.

K.4 Task 2.4


Compute the SGS stress 12 from the Smagorinsky model, which reads ij = 2sgs sij , sgs = (Cs f ) sij = 1 2 i v j v + xj xi
2

2ij sij s (K.2)

f = 1 exp(x+ /26) 2 The ltered velocities, vi , are taken from Task 2.3 using the 2D lter (in x1 and x3 ); we should really have used a 3D lter, but in order to keep it simple, we use the 2D lter. Before doing the 2D lter, look in the Lecture Notes how a 3D lter is done. The constant Cs = 0.1. Compare the SGS stress 12 with the resolved stress u v and compare the SGS viscosity with the physical one. Plot them across the channel. Any thoughts? As an alternative to the damping function, f , compute the lter length as = min{n, } (K.3)

where n is the distance to the nearest wall and = 0.4 (von K` rm` n constant). In this a a case you should set f = 1.

K.5. Task 2.5

216

K.5 Task 2.5


Repeat the Task 2.4, but now for the WALE model by [160], which reads gij = i 2 v , g = gik gkj xj ij 1 1 2 2 2 g + gji ij gkk sd = ij 2 ij 3
3/2 sd sd ij ij 2 (Cm ) 5/4 5/2 (ij sij ) + sd sd s ij ij

(K.4)

sgs =

with Cm = 0.106 which correspond to Cs = 0.1.

K.6 Task 2.6


Compute the dissipation =
vi vi xj xj

and plot across the channel. In LES we introduce a lter which is assumed to cut off the spectrum at c in the inertial region, see Fig. K.1. At cut-off, kinetic energy is extracted from the resolved ow by the SGS dissipation sgs . Since the cut-off is assumed to be located in the inertial sub-range (II), the SGS dissipation is at high Re numbers equal to the dissipation. Introduce a 2D lter (2x1 and 2x3 ) as in Tasks 2.3 & 2.4 and lter all velocities to obtain v1 , v2 and v3 . Compute the SGS stresses from the denition ij = vi vj vi vj and compute the SGS dissipation sgs = ij i v xj (K.6) (K.5)

Now, what is the relation between sgs and ? Considering the cascade process, what did you expect? Recall that when we do traditional Reynolds decomposition, the production term in the equation for turbulent kinetic energy appears as a sink term in the equation for the mean kinetic energy, see Eq. 8.30. This is the case also in LES, but now we have a decomposition into time-averaged ltered velocity, vi , resolved uctuation, vi , and SGS uctuation, vi , i.e.
vi = vi + vi = vi + vi + vi

(K.7)

1 Now we have three equations for kinetic energy: K = 2 vi vi , k = 1 vi vi and 2 1 ksgs = 2 vi vi . The ow of kinetic energy can be illustrated as in Fig. K.2 (cf. Fig. 20 in [73]) The transport equation for 1 vi vi is derived in [14]. (can be downloaded from 2 www.tfd.chalmers.se/lada). When deriving the ksgs equation, no decomposition into time-averaged, vi , and resolved uctuations, vi , is made. Hence the SGS dissipation in Eq. K.6 appears as an instantaneous production term in the equation for ksgs [74, 75, ] (can be downloaded from www.tfd.chalmers.se/lada). Plot (along x2 ), compare and discuss the four dissipations (see Fig. K.2)

K.7. Task 2.7

217

ksgs
s ij

s sg

s ij

s
vi xj

gs

vi vj

K
v x i v j x i j

vi v i xj xj

T
1 1 Figure K.2: Transfer of kinetic turbulent energy. K = 2 vi vi and k = 2 vi vi denote time-averaged kinetic and resolved turbulent kinetic energy, respectively. T denotes increase in internal energy, i.e. dissipation.

v1 v2 = sgs

v1 : dissipation (minus production) by resolved turbulence in the K equation x2 i v v i i v v sgs i : SGS dissipation term in the k equaxj xj xj xj tion. This is the modelled SGS dissipation; compare it with the exact SGS dissiv v pation ij xi + ij xi , cf. Eq. K.6 j j sgs

i i v v : viscous dissipation term in the k equation xj xj sgs 1 v x2 v1 sgs x2 v1 x2


2

: SGS dissipation term in the K equation.

K.7 Task 2.7


Above the ltered velocities were computed using the lter width = 2x1 . In dynamic models, we often dene the test lter as twice the usual lter, i.e. = 2. Use this denition (1D lter, i.e. = 4x1 ) to compute the dynamic Leonard stress L12 from the denition Lij = vi vj v i v j

(K.8)

and compare it (across the channel) with the resolved stress v1 v2 and the SGS stress 12 dened in Eq. K.5. Do you expect the magnitude of stresses to be similar?

K.8. Task 2.9

218

K.8 Task 2.9


2 2 What is the near-wall behavior of v1 , v1 and v2 (i.e., for v1 , what is m in v1 = m O(x2 )). In order to estimate m, plot the quantities in log-log coordinates. Do the quantities exhibit the near-wall behaviour that you expected?

K.9 Task 2.10


The two-point correlation for u B11 (x2 , m ) = 1 96 96
96 96 v1 (xI , x2 , xK )v1 (xI , x2 , xK m ) 1 3 1 3

(K.9)

I=1 K=1

where xK and m are the spanwise locations of the two points. Take advantage of the 3 fact that the ow is periodic, but be careful when integrating the correlation above in the x3 direction. We have 96 cells in the x3 direction. If, for example, m = 2x3 , and one of the points (x1 ) is at K = 1 then the other (x1 2) is at K = 95. 3 3 Plot the two-point correlation at a couple of x2 positions. When plotting two-point correlations, it is no point showing both symmetric parts; show only half of it (cf. the two-point correlations in Section 10.1 and Fig. D.1). Compute and plot the integral length scale, L1 , which is dened by L1 (x2 ) = 1
2 v1,rms 0

B11 (x2 , )d

(K.10)

Compute also L3 . Whats the difference between L1 and L3 ?

K.10 Task 2.11


The energy spectrum of any second moment can be obtained by taking the FFT of the corresponding two-point correlation. The energy spectrum of any second moment can be obtained by taking the FFT of the corresponding two-point correlation. You can nd some details on how to use Matlabs FFT in Appendix D. If you have computed the Fourier coefcients properly, the sum of all coefcients should give the energy. The reason is that the Fourier coefcients correspond to the energy spectrum, and if we integrate the energy spectrum over all wave numbers we get the total energy. When we take the FFT of Eq. K.9, for example, we get B11 (z ) = F F T (B11 ) and summation gives
N 2 v1,rms = 1

B11 /N

(K.11)

see Appendix D Plot the energy spectra at a couple of x2 locations. Conrm that Eq. K.11 is satised. When plotting two energy spectrum, it is no point showing both symmetric parts; show only half of it (cf. the energy spectrum in Fig. D.5 b).

K.11 Task 2.12


Think of an interesting turbulent quantity and plot it and analyze it!

L. MTF270, Assignment 4: Hybrid LES-RANS

219

MTF270, Assignment 4: Hybrid LES-RANS

In this exercise you will use data from a Hybrid LES-RANS for fully developed channel ow. The data are taken from [73] with isotropic forcing. The Re number based on the friction velocity and the half channel width is Re = u h/ = 2000. Eleven cells (twelve nodes including the boundary) are located in the URANS region at each wall. The matching line is located at x2 / = 0.0306. A 32 64 32 mesh has been used. The cell size in x1 and x3 directions are x1 = 0.3925 and x3 = 0.1964. Periodic boundary conditions were applied in the x1 and x3 direction (homogeneous directions). All data have been made nondimensional by u and .

L.1

Time history

At the www-page you nd a le u v time 4nodes hybrid.dat with the time history of v1 and v2 . The le has eight columns of v1 and v2 at four nodes: x2 / = 0.0032, x2 / = 0.0118, x2 / = 0.114 and x2 / = 0.57. Hence, two nodes are located in the URANS region and two nodes in the LES region. With u = 1 and = 1/Re = 1/2000, this correspond to x+ = 6.4, x+ = 23.6, x+ = 228 and 2 2 2 x+ = 1140, respectively. The sampling time step is 0.0049 (every time step). Use the 2 Matlab program pl time hybrid to load and plot the time history of v1 and v2 . Recall that the velocities have been scaled with the friction velocity u , and thus what you see is really v1 /u . The time history of v1 at x2 / = 0.0118 and x2 / = 0.57 are shown. To study the proles in closer detail, use the axis-command in the same way as when you studied the DNS data. Plot v1 for all four nodes. How does the time variation of v1 differ for different positions? Recall that the two points closest the wall are located in the URANS region and the other two are located in the LES region. In the URANS region the turbulent viscosity is much larger than in the LES region. How do you expect that the difference in t affects the time history of v1 . Does the time history of v1 behave as you expect? What about v2 ?

L.2

Mean velocity prole

After having performed a hybrid LES-RANS, we want to look at the time-averaged results. Use the le pl uvw hybrid.m to look at the mean velocity proles. pl uvw hybrid.m reads the instantaneous v1 eld and performs an averaging in the homogeneous direc tions x1 and x3 . The time averaged velocity prole is compared with the log prole (markers). The agreement is not as good as in the paper [73], but the reason is that you have used too few samples. There are four les with instantaneous values of v1 . Use more than one le to perform a better averaging.

L.3

Resolved stresses

We want to nd out how much of the turbulence that has been resolved and how much that has been modelled. Compute rst vmean (this quantity should be very small, but if you use only one le this may not be the case due to too few samples). Now compute v1 v2 . Heres an example how to do: uv=zeros(nj,1);

L.4. Turbulent kinetic energy


URANS region 2.5n[1 exp(0.2k 1/2 y/)] 0.09 2.5k 1/2 n[1 exp(0.014k 1/2y/)] LES region = = (V )1/3 1/2 0.07ksgs

220

Table L.1: Expressions for and T in the LES and URANS regions. n denotes the distance from the wall. for k=1:nk for j=1:nj for i=1:ni ufluct=u3d(i,j,k)-umean(j); vfluct=v3d(i,j,k)-vmean(j); uv(j)=uv(j)+ufluct*vfluct; end end end uv=uv/ni/nk; Plot it in a new gure (a gure is created by the command figure(2)). Compute also the resolved turbulent kinetic energy kres = 0.5 and plot it in a new gure.
2 2 2 v1 + v2 + v3

L.4

Turbulent kinetic energy

Now plot and compare the resolved and modelled turbulent kinetic energies. Which is largest? Which is largest in the URANS region and in the LES region, respectively? What about the sum? The magnitude of resolved and modelled turbulent kinetic energies is discussed in the last subsection in [73].

L.5

The modelled turbulent shear stress

We have computed the resolved shear stress. Lets nd the modelled shear stress. Load a le with the modelled turbulent kinetic energy, kT (le te1 hybrid.mat). Recall that = 1/2000. Compute the turbulent viscosity according to Table L.1 and do the usual averaging. When computing , you need the volume, V , of the cells. It is computed as V = (x1 x2 x3 ); x1 and x3 are constant and x2 is stored in the array dy(j), look at the beginning of the m-le. Plot T /. Where is it large and where is it small? (Recall that the URANS region is located in the rst ten cells). Is it smooth? Do you need more samples? If so, use more les. Compute the modelled shear stress from the Boussinesq assumption 12 = 2T s12 = T 1 v 2 v + x2 x1

Plot it and compare with the resolved shear stress. Are they smooth across the interface? (recall that forcing is used) Is the resolved shear stress large in the URANS region? Should it be large? Why/why not?

L.6. Turbulent length scales

221

L.6

Turbulent length scales

Compute and plot the turbulent length scales given in Table L.1. Plot the SGS and URAN S length scales in both regions. Which is largest? Any surprises? Compare them with x2 and (x1 x2 x3 )1/3 . One would expect that (x1 x2 x3 )1/3 < URAN S everywhere. Is this the case?

L.7

SAS turbulent length scales


v1 /x2 2 v /x2 1 2

Compute the 1D von K rm n length scale dened as a a LvK,1D = (L.1)

Note that you should take the derivatives of the averaged v1 velocity, i.e. of v1 . Zoom up near the wall. How does it behave (i.e. what is n in O(xn )? What should n be? 2 Compare that with the von K rm n length scale dened from instantaneous v1 , i.e. a a LvK,1D,inst = 1 /x2 v 2 v /x2 1 2 (L.2)

How does it compare with LvK,1D ? When were doing real 3D simulations, the rst and second derivative must be dened in 3D. One way of dening the von K rm n length scale in 3D is [141, 142] a a LvK,3D,inst = S U
0.5 2 0.5

S = (2t sij sij ) U =


2

(L.3)

vi vi xj xj xj xj
2

The second derivative is then computed as U


2

= + +

2 v1 2 v1 2 v1 2 + x2 + x2 x1 2 3 2 v2 2 v2 2 v2 2 + x2 + x2 x1 2 3 2 w3 2 w3 2 w3 + + x2 x2 x2 1 2 3

(L.4)
2

Plot the von K rm n length scale using Eqs. L.3 and L.4. Compare them with Eq. L.1. a a Whats the difference? What effect do the different length scales give for PSAS (i.e. T1 in Eq. 21.5) and what effect does it give to ? Another way to compute the second derivative is U 2 = + + 2v x2 2v x2 2w x2
2

+
2

2v y 2 2v y 2 2w 2 y

+
2

2v z 2 2v z 2 2w 2 z

+
2

+
2

(L.5)
2

L.7. SAS turbulent length scales

222

Plot and compare the von K rm n length scales using the second derivatives dened in a a Eqs. L.4 and L.5.

M. MTF270: Transformation of a tensor

223

M MTF270: Transformation of a tensor


The rotation of a vector from the xi coordinate system to xi reads (see, e.g., Chapter 1 in [19]) ui = bij uj (M.1) where bij denotes the cosine between the axis bij = cos (xi , xj ) In Fig. M.1, the bij is given by b11 = cos , b12 = cos = sin b21 = cos(/2 ) = sin , b22 = cos The relations bik bjk = bki bkj = ij are fullled as they should. For a second-order tensor, the transformation reads uij = bik bjm ukm As an example, set = /4. Equation M.3 gives b11 = 1/ 2, b12 = 1/ 2, b21 = 1/ 2, Inserting Eq. M.5 into Eq. M.4 gives u11 = b11 b11 u11 + b12 b11 u21 + b11 b12 u12 + b12 b12 u22 1 = (u11 u21 u12 + u22 ) 2 u12 = b11 b21 u11 + b12 b21 u21 + b11 b22 u12 + b12 b22 u22 1 = (u11 u21 + u12 u22 ) 2 u21 = b21 b11 u11 + b22 b11 u21 + b21 b12 u12 + b22 b12 u22 1 = (u11 + u21 u12 u22 ) 2 u22 = b21 b21 u11 + b22 b21 u21 + b21 b22 u12 + b22 b22 u22 1 = (u11 + u21 + u12 + u22 ) 2 x2 x2 x1 (M.6a) b22 = 1/ 2 (M.4) (M.3) (M.2)

(M.5)

(M.6b)

(M.6c)

(M.6d)

x1 Figure M.1: Transformation between the coordinate systems (x1 , x2 ) and (x1 , x2 ).

M.1. Rotation to principal directions

224

M.1 Rotation to principal directions


Consider fully developed ow in a channel, see Appendix B. The strain-rate tensor, sij , reads 1 v1 s11 = 0, s12 = , s21 = s12 , s22 = 0 (M.7) 2 x2 Assume that the x1 and x2 coordinates in Fig. M.1 correspond to the streamwise and wall-normal directions, respectively. Let the x1 x2 coordinate system denote the eigenvectors. The transformation from x1 x2 to x1 x2 reads sij = cik skm , where c11 = cos , c21 = cos = cos , c12 = cos(/2 ) = sin c22 = cos (M.9) cij = cos (xi , xj ) (M.8)

see Fig. M.1. It can be seen that the relation cji = bij is satised as it should. The eigenvectors for Eq. M.7 are any two orthogonal vectors with angles /4, 3/4. Let us choose /4 and 3/4 for which the transformation in Eq. M.8 reads ( = /4) s11 = c11 c11 s11 + c12 c11 s21 + c11 c12 s12 + c12 c12 s22 1 = (s11 + s21 + s12 + s22 ) 2 s12 = c11 c21 s11 + c12 c21 s21 + c11 c22 s12 + c12 c22 s22 1 = (s11 s21 + s12 + s22 ) 2 s21 = c21 c11 s11 + c22 c11 s21 + c21 c12 s12 + c22 c12 s22 1 = (s11 + s21 s12 + s22 ) 2 s22 = c21 c21 s11 + c22 c21 s21 + c21 c22 s12 + c22 c22 s22 1 = (s11 s21 s12 + s22 ) 2 The fully developed channel ow is obtained by inserting Eq. M.7 s11 = s12 , s12 = 0, s21 = 0, s22 = s21 (M.11) (M.10a)

(M.10b)

(M.10c)

(M.10d)

Since the diagonal elements are zero it conrms that the coordinate system x1 x2 with = /4 is indeed a principal coordinate system. The eigenvalues, (k) , of sij correspond to the diagonal elements in Eq. M.11, i.e. (1) s11 = s12 = 1 v1 , 2 x2 (2) s22 = s12 = 1 v1 2 x2 (M.12)

M.2. Transformation of a velocity gradient

225

M.2 Transformation of a velocity gradient


Consider the velocity gradient Aij = vi /xj . Apply the transformation from the x1 x2 system to the x1 x2 in Eqs. M.10a-M.10d with = /4 A11 = A12 A21 A22 1 (A11 + A21 + A12 + A22 ) 2 1 = (A11 A21 + A12 + A22 ) 2 1 = (A11 + A21 A12 + A22 ) 2 1 = (A11 A21 A12 + A22 ) 2

(M.13)

Insert Eq. M.9 with = /4 and replace Aij by the velocity gradient v1 v1 1 v1 = = , x1 x2 2 x2 v2 v2 1 v1 = = , x1 x2 2 x2 (M.14)

It can be seen that v1 /x1 = s11 and v1 /x2 + v2 /x1 = 2s12 = 0 (see Eqs. M.12 and M.14) as it should.

N. MTF270: Greens formulas

226

MTF270: Greens formulas

In this appendix we will derive Greens three formulas from Gauss divergence law. In the last section we will derive the analytical solution to the Poisson equation. The derivations below are partly taken from [161].

N.1 Greens rst formula


Gauss divergence law reads Fi dV = xi Fi ni dS
S

(N.1)

where S is the bounding surface of the volume, V , and ni is the normal vector of S pointing out of V . Replacing Fi by gives xi xi xi dV =
S

ni dS xi

(N.2)

The left side is re-written as xi xi = 2 + xi xi xi xi (N.3)

which inserted in Eq. N.2 gives


V

2 dV + xi xi

dV = xi xi

ni dS xi

(N.4)

This is Greens rst formula.

N.2 Greens second formula


Switching and in Eq. N.4 gives
V

2 dV + xi xi

dV = xi xi

ni dS xi

(N.5)

Subtract Eq. N.5 from N.4 gives


V

2 2 xi xi xi xi

dV =
S

xi xi

ni dS

(N.6)

This is Greens second formula.

N.3 Greens third formula


In Greens second formula, Eq. N.6, set (r) = 1 |r rP | (N.7)

As usual we are considering a volume V with bounding surface S and normal vector ni . Since function (r) is singular for r = rP , consider a small sphere in V , see Fig. N.1.

N.3. Greens third formula


ni S

227

V r S rP

n i

x2

x1 Figure N.1: Greens third formula. A volume V with bounding surface S with normal vector ni . In V there is a small sphere S located at rP with radius and normal vector n . i In Eq. N.6 we need the rst and the second derivative of . The rst derivative of 1/ri is computed as r/xi ri 1 = = 3 (N.8) 2 xi r r r since the derivative of a distance X is a vector along the increment of the distance, i.e. X/xi = Xi /X where X = |Xi |. The second derivative is obtained as 2 xi xi 1 r = xi 1 r3 ri r3 = ri + r ri xi 3ri r4 r 3ri xi r4 3 r2 3 = 3 + r r r4 1 r3 + (N.9) =0

= 3

To get the right side on the second line we used the fact that ri ri = r2 . Now we replace ri = r by r rP = ri rP,i in Eqs. N.8 and N.9 which gives xi 2 xi xi 1 |r rP | 1 |r rP | = =0 ri rP,i |r rP |3 (N.10)

P for ri = ri , i.e. for V excluding the sphere S , see Fig. N.1. Apply Greens second formula for this volume which has the bounding surfaces S and S with normal vectors ni (outwards) and n (inwards), respectively. We get i

V S

1 2 dV = |r rP | xi xi +
S

where the volume integral is taken over the volume V but excluding the sphere S , i.e. V S . Note the minus sign in front of the normal vector in the S integral; this is

P 1 ri ri 3 |r rP | |r rP | xi

P ri ri 1 |r rP |3 |r rP | xi

ni dS (N.11)

(n )dS i

N.3. Greens third formula

228

because the normal vector must point out of the volume V S , i.e. into the sphere, S . In the sphere the normal vector, n , is the direction from point rP , i.e. i n = i ri rP,i r rP = |r rP | |ri rP,i | (N.12)

where we have normalized the vector ri rP,i in order to make its length equal to one. The length of the vector ri rP,i is the radius of sphere S , i.e. |r rP | = dS = 2 = 2 sin dd (N.13) The surface area, dS, for sphere S can be expressed in spherical coordinates as (N.14)

where is the solid angle. Inserting Eqs. N.12, N.13 and N.14 in the last integral in Eq. N.11 gives IS =
S

ri rP,i + 2 2 xi

2 d =
S

+ (ri rP,i )

xi

d (N.15)

To re-write this integral we will use the mean value theorem for integrals. In one dimension this theorem simply states that for the integral of a function, g(x), over the interval [a, b], there exists (at least) on point for which the the relation
b a

g(x)dx = (a b)g(xQ )

(N.16)

holds, where xQ denotes a point on [a, b]. Applying this theorem to the integral in Eq. N.15 gives IS = (rQ )
S

d + rQ,i rP,i )

(rQ ) xi

d
S

(N.17)

where rQ rQ,i denotes a point on S . As we let Q P , the radius, , of sphere S goes to zero so that the integral in Eq. N.17 reads
0

lim IS = 4(rQ )

(N.18)

since

d = 4. Inserted in Eq. N.18 gives 2 1 dV V |r rP | xi xi P 1 1 ri ri 1 + ni dS + ni dS 4 S |r rP |3 4 S |r rP | xi (rP ) = 1 4

(N.19)

This is Greens third formula. The singularity 1/|r rP | in the volume integral in Eq. N.19 is not a problem. Consider a small sphere with radius r1 = |r rP | centered at point P . In spherical coordinates the volume element can then be expressed as
2 2 dV = r1 sin dr1 dd = r1 dr1 d

(N.20)

Hence it is seen that the volume element dV goes to zero faster than the singularity 1/|r rP |.

N.4. Analytical solution to Poissons equation

229

N.4 Analytical solution to Poissons equation


Poissons equation reads 2 =f xj xj (N.21)

where we assume that goes to zero at innity and that the right side is limited. Greens third formula, Eq. N.19, gives (rP ) = f (r) dV |r rP | V P 1 1 ri ri + ni dS + 4 S |r rP |3 4 1 4

1 ni dS |r rP | xi

(N.22)

We choose the volume as a large sphere with radius R. Using Eqs. N.12, N.13 and N.14, the rst surface integral can be written as 1 4
S P 1 ri ri ni dS = |r rP |3 4R2

ni ni dS =
S

1 4

d
S

(N.23)

using ni ni = 1. This integral goes to zero since 0 as R . The second integral in Eq. N.22 can be re-written using Eq. N.13, Gauss divergence law and Eq. N.21 as 1 4 1 1 ni dS = |r rP | xi 4R S 1 2 1 dV = = 4R S xi xi 4R ni dS xi f dV
V

(N.24)

This integral also goes to zero for large R since we have assumed that f is limited. Hence the nal form of Eq. N.22 reads (rP ) = 1 4 f (r) dV |r rP | (N.25)

This is the analytical solution to Poissons equation, Eq. N.21.

O. MTF270: Learning outcomes

230

MTF270: Learning outcomes

Week 1
1. How is the buoyancy term, gi , re-written in incompressible ow? 2. Given the transport equation for the temperature, , derive the transport equation for . Derive the transport equation for vi (Eq. 11.19). Discuss the physical meaning of the different terms. Which terms need to be modelled? 3. What is the expression for the total heat ux which appear in the equation?
4. Which terms in the vi vj equation need to be modelled? Explain the physical meaning of the different terms in the vi vj equation.

O. MTF270: Learning outcomes

231

Week 2
1. Discuss and show how the dissipation term, ij , is modelled.
2. How are the Reynolds stress, vi vj , and the turbulent heat ux, vi , modelled in the Boussinesq approach?

3. How is the turbulent diffusion ux for k modeled in the k model (dk )? What i,t is the nal expressions of the modelled diffusion terms in the k and equations? 4. How is the production term in the k equation modelled (Boussinesq)? 5. Derive the transport equation for (start from the k equation) 6. Use physical reasoning to derive a model for the diagonal components of the pressure-strain term (slow part).
2 7. The slow pressure-strain model reads ij,1 = c1 k vi vj 3 ij k . The
i anisotropy tensor is dened as bij = 2kj 1 ij . Show that for decaying grid 3 turbulence, the model for the slow pressure-strain model indeed acts as to make the turbulence more isotropic if c1 > 1.

v v

8. Derive the exact Poisson equation for the pressure uctuation, Eq. 11.57. For a Poisson equation 2 =f xj xj there exists an exact analytical solution (x) = 1 4 f (y)dy1 dy2 dy3 |y x| (O.1)

Using Eq. O.1, give the exact analytical solution for the pressure-strain term. What are the slow and rapid terms? Which two assumptions are used to simplify this equation? Give the simplied analytical solution for p (Eq. 11.61).

O. MTF270: Learning outcomes

232

Week 3
1. Derive the algebraic stress model (ASM). What main assumption is made? 2. Describe the physical effect of the pressure-strain term in the near-wall region. What sign must hence 22,1w have?
2 3. The modelled slow and rapid pressure strain term read ij,1 = c1 k vi vj 3 ij k 2 and ij,2 = c2 Pij 3 ij P k , respectively. Give the expression for the production terms, modelled pressure-strain terms and modelled dissipation terms for a simple shear ow. In some stress equations there is no production terms nor any dissipation term. How come? Which is the main source term (or sink term) in these equatiions?

4. Describe the effect of stable stratication and unstable stratication on turbulence 5. Consider buoyancy-dominated ow with x3 vertically upwards. The production term for the vi vj and the vi equations read
Gij = gi vj gj vi , Pi = vi vk

xk

respectively. Show that the Reynolds stress model dampens and increases the vertical uctuation in stable and unstable stratication, respectively. Show also that k in the k model is affected in the same way. 6. Consider streamline curvature for a streamline formed as a circular arc (convex curvature). Show that the turbulence is dampened if v /r > 0 and that it is enhanced if the sign of v /r is negative. Now consider a boundary layer where the streamlines are curved away from the wall (concave curvature). Show that the Reynolds stress model gives an enhanced turbulence production (as it should) because of positive feedback between the production terms. Show that the effect of streamline curvature in the k model is much smaller. 7. Consider stagnation ow. Show that in the Reynolds stress model, there is only a small production of turbulence whereas eddy-viscosity models (such as the k model) give a large production of turbulence.

O. MTF270: Learning outcomes

233

Week 4
1. What is a realizability constraint? There are two main realizability constraints on the normal and the shear stresses: which ones? Show that the Boussinesq assumption may give negative normal stresses. In which coordinate system is the risk largest for negative normal stresses? Derive an expression (2D) how to avoid negative normal stresses by reducing the turbulent viscosity (Eq. 13.12). 2. What is the two-component limit? What requirement does it put on the pressurestrain models? Show that the standard IP model and the Rotta model do not satisfy this requirement. 3. What is a non-linear eddy-viscosity model? When formulating a non-linear model, the anisotropy tensor aij = 2t sij /k is often used. Show the three rst terms (S 2 , 2 , S) in the non-linear model in the lecture notes. Show that each term is traceless and symmetric (as aij ). 4. What is the modeling idea of the V2F model? Which equations are solved?
2 5. The transport equation for v2 reads (the turbulent diffusion terms are modelled)

1 v2 v 2 v 2 v2 v 2 ( + t ) 2 2v2 p /x2 22 + = x1 x2 x2 x2
22

Show how this equation is re-written in the V2F model.

O. MTF270: Learning outcomes

234

Week 5
1. The f equation in the V2F model reads L2 2f 22 1 2 f = k T x2
2 2 v2 k 3

k ,

k 3/2

Show how the magnitude of the right side and L affect f . How does f enter into 2 the v2 equation? What it the physical meaning of f ? Show that far from the 2 2 walls, the V2F model (i.e. the f and the v2 equation) returns to the v2 equation in the Reynolds stress model. 2. Consider the boundary condition for the f equation. Very close to the wall, 2 how is the v2 equation simplied? Use this equation to derive a wall boundary condition for f . 3. What does the acronym SST mean? The SST model is a combination of the k and the k model. In which region is each model being used and why? How is expressed in k and ? 4. Derive a transport equation for from the k and transport equations. In the SST model, a blending function F1 is used; what does this function do? In which region is each model being used and why? 5. Describe the shear stress limiter. Show that the eddy-viscosity assumption gives too high shear stress in APG since P k / 1. 6. Show the difference between volume averaging (ltering) in LES and timeaveraging in RANS. 7. Consider the spatial derivative of the pressure in the ltered Navier-Stokes: show that the derivative can be moved outside the ltering integral (it gives an additional second-order term). The ltered non-linear term has the form vi vj xj Show that it can be re-written as i vj v xj giving an additional term on the right side. 8. Consider a 1D nite volume grid. Carry out a second ltering of v at node I and show that v I = vI . 9. Consider the energy spectrum. Show the three different regions (the large energycontaining scales, the 5/3 range and the dissipating scales). Where should the cut-off be located? What does cut-off mean? Show where the SGS scales, grid (i.e resolved) scales and the cut-off, c are located in the spectrum. ij (vi vj ) + (i vj ) = v xj xj xj

O. MTF270: Learning outcomes

235

Week 6
1. Show how a sinus wave sin(c x) corresponding to cut-off is represented on a grid with two and four nodes, respectively. How is c related to the grid size x for these cases? 2. Show the Smagorinsky model. 3. Discuss the energy path in connection to the source and sink terms in the k, K and ksgs computed from the energy spectrum? and the ksgs equations. How are k 4. What is a test lter? Grid and test lter Naiver-Stokes equation and derive the relation (O.2) vi vj v i v j + ij = Tij Draw an energy spectrum and show which wavenumber range k, ksgs , ksgs,test cover.

5. Formulate the Smagorinsky model for the grid lter SGS stress, ij , and the test lter SGS stress, Tij . Use Eq. O.2 and derive the relation 1 Lij ij Lkk = 2C 3
2

ss | s | s ij 2 ||ij

This equation is a tensor equation for C. Use this relation and derive the nal expression for the dynamic coefcient, C, Eq. 17.33. 6. Show that when a rst-order upwind schemes is used for the convection term, an additional diffusion term and dissipation terms appear because of a numerical SGS viscosity. 7. What is a scale-similarity model? How are the cross terms, Cij , modelled? What is the exact form of the the cross term, Cij ? 8. Show that the Bardina scale-similarity model is not Galilean invariant. How is the the model modied to make it Galilean invariant? 9. What are the four main differences between a RANS nite volume CFD code and a LES nite volume CFD code? What do you need to consider in LES when you want to compute time-averaged quantities?

O. MTF270: Learning outcomes

236

Week 7
1. When doing LES, how ne does the mesh need to be in the wall region be? Why does it need to be that ne?
2. What is DES? The length scale in the RANS S-A model reads t ; how is it d computed in the corresponding DES model? How is the length scale computed in a k two-equation DES model? 2

3. The modied (reduced) length scale in two-equation DES models can be introduced in either the dissipation term in the k equation and/or in the expression for the turbulent viscosity. Show the two different methods. What is the effect on the modelled, turbulent quantities? 4. Describe hybrid LES-RANS based on a one-equation model. What is the difference between DES and hybrid LES-RANS? 5. Describe URANS. How is the instantaneous velocity decomposed? What turbulence models are used? What is scale separation? 6. Describe the SAS model. How is the von K rm n length scale dened? An a a additional source term is introduced in the equation: what is the form of this term? What is the object of this term? When is it large and small, respectively? 7. Give a short description of the method to generate synthetic turbulent inlet uctuations. What form on the spectrum is assumed? How are the maximum and minimum wavelengths, max , min , determined. With this method, the generated shear stress is zero: why? How is the correlation in time achieved?

P. References

237

References
[1] L. Davidson, P.V. Nielsen, and A. Sveningsson. Modications of the v 2 f model for computing the ow in a 3D wall jet. In K. Hanjali , Y. Nagano, and c M.J. Tummers, editors, Turbulence Heat and Mass Transfer 4, pages 577584, New York, Wallingford (UK), 2003. begell house, inc. [2] J.N. Reddy. An Introduction to Continuum Mechanics. Cambridge University Press, New York, 2008. [3] S. Toll and M. Ekh. Mechanics of solids & uids. Part I: Fundamentals. Report, Div. of Material and Computational Mechanics, Dept. of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, 2011. o [4] H. Abedi. Aerodynamic loads on rotor blades. MSc Thesis 2011:18, Division of Fluid Dynamics, Department of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, 2011. o [5] L. Davidson. An introduction to turbulence models. Technical Report 97/2, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1997. [6] J. Wojtkowiak and C.O. Popiel. Inherently linear annular-duct-type laminar owmeter. Journal of Fluids Engineering, 128(1):196198, 2006. [7] S.B. Pope. Turbulent Flow. Cambridge University Press, Cambridge, UK, 2001. [8] H. Tennekes and J.L. Lumley. A First Course in Turbulence. The MIT Press, Cambridge, Massachusetts, 1972. [9] P. Bradshaw. Turbulence. Springer-Verlag, Berlin, 1976.

[10] S. Hoyas and J. Jim nez. Scaling of the velocity uctuations in turbulent chane nels up to Re = 2003. Physics of Fluids A, 18(011702), 2006. [11] S. Hoyas and J. Jim nez. http://torroja.dmt.upm.es/ftp/channels/data/statistics/. e 2006. [12] Th. von K rm n. Mechanische ahnlichkeit und Turbulenz. Nachrichten von a a der Gesellschaft der Wissenchaften zu G ttingen, Fachgruppe 1 (Mathematik), o 5:5876, 1930. [13] P.R. Spalart. Direct simulation of a turbulent boundary layer up to R = 1410. Journal of Fluid Mechanics, 187:6198, 1988. [14] L. Davidson. Transport equations in incompressible URANS and LES. Report 2006/01, Div. of Fluid Dynamics, Dept. of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, 2006. o [15] K. Hanjali . Advanced turbulence closure models: A view of current status and c future prospects. Int. J. Heat and Fluid Flow, 15:178203, 1994. [16] D.C. Wilcox. Turbulence Modeling for CFD. DCW Industries, Inc., 5354 Palm Drive, La Ca ada, California 91011, 2 edition, 1998. n

P. References

238

[17] K. Hanjali and B.E. Launder. A Reynolds stress model of turbulence and its c application to thin shear ows. Journal of Fluid Mechanics, 52:609638, 1972. [18] J.C. Rotta. Statistische theorie nichthomogener turbulenz. Zeitschrift f r Physik, u 129:547572, 1951. [19] G.E. Mase. Continuum Mechanics. Schaums Outline Series. McGraw-Hill, 1970. [20] S. Fu, B.E. Launder, and D.P. Tselepidakes. Accomodating the effects of high strain rates in modelling the pressure-starin correlation. Technical Report TFD/87/5, UMIST, Mechanical Engineering Dept., 1987. [21] B.E. Launder, G.J. Reece, and W. Rodi. Progress in the development of a Reynolds-stress turbulence closure. Journal of Fluid Mechanics, 68(3):537 566, 1975. [22] D. Naot, A. Shavit, and M. Wolfshtein. Interactions between components of the turbulent velocity correlation tensor. Israel J. Techn., 8:259, 1970. [23] B. E. Launder. An introduction to single-point closure methodology. In T.B. Gatski, M.Y. Hussaini, and J.L. Lumpley, editors, Simulation and Modeling of Turbulent Flows, New York, 1996. Oxford University Press. [24] M.M. Gibson and B.E. Launder. Ground effects on pressure uctuations in the atmospheric boundary layer. Journal of Fluid Mechanics, 86:491511, 1978. [25] S.B. Pope. A more general effective-viscosity hypothesis. Journal of Fluid Mechanics, 472:331340, 1975. [26] P. Bradshaw. Effects of streamline curvature on turbulent ow. Agardograph progress no. 169, AGARD, 1973. [27] W. Rodi and G. Scheuerer. Calculation of curved shear layers with two-equation turbulence models. Physics of Fluids, 26:14221435, 1983. [28] B.E. Thompson and J.H. Whitelaw. Characteristics of a trailing-edge ow with turbulent boundary-layer separation. Journal of Fluid Mechanics, 157:305326, 1985. [29] S. Wallin and A.V. Johansson. Modelling streamline curvature effects in explicit algebraic reynolds stress turbulence models. International Journal of Heat and Fluid Flow, 23(5):721730, 2002. [30] L. Davidson. Prediction of the ow around an airfoil using a Reynolds stress transport model. Journal of Fluids Engineering, 117:5057, 1995. [31] P.A. Durbin. On the k 3 stagnation point anomaly. International Journal of Heat and Fluid Flow, 17:8990, 1996. [32] J. Hult. Cartesiska tensorer: Ett praktiskt konspr k. Almqvist & Wiksell, a 1972. [33] J.L. Lumley. Computational modeling of turbulent ows. Advances in Applied Mechanics, 18:123176, 1978.

P. References

239

[34] T.J. Craft and B.E. Launder. A Reynolds stress closure designated for complex geometries. International Journal of Heat and Fluid Flow, 17(3):245254, 1996. [35] T.J. Craft, B.E. Launder, and K. Suga. Prediction of turbulent transitional phenomena with a nonlinear eddy-viscosity model. International Journal of Heat and Fluid Flow, 18:1528, 1997. [36] T.B. Gatski and C.G. Speziale. On explicit algebraic stress models for complex turbulent ows. Journal of Fluid Mechanics, 154:5978, 1993. [37] T.-H. Shih, J. Zhu, and J.L. Lumley. A new Reynolds stress algebraic equation model. Comput. Methods Appl. Engng., 125:287302, 1995. [38] S. Wallin and A.V. Johansson. A new explicit algebraic Reynolds stress model for incompressible and compressible turbulent ows. Journal of Fluid Mechanics, 403:89132, 2000. [39] P.A. Durbin. Near-wall turbulence closure modeling without damping functions. Theoretical and Computational Fluid Dynamics, 3:113, 1991. [40] P.A. Durbin. Application of a near-wall turbulence model to boundary layers and heat transfer. International Journal of Heat and Fluid Flow, 14(4):316 323, 1993. [41] A. Sveningsson. Analysis of the performance of different v 2 f turbulence models in a stator vane passage ow. Thesis for licentiate of engineering, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, o Sweden, 2003. [42] K. Hanjalic, M. Popovacx, and M. Habziabdic. A robust near-wall elliptic relaxation eddy-viscosity turbulence model for CFD. International Journal of Heat and Fluid Flow, 2004. [43] F-S Lien and G. Kalitzin. Computations of transonic ow with the v2f turbulence model. International Journal of Heat and Fluid Flow, 22(1):5361, 2001. [44] A. Sveningsson and L. Davidson. Assessment of realizability constraints in v 2 f turbulence models. International Journal of Heat and Fluid Flow, 25(5):785 794, 2004. [45] A. Sveningsson and L. Davidson. Computations of ow eld and heat transfer in a stator vane passage using the v 2 f turbulence model. Paper GT2004-53586, ASME TURBO EXPO 2004, Vienna, 2004. [46] A. Sveningsson and L. Davidson. Computations of ow eld and heat transfer in a stator vane passage using the v 2 f turbulence model. Journal of Turbomachinery, 127(3):627634, 2005. [47] A. Sveningsson. Turbulence Transport Modelling in Gas Turbine Related Applications. PhD thesis, Div. of Fluid Dynamics, Dept. of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, 2006. o [48] F.R. Menter. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA Journal, 32:15981605, 1994.

P. References

240

[49] D.C. Wilcox. Reassessment of the scale-determining equation. AIAA Journal, 26(11):12991310, 1988. [50] F.R. Menter. Inuence of freestream values on k turbulence model predictions. AIAA Journal, 30(6):16571659, 1992. [51] J. Bredberg. On two-equation eddy-viscosity models. Report 01/8, can be downloaded from www.tfd.chalmers.se/lada/allpaper.html, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, 2001. [52] P. Bradshaw, D.H. Ferriss, and N.P. Atwell. Calculation of boundary-layer development using the turbulent energy equation. Journal of Fluid Mechanics, 28:593616, 1967. [53] F.R. Menter, M. Kuntz, and R. Langtry. Ten years of industrial experience of the SST turbulence model. In K. Hanjali , Y. Nagano, and M.J. Tummers, editors, c Turbulence Heat and Mass Transfer 4, pages 624632, New York, Wallingford (UK), 2003. begell house, inc. [54] J. Smagorinsky. General circulation experiments with the primitive equations. Monthly Weather Review, 91:99165, 1963. [55] P. Moin and J. Kim. Numerical investigation of turbulent channel ow. Journal of Fluid Mechanics, 118:341377, 1982. [56] W.P. Jones and M. Wille. Large eddy simulation of a jet in a cross-ow. In 10th Symp. on Turbulent Shear Flows, pages 4:1 4:6, The Pennsylvania State University, 1995. [57] J.O. Hinze. Turbulence. McGraw-Hill, New York, 2nd edition, 1975. [58] H. Schlichting. Boundary-Layer Theory. McGraw-Hill, New York, 7 edition, 1979. [59] M. Germano, U. Piomelli, P. Moin, and W.H. Cabot. A dynamic subgrid-scale eddy viscosity model. Physics of Fluids A, 3:17601765, 1991. [60] Y. Zang, R.L. Street, and J.R. Koseff. A dynamic mixed subgrid-scale model and its application to turbulent recirculating ows. Physics of Fluids A, 5:3186 3196, 1993. [61] D.K. Lilly. A proposed modication of the Germano subgrid-scale closure method. Physics of Fluids A, 4:633635, 1992. [62] J.P. Boris, F.F. Grinstein, E.S. Oran, and R.L. Kolbe. New insights into large eddy simulation. Fluid Dynamic Research, 10:199228, 1992. [63] J. Bardina, J.H. Ferziger, and W.C. Reynolds. Improved subgrid scale models for large eddy simulation. AIAA 80-1357, Snomass, Colorado, 1980. [64] C.G. Speziale. Galilean invariance of subgrid-scale stress models in the largeeddy simulation of turbulence. Journal of Fluid Mechanics, 156:5562, 1985. [65] M. Germano. A proposal for a redenition of the turbulent stresses in the ltered Navier-Stokes equations. Phys. Fluids A, 7:23232324, 1986.

P. References

241

[66] L. Davidson. A dissipative scale-similarity model. In V. Armenio, B. Geurts, and J. Fr hlich, editors, DLES7: Direct and Large-Eddy Simulations 7, volume 13 o of ERCOFTAC series, pages 261266, 2010. [67] C. Hirsch. Numerical Computation of Internal and External Flows: Fundamentals of Numerical Discretization, volume 1. John Wiley & Sons, Chichester, UK, 1988. [68] L. Davidson. Hybrid LES-RANS: back scatter from a scale-similarity model used as forcing. Phil. Trans. of the Royal Society A, 367(1899):29052915, 2009. [69] P. Emvin. The Full Multigrid Method Applied to Turbulent Flow in Ventilated Enclosures Using Structured and Unstructured Grids. PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, o 1997. [70] L. Davidson and S.-H. Peng. Hybrid LES-RANS: A one-equation SGS model combined with a k model for predicting recirculating ows. International Journal for Numerical Methods in Fluids, 43:10031018, 2003. [71] S.V. Patankar. Numerical Heat Transfer and Fluid Flow. McGraw-Hill, New York, 1980. [72] H.K. Versteegh and W. Malalasekera. An Introduction to Computational Fluid Dynamics - The Finite Volume Method. Longman Scientic & Technical, Harlow, England, 1995. [73] L. Davidson and M. Billson. Hybrid LES/RANS using synthesized turbulence for forcing at the interface. International Journal of Heat and Fluid Flow, 27(6):10281042, 2006. [74] L. Davidson. Large eddy simulation: A dynamic one-equation subgrid model for three-dimensional recirculating ow. In 11th Int. Symp. on Turbulent Shear Flow, volume 3, pages 26.126.6, Grenoble, 1997. [75] L. Davidson. Large eddy simulations: A note on derivation of the equations for the subgrid turbulent kinetic energies. Technical Report 97/11, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1997. [76] A. Sohankar, L. Davidson, and C. Norberg. Large eddy simulation of ow past a square cylinder: Comparison of different subgrid scale models. Journal of Fluids Engineering, 122(1):3947, 2000. [77] A. Sohankar, L. Davidson, and C. Norberg. Erratum. Journal of Fluids Engineering, 122(3):643, 2000. [78] S. Krajnovi and L. Davidson. A mixed one-equation subgrid model for largec eddy simulation. International Journal of Heat and Fluid Flow, 23(4):413425, 2002. [79] S. Krajnovi and L. Davidson. Large eddy simulation of the ow around a bluff c body. AIAA Journal, 40(5):927936, 2002.

P. References

242

[80] S. Krajnovi . Large Eddy Simulations for Computing the Flow Around Vehic cles. PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, Sweden, 2002. o [81] J. Pallares and L. Davidson. Large-eddy simulations of turbulent ow in a rotating square duct. Physics of Fluids, 12(11):28782894, 2000. [82] J. Pallares and L. Davidson. Large-eddy simulations of turbulent heat transfer in stationary and rotating square duct. Physics of Fluids, 14(8):28042816, 2002. [83] S. Krajnovi and L. Davidson. Numerical study of the ow around the busc shaped body. Journal of Fluids Engineering, 125:500509, 2003. [84] S. Krajnovi and L. Davidson. Flow around a simplied car. part I: Large eddy c simulations. Journal of Fluids Engineering, 127(5):907918, 2005. [85] S. Krajnovi and L. Davidson. Flow around a simplied car. part II: Understandc ing the ow. Journal of Fluids Engineering, 127(5):919928, 2005. [86] S. Krajnovi and L. Davidson. Inuence of oor motions in wind tunnels on c the aerodynamics of road vehicles. Journal of Wind Engineering and Industrial Aerodynamics, 92(5):677696, 2005. [87] S. Dahlstr m and L. Davidson. Large eddy simulation applied to a higho Reynolds ow around an airfoil close to stall. AIAA paper 2003-0776, 2003. [88] S. Dahlstr m. Large Eddy Simulation of the Flow Around a High-Lift Airfoil. o PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, Sweden, 2003. o [89] J. Ask and L. Davidson. Flow and dipole source evaluation of a generic SUV. Journal of Fluids Engineering, 132(051111), 2010. [90] H. Hemida and S. Krajnovi . LES study of the impact of the wake structures on c the aerodynamics of a simplied ICE2 train subjected to a side wind. In Fourth International Conference on Computational Fluid Dynamics (ICCFD4), 10-14 July, Ghent, Belgium, 2006. [91] H. Hemida and S. Krajnovi . Numerical study of the unsteady ow structures c around train-shaped body subjected to side winds. In ECCOMAS CFD 2006, 5-8 September, Egmond aan Zee, The Netherlands, 2006. [92] S. Krajnovi . Numerical simulation of the ow around an ICE2 train under the c inuence of a wind gust. In International Conference on Railway Engineering 2008 (IET ICRE2008), 25-28 March, Hong Kong, China, 2008. [93] S.-H. Peng and L. Davidson. On a subgrid-scale heat ux model for large eddy simulation of turbulent thermal ow. International Journal of Heat and Mass Transfer, 45:13931405, 2002. [94] S.-H. Peng and L. Davidson. Large eddy simulation for turbulent buoyant ow in a conned cavity. International Journal of Heat and Fluid Flow, 22:323331, 2001.

P. References

243

[95] S.-H. Peng. Modeling of Turbulent ow and Heat Transfer for Building Ventilation. PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, 1998. o [96] D.G. Barhaghi and L. Davidson. Natural convection boundary layer in a 5:1 cavity. Physics of Fluids, 19(125106), 2007. [97] D. Barhaghi. A Study of Turbulent Natural Convection Boundary Layers Using Large-Eddy Simulation. PhD thesis, Div. of Fluid Dynamics, Dept. of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, 2007. o [98] D.G. Barhaghi, L. Davidson, and R. Karlsson. Large-eddy simulation of natural convection boundary layer on a vertical cylinder. International Journal of Heat and Fluid Flow, 27(5):811820, 2006. [99] D.G. Barhaghi and L. Davidson. Les of mixed convection boundary layer between radiating parallel plates. In 5th International Symposium on Turbulence, Heat and Mass Transfer, September 25-29, 2006, Dubrovnik, Croatia, 2006. [100] S.R. Robinson. Coherent motions in the turbulent boundary layer. Annual Review of Fluid Mechanics, 23:601639, 1991. [101] L. Davidson. Large eddy simulations: how to evaluate resolution. International Journal of Heat and Fluid Flow, 30(5):10161025, 2009. [102] L. Davidson. How to estimate the resolution of an LES of recirculating ow. In M. V. Salvetti, B. Geurts, J. Meyers, and P. Sagaut, editors, ERCOFTAC, volume 16 of Quality and Reliability of Large-Eddy Simulations II, pages 269 286. Springer, 2010. [103] L. Davidson. Inlet boundary conditions for embedded LES. In First CEAS European Air and Space Conference, 10-13 September, Berlin, 2007. [104] L. Davidson, D. Cokljat, J. Fr hlich, M.A. Leschziner, C. Mellen, and W. Rodi, o editors. LESFOIL: Large Eddy Simulation of Flow Around a High Lift Airfoil, volume 83 of Notes on Numerical Fluid Mechanics. Springer Verlag, 2003. [105] C.P. Mellen, J. Fr hlich, and W. Rodi. Lessons from LESFOIL project on large o eddy simulation of ow around an airfoil. AIAA Journal, 41(4):573581, 2003. [106] S. Johansson, L. Davidson, and E. Olsson. Numerical simulation of vortex shedding past triangular cylinders at high reynolds numbers using a k turbulence model. International Journal for Numerical Methods in Fluids, 16:859878, 1993. [107] S. Perzon and L. Davidson. On CFD and transient ow in vehicle aerodynamics. In SAE Technical Paper 2000-01-0873, Detroit, 2000. [108] S. Perzon and L. Davidson. On transient modeling of the ow around vehicles using the Reynolds equation. In J.-H. Wu, Z.-J. Zhu, F.-P. Jia, X.-B. Wen, and W. Hu, editors, ACFD 2000, Bejing, 2000. [109] T.H. Shih, J. Zhu, and J. L. Lumley. A realizable Reynolds stress algebraic equation model. Technical memo 105993, NASA, 1993.

P. References

244

[110] P. Spalart and S. Allmaras. A one equation turbulence model for aerodynamic ows. AIAA paper 92-0439, Reno, NV, 1992. [111] P.R. Spalart, W.-H. Jou, M. Strelets, and S.R. Allmaras. Comments on the feasability of LES for wings and on a hybrid RANS/LES approach. In C. Liu and Z. Liu, editors, Advances in LES/DNS, First Int. conf. on DNS/LES, Louisiana Tech University, 1997. Greyden Press. [112] F.R. Menter, M. Kuntz, and R. Bender. A scale-adaptive simulation model for turbulent ow prediction. AIAA paper 20030767, Reno, NV, 2003. [113] A.K. Travin, M. Shur, , M. Strelets, and P.R. Spalart. Physical and numerical upgrades in the detached-eddy simulations of complex turbulent ows. In R.Friederich and W.Rodi, editors, Fluid Mechanics and its Applications. Advances in LES of Complex Flows, volume 65 of Euromech Colloquium 412. Fluid Mechanics and its Applications. Advances in LES of Complex Flows, pages 239254. Academic Publishers, Dordrecht, 2002. [114] C. De Langhe, B. Merci, K. Lodeer, and E. Dick. Hybrid RANS-LES modelling with the renormalization group. In K. Hanjali , Y. Nagano, and M.J. c Tummers, editors, Turbulence Heat and Mass Transfer 4, pages 697704, New York, Wallingford (UK), 2003. begell house, inc. [115] J.C. Kok, H.S. Dol, B. Oskam, and H. van der Ven. Extra-large eddy simulation of massively separated ows. AIAA paper 2004-264, Reno, NV, 2004. [116] J. Yan, C. Mocket, and F. Thiele. Investigation of alternative length scale substitutions in detached-eddy simulation. Flow, Turbulence and Combustion, 74:85 102, 2005. [117] S.S. Girimaji. Partially-Averaged Navier-Stokes model for turbulence: A Reynolds-averaged Navier-Stokes to direct numerical simulation bridging method. Journal of Fluids Engineering, 73(2):413421, 2006. [118] J. Ma, S.-H. Peng, L. Davidson, and F. Wang. A low Reynolds number partiallyaveraged Navier-Stokes model for turbulence. In 8th International ERCOFTAC Symposium on Engineering Turbulence, Modelling and Measurements, Marseille, France, 9-11 June, 2010. [119] J. Ma, S.-H. Peng, L. Davidson, and F. Wang. A low Reynolds number variant of Partially-Averaged Navier-Stokes model for turbulence. International Journal of Heat and Fluid Flow, 32:652669, 2011. 10.1016/j.ijheatuidow.2011.02.001. [120] R. Schiestel and A. Dejoan. Towards a new partially integrated transport model for coarse grid and unsteady turbulent ow simulations. Theoretical and Computational Fluid Dynamics, 18:443468, 2005. [121] K. Abe, T. Kondoh, and Y. Nagano. A new turbulence model for predicting uid ow and heat transfer in separating and reattaching ows - 1. Flow eld calculations. Int. J. Heat Mass Transfer, 37:139151, 1994. [122] P. Batten, U. Goldberg, and S. Chakravarthy. Interfacing statistical turbulence closures with large-eddy simulation. AIAA Journal, 42(3):485492, 2004.

P. References

245

[123] U. Piomelli, E. Balaras, H. Pasinato, K.D. Squire, and P.R. Spalart. The innerouter layer interface in large-eddy simulations with wall-layer models. International Journal of Heat and Fluid Flow, 24:538550, 2003. [124] L. Davidson and S. Dahlstr m. Hybrid LES-RANS: An approach to make LES o applicable at high Reynolds number. International Journal of Computational Fluid Dynamics, 19(6):415427, 2005. [125] J. Larsson, F.S. Lien, and E. Yee. The articial buffer layer and the effects of forcing in hybrid LES/RANS. International Journal of Heat and Fluid Flow, 28(6):14431459, 2007. [126] F.R. Menter and M. Kuntz. Adaption of eddy-viscosity turbulence models to unsteady separated ows behind vehicles. In Rose McCallen, Fred Browand, and James Ross, editors, The Aerodynamics of Heavy Vehicles: Trucks, Buses, and Trains, volume 19 of Lecture Notes in Applied and Computational Mechanics. Springer Verlag, 2004. [127] M. Strelets. Detached eddy simulation of massively separated ows. AIAA paper 20010879, Reno, NV, 2001. [128] X. Xiao, J.R. Edwards, and H.A. Hassan. Inow boundary conditions for LES/RANS simulations with applications to shock wave boundary layer interactions. AIAA paper 20030079, Reno, NV, 2003. [129] F. Hamba. An approach to hybrid RANS/LES calculation of channel ow. In W. Rodi and N. Fueyo, editors, Engineering Turbulence Modelling and Experiments 5, pages 297305. Elsevier, 2003. [130] P. Tucker and L. Davidson. Zonal k-l based large eddy simulation. Computers & Fluids, 33(2):267287, 2004. [131] P. Tucker. Differential equation based length scales to improve DES and RANS simulations. AIAA paper 2003-3968, 16th AIAA CFD Conference, 2003. [132] L. Temmerman, M. Had iadbic, M.A. Leschziner, and K. Hanjali . A hybrid z c two-layer URANS-LES approach for large eddy simulation at high Reynolds numbers. International Journal of Heat and Fluid Flow, 26:173190, 2005. [133] A. Dejoan and R. Schiestel. Large-eddy simulation of non-equilibrium pulsed turbulent ow using transport equations subgrid model. In E. Lindborg, A. Johansson, J. Eaton, J. Humphrey, N. Kasagi, M. Leschziner, and M. Sommerfeld, editors, The Second International Symp. on Turbulence and Shear Flow Phenomena, volume 2, pages 341346, Stockholm, 2001. [134] S. Kenjere and K. Hanjali . LES, T-RANS and hybrid simulations of thermal s c convection at high RA numbers. In W. Rodi and M. Mulas, editors, Engineering Turbulence Modelling and Measurements 6, pages 369378. Elsevier, 2005. [135] P.R. Spalart. Strategies for turbulence modelling and simulations. International Journal of Heat and Fluid Flow, 21:252263, 2000. [136] N.V. Nikitin, F. Nicoud, B. Wasistho, K.D. Squires, and P. Spalart. An approach to wall modeling in large-eddy simulations. Physics of Fluids, 12(7):16291632, 2000.

P. References

246

[137] L. Temmerman, M.A. Leschziner, and K. Hanjali . A-priori studies of near-wall c RANS model within a hybrid LES/RANS scheme. In W. Rodi and N. Fueyo, editors, Engineering Turbulence Modelling and Experiments 5, pages 317326. Elsevier, 2002. [138] F.R. Menter and Y. Egorov. Revisiting the turbulent length scale equation. In IUTAM Symposium: One Hundred Years of Boundary Layer Research, G ttingen, o 2004. [139] F.R. Menter, M. Kuntz, and R. Bender. A scale-adaptive simulation model for turbulent ow predictions. AIAA-2003-0767, Reno, 2003. [140] J.C. Rotta. Turbulente Str mungen. Teubner Verlag, Stuttgart, 1972. o [141] F.R. Menter and Y. Egorov. A scale-adaptive simulation model using twoequation models. AIAA paper 20051095, Reno, NV, 2005. [142] L. Davidson. Evaluation of the SST-SAS model: Channel ow, asymmetric diffuser and axi-symmetric hill. In ECCOMAS CFD 2006, September 5-8, 2006, Egmond aan Zee, The Netherlands, 2006. [143] S.S. Girimaji. Partially-averaged Navier-Stokes model for turbulence: A Reynolds-averaged Navier-Stokes to direct numerical simulation bridging method. ASME Journal of Applied Mechanics, 73(2):413421, 2006. [144] S.S. Girimaji. Partially-Averaged Navier-Stokes model for turbulence: Implementation and Validation. In 43rd AIAA Aerospace Sciences Meeting and Exhibt, 10-13 January, Reno, Nevada, 2005. [145] M. L. Shur, P.R. Spalart, M. Kh. Strelets, and A.K. Travin. A hybrid RANS-LES approach with delayed-des and wall-modelled LES capabilities. International Journal of Heat and Fluid Flow, 29:16381649, 2008. [146] L. Davidson. Hybrid LES-RANS: Inlet boundary conditions for ows including recirculation. In 5th International Symposium on Turbulence and Shear Flow Phenomena, volume 2, pages 689694, 27-29 August, Munich, Germany, 2007. [147] N. Jarrin, S. Benhamadouche, D. Laurence, and R. Prosser. A synthetic-eddymethod for generating inow conditions for large-eddy simulations. International Journal of Heat and Fluid Flow, 27(4):585593, 2006. [148] M. Billson. Computational Techniques for Turbulence Generated Noise. PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, G teborg, Sweden, 2004. o [149] M. Billson, L.-E. Eriksson, and L. Davidson. Jet noise prediction using stochastic turbulence modeling. AIAA paper 2003-3282, 9th AIAA/CEAS Aeroacoustics Conference, 2003. [150] L. Davidson. Hybrid LES-RANS: Inlet boundary conditions. In B. Skallerud and H.I. Andersson, editors, 3rd National Conference on Computational Mechanics MekIT05 (invited paper), pages 722, Trondheim, Norway, 2005.

P. References

247

[151] L. Davidson. Hybrid LES-RANS: Inlet boundary conditions for ows with recirculation. In Second Symposium on Hybrid RANS-LES Methods, Corfu island, Greece, 2007. [152] L. Davidson. Using isotropic synthetic uctuations as inlet boundary conditions for unsteady simulations. Advances and Applications in Fluid Mechanics, 1(1):135, 2007. [153] J.R. Welty, C.E.Wicks, and R.E. Wilson. Fundamentals of Momentum, Heat, and Mass Transfer. John Wiley & Sons, New York, 3 edition, 1984. [154] L. Davidson. HYBRID LES-RANS: Inlet boundary conditions for ows with recirculation. In Advances in Hybrid RANS-LES Modelling, volume 97 of Notes on Numerical Fluid Mechanics and Multidisciplinary Design, pages 5566. Springer Verlag, 2008. [155] L. Davidson and B. Farhanieh. CALC-BFC: A nite-volume code employing collocated variable arrangement and cartesian velocity components for computation of uid ow and heat transfer in complex three-dimensional geometries. Rept. 95/11, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1995. [156] F. Eriksson. Flerdimensionell analys (in Swedish). Studentlitteratur, Lund, Sweden, 4 edition, 1977. [157] L. Davidson. How to make energy spectra with Matlab (available at www.tfd.chalmers/lada). Report 2006/05, Div. of Fluid Dynamics, Dept. of Applied Mechanics, Chalmers University of Technology, G teborg, Sweden, o 2006. [158] R.L. Panton. Incompressible Flow. John Wiley & Sons, New York, 1984. [159] B.J. Daly and F.H. Harlow. Transport equation in turbulence. Physics of Fluids, 13:26342649, 1970. [160] F. Nicoud and F. Ducros. Subgrid-scale stress modelling based on the square of the velocity gradient tensor. Flow, Turbulence and Combustion, 62:183200, 1999. [161] O. Brander. Vektoranalys (in Swedish). Studentlitteratur, Lund, Sweden, 2007.

You might also like