You are on page 1of 88

Chapter 3:

Turbulence and its modeling

Ibrahim Sezai
Department of Mechanical Engineering
Eastern Mediterranean University

Fall 2018-2019

Is the Flow Turbulent?

External Flows UL


where Re L 

Rex  510 5 along a surface
L = x, D, Dh, etc.
ReD  20,000 around an obstacle
Other factors such as free-stream
turbulence, surface conditions, and
Internal Flows disturbances may cause earlier
transition to turbulent flow.
ReDh  2,300

Natural Convection
where

is the Rayleigh number

ME555 : Computational Fluid Dynamics 2 I. Sezai – Eastern Mediterranean University

1
Introduction
For Re > Recritical => Flow becomes Turbulent.
Chaotic and random state of motion develops.
Velocity and pressure change continuously with
time
The velocity fluctuations give rise to additional
stresses on the fluid
=> these are called Reynolds stresses
We will try to model these extra stress terms

ME555 : Computational Fluid Dynamics 3 I. Sezai – Eastern Mediterranean University

3.1 What is turbulence?

u (t )  U  u(t )

ME555 : Computational Fluid Dynamics 4 I. Sezai – Eastern Mediterranean University

2
Two Examples of Turbulent Flow
In turbulent flows there are rotational flow structures called
turbulent eddies, which have a wide range of length scales.

Smaller Structures Larger Structures

ME555 : Computational Fluid Dynamics 5 I. Sezai – Eastern Mediterranean University

In turbulent flow:
A streak of dye which is introduced at a point will rapidly
break up and dispersed  effective mixing

Give rise to high values of diffusion coefficient for;


- Mass
- Momentum
- and heat
All fluctuating components contain energy across a wide
range of frequencies or wave numbers
2 f
wave number  f  frequency
U
ME555 : Computational Fluid Dynamics 6 I. Sezai – Eastern Mediterranean University

3
Fig.3.3 Energy
Spectrum of
turbulence behind a
grid

ME555 : Computational Fluid Dynamics 7 I. Sezai – Eastern Mediterranean University

Scales of turbulence
 Largest eddies break up due to inertial forces
 Smallest eddies dissipate due to viscous forces
 Richardson Energy Cascade (1922)

ME555 : Computational Fluid Dynamics 8


8
I. Sezai – Eastern Mediterranean University

4
Large Eddies:
- have large eddy Reynolds number, vl 
- are dominated by inertia effects
- viscous effects are negligible
- are effectively inviscid
Small Eddies:
- motion is dictated by viscosity
- Re ≈ 1
- length scales : 0.1 – 0.01 mm
- frequencies : ≈ 10 kHz
Energy associated with eddy motions is dissipated and
converted into thermal internal energy
 increases energy losses.
- Largest eddies  anisotropic
- Smallest eddies  isotropic
ME555 : Computational Fluid Dynamics 9 I. Sezai – Eastern Mediterranean University

3.2 Transition from laminar to turbulent flow


Can be explained by considering the stability of
laminar flows to small disturbances
 Hydrodynamic instability
Hydrodynamic stability of laminar flows;
a) Inviscid instability: flows with velocity profile having a
point of inflection.
1. jet flows
2. mixing layers and wakes
3. boundary layers with adverse pressure gradients
b) Viscous instability: flows with laminar profile having no
point of inflection
- occurs near solid walls
ME555 : Computational Fluid Dynamics 10 I. Sezai – Eastern Mediterranean University

5
Fig. 3.4 Velocity profiles
susceptible to
a) Inviscid instability and
b) Viscous instability

ME555 : Computational Fluid Dynamics 11 I. Sezai – Eastern Mediterranean University

Transition to Turbulence
Jet flow: (example of a flow with point of inflection in velocity
profile)

Fig. 3.5. transition in a


jet flow

ME555 : Computational Fluid Dynamics 12 I. Sezai – Eastern Mediterranean University

6
Boundary layers on a flat plate (Example with no
inflection point in the velocity profile)
The unstable two-dimensional disturbances are called Tolmien-
Schlichting (T-S) waves

Fig. 3.6 plan view


sketch of transition
processes in boundary
layer flow over a flat
plate

ME555 : Computational Fluid Dynamics 13 I. Sezai – Eastern Mediterranean University

Fig. 3.7 merging


of turbulent spots
and transition to
turbulence in a
natural flat plate
boundary layer

ME555 : Computational Fluid Dynamics 14 I. Sezai – Eastern Mediterranean University

7
Common features in the transition process:
(i) The amplification of initially small disturbances
(ii) The development of areas with concentrated rotational
structures
(iii) The formation of intense small scale motions
(iv) The growth and merging of these areas of small scale motions
into fully turbulent flows
Transition to turbulence is strongly affected by:
- Pressure gradient
- Disturbance levels
- Wall roughness
- Heat transfer
The transition region often comprises only a very small fraction of
the flow domain
 Commercial CFD packages often ignore transition entirely
(classify the flow as only laminar or turbulent)
ME555 : Computational Fluid Dynamics 15 I. Sezai – Eastern Mediterranean University

3.3Effect of turbulence on time-averaged Navier-Stokes eqns


In turbulent flow there are eddying motions of a wide range
of length scales
A domain of 0.1×0.1m contains smallest eddies of 10-100
μm size
We need 109  1012 mesh points
The frequency of fastest events ≈ 10 kHz Δt ≈ 100μs
needed
DNS of turbulent pipe flow of Re = 105 requires a computer
which is 10 million times faster than CRAY supercomputer
Engineers need only time-averaged properties of the flow
Let’s see how turbulent fluctuations effect the mean flow
properties
ME555 : Computational Fluid Dynamics 16 I. Sezai – Eastern Mediterranean University

8
Descriptors of Turbulent Flow
Time average or mean
First we define the mean Φ of a flow property φ as follows:
t
1

t 0  (t )dt (3.2)

The property φ can be thought of as the sum of a steady mean component


Φ and fluctuation component φ'
 (t )     (t )

The time average of the fluctuations φ' is , by definition, zero:


t
1
t 0
   (t )dt  0 (3.3)

ME555 : Computational Fluid Dynamics 17 I. Sezai – Eastern Mediterranean University

Variance, r.m.s. and turbulence kinetic energy


The spread of the fluctuations φ' about the mean Φ are measured by
t
1
       dt (3.4a)
2 2
variance
t 0
1/ 2
 1 t 
and root mean square (r.m.s.)  rms           dt 
2 2
(3.4b)
 t 0 
The rms values of velocity components can be measured (e.g. by hot-
wire anemometer)
The kinetic energy k (per unit mass) associated with the turbulence is
defined as
1 (3.5)
k  (u 2  v2  w2 )
2
The turbulence intensity Ti is linked to the kinetic energy and a
reference mean flow velocity Uref as follows:
(2 / 3k )1/ 2 (3.6)
Ti 
U ref
ME555 : Computational Fluid Dynamics 18 I. Sezai – Eastern Mediterranean University

9
Moments of different fluctuating variables
The variance is also called the second moment of the fluctuations.
If       and      with       0
Their second moment is defined as
t
1
t 0
      dt (3.7)

If velocity fluctuations in different directions were independent


random fluctuations, then u v, u w and vw would be equal to zero.

However, although u , v and w are chaotic, they are not independent.


As a result the second moments uv, u w and vw are non-zero.

ME555 : Computational Fluid Dynamics 19 I. Sezai – Eastern Mediterranean University

Higher order moments


Third moment:
t
1
( )   ( )3 dt
3
(3.8)
t 0
Third moment is related to skewness (asymmetry) of the
distribution of the fluctuations:

Fourth moment:
t
1
( )   ( ) 4 dt
4
(3.9)
t 0
Fourth moment is related to kurtosis (peakedness) of the
distribution of the fluctuations:
ME555 : Computational Fluid Dynamics 20 I. Sezai – Eastern Mediterranean University

10
Correlation functions – time and space
The autocorrelation function R   ( ) based on two
measurements shifted by time τ is defined as
t
1
t 0
R   ( )   (t ) (t   )   (t ) (t   )dt (3.10)

Similarly, the autocorrelation function R   (ξ ) based on two


measurements shifted by a certain distance in function space
is defined as t t
1
R   ( )   (x, t ) (x  ξ, t ) 
t   (x, t ) (x  ξ, t )dt 
t
(3.11)

ME555 : Computational Fluid Dynamics 21 I. Sezai – Eastern Mediterranean University

When τ = 0, or ξ = 0 R   (0)   2 and it is maximum because the


two correlations are perfectly correlated.
Since φ' is chaotic we expect that the fluctuations become
increasingly decorrelated as τ or ξ → ∞ so, Rφ'φ'(∞) → 0.
The eddies at the root of turbulence cause a certain degree of local
structure in the flow, so there will be a correlation between the values
of φ' at time t and a short time later or at a given location x and a
small distance away.
The decorrelation process will take place gradually over the lifetime
(or size scale) of a typical eddy.
The integral time and length scale represent concrete measures of the
average period or size of a turbulent eddy. They can be computed
from integrals of the autocorrelation functions Rφ'φ'(τ) or Rφ'φ'(ξ).
By analogy it is also possible to define cross-correlation functions
Rφ'ψ'(τ) or Rφ'ψ'(ξ) between pairs of different fluctuations.

ME555 : Computational Fluid Dynamics 22 I. Sezai – Eastern Mediterranean University

11
Probability density function
Probability density function P(φ*) is related to the fraction of
time that a fluctuating signal spends between φ* and φ*+d φ*:
P( * )d *  Prob( *     *  d * ) (3.12)
The average, variance and higher moments of the variables
and its fluctuations are related to the probability density
functions as follows:

    P( )d (3.13a)


( ) n   ( ) P( )d  (3.13b)
n



If n = 2  we get variance, if n = 3, 4,… we get higher order


moments.
ME555 : Computational Fluid Dynamics 23 I. Sezai – Eastern Mediterranean University

3.4.1 Free turbulent flows

U  y U  U min  y U max  U  y
 g   f   h 
U max b U max  U min b U max  U min b
for jets for mixing layers for wakes
b= cross sectional layer width, y=distance in cross-stream direction
x=distance downstream the source
ME555 : Computational Fluid Dynamics 24 I. Sezai – Eastern Mediterranean University

12
Eddies of a wide range of
length scales are visible
Fluid from surroundings is
entrained into the jet.

ME555 : Computational Fluid Dynamics 25 I. Sezai – Eastern Mediterranean University

If x is large enough  functions f, g and h are independent


of x.  such flows are called self – preserving.
The turbulence structure also reaches a self-preserving state
albeit after a greater distance from the flow source than the
mean velocity. Then
u2  y v2  y w2  y uv  y
2
 f1   2
 f2   2
 f3   2
 f4  
Uref b Uref b Uref b Uref b

Uref= Umax – Umin for a mixing layer and wakes


Uref= Umax for jets
Form of functions f, g, h and fi varies from flow to flow
See Fig.3.10

ME555 : Computational Fluid Dynamics 26 I. Sezai – Eastern Mediterranean University

13
Fig. 3.10 mean velocity
distributions and turbulence
properties for
(a) two- dimensional mixing layer,
(b) planar turbulent jet and (c)
wake behind a solid strip

ME555 : Computational Fluid Dynamics 27 I. Sezai – Eastern Mediterranean University

In Fig.3.10:
u
u2 , v2 and w2 and  u v are maximum when is maximum
y
u' gives the largest of the normal stresses (v´and w´) .
 u2
 0.15  0.40
umax
Fluctuating velocities are not equal anisotropic structure of
turbulence
As mean velocity gradients tend to zero turbulence quantities tend
to zero turbulence can’t be sustained in absence of shear
The mean velocity gradient is also zero at the centerline of jets and
wakes.No turbulence there.
The value of  uv is zero at the centerline of a jet and wake since
shear stress must change sign here.

ME555 : Computational Fluid Dynamics 28 I. Sezai – Eastern Mediterranean University

14
3.4.2 Flat plate boundary and pipe flow
inertia forces
Re 
viscous forces
Uy
Re  y: distance away from the wall
v
Near the wall (y small) Re y is small viscous forces dominate
Away from the wall (y large) Re y is large inertia forces dominate.
Near the wall U only depends on y, ρ, μ and τ (wall shear stress), so
U  f ( y ,  ,  , w )
Dimensional analysis shows that
U  u y
u   f     f ( y ) (3.16)
u   
Formula (3.18) is called the law of the wall
u   w  
1/ 2
 V (C f / 2)1/ 2 friction velocity

ME555 : Computational Fluid Dynamics 29 I. Sezai – Eastern Mediterranean University

Far away from the wall:


U  g ( y ,  ,  , w ) independent of 
Dimensional analysis yields U  y
u   g 
u  
The most useful form emerges if we view the wall shear stress
as the cause of a velocity deficit Umax – U which decreases
the closer we get to the edge of the boundary layer or the
pipe centerline. Thus
U max  U  y (3.17)
 g 
u  
This formula is called the velocity – defect law.
ME555 : Computational Fluid Dynamics 30 I. Sezai – Eastern Mediterranean University

15
Linear sub layer – the fluid layer in contact with smooth wall
Very near the wall there is no turbulent (Reynolds) shear stresses
 flow is dominated by viscous shear

For ( y  5) shear stress is approximately constant,
U
 ( y)   w
y
Integrating and using U = 0 at y = 0,
w y
U

After some simple algebra and making use of the definitions of u+
and y+ this leads to
u  y (3.18)

Because of the linear relationship between velocity and distance


from the wall the fluid layer adjacent to the wall is often known
as the linear sub-layer
ME555 : Computational Fluid Dynamics 31 I. Sezai – Eastern Mediterranean University

Log-law layer – the turbulent region close to smooth wall

Outside the viscous sublayer (30 < y+ < 500) a region exists
where viscous and turbulent effects are both important.
The shear stress τ varies slowly with distance; it is assumed to
be constant and equal to τw
Assuming a length scale of turbulence lm = κy where lm is
mixing length, the following relationship can be derived:
1 1
u   ln y   B  ln( Ey  ) (3.19)
 
κ = 0.4, B=5.5, E=9.8; (for smooth walls)
B decreases with roughness
κ and B are universal constants valid for all turbulent flows past smooth
walls at high Reynolds numbers.
(30 < y+ < 500)  log – law layer

ME555 : Computational Fluid Dynamics 32 I. Sezai – Eastern Mediterranean University

16
Outer layer – the inertia dominated region far from the wall

Experimental measurement show that the log-law is valid in


the region 0.02 < y ∕ δ < 0.2.
For larger values of y the velocity-defect law (3.19) provides
the correct form.
In the overlap region the log-law and velocity-defect law have
to become equal.
Tennekes and Lumley (1972) show that a matched overlap is
obtained by assuming the following logarithmic form:
U max  U 1  y 
 ln    A Law of the Wake (3.20)
u   
where A is constant
See Fig.3.11
ME555 : Computational Fluid Dynamics 33 I. Sezai – Eastern Mediterranean University

Fig.3.11

- The inner region: 10 to 20% of the total thickness of the wall layer; the shear stress is
(almost) constant and equal to the wall shear stress τw. Within this region there are three
zones
- the linear sub-layer: viscous stresses dominate the flow adjacent to the surface
- the buffer layer: viscous and turbulent stresses are of similar magnitude
- the log-law layer: turbulent (Reynolds) stresses dominate.
- The outer region or law-of-the-wake layer: inertia dominated core flow far from wall; free
from direct viscous effects.
ME555 : Computational Fluid Dynamics 34 I. Sezai – Eastern Mediterranean University

17
For y ∕ δ > 0.8 fluctuating velocities become almost equal
isotropic turbulence structure here. (far away the wall)
For y ∕ δ < 0.2 large mean velocity gradients
high values of u2 , v2 and w2 and  uv .(high turbulence
production) .
Turbulence is anisotropic near the wall.
ME555 : Computational Fluid Dynamics 35 I. Sezai – Eastern Mediterranean University

Fluid entering from top into the CV


will bring in higher momentum fluid
and will accelerate the slower
moving layer. As a result, the fluid
layer will experience additional
turbulent shear stresses which are
Control volume in a 2D turbulent known as the Reynolds stresses.
shear flow

Rules which govern the time averages of fluctuation properties


      and      :
 
      0;   ; 
s s
;   ds   ds
.
      ;       ;   ;    0

ME555 : Computational Fluid Dynamics 36 I. Sezai – Eastern Mediterranean University

18
Since div and grad are both differentiations the above rules can be
extended to a fluctuating vector quantity a = A + a' and its
combinations with a fluctuating scalar       :
div a  div A; div  a   div  a   div     div  a ;   (3.22)
div grad  div grad 
To illustrate the influence of turbulent fluctuations on mean flow we consider
div u  0 (3.23)
u 1 p
 div  uu     v div grad u (3.24a)
t  x
v 1 p
 div  vu     v div grad v (3.24b)
t  y
w 1 p (3.24c
 div  wu     v div grad w
t  z )
Substitute u  U  u; u  U  u ; v  V  v; w  W  w; p  P  p
ME555 : Computational Fluid Dynamics 37 I. Sezai – Eastern Mediterranean University

Consider continuity equation: note that divu  divU this yields the
continuity equation
divU  0 (3.25)
A similar process is now carried out on the x-momentum equation (3.9a).
The time averages of the individual terms in this equation can be
written as follows:
u U
 ; div (uu)  div UU   div  u u 
t t
1 p 1 P
  ; vdiv grad u  v div grad U
 x  x
Substitution of these results gives the time-average x-momentum
equation
U 1 P
 div(UU)  div(uu)    v div grad U (3.26a)
t  x
(I ) ( II ) ( III ) ( IV ) (V )
ME555 : Computational Fluid Dynamics 38 I. Sezai – Eastern Mediterranean University

19
Repetition of this process on equations (3.9b) and (3.9c) yields the
time-average y-momentum and z-momentum equations

V 1 P
 div(VU)  div(vu)    v div grad V (3.26b)
t  y
(I) (II) (III) (IV) (V)
W 1 P
 div(WU)  div( wu)    v div grad W (3.26c)
t  z
(I) (II) (III) (IV) (V)

The process of time averaging has introduced new terms (III).


Their role is additional turbulent stresses on the mean velocity
components U, V and W.
ME555 : Computational Fluid Dynamics 39 I. Sezai – Eastern Mediterranean University

Placing these terms on the rhs:

U 1 P  u 2 u v uw 
 div(UU )    v div grad U      
t  x  x y z 
(3.27a)
V 1 P  u v v2 vw 
 div(VU)    v div grad V      
t  y  x y z 
Reynolds
equations (3.27b)
W 1 P  uw vw w2 
 div(WU)    v div grad W      
t  z  x y z 
(3.27c)

ME555 : Computational Fluid Dynamics 40 I. Sezai – Eastern Mediterranean University

20
The extra stress terms result from six additional stresses, three normal stresses
and three shear stresses:
 xx    u2  yy    v2  zz    w2 (3.28a)
 xy   yx    uv  xz   zx    uw  yz   zy    vw (3.28b)
-These extra turbulent stresses are termed the Reynolds stresses.
-The normal stresses   u 2 ,   v2 and   w2 are always non-zero
-The shear stresses   u v,   u w and   vw are also non-zero
If, for example, u' and v' were statistically independent fluctuations, the time
average of their product uv would be zero
Similar extra turbulent transport terms arise when we derive a transport
equation for an arbitrary scalar quantity. The time average transport
equation for scalar φ is
  u   v  w  
 div(U)  div( * grad )        S (3.29)
t  x y z 

ME555 : Computational Fluid Dynamics 41 I. Sezai – Eastern Mediterranean University

In compressible flow density fluctuations are usually negligible


The density-weighted averaged (or Favre-averaged) form of the compressible
turbulent flow are:

ME555 : Computational Fluid Dynamics 42 I. Sezai – Eastern Mediterranean University

21
Closure problem – the need for turbulence modeling

In the continuity and Navier – Stokes equations (3.8)


and (3.9 a-c) there are 6 additional unknowns
 the Reynolds stresses
Similarly in scalar transport eqn (3.14) there are 3
extra unknowns:
u , v  and w 
Main task of turbulence modeling:
is to develop computational procedures to predict the
9 extra terms. (Reynolds stresses and scalar
transport terms).
ME555 : Computational Fluid Dynamics 43 I. Sezai – Eastern Mediterranean University

3.4. Characteristics of simple turbulent flows


In turbulent thin shear layers,
 

x y

 1
The picture can't be display ed.

L
The following 2D incompressible turbulent flows with constant P will be
considered:
Free turbulent flows
- Mixing layers
- Jet
- Wake
Boundary layer near solid walls
- Flat plate boundary layer
- Pipe flow
Data for U,   u2 ,   v2 ,   w2 and   uv will be reviewed.
These can be measured by 1) Hot wire anemometry 2) Laser doppler
anemometers
ME555 : Computational Fluid Dynamics 44 I. Sezai – Eastern Mediterranean University

22
3.5 Turbulence models

We need expressions for :


- Reynolds stresses: u '2 , u v, uw, v2 , vw, w2
- Scalar transport terms : u  , v  and w 
ME555 : Computational Fluid Dynamics 45 I. Sezai – Eastern Mediterranean University

Classical models: use the Reynolds eqn’s.(all commercial CFD codes)


Large eddy simulation: the flow eqn’s are solved for the
- mean flow
- and largest eddies
but the effect of the smaller eddies are modeled
- are at the research state
- calculations are too costly for engineering use.

ME555 : Computational Fluid Dynamics 46 I. Sezai – Eastern Mediterranean University

23
Turbulence Scales and Prediction Methods
energy cascade
(Richardson, 1922)

ME555 : Computational Fluid Dynamics 47 I. Sezai – Eastern Mediterranean University

The mixing length and k-ε models are the most widely used and
validated.
They are based on the presumption that
“there exists an analogy between the action of viscous stresses
and Reynolds stresses on the mean flow”
Viscous stresses are proportional to the rate of deformation. For
incompressible flow:
 ui u j 
 ij  2 sij      (2.31)
 x x
 j i 
Notation:
i = 1 or j = 1  x-direction
i = 2 or j = 2  y-direction
i = 3 or j = 3  z-direction

ME555 : Computational Fluid Dynamics 48 I. Sezai – Eastern Mediterranean University

24
For example  u1 u2   u v 
 12   xy         
 x2 x1   y x 
Turbulent stresses are found to increase as the mean rate of
deformation increases.
It was proposed by Boussinesq in 1877 that
Reynolds stress could be linked to mean rate of deformation
 U i U j  2
 ij    uiuj  t      k ij (3.33)
 x j xi
  3
where k  0.5(u2  v2  w2 )
This is similar to viscous stress equation
 U U j 
 ij  2 sij    i  
 x xi 
 j
except that μ is replaced by μt, where μt = turbulent (eddy) viscosity
ME555 : Computational Fluid Dynamics 49 I. Sezai – Eastern Mediterranean University

Similarly; vt  t /  kinematic turbulent or eddy viscosity.


Eqn.(3.23) shows that
 U U j 
 ij    uiuj  t  i  
 x x
 j i 
Turbulent momentum transport is assumed to be proportional to mean
gradients of velocity
By analogy turbulent transport of a scalar is taken to be proportional to
the gradient of the mean value of the transported quantity. In suffix
notation we get

  ui   t (3.34)
xi
Where Γt is the turbulent diffusivity.
We introduce a turbulence Prandtl / Schmidt number as
t
t  (3.35)
t
ME555 : Computational Fluid Dynamics 50 I. Sezai – Eastern Mediterranean University

25
Experiments in many flows have shown that
t  1
Most CFD procedures use  t  1
Mixing length models:
Attempts to describe the turbulent stresses by means of simple algebraic
formulae for μt as a function of position
The k-ε models:
Two transport eqn’s (PDE’s) are solved:
1. For the turbulent kinetic energy, k
2. For the rate of dissipation of turbulent kinetic energy, ε.
t  C     C k 2 /  (C  constant)
Both models assume that μt is isotropic (same for u, v, and w equations)
(i.e. the ratio between Reynolds stresses and mean rate of deformation is the same in all
directions)
ME555 : Computational Fluid Dynamics 51 I. Sezai – Eastern Mediterranean University

This assumption fails in many categories of flow.


It is necessary to derive and solve transport eqn’s for the 6
Reynolds stresses themselves. ( u '2 , u v, u w, v2 , vw, w2 ).
These eqn’s contain:
1. Diffusion
2. Pressure strain
3. Dissipation
terms whose individual effects are unknown and cannot be measured
Reynolds stress equation models:
Solves the 6 transport equations (PDE’s) for the Reynolds stress terms
together with k and ε eqn’s., by making assumptions about
- diffusion
- pressure strain
- and dissipation terms.
ME555 : Computational Fluid Dynamics 52 I. Sezai – Eastern Mediterranean University

26
Algebraic stress models:

Approximate the Reynolds stresses in terms of


algebraic equations instead of PDE type transport
equations.
Is an economical form of Reynolds stress model.
Is able to introduce anisotropic turbulence effects
into CFD simulations

ME555 : Computational Fluid Dynamics 53 I. Sezai – Eastern Mediterranean University

3.5.1 Mixing length model


m2 m
t   m  : turbulent velocity scale
s s
 : length scale of largest eddies
 
vt  C  (3.36)
where C is a dimensionless constant of proportionality.
Turbulent viscosity is given by t  C  
This works well in simple 2-D turbulent flows where the only significant
Reynolds stress is  xy   yx    u v and only significant velocity
U
gradient is
y
U
For such flows   c
y (3.37)

where c is dimensionless constant

ME555 : Computational Fluid Dynamics 54 I. Sezai – Eastern Mediterranean University

27
Combining (3.26) and (3.27) and absorbing C and c into a new length scale  m
U
vt   2m (3.38)
y
This is Prandtl’s mixing length model.
The Reynolds stress is
2 U U
 xy   yx    u v    m (3.39)
y y

ME555 : Computational Fluid Dynamics 55 I. Sezai – Eastern Mediterranean University

The transport of scalar quantity φ is modeled as


 (3.40)
  v   t
y
where Γt = μt ∕ σt and vt is found from (3.28).
σt = 0.9 for near wall flows
σt = 0.5 for jets and mixing layers Rodi(1980)
σt = 0.7 in axisymmetric jets
In table 3.3:
y : distance from the wall
κ : 0.41 von Karman’s constant
ME555 : Computational Fluid Dynamics 56 I. Sezai – Eastern Mediterranean University

28
ME555 : Computational Fluid Dynamics 57 I. Sezai – Eastern Mediterranean University

ME555 : Computational Fluid Dynamics 58 I. Sezai – Eastern Mediterranean University

29
3.5.2. The k – ε model
If convection and diffusion of turbulence properties
are not negligible (as in the case of recirculating
flows), then the mixing length model is not
applicable.
The k-ε model focuses on the mechanisms that affect
the turbulent kinetic energy.
Some preliminary definitions:
1 2
K
2
U V 2 W 2  mean kinetic energy

1

k  u 2  v2  w2
2
 turbulent kinetic energy

k (t )  K  k instantaneous kinetic energy


ME555 : Computational Fluid Dynamics 59 I. Sezai – Eastern Mediterranean University

To facilitate the subsequent calculations it is common to write the components


of the rate of deformation sij and the stresses τij in tensor (matrix) form:
 sxx sxy sxz    xx  xy  xz 
   
sij   s yx s yy s yz  and  ij   yx  yy  yz 
   
- Using  gives  szx szy szz 
sij (t )  Sij  sij  zx  zy  zz 
U u  V v
sxx (t )  S xx  sxx   ; s yy (t )  S yy  syy   ;
x x y y
W w
szz (t )  S zz  szz   ;
z z
1  U V  1  u  v 
sxy (t )  S xy  sxy  s yx (t )  S yx  syx    
2  y x  2  y x 
1  U W  1  u  w 
sxz (t )  S xz  sxz  szx (t )  S zx  szx     
2  z x  2  z x 
1  V W  1  v w 
s yz (t )  S yz  syz  szy (t )  S zy  szy     
2  z y  2  z y 
ME555 : Computational Fluid Dynamics 60 I. Sezai – Eastern Mediterranean University

30
The product of a vector a and a tensor bij is a vector c
 b11 b12 b13 
 
abij  ai bij   a1 a2 a3   b21 b22 b23 
b b33 
 31 b32
T T
 a1b11  a2b21  a3b31   c1 
   
  a1b12  a2b22  a3b32    c2   c j  c
a b  a b  a b  c 
 1 13 2 23 3 33   3 
The scalar product of two tensors aij and bij is evaluated as follows
aij  bij  a11b11  a12b12  a13b13  a21b21  a22b22  a23b23
 a31b31  a32b32  a33b33
Suffix notation:
1 x – direction
2 y – direction
3 z – direction
ME555 : Computational Fluid Dynamics 61 I. Sezai – Eastern Mediterranean University

Governing equation for mean flow kinetic energy K


Reynolds equations: (3.27a), (3.27b), (3.27c)

Multiply x – component Reynolds eqn. (3.27a) by U


Multiply y – component Reynolds eqn. (3.27b) by V
Multiply z – component Reynolds eqn. (3.27c) by W

Then add the result together.


After some algebra the time-average eqn. for mean
kinetic energy of the flow is

ME555 : Computational Fluid Dynamics 62 I. Sezai – Eastern Mediterranean University

31
Or in words, for the mean kinetic energy K, we have
(  K )
 div(  KU )  div   PU  2 USij   Uuiu j   2  Sij  Sij   uiu j  Sij
t
(I ) ( II ) ( III ) ( IV ) (V ) (VI ) (VII )
(3.41)

ME555 : Computational Fluid Dynamics 63 I. Sezai – Eastern Mediterranean University

Governing equation for turbulent kinetic energy k


1- Multiply x, y and z momentum eqns (3.24a-c) by u´, v´ and w´,respectively
2- Multiply x, y and z Reynolds eqns (3.27a-c) by u´, v´ and w´, respectively
3- subtract the two of the resulting equations
Rearrange:
 yields the equation for turbulent kinetic energy k:

(  k )  1 
 div(  kU )  div   pu  2  usij   ui  uiu j   2  sij  sij   uiu j  Sij
t  2 
(I ) ( II ) ( III ) ( IV ) (V ) (VI ) (VII )
(3.42)

ME555 : Computational Fluid Dynamics 64 I. Sezai – Eastern Mediterranean University

32
The viscous term (VI)

2  sij  sij  2  s112  s22
2  s33
2  2 s122  2 s132  2 s23
2 
Gives a negative contribution to (3.32) due to the appearance of the sum
of squared fluctuating deformation rate s'ij
The dissipation of turbulent kinetic energy is caused by work done by the
smallest eddies against viscous stresses.
The rate of dissipation per unit mass (m2/s3) is denoted by

  2vsij  sij (3.43)

The k – ε model equations


It is possible to develop similar transport eqns. for all other turbulence
quantities
The exact ε – equation, however, contains many unknown and unmeasurable
terms
The standard k – ε model(Launder and Spalding,1974) has two model eqns
(a) one for k and (b) one for ε
ME555 : Computational Fluid Dynamics 65 I. Sezai – Eastern Mediterranean University

We use k and ε to define velocity scale  and length scale


 representative of the large scale turbulence as follows:
k 3/ 2
  k 1/ 2 

Applying the same approach as in the mixing length model
we specify the eddy viscosity as follows
k2
t  C     C (3.44)

where Cμ is a dimensionless constant

ME555 : Computational Fluid Dynamics 66 I. Sezai – Eastern Mediterranean University

33
The standard model uses the following transport equations used for k and ε
(  k )  
 div(  kU)  div  t grad k   2 t Sij  Sij   (3.45)
t k   P

k

 (  )    2
 div(  U)  div  t grad    C1 2t Sij  Sij  C2  (3.46)
t
In words the equations are    k k

The equations contain five adjustable constants C ,  k ,   , C1 and C2 . The
standard k – ε model employs values for the constants that are arrived at by
comprehensive data fitting for a wide range of turbulent flows:
C  0.09;  k  1.00;    1.30; C1  1.44; C2  1.92
(3.47)
ME555 : Computational Fluid Dynamics 67 I. Sezai – Eastern Mediterranean University

To compute the Reynolds stresses with the k – ε model (3.44-3.46)


Boussinesq relationship is used:
 U U j  2 2
  uiu j  t  i     k ij  2 t Sij   k ij (3.48)
 x xi  3 3
 j
δij = 1 if i = j, δij = 0 if i ≠ j
U i 2 U i 2
Contraction gives   uiui  2 t   ii  k  2 t  k
xi 3 xi 3
U U U U
flow x  x  x  x  0 due to continuity
i 1 2 3
For incompressible
3
Then  uiui  2k which is the definition of k.
i 1 2 3

i 1
 1

k  u1u1  u2 u2  u3u3
 2
 
where uiui  uiu i 

So an equal 1/3 is allocated to each normal stress component in eqn
(3.48) to have –2ρk when summed. This implies an isotropic
assumption for the normal Reynolds stresses.
ME555 : Computational Fluid Dynamics 68 I. Sezai – Eastern Mediterranean University

34
B.c.’s for high Re model k and ε-equations
The following boundary conditions are needed

For inlet, if no k and ε is available crude approximations can be obtained


from the
- Turbulence intensity, Ti
- Characteristic length, L (equivalent pipe diameter) :
2 k 3/ 2
 
2
k  U ref Ti ;   C3/ 4 ;   0.07 L
3 
The formulae are closely related to the mixing length formulae in section
3.5.1 and the universal distributions near a solid wall given below.

ME555 : Computational Fluid Dynamics 69 I. Sezai – Eastern Mediterranean University

Wall boundary conditions for high Re models


Near the wall, at high Reynolds numbers, equations (3.44-3.46) of the
standard k – ε model are not integrated right to the wall. Instead,
universal behaviour of near wall flows is used.
1 (3.19)
u  ln Ey  for 30  y   500

In this region measurements of the budgets indicates that:
the rate of turbulence production = rate of dissipation
Using this assumption and t   C k 2 /  , the following wall functions
can be derived for a point P in the region 30 < y+ < 500:
UP 1 u 2 u 3
uP   ln  EyP  ; k ;  for 30  y   500 (3.49)
u  C y
u y
where u   w   friction velocity, y   
1/2


κ = 0.41 (Von Karman’s constant)
E = 9.8 (wall roughness parameter) for smooth walls
ME555 : Computational Fluid Dynamics 70 I. Sezai – Eastern Mediterranean University

35
Wall b.c.’s for high Re models: Momentum Eqns
2
Using k from Eqn. (3.49), k P  u  u  k P1/ 2C1/4
C
u y k C y 1/2 1/4

y   P  P  P 

 P

In the viscous sublayer, near the wall (y+ < 5) the flow is laminar and
the velocity is given by
u  y (3.18)
The position of the interface between the laminar sublayer and the
log law layer can be found by equating Eqns. (3.18) and (3.19):
 1  
yint  ln( Eyint ),  yint  11.63

Then, the velocity at a point P near the wall is found from
 yP if yP  11.63
 
u  1
P  
 ln( EyP ) if yP  11.63

ME555 : Computational Fluid Dynamics 71 I. Sezai – Eastern Mediterranean University

Wall b.c.’s for high Re models: k and ε-Eqns


k-equation, Eqn. (3.45), is integrated right to the wall.
However, the source term of the k-equation given by
sk  2t Sij  Sij    Pk  
is calculated at a near wall point P by assuming local
equilibrium. Under this assumption,
the rate of turbulence production = rate of dissipation.
Thus, the rate of turbulence production is calculated from
U w
Pk   w w , where  w   u2 , u  k P1/2C1/ 4
y  C k P yP
1/4 1/2

and ε is calculated from


C3/4 k P3/2
P 
 yP
ME555 : Computational Fluid Dynamics 72 I. Sezai – Eastern Mediterranean University

36
Wall b.c.’s for high Re models: k and ε-Eqns
The boundary condition for k imposed at the wall is
k / n  0
where n is normal to the wall.
The ε-equation is not solved at the wall-adjacent cells but is computed
from
C3/ 4 k P3/ 2
P 
 yP

ME555 : Computational Fluid Dynamics 73 I. Sezai – Eastern Mediterranean University

Wall b.c.’s for high Re models: Energy equation


The wall heat flux is given by
qw  CP  C1/4 k P1/2 (TP  Tw ) / T 
where T+ is calculated from the universal near wall temperature
distribution valid at high Reynolds numbers (Launder and Spalding,
1974):
T  Tw  CP  u   u   P   T ,l   (3.50)
T      
qw
T ,t  
  T ,t  
with Tp  temperature at near wall point yP  T ,t  turbulent Prandtl number
Tw  wall temperature  T ,l   CP / T  Prandtl number
qw  wall heat flux T  thermal conductivity
C p  fluid specific heat and constant pressure
Finally P is the “pee-function”, a correction function dependent on the
ratio of laminar to turbulent Prandtl numbers (Launder and Spalding,
1974)
ME555 : Computational Fluid Dynamics 74 I. Sezai – Eastern Mediterranean University

37
Low Reynolds number k-ε models
At low Reynolds numbers (which is the case near the wall) the constants
Cμ, C1ε and C2ε in Eqns. (3.52-3.53) are not valid so these equations cannot
be integrated right to the wall. To integrate the governing equations right to
the wall, modifications to the standard k-ε model is necessary.
The equations of the low Reynolds number k – ε model, which replace
(3.44-3.46), are given below:

k2
t   C f  (3.51)

(  k )    
 div(  kU)  div     t grad k   2 t Sij  Sij   (3.52)
t  k 
 (  )    
 div(  U)  div     t  grad  
t    
 2
C1 f1 2 t Sij  Sij  C2 f 2  (3.53)
k k
ME555 : Computational Fluid Dynamics 75 I. Sezai – Eastern Mediterranean University

Modifications:
A viscous contribution is included
Cμ, C1ε and C2ε are multiplied by wall damping functions fμ, f1 , f2
which are functions of turbulence Reynolds number or similar
functions
As an example we quote the Lam and Bremhorst (1981) wall-
damping functions which are particularly successful:
2 20.5 
f   1  exp(0.0165 Re y )   1  ;
 Ret 
3
(3.54)
 0.05 
f1  1   f 2  1  exp(  Ret2 )
 f 
;
  
In function f  the parameter Re y is defined by k 1/ 2 y / v . Lam and
Bremhorst use  y  0 as a boundary condition.
ME555 : Computational Fluid Dynamics 76 I. Sezai – Eastern Mediterranean University

38
Assessment of performance

Table 3.4 k-ε model assessment

ME555 : Computational Fluid Dynamics 77 I. Sezai – Eastern Mediterranean University

k-ε Model Implementation Details


The flow equations (3.27) can be written in tensor notation as
  U i      U U j   P
t

x j
 U iU j  
x j
   i 
  x j xi
   
 uiuj 
  x j xi (3.54a)
   U i U j   P
        uiuj  
x j   x j xi   xi
Inserting the Boussinesq relation (3.38)
 U i U j  2
 ij    uiuj  t      k ij
 x xi
 j  3
  U i     U U j   P*
t

x j
 U iU j   x (   t )  i     (3.54b)
j   x j xi   xi
where P* = P + (2/3)ρk
Note that Eqn. (3.54b) is the same as that of Navier-Stokes Eqns.
given by (2.32a,b,c) with μ replaced by (μ + μt) and p by P*.
ME555 : Computational Fluid Dynamics 78 I. Sezai – Eastern Mediterranean University

39
Rearranging
  U i      U U j   P*
t

x j
 U iU j   (   t )  i 
x j 
  
 x j xi   xi
  U i    U j  P*
  (   t )   (   t ) 
x j  x j  x j  xi  xi

  U i    U j    U j  P*
  (   t )   ( t )  (  ) 
x j  x j  x j  xi  x j  xi  xi

  U i    U j  P*
  (   t )   ( t ) 
x j  x j  x j  xi  xi (3.54c)

source term
Note that we have used:
  U j    U j    U j  
(  )         0  0
x j  xi  x j  xi  xi  x j  xi
(due to continuity if  = constant, and the fluid is incompressible)
The terms after the first parenthesis on the RHS of Eqn. (3.54c) are treated as a source .
ME555 : Computational Fluid Dynamics 79 I. Sezai – Eastern Mediterranean University

k-ε Model Equations in Generic Form


The low Reynolds number k-ε model equations can be written in generic form as
 (  )
 div(  U )  div( grad  )  s
t
Equation   s
Continuity 1 0 0
  U    V    W  P*
x-Momentum U eff  eff    eff    eff 
x  x  y  x  z  x  x
  U    V    W  P*
y-Momentum V eff  eff    eff    eff 
x  y  y  y  z  y  y
  U    V    W  P*
z-Momentum W eff  eff    eff    eff 
x  z  y  z  z  z  z
 t
Energy T  0
 t

Turbulence ke k  t Pk  
k
t  2
Dissipation of ke   C1 f1 Pk  C2 f 2 
 k k
2 2 2
1  U U j 
2 2 2
 U   V   W  1  U V  1  W U  1  V W 
Sij Sij                    , Sij   i  
 x   y   z  2  y x  2  x z  2  z y  2  x j xi 
 U U j  U i
f1  1   0.05 / f   , Pk  2 t Sij Sij  t  i 
3
P*  P  (2 / 3)  k , eff    t , t   C f  k 2 /  , 
 x j xi  x j

f   1  exp  0.0165 Re y   1  20.5 / Ret  , f 2  1  exp( Ret ), Ret   k /  , y = minimum distance to the nearest wall
2
2 2

Re y   k 1/2 y /  , C  0.09,  k  1.00,    1.30,  t  0.9, C1  1.44, C2  1.92,   Pr


ME555 : Computational Fluid Dynamics 80 I. Sezai – Eastern Mediterranean University

40
Source term linearization in k-ε models
When the source term is linearized as
s  sC  sPP
1) sP must be negative for a convergent iterative solution,
which ensures diagonal dominance of the coefficient
matrix. This is a requirement for the boundedness
criteria discussed in chapter 5.
2) sC must be positive (and sP must be negative ) to obtain
all positive  values. (Patankar, 1980)
Both k and ε are strictly positive quantities. So, to obtain
always positive results for k and ε, the source terms should
be formulated such that sC is positive and sP is negative for
k and ε equations.
ME555 : Computational Fluid Dynamics 81 I. Sezai – Eastern Mediterranean University

Source term linearization in k equation


k-equation: sk  Pk  
Consider first term, ρPk. Inserting t   C f  k 2 /  into Pk
k2
Pk  2 t Sij Sij  2  C f  Sij Sij  C3 k 2

Linearization of ρPk in the form of sk = sPP + sC would
give sP  C3 k and sC  0 . However, since C3 > 0 and k > 0,
then sP > 0, which violates the boundedness criteria. We can
conclude that positive terms (e.g. Pk) cannot be linearized
using such an implicit manner.
As a result, Pk should be put into sC.
Consider the second term, –ρε. From t   C f  k 2 / 
   C f  k 2 / t . Inserting Pk and ε into the source term
sk  Pk   , gives:

ME555 : Computational Fluid Dynamics 82 I. Sezai – Eastern Mediterranean University

41
Source term linearization in k-equation
k-equation:
 2 C f  2  2 C f 
sk  Pk    Pk  k  b  ak , b  Pk , a 
2

t t
For the second term, ak2, the linearization proposed by Patankar (1980) can be used:
*
 ds 
     
s  s*    k P  k P*  b  a (k P* ) 2  2ak P* k P  k P*  b  a(k P* ) 2   2ak P* k P
 dk 
 2 C f  * 2  2 C f  *
Then, s  b  a ( k * 2
)  P  ( k ) and s  2 ak *
 2 kP
C P k
t P P P
t
where the superscript * refers to the previous iteration values. However, this
linearization does not yield a robust algorithm. A better approach is to put the
coefficient of k to the sP term. Comparing sk relation with s  sC  sPP gives:

( sC ) k  b  Pk
 2C f  *
( sP ) k  ak *   k
t

ME555 : Computational Fluid Dynamics 83 I. Sezai – Eastern Mediterranean University

Positivity of sC term in k-equation


For consistency, the turbulent viscosity, t, should be a positive
quantity.
However,
 C f  k 2
t 

t may become negative if ε becomes negative during iterations.
t should be enforced to take positive values during iterations. The
limiter used for enforcing positive values for t is given later in Eqn.
(3.51b).

ME555 : Computational Fluid Dynamics 84 I. Sezai – Eastern Mediterranean University

42
Source term linearization in ε-equation
ε-equation:
 2
s  C1 f1 Pk  C2  f 2   b  a 2
k k

where b  C1 f1 Pk , a  C2  f 2  / k
k
Note that, since b > 0, it cannot be linearized impicitly as it gives a
positive sP.
The second term, aε2, can be linearized by transferring the coefficient
of the variable ε to the sP term, in accordance with s  sC  sPP
Then, for  =  we obtain:
*
( sC )  b  C1 f1 Pk
k*
*
( sP )  a  C2 f 2 
k*
where the superscript * refers to the previous iteration values.
ME555 : Computational Fluid Dynamics 85 I. Sezai – Eastern Mediterranean University

Source term linearization in ε-equation


The source terms sP and sC of the ε-equation can be written
in terms of eddy viscosity μt using
 C f  k 2   C f  k
t   
 k t
which yields
 C f  k *
( sC )  b  C1 f1 Pk
t
 C f  k *
( sP )  a  C2 f 2 
t

ME555 : Computational Fluid Dynamics 86 I. Sezai – Eastern Mediterranean University

43
Positivity of sC term in k- and ε-equations
sC term should be positive to obtain positive values during the
iteration of a general variable . Rewriting the sC term for the k-
equation:
( sC ) k  Pk  2 t Sij Sij

We observe that, (sC)k is positive if μt is enforced to take positive


values since Pk is always positive.
Rewriting the sC term for the -equation:
 C f  k *
( sC )  C1 f1 Pk
t
Again, we observe that, (sC) is positive if μt is enforced to take
positive values.
Then, enforcing μt to take positive values is necessary to obtain
positive results during iterations of both k- and -equations.

ME555 : Computational Fluid Dynamics 87 I. Sezai – Eastern Mediterranean University

Limiting of eddy viscosity μt


Note that both k and ε are strictly positive quantities.
However, during iterations, undershoots may develop which
will result in negative values for either k or ε which will
give negative μt values in accordance with Eqn. (3.51):.
t   C f  k 2 /  (3.51)
Also if ε → 0, division by zero will occur in accordance
with Eqn. (3.51).
Also, sP and sC terms of both k and ε-equations contain 1/μt
term so that if μt → 0, division by zero will occur. To
overcome these difficulties μt is limited to a small fraction
of laminar viscosity μl by using 2
 k 
t  max   C f  ,104 l  (3.51b)
  
ME555 : Computational Fluid Dynamics 88 I. Sezai – Eastern Mediterranean University

44
3.5.3 Reynolds stress equation models
The most complex classical turbulence model is Reynolds
stress equation model (RSM),
Also called :
- second order
- or second – moment closure model
Drawbacks of k – ε model emerge when it is attempted to
predict
- flows with complex strain field
- flows with significant body forces
The exact Reynolds stress transport eqn can account for the
directional effects of the Reynolds stress field.

Let R ij   ij  u iu j (Reynolds stress)

ME555 : Computational Fluid Dynamics 89 I. Sezai – Eastern Mediterranean University

The exact solution for the transport of Rij takes the following form
DRij (3.55)
 Pij  Dij   ij   ij  ij
Dt

Equation (3.45) describes six partial differential equations: one for


the transport of each of the six independent Reynolds stresses
( u12 , u22 , u32 , u1u2 , u1u3 and u2 u3 ,since u2 u1  u1u2 , u3u1  u1u3 and u3u2  u2 u3 )
Two new terms appear compared with ke eqn (3.32)
1. . ij : pressure strain correction term
2. . ij : rotation term
ME555 : Computational Fluid Dynamics 90 I. Sezai – Eastern Mediterranean University

45
The convective term is
 ( U k uiuj )
Cij   div(  uiu j U) (3.56)
xk

The production term is

 U j U i 
Pij    Rim  R jm  (3.57)
 xm xm 

The rotation term is


ij  2k (uj um eikm  uium e jkm ) (3.58)
where k  rotation vector
eijk  1 if i, j and k are different and in cyclic order
eijk  1 if i, j and k are different and in anti-cyclic order
eijk  0 if any two indices are the same
ME555 : Computational Fluid Dynamics 91 I. Sezai – Eastern Mediterranean University

The diffusion term Dij can be modelled by the assumption that the rate of
transport of Reynolds stresses by diffusion is proportional to gradient of
Reynolds stress.
  vt Rij   vt 
Dij     div  grad  Rij   (3.59)
xm   k xm   k 
k2
with vt  C ; C  0.09 and  k  1.0

The dissipations rate εij is modeled by assuming isotropy of the small
dissipative eddies. It is set so that it effects the normal Reynolds stresses (i =
j) only and in equal measure. This can be achieved by
2 (3.60)
 ij   ij
3
where ε is the dissipation rate of turbulent kinetic energy defined by (3.43).
The Kronecker delta, δij is given by δij = 1 if i = j and δij = 0 if i ≠ j

ME555 : Computational Fluid Dynamics 92 I. Sezai – Eastern Mediterranean University

46
For Πij term:
The pressure – strain interactions are the most difficult to model
Their effect on the Reynolds stresses is caused by two distinct
physical processes:
1. Pressure fluctuations due to two eddies interacting with each
other
2. Pressure fluctuations due to the interactions of an eddy with a
region of flow of different mean velocity
Its effect is to make Reynolds stresses more isotropic and to
reduce the Reynolds shear stresses
Measurements indicate that;
The wall effect increases the isotropy of normal Reynolds
stresses by damping out fluctuations in the directions normal to
the wall and decreases the magnitude of the Reynolds shear
stresses.

ME555 : Computational Fluid Dynamics 93 I. Sezai – Eastern Mediterranean University

A comprehensive model that accounts for all these effects is given in


Launder et al (1975). They also give the following simpler form favoured by
some commercial available CFD codes:

 2   2  (3.61)
 ij  C1  Rij  k ij   C2  Pij  P ij 
k 3   3 
with C1  1.8; C2  0.6

Turbulent kinetic energy k is

k
1
2
1

 R11  R22  R33   u12  u22  u32
2

ME555 : Computational Fluid Dynamics 94 I. Sezai – Eastern Mediterranean University

47
The six equations for Reynolds stress transport are solved along with a
model equation for the scalar dissipation rate ε. Again a more exact form is
found in Launder et al (1975), but the equation from the standard k – ε
model is used in commercial CFD for the sake of simplicity
D v   2
 div  t grad    C1 f1 2vt Sij  Sij  C2 (3.62)
Dt    k k
where C1  1.44 and C2  1.92

ME555 : Computational Fluid Dynamics 95 I. Sezai – Eastern Mediterranean University

Boundary conditions for the RSM model


The usual boundary conditions for elliptic flows are required for the
solution of the Reynolds stress transport equations:
inlet: specified distributions of Rij and ε
outlet and symmetry: ∂Rij /∂n = 0 and ∂ ε/∂n = 0
free stream: Rij = 0 and ε = 0 are given
or, ∂Rij /∂n = 0 and ∂ ε/∂n = 0
solid wall: use wall functions relating Rij to either k or (uτ)2
e.g. u12  1.1k , u2 2  0.25k , u32  0.66k ,  u1u2  0.26k
In the absence of any information, inlet distributions are calculated from
2 k 3/2
U re f T i  ;
2
k    C 3 / 4 ;   0 .0 7 L ;
3 
1
u 1 2  k ; u 2 2  u 3 2  k ; u iu j  0 ( i  j )
2
where ℓ is the characteristic length of the equipment (e.g. pipe diameter)
ME555 : Computational Fluid Dynamics 96 I. Sezai – Eastern Mediterranean University

48
Boundary conditions for RSM model
For computations at high Reynolds numbers wall-function-type boundary
conditions can be used which are very similar to those of the k-ε model .
Near wall Reynolds stress values are computed from formulae such as
Rij  uiuj  cij k
where cij are obtained from measurements.

Low Reynolds number modifications to the model can be incorporated to add


the effects of molecular viscosity to the diffusion terms and to account for
anisotropy in the dissipation rate in the Rij-equations.

Wall damping functions to adjust the constants of the ε-equation and a modified
dissipation rate variable
 (   2v(k 1/ 2 / y ) 2 )
give more realistic modeling near solid walls.

ME555 : Computational Fluid Dynamics 97 I. Sezai – Eastern Mediterranean University

Similar models exist for the 3 scalar transport terms ui  of eqn (3.32)
in the form of PDE’s.
In commercial CFD codes a turbulent diffusion coefficient
 t  t /  
is added to the laminar diffusion coefficient, where
   0.7
for all scalars.

ME555 : Computational Fluid Dynamics 98 I. Sezai – Eastern Mediterranean University

49
Assessment of RSM model

Table 3.6 Reynolds stress equation model assessment.

ME555 : Computational Fluid Dynamics 99 I. Sezai – Eastern Mediterranean University

Asssesment of performance of RSM model

Figure 3.16
Comparison of predictions of RSM and standard k-ε model with measurements on a
high-lift Aerospatiale aerofoil: (a) pressure coefficient; (b) skin friction coefficient
Source: Leschziner, in Peyret and Krause (2000)

ME555 : Computational Fluid Dynamics 100 I. Sezai – Eastern Mediterranean University

50
Incorrect usage of the terms: High and low Reynolds number
High or low Reynolds number turbulence models do not necessarily
refer to models which are used to simulate flows having high or low
speeds. In this usage of the term, the Reynolds number is based on the
distance from the wall. In the near-wall region the flow velocity and
distance to the wall is low so that the Reynolds number is low. Then,
low Reynolds number models refer to turbulence models which can
simulate the flow in near-wall region where the viscous effects
dominate, by integrating the equations right to the wall, without
resorting to wall functions.
High Reynolds number models on the other hand refer to models
which can simulate the flow in the fully turbulent region only, where
the Reynolds number is high. In this region the turbulence effects
dominate over the viscous effects. The model equations are not valid
in the near wall region. As a result, high Reynolds number model
equations are not integrated right to the wall. These models use wall
functions to find the variables in the near wall region.
ME555 : Computational Fluid Dynamics 101 I. Sezai – Eastern Mediterranean University

Shortcomings of two-equation models such as k-ε model


Low Reynolds number flows: Very rapid changes occur in the
distribution of k and ε as we reach the buffer layer between the fully
turbulent region and the viscous sublayer. To force the rapid changes
to k and ε, the constants of the high Reynolds number model are
multiplied by exponential damping functions which require large
number of grid points to resolve the changes. As a result, the system
of equations become stiff which makes the numerical solution difficult
to converge.
Rapidly changing flows: The Reynolds stress   uiuj is proportional
to Sij in two-equation models. This only holds in equilibrium flows
where the rates of production and dissipation of k are roughly in
balance. In rapidly changing flows this is not true.

ME555 : Computational Fluid Dynamics 102 I. Sezai – Eastern Mediterranean University

51
Shortcomings of two-equation models such as k-ε model
Stress anisotropy: Two equation model predicts normal stresses   ui2
which are all approximately equal to –⅔ρk if a thin shear layer is
simulated. Experimental data presented in section 3.4 showed that this
is not correct. In spite of this the k-e model performs well in such
flows because the gradients of normal turbulent stresses   ui2 are
small compared with the gradient of the dominant turbulent shear
stress   uv . In more complex flows the gradients of normal
turbulent stresses are not negligible and can drive significant flows.
These effects can not be predicted by the two equation models.
Strong adverse pressure gradients and recirculation regions: This
problem is also attributable to the isotropy of the predicted normal
stresses of the k-ε model. The k-ε model overpredicts the shear stress
and suppresses separation in flows over curved walls. This is a
significant problem in flows over airfoils, e.g. in aerospace
applications.
ME555 : Computational Fluid Dynamics 103 I. Sezai – Eastern Mediterranean University

Shortcomings of two-equation models such as k-ε model


Extra strains: Streamline curvature, rotation and extra body forces all
give rise to additional interactions between the mean strain rate and
the Reynolds stresses. These physical effects are not captured by
standard two-equation models.
RSM model addresses most of these problems adequately, but at the
cost of additional storage and computer time.
Below we consider some of the more recent advances in turbulence
modeling that seek to address some or all of the above problems.

ME555 : Computational Fluid Dynamics 104 I. Sezai – Eastern Mediterranean University

52
Advanced turbulence models
Advanced treatment of the near-wall region: two-layer k-ε
model
The numerical instability problems associated with the wall damping
functions used in low Reynolds number k- ε models are avoided by
subdividing the boundary layer into two regions (Jongen,1997):
1) Fully turbulent region: Re d  y k /   Re*d , Re*d  50  200
In this region the standard k-ε model is used where the eddy
viscosity is defined by Eqn. (3.44); t ,t  C  k 2 / 
2) Viscous region, Red < Red*:
Only the momentum equations and the k-equation is solved
(Eqn.3.45). Turbulent viscosity in this viscous region is calculated
from
t , v  C    k
1/ 2

and ε is calculated from   k 3/ 2 /  


where
ME555 : Computational Fluid Dynamics 105 I. Sezai – Eastern Mediterranean University

Two-layer k-ε model


The length scales ℓμ and ℓε contain necessary damping effects in the
near wall region and are taken from the one–equation model of
Wolfshtein (1969):
   Cl y 1  exp( Re d / A 
   Cl y 1  exp( Re d / A 
where
Cl   C3/4 , A  2Cl , Re d  k 1/ 2 d / 

y is the normal distance to the wall.


In order to avoid instabilities associated with differences between μt,t
and μt,v at the join between the fully turbulent and viscous regions, a
blending formula is used to evaluate the eddy viscosity in
 ij    uiu j  2 t Sij  2 / 3 k ij

ME555 : Computational Fluid Dynamics 106 I. Sezai – Eastern Mediterranean University

53
The turbulent viscosity is calculated from
t   t ,t  (1   ) t ,v
1  Re d  Re*d  
where   1  tanh  
2  A 

The blending function λε is zero at the wall and tends to 1 in the fully
turbulent region when Red >> Red*.
The constant A can be adjusted in order to control the sharpness of the
transition from one model to the other. For example A =1,…,10 leads
to transitions occurring within a few cells, smaller values giving
sharper transitions.

ME555 : Computational Fluid Dynamics 107 I. Sezai – Eastern Mediterranean University

Geometry of the diffuser.


(Jongen, 1997)

Comparison of two-layer k-ε model with various


turbulence models in predicting Cp = (P2–P1) / (ρU12)
in a 2-D diffuser, (Jongen, 1997).

Profiles of ε at the exit section of the


diffuser (θ = 10), for different values of the
switching parameter Red* between the two
layers, (Jongen, 1997).
ME555 : Computational Fluid Dynamics 108 I. Sezai – Eastern Mediterranean University

54
Strain sensitivity: RNG k-ε model
RNG k - ε model (Renormalization Group). The RNG procedure:
systematically removes the small scales of motion from the governing
equations by expressing their effects in terms of large scale motions and a
modified viscosity. (Yakhot et al 1992):
RNG k-ε model equations for high Reynolds number flows:
(  k )
 div(  kU )  div  k eff grad k    ij  Sij   (3.65)
t
 (  )  2
 div(  U)  div   eff grad    C1* f1  ij  Sij  C2 f 2  (3.66)
t k k
with  ij    uiuj  2t Sij  2 / 3 k ij
k2
and eff    t ; t   C (3.67)

and C  0.0845;  k     1.39; C1  1.42; C2  1.68
 1   / 0  k
;    2Sij  Sij  ; 0  4.377;   0.012
1/2
and C1 *  C1 
1   3

ME555 : Computational Fluid Dynamics 109 I. Sezai – Eastern Mediterranean University

Only the constant β is adjustable; the above value is


calculated from near wall turbulent data. All other constants
are explicitly computed as part of the RNG process.
The ε-equation has long been suspected as one of the main
sources of accuracy limitations for the standard version of
the k-ε model and the RSM.
It is, therefore, interesting to note that the model contains a
strain-dependent correction term in the constant C1ε of the
production term in the RNG model ε-equation.
Its performance is better than the k-ε model for an
expanding duct, but actually worse for a contraction with
the same ratio.

ME555 : Computational Fluid Dynamics 110 I. Sezai – Eastern Mediterranean University

55
Spalart-Allmaras model
Has only one transport equation for kinematic eddy viscosity
parameter  . Turbulent eddy viscosity is computed from
t   f v1 (3.68)
where fv1 is the wall damping function given by
3 
f v1  3 3 ;  
  cv1 
which tends to unity for high Reynolds number flows (   t ).
and fv1 → 0 at the wall.
The Reynolds stresses are computed from

 U i U j 
 ij    uiuj  2 t Sij   f v1    (3.69)
 x j xi
 

ME555 : Computational Fluid Dynamics 111 I. Sezai – Eastern Mediterranean University

The transport equation for  is as follows:


2
 (  )        
 div( U )  div (    ) grad ( )  Cb 2    Cb1   Cw1    fw
t  xk xk  y 
(I) (II) (III) (IV) (V) (VI) (3.70)
Transport Transport of Rate of Rate of
Rate of change  of  by   by turbulent  production  dissipation
of  convection diffusion of  of 

     f

( y ) 2
v2

where   2ij ij  mean vorticity

1  U U j 
ij   i    mean vorticity tensor
2  x j xi  1/6
  1  cw6 3 
fv 2  1  , fw  g  6 6  wall damping functions
1   f v1  g  cw3 
ME555 : Computational Fluid Dynamics 112 I. Sezai – Eastern Mediterranean University

56
In the k-ε model the length scale is ℓ = k3/2/ε.
In a one-equation model ℓ cannot be computed since there is no
transport equation for k and ε. However, ℓ should be specified to
determine the dissipation rate. Inspection of Eqn. (3.70) reveals that ℓ
= κy has been used as a length scale. This is also the mixing length
used in developing the log-law for wall boundary layers.
The constants are:
1  Cb 2
 v  2 / 3,   0.4187, Cb1  0.1355, Cb 2  0.622, Cw1  Cb1   2
v
  
g  r  Cw 2 (r 6  r ), r  min  ,10  , Cw 2  0.3, Cw3  2
 
  2 2
y 
The model gives good performance for flows with adverse pressure
gradients. In complex geometries it is difficult to determine the length
scale, so the model is unsuitable for more general internal flows.
ME555 : Computational Fluid Dynamics 113 I. Sezai – Eastern Mediterranean University

Wilcox k-ω model


In the k-ε model the kinematic eddy viscosity is expressed as
t   ,   k velocity scale,   k 3/2 / length scale
However, ε is not the only possible length scale determining variable.
In fact, k-ω model (Wilcox,1994) uses ω = ε / k as the second
variable. Then, the length scale is ℓ = (k)1/2 / ε and
t   k /  (3.71)
The Reynolds stresses are computed from Boussinesq expression:
2  U U j  2
 ij    uiuj  2 t Sij   k ij  t  i     k ij
3  x j xi  3 (3.72)

The transport equation for k and ω for turbulent flow at high Reynolds
is:
(  k )    
 div(  kU)  div    t  grad (k )   Pk   *  k
t  k  

ME555 : Computational Fluid Dynamics 114 I. Sezai – Eastern Mediterranean University

57
where 2 U i
Pk  2 t Sij  Sij   k  ij
and 3 x j
 (  )    
 div(  U)  div    t  grad ( ) 
t     (3.74)
 2 U i 
  1  2  Sij  Sij    ij   1  2
 3 x j 
 
Transport Transport of k Rate of Rate of
Rate of change  of k or  by  or by turbulent  production  dissipation
of k or  convection diffusion of k or  of k or 

The model constants are


 k  2.0,    2.0,  1  0.553, 1  0.075,  *  0.09

ME555 : Computational Fluid Dynamics 115 I. Sezai – Eastern Mediterranean University

The k-ω model initially attracted attention because integration to the


wall does not require wall-damping functions in low Reynolds
number applications.
The value of k is set to zero at the wall.
ω tends to infinity at the wall, → use very large value or use
P  6 / ( 1 yP2 ) at the near wall grid point P.
At inlet boundaries k and ω must be specified.
At outlet boundaries zero gradient conditions are used.
In a free stream k→0 and ω→0, then μt→∞ from Eqn. (3.71). This
causes problems in free stream regions. So a non-zero value of ω
should be specified. Unfortunately, results of the model tend to be
dependent on the assumed free stream value of ω.

ME555 : Computational Fluid Dynamics 116 I. Sezai – Eastern Mediterranean University

58
Menter SST k-ω model
Menter (1992) noted that the results of the k-ε model are much less
sensitive to the (arbitrary) assumed values in the free stream, but its
near-wall performance is unsatisfactory for boundary layers with
adverse pressure gradients. He suggested a hybrid model using
(i) a transformation of the k-ε model into k-ω model in the near-wall
region
(ii) the standard k-ε model in the fully turbulent region far from the
wall.
The calculation of τij and the k-equation are the same as in Wilcox’s
original k-ω model, but the ε-equation is transformed into an ω-
equation by substituting ε = kω. This yields

 (  )      2 U i   k 
 div(  U)  div    t  grad ( )    2  2  Sij  Sij    ij    2  2  2
     
t   ,1    3 x j   ,2 xk xk

ME555 : Computational Fluid Dynamics 117 I. Sezai – Eastern Mediterranean University

 (  )    
 div(  U)  div    t  grad ( ) 
t      
  ,1  (3.75)
 (II) 
 

(I)
(III)

 2 U i   k 
  2  2  Sij  Sij    ij    2  2  2
 3 x j     ,2 xk xk

   (V)  
(IV) (VI)

Comparison with Eqn. (3.74) shows that (3.75) has an extra source
term (VI): the cross-diffusion term, which arises during the ε = kω
transformation of the diffusion term in the ε-equation.

ME555 : Computational Fluid Dynamics 118 I. Sezai – Eastern Mediterranean University

59
Revised Menter SST k-ω model (2003)
Model constants:
 k  1.0,   ,1  2.0,   ,2  1.17,  2  0.44,  2  0.083,  *  0.09
Blending functions:
The differences between the μt values computed from the standard k-ε
model in the far field and the k-ω model near the wall may cause
instabilities. To prevent this, blending functions are introduced in the
equation to modify the cross diffusion term. Blending functions are
also used for model constants:
C  F1C1  (1  F1 )C2 (3.76)
where C1 = constants in original k-ω model (inner constants)
C2 = constants in Menter’s transformed k-ε model (outer
constants)
F1 = a blending function

ME555 : Computational Fluid Dynamics 119 I. Sezai – Eastern Mediterranean University

e.g. FC = FC(ℓt / y, Rey) is a function of the ratio of turbulence ℓt =


(k)1/2/ω and distance y to the wall and a turbulence Reynolds number
Rey = y2ω/ν.
The functional form of FC is chosen such that
(i) it is zero at the wall
(ii) tends to 1 in the far field
(iii) produces a smooth transition around a distance half way between
the wall and the edge of the boundary layer.
In this way, the model combines the good near-wall behaviour of the
k-ω model with the robustness of the k-ε model in the far field.

ME555 : Computational Fluid Dynamics 120 I. Sezai – Eastern Mediterranean University

60
Limiters
Eddy viscosity models tend to overpredict k in regions where
Sij is high. This problem is sometimes referred to as
“stagnation point anomaly” [Durbin, 1996]. To avoid such
overprediction, turbulent viscosity μt should be limited. The
limiter proposed by Medic and Durbin [2002] is

t k  
 min  , 
C  k   6C S 
  

where α = 0.6
S  ( Sij Sij )1/ 2

ME555 : Computational Fluid Dynamics 121 I. Sezai – Eastern Mediterranean University

Limiter in SST k-ω model


Limiters:
The eddy viscosity in SST k-ω model is limited to give improved
performance in flows with adverse pressure gradients and wake
regions:
a1  k (3.77a)
t 
max(a1 , SF2 )
where S  2Sij Sij , a1  a constant, F2  a blending function

The k-production is limited to prevent the build-up of turbulence in


stagnation regions:
 2 U i 
Pk  min  10  *  k , 2 t Sij  Sij   k  ij  (3.77b)
 3 x 
 j 

ME555 : Computational Fluid Dynamics 122 I. Sezai – Eastern Mediterranean University

61
Menter SST k-ω model: summary

ME555 : Computational Fluid Dynamics 123 I. Sezai – Eastern Mediterranean University

Menter SST k-ω model: summary

C1 = 5/9, C2 = 0.44

ME555 : Computational Fluid Dynamics 124 I. Sezai – Eastern Mediterranean University

62
Menter SST k-ω model: summary
Each of the constants is a blend of an inner (1) and outer (2) constants,
blended via:
C  F1C1  (1  F1 )C2
where C1 → inner (1) constants, and C2 → outer (2) constants.
Note that it is generally recommended to use a production limiter
where P in the k-equation is replaced with
P  min P,10  *  k  
The boundary conditions recommended are:
6
wall  10
1 (y1 ) 2
k wall  0

ME555 : Computational Fluid Dynamics 125 I. Sezai – Eastern Mediterranean University

Assessment of turbulence models for aerospace applications


External Aerodynamics:
The Spalart-Allmaras, k-ω and SST k-ω models are all suitable
The SST k-ω model is most general; it gives superior performance for
zero pressure gradient and adverse pressure gradient boundary layers.
General purpose CFD:
The Spalart-Allmaras model is unsuitable, but the k-ω and SST k-ω
models can both be applied.
They both have a similar range of strengths and weaknesses as the k-ε
model and fail to include accounts of more subtle interactions between
turbulent stresses and mean flow compared with RSM.

ME555 : Computational Fluid Dynamics 126 I. Sezai – Eastern Mediterranean University

63
The k-ε-v2-f turbulence model
In the k-ε-v2-f model of Durbin, (1991, 1993, 1995) two additional
equations, apart from the k and ε-equations, are solved:
(i) the normal stress, v22
(ii) a relaxation function f for the production of k, (f = Pk /k)
In usual eddy-viscosity models wall effects are accounted for through
wall functions.
In the k-ε-v2-f model the problem of accounting for the wall damping
of v22 is simply resolved by solving its transport equation similar to
RSM model:
 (  v22 )     v22
 div(  v22 U)  div    t  grad v 
2 
2
  kf   
t  k   k

ME555 : Computational Fluid Dynamics 127 I. Sezai – Eastern Mediterranean University

An equation for the production of k is formulated in terms of a


function f = Pk /k:
 2 / 3  v22 / k 
L div( grad ( f ))  f  (1  C1 ) 
2   C Pk
2
T k
where  k    1/2

T  max  , 6    time scale


     
 k3  3  
1/2

L  CL ,  2  max  2 , C2    ,   length scale


    
Note that: divgrad        / x   / y 2   / z 2
2 2

The eddy viscosity is given by:  t   /   C v22T


Constants: C  0.19, C1  1.4, C2  0.3, CL  0.3, C  70.0
Boundary conditions at the wall: 2 2
20 v2
k  0, v22  0,   2 k / y 2 , f   , y  normal distance to the wall
 y4
ME555 : Computational Fluid Dynamics 128 I. Sezai – Eastern Mediterranean University

64
A variant of the k-ε-v2-f turbulence model: ς-f model
In the k-ε-v2-f model of Durbin (1995), f is proportional to y4 near the
wall. As a result, the boundary condition for f makes the equation
system numerically unstable.
An eddy-viscosity model based on k-ε-v2-f model of Durbin is
proposed by Hanjalic et.al. (2004), which solves a transport equation
for the velocity scales ratio   v2 / k instead of v2
2 2

Thus, the resulting model is more robust and less sensitive to grid
non-uniformities.
Eddy viscosity is defined as
 t  C  kT
where   v22 / k
can be interpreted as the ratio of the wall-normal velocity time scale Tv
and the general (scalar) turbulence time scale T:
Tv  v22 /  and T  k / 
ME555 : Computational Fluid Dynamics 129 I. Sezai – Eastern Mediterranean University

A variant of the k-ε-v2-f turbulence model: ς-f model


The complete set of the model equations are:
( )
 div(U)  0
t
(U )
 div(UU)  div  eff gradU   sMx
t
(V )
 div(VU)  div  eff gradV   sMy
t
(W )
 div(WU)  div  eff gradW   sMz
t
( )     
 div( U)  div    t  grad     f  Pk
t    k
   
(k )    
 div(kU)  div    t  grad k   Pk  
t   k  
( )      C P  C 2
 div( U)  div    t  grad    1 k
t     T
1 P  2
L2div( grad ( f ))  f   C1 1 C2 k     
T   3 
ME555 : Computational Fluid Dynamics 130 I. Sezai – Eastern Mediterranean University

65
where t   C  kT
 k a    
1/2

T  max  min  ,  , CT   
  6C S  
   
   
  k 3/2 k 1/ 2   3  
1/4

L  CL max  min  ,  , C   
 6C S  
     


a  0.6, C =0.22, C 1 =1.4(1+0.012/ ), C 2 =1.9, C2 =0.65,  k =1,   =1.3,
   1.2, CT  6.0, CL  0.36, C  85
  U    V    W  P
*
sMx   eff    eff    eff 
x  x  y  x  z  x  x
  U    V    W  P*
sMy   eff    eff    eff 
x  y  y  y  z  y  y
  U    V    W  P*
sMz   eff    eff    eff 
x  z  y  z  z  z  z
eff    t , P*  P  (2 / 3)  k , Pk  2t Sij Sij , S  2 Sij Sij
2 2 2 2 2 2
 U   V   W  1  U V  1  W U  1  V W 
Sij Sij                   
 x   y   z  2  y x  2  x z  2  z y 
ME555 : Computational Fluid Dynamics 131 I. Sezai – Eastern Mediterranean University

Boundary conditions at the wall:


2
k  0,   0,   2 k / y 2 , f   , y  normal distance to the wall
y2
Note that both the nominator and the denominator of fw ( f at the wall)
are proportional to y2 instead of y4 as in the Durbin’s v2-f model.
This brings improved stability to computational scheme.

ME555 : Computational Fluid Dynamics 132 I. Sezai – Eastern Mediterranean University

66
3.5.4 Algebraic stress equation models (ASM)
- ASM is an economical way of accounting for the
anisotropy of Reynolds stresses.
- Avoids solving the Reynolds stress transport eqns.
- Instead, uses algebraic eqns to model Reynolds
stresses.
The Simplest method is:
To neglect the convection and diffusion terms
altogether
A more generally applicable method is:
To assume that the sum of the convection and diffusion
terms of the Reynolds stresses is proportional to the
sum of the convection and diffusion terms of turbulent
kinetic energy
ME555 : Computational Fluid Dynamics 133 I. Sezai – Eastern Mediterranean University

Duiu j uiuj  Dk  (3.78)


 Dij     transport of k (i.e. div)terms  
Dt k  Dt 
uiuj

k

 uiu j  Sij   
Introducing approximation (3.78) into the Reynolds stress transport equation
(3.55) with production term Pij (3.57), modeled dissipation rate term (3.60)
and pressure – strain correction term (3.61) on the right hand side yields
after some arrangement the following algebraic stress model:

2  CD  2 k
Rij  uiu j  k ij     Pij  P ij  (3.79)
3  1  P/
C  1  3 

- uiu j appear on both sides (on rhs within Pij ). For swirling flows, CD=0.55,
C1= 2.2
- eqn(3.79) is a set of 6 simultaneous algebraic eqns. for 6 unknowns, uiu j
- solved iteratively if k and ε are known.
- The standard
ME555 k-ε model
: Computational eqns has
Fluid Dynamics 134to be solved also (3.44
I. Sezai - 3.47)
– Eastern Mediterranean University

67
Algebraic stress model assessment

ME555 : Computational Fluid Dynamics 135 I. Sezai – Eastern Mediterranean University

Non-linear k-ε models


The standard k-ε model uses the Boussinesq approximation (3.33)
 U U  2
    uu    
 x
    k
i j
(3.33)
x  3
ij i j t ij
 j i
and eddy viscosity expression (3.44).
t  C     C k 2 /  (3.44)
Hence   uiu j   ij   ij ( Sij , k ,  ,  ) (8.80)
This relationship implies that the turbulence adjusts itself
instantaneously as it is convected through the flow domain. However,
the viscoelastic analogy holds that the adjustment does not take place
immediately. → τij should also be a function DSij / Dt. So,
DSij
  uiu j   ij   ij ( Sij , , k, , )
Dt
In RSM, τij is actually regarded as a transported quantity.
Bringing in a dependence on Dsij /Dt can be regarded as a partial
account of Reynolds stress transport, which recognises that the state
of turbulence lags behind the rapid changes.
ME555 : Computational Fluid Dynamics 136 I. Sezai – Eastern Mediterranean University

68
Elaborating this idea, a non-linear k-ε model is proposed. This
approach involves the derivation of asymptotic expansions for the
Reynolds stresses which maintain terms that are quadratic in
velocity gradients.
The nonlinear k-ε model (Speziale, 1987):

2 k2
 ij    uiuj    k ij   C 2Sij (3.82)
3 
k3  1 1 o 
 4CD C2 2  Sim  S mj  Smn  Smn ij  Sijo  Smm  ij 
  3 3 
Sij  U U j 
where Sijo   U  grad ( Sij )   i  S mj   S mi  and CD  1.68
t  xm xm 

The nonlinear k - ε model accounts for anisotropy.


Accounts for the secondary flow in non-circular duct flows.

ME555 : Computational Fluid Dynamics 137 I. Sezai – Eastern Mediterranean University

The simplest non-linear eddy viscosity model


The simplest non-linear eddy viscosity model relates the Reynolds
stresses to quadratic tensor products of Sij and Ωij:
2
 ij    uiuj  2 t Sij   k ij (3.83)
3
k 1  
C1t  Sik  S jk  S kl  Skl ij  
 3  
k 
C2 t  Sik   jk  S jk  ik   quadratic terms
 
k 1  
C3 t  ik   jk   kl   kl ij  
 3  

where Sij  1/ 2(U i / x j  U j / xi ), ij  1/ 2(U i / x j  U j / xi )

Quadratic terms allow Reynolds stresses to be different.


The model has the potential to capture anisotropy effects.
C1, C2 and C3 are constants in addition to the 5 constants of the
original k-ε model.

ME555 : Computational Fluid Dynamics 138 I. Sezai – Eastern Mediterranean University

69
Cubic k-ε model
Craft et al. (1996) demonstrated that it is necessary to introduce cubic
tensor products to obtain the correct sensitising effect for interactions
between Reynolds stress production and streamline curvature.
They also included:
Variable Cμ with functional dependence on Sij and Ωij
Ad hoc modification of the ε-equation to reduce the overprediction of
the length scale, leading to poor shear stress predictions in separated
flows.
Wall damping functions to enable integration of the k- and ε-equations
to the wall through the viscous sub-layer.
The performance of cubic k-ε model is very close to that of RSM.

ME555 : Computational Fluid Dynamics 139 I. Sezai – Eastern Mediterranean University

Large Eddy Simulation (LES)


Up to now, developing a general-purpose RANS turbulence model
suitable for a wide range of practical applications have not been
possible.
This is due to differences in the behaviour of large and small eddies.
Small eddies are nearly isotropic, but large eddies are anisotropic and
their behaviour is dictated by the geometry, boundary conditions and
body forces.
Problem dependence of the largest eddies complicates the search for
widely applicable models using RANS.
An alternative approach to the problem is large eddy simulation
(LES) which computes the larger eddies with a time dependent
simulation, but tries to model only the smaller eddies.
The universal behaviour of the smaller eddies is easier to capture with
a compact model.

ME555 : Computational Fluid Dynamics 140 I. Sezai – Eastern Mediterranean University

70
Instead of time averaging, LES uses a spatial filtering operation to
separate the larger and smaller eddies.
To resolve all those eddies with a length scale greater than a certain
width, the method selects a filtering function and a certain cutoff
width.
Then, the filtering operation is performed on the time-dependent flow
equations.
During spatial filtering, information relating to the smaller, filtered-
out turbulent eddies is destroyed.
This, and interaction effects between the larger, resolved eddies and
the smaller unresolved ones, gives rise to sub-grid-scale stresses or
SGS stresses.
Their effect on the resolved flow must be described by means of an
SGS model.

ME555 : Computational Fluid Dynamics 141 I. Sezai – Eastern Mediterranean University

In LES,
- the time-dependent, space filtered flow equations
- together with the SGS model of the unresolved stresses
are solved on a grid of control volumes
The solution gives:
- the mean flow and
- all turbulent eddies at scales larger than the cutoff width.

ME555 : Computational Fluid Dynamics 142 I. Sezai – Eastern Mediterranean University

71
Spatial filtering of unsteady Navier-Stokes equations
Filters are familiar separation devices in electronics that are designed
to split an input into a desirable, retained part and an undesirable,
rejected part.
Filtering Functions:
In LES spatial filtering is done by using a filter function G(x, x′,Δ):
  
 (x, t )     G(x, x, ) (x, t )dxdx dx
  
1 2 3
(3.84)
where  (x, t )  filtered function
 (x, t )  original (unfiltered) function
  filter cutoff width
In this section overbar indicates spatial filtering, not time averaging.
Only in LES, the integration is not carried out in time but in three-
dimensional space.

ME555 : Computational Fluid Dynamics 143 I. Sezai – Eastern Mediterranean University

Common filtering functions in LES


Top hat filter:
1/  if x  x   / 2
3

G (x, x, )   (3.85a)


0 if x  x   / 2

Gaussian filter:
  
3/2
 x  x 
2

G (x, x, )   2  exp   ,   6


     2  (3.85b)

Spectral cutoff:
3
sin[( xi  xi) /  (3.85c)
G (x, x, )  
i 1 ( xi  xi)

ME555 : Computational Fluid Dynamics 144 I. Sezai – Eastern Mediterranean University

72
The top-hat filter is used in finite volume implementations of LES.
Gaussian and spectral cutoff filters are preferred in research.
Δ is intended as an indicative measure of the size of eddies that are
retained in the computations and the eddies that are rejected.
In principle we can choose Δ to be any size, but in finite volume
method it is pointless to choose Δ < grid size, since only a single
nodal value of the variable is retained over the control volume, so all
finer detail is lost anyway.
The most common selection is

  3 xyz  (Vcell )1/3 (3.86)


where Δx, Δy and Δz are the length of a cubic cell in x, y and z-
directions.

ME555 : Computational Fluid Dynamics 145 I. Sezai – Eastern Mediterranean University

Filtering in 1-D

In LES we filter (volume average) the equations. In 1-D we get


1 x  0.5 x
x x 0.5 x
 ( x, t )   ( , t )d 

ME555 : Computational Fluid Dynamics 146 I. Sezai – Eastern Mediterranean University

73
Filtered unsteady Navier-Stokes equations
The unsteady Navier-Stokes equations for a fluid with constant
viscosity μ are
 (2.4)
 div (  u)  0
t
( u) p
 div   uu      div ( grad ( u ))  Su (2.37a)
t x
 (  v) p
 div   vu      div ( grad ( v))  Sv (2.37b)
t y
 (  w) p (2.37c)
 div   wu      div ( grad ( w))  S w
t z
If the flow is also incompressible → div(u) = 0, → Su, Sv, Sw = 0

ME555 : Computational Fluid Dynamics 147 I. Sezai – Eastern Mediterranean University

Filtering of the equations (2.4) and (2.37) yields:


 (3.87)
 div (  u )  0
t
( u ) p
t
 
 div  uu     div ( grad ( u ))
x
(3.88a)

(  v ) p
t
 
 div  vu     div ( grad ( v ))
y
(3.88b)
 (  w) p
t
 z

 div  wu     div ( grad ( w)) (3.88c)
The overbar indicates a filtered flow variable.
The problem is, how to compute div(  u)
div(  u)  div( u )  [div(  u)  div( u)]
The second term on the rhs is modeled.

ME555 : Computational Fluid Dynamics 148 I. Sezai – Eastern Mediterranean University

74
Substitution of the modeled div(  u) into (3.88a-c) yields the LES
momentum equations:
( u ) p
 div   u u      div ( grad ( u ))  (div(  uu)  div(  u u ))
t x
(3.89a)
(  v ) p
 div   v u      div ( grad ( v ))  (div(  vu)  div(  v u ))
t y
 (  w) p (3.89b)
 div   w u      div ( grad ( w))  ( div(  wu)  div(  wu ))
t z (3.89c)
(I) (II) (III) (IV) (V)

The filtered momentum eqns. are similar to RANS momentum eqns.


(3.26a-c) or (3.27a-c).
The last terms (V) are caused by the filtering operation similar to the
Reynolds stresses in RANS equations that arose as a result of time
averaging.
They can be considered as a divergence of a set of stresses τij
ME555 : Computational Fluid Dynamics 149 I. Sezai – Eastern Mediterranean University

The i-component of these terms can be written as follows:


 (  ui u   ui u )  (  ui v   ui v )  (  ui w   ui w)
div(  ui u   ui u )   
x y z
 ij (3.90a)

x j
where  ij   ui u   ui u   ui u j   ui u j (3.90b)
τij are termed as sub-grid-scale stresses.
Remembering that a flow variable (x, t) can be decomposed as the
sum of
(i) the filtered function  (x, t ) with spatial variations that are larger
than the cutoff width and are resolved by the LES computations and
(ii) ′(x, t), which contains unresolved spatial variations at length
scales smaller than the filter cutoff width,
we can write:
ME555 : Computational Fluid Dynamics 150 I. Sezai – Eastern Mediterranean University

75
 (x, t )   (x, t )   (x, t ) (3.91)
Using this decomposition in eqn. (3.90b) we can write the first term
on the far rhs as
 ui u j   (ui  ui)(u j  u j )   ui u j   ui u j   uiu j   uiu j
  ui u j  (  ui u j   ui u j )   ui uj   uiu j   uiu j
Then, we can write SGS stresses as
 ij   ui u j   ui u j  (  ui u j   ui u j )   ui uj   uiu j   uiuj (3.92)
 

 
Leonard stresses, Lij cross-stresses, Cij LES Reynolds
stresses, Rij

Leonard stresses, Lij, are due to effects at resolved scale. They are
caused due to the fact that    for space filtered variables, unlike in
time averaging, where  (t )       (t )

ME555 : Computational Fluid Dynamics 151 I. Sezai – Eastern Mediterranean University

The cross stresses Cij are due to interactions between the SGS eddies
and the resolved flow.
LES Reynolds stresses Rij are caused by convective momentum
transfer due to interactions of SGS eddies and are modeled with a so-
called SGS turbulence model.
The SGS stresses (3.92) must be modeled.

ME555 : Computational Fluid Dynamics 152 I. Sezai – Eastern Mediterranean University

76
Smagorinski-Lilly SGS model
Smagorinski (1963) suggested that, since the smallest turbulent eddies
are almost isotropic, we expect that the Boussinesq hypothesis (3.33)
might provide a good description of the effects of the unresolved
eddies on the resolved flow.
Thus in Smagorinsky’s model the local SGS stresses Rij are taken to
be proportional to the local rate of strain of the resolved flow:
Rij  Sij  1/ 2(ui / x j  u j / xi ) :
1  u u  1 (3.93)
Rij  2  SGS Sij  Rkk  ij    SGS  i  j   Rkk  ij
 x 
3  j xi  3
⅓Rkkδij performs the same function as the term ⅔ρkδij in eqn (3.33).
It ensures that the sum of the modeled normal SGS stresses is equal to
the kinetic energy of the of the SGS eddies.
Above model is used together with approximate forms of Leonard
stresses Lij and cross stresses Cij.
ME555 : Computational Fluid Dynamics 153 I. Sezai – Eastern Mediterranean University

In spite of the different nature of Lij and Cij, they are lumped together
with Rij in the current versions of the finite volume method.
Then, the whole stress τij is modeled as a single entity by means of a
single SGS turbulence model:
1  u u  1 (3.94)
 ij  2 SGS Sij   kk  ij    SGS  i  j    kk  ij
3  x j xi  3
 
Smagorinsky-Lilly model is based on Prandtl’s mixing length model:
 SGS  C 
Length scale: ℓ = Δ is used since Δ fixes the size of SGS eddies.
Velocity scale:     S where S  2Sij Sij
Substituting  and  into  SGS and defining  SGS   SGS
 SGS   (CSGS ) 2 S   (CSGS ) 2 2Sij Sij (3.95)
where 1  ui u j 
S  
 
2  x j xi 
ij

ME555 : Computational Fluid Dynamics 154 I. Sezai – Eastern Mediterranean University

77
Near the wall, μSGS becomes quite large because velocity gradients are
high there.
However, since the SGS turbulent fluctuations near a wall go to zero,
so must μSGS .
To ensure this, μSGS is multiplied with a damping function fμ, so that
 SGS   ( f  CSGS )2 S
where f   1  exp( y  / 26)
A more convenient way to dampen μSGS near the wall is simply to use
the RANS length scale as an upper limit, i.e.
  min (Vcell )1/3 ,  n 
n = distance to the nearest wall
Disadvantage of the Smagorinsky model: the “constant” CSGS is not
constant, but it is flow dependent. It is found to vary in the range from
CSGS = 0.065 (Moin, 1982) to CSGS = 0.25 (Jones, 1995).
ME555 : Computational Fluid Dynamics 155 I. Sezai – Eastern Mediterranean University

The difference in CSGS values is attributable to the effect of the mean


flow strain or shear.
This indicates that the behavior of the small eddies is not as universal
as was surmised at first.
→ Successful LES turbulence modeling might require
(i) a case by case adjustment of CSGS, or
(ii) a more sophisticated approach.

ME555 : Computational Fluid Dynamics 156 I. Sezai – Eastern Mediterranean University

78
Higher order SGS models
Boussinesq eddy viscosity hypothesis for Reynolds stresses given by
Eqn. (3.93) assumes that changes in the resolved flow take place
sufficiently slowly that the SGS eddies can adjust themselves
instantaneously to the rate of strain of the resolved flow field.
Instead of using a case by case adjustment of CSGS for different
applications, another approach is to use the idea of RANS modeling to
account for the transport effects. In this approach we define
Length scale: ℓ = Δ
Velocity scale:   kSGS where kSGS = SGS turbulent kinetic energy
Then  SGS   CSGS
  k SGS (3.96)
where   constant
CSGS
To account for the effects of convection, diffusion, production and
destruction on the SGS velocity scale we solve a transport equation to
determine the distribution of kSGS:
ME555 : Computational Fluid Dynamics 157 I. Sezai – Eastern Mediterranean University

 (  kSGS )  
 div(  kSGS u)  div  SGS grad (k SGS )   2  SGS Sij  Sij   SGS (3.96)
t  k   PkSGS

Dimensional analysis shows that the rate of dissipation εSGS of SGS


turbulent kinetic energy is related to length and velocity scales as
k 3/2
 SGS  C SGS
 (3.97)
C  constant
This model is implemented in commercial CFD code STAR-CD.

ME555 : Computational Fluid Dynamics 158 I. Sezai – Eastern Mediterranean University

79
Advanced SGS models
The Smagorinsky model is purely dissipative: the direction of energy
flow is from eddies at the resolved scales towards the sub-grid scales.
Quarini (1979) have shown that there is also energy flow in reverse
direction.
Furthermore, modeled SGS stresses using the Smagorinsky-Lilly
model do not correlate strongly with actual SGS stresses computed by
accurate DNS. (Clark et al. (1979), McMillan and Ferziger, (1979)).
These authors suggested that the SGS stresses should not be taken as
proportional to the strain rate of the whole resolved flow field, but
should be estimated from the strain rate of the smallest resolved
eddies.
Bardina et al. (1980) proposed a method to compute local values of
CSGS based on the application of two filtering operations, taking the
SGS stresses to be proportional to the stresses due to eddies at the
smallest resolved scale. They proposed:
ME555 : Computational Fluid Dynamics 159 I. Sezai – Eastern Mediterranean University

The Bardina model



 ij   C  ui u j  ui u j 
where C′ is an adjustable constant.
The term in the brackets can be evaluated from twice-filtered resolved
flow field information. The cutoff width of the test filter may be equal
to that of first filter or it may be coarser.
Bardina model has improved performance. But, the appearance of
negative viscosities generated stability problems.
They proposed adding a damping term with the form of the
Smagorinsky model (3.94)-(3.95) to stabilise calculations, which
yields a mixed model:
 
 ij   C  ui u j  ui u j  2  CSGS
2
 2 S Sij (3.99)

C′ depends on the cutoff width used for the second filtering, but C′ ≈ 1
ME555 : Computational Fluid Dynamics 160 I. Sezai – Eastern Mediterranean University

80
Twice filtering
Let’s filter velocity component u once more at a node I using a
cutoff width Δx. For simplicity we do it in 1D.

Box filter for a control volume.


(Lars Davidson, Lecture notes
Chalmers Univ.)

uI 
1 x / 2
x x / 2
u ( )d  
1
x  0

x / 2
u ( ) d  
0
x / 2

u ( )d  
1  x
 uA 
x  2
x 
2
uB 

Estimating u at locations A and B by linear interpolation gives
1  1 3  3 1  1
uI   uI 1  u I    uI  u I 1     uI 1  6u I  uI 1   uI
2  4 4  4 4  8
ME555 : Computational Fluid Dynamics 161 I. Sezai – Eastern Mediterranean University

Dynamic SGS model


Germano (1986) proposed a different decomposition of the turbulent
stresses for 2 different filtering operations with cutoff widths Δ1and Δ2
 ij(2)   ij(1)   Lij   ui u j  ui u j (3.100)
Second filter is coarser and is called the test filter with Δ2 = 2Δ1
Eqn. (3.100) is evaluated from the resolved flow data.
The SGS stresses are modeled using Smagorinsky’s model (3.94) –
(3.95) assuming same CSGS for both filtering operations. This yields:
1 (3.101a)
Lij  Lkk  ij  CSGS
2
M ij
3
where M ij  2 22 S Sij  212 S Sij (3.101b)
Lilly suggested a least squares approach to evaluate local values of
CSGS : LL
2
CSGS  ij ij
M ij M ij (3.102)
Note that the product of two tensors is a scalar.
ME555 : Computational Fluid Dynamics 162 I. Sezai – Eastern Mediterranean University

81
The test filter

Control volume for grid and test


filter

The test-filtered variables are computed by integration over the test


filter. The test filter is twice the size of grid filter, i.e. Δ2 = 2Δ1.
For example in 1D, u is computed as
u
1 E
2x W
udx 
1
2x W
P
 E
udx   udx
P 
1 u u u u
 uw x  ue x    W P  P E 
1

2x 2 2 2 
1
  uW  2uP  uE 
4
ME555 : Computational Fluid Dynamics 163 I. Sezai – Eastern Mediterranean University

The test filter in 3D

A 2D test filter control volume.


(Lars Davidson, Lecture notes at
Chalmers Univ Tech., Sweden)

In 3D, filtering at the test level is carried out in the same way by integrating over the
test cell assuming linear variation of the variables (Zang, et. al., 1993)
1
uI , J , K  (u I 1/2, J 1/2, K 1/2  uI 1/ 2, J 1/2, K 1/ 2
8
 u I 1/2, J 1/2, K 1/ 2  uI 1/ 2, J 1/ 2, K 1/2
 u I 1/2, J 1/ 2, K 1/ 2  uI 1/ 2, J 1/ 2, K 1/2
 u I 1/2, J 1/2, K 1/ 2  uI 1/2, J 1/2, K 1/ 2 )
ME555 : Computational Fluid Dynamics 164 I. Sezai – Eastern Mediterranean University

82
Initial and boundary conditions for LES
Initial conditions
For steady flows it is adequate to specify an initial field that conserves
mass with superimposed Gaussian random fluctuations with the
correct turbulence level or spectral content.
Solid walls
No-slip b.c. is used if equations are integrated to the wall. This
requires fine grids with near-wall grid points y+ ≤ 1
For high Re flows with thin boundary layers non-uniform grids
clustered near the walls is necessary.
Alternatively, wall functions can be used.

ME555 : Computational Fluid Dynamics 165 I. Sezai – Eastern Mediterranean University

Implementation of Wall functions in LES


The wall function models used in LES are:
1) The universal near wall model
If the first near-wall node P is in the laminar sublayer use
u  u y u y
for y   5 where u   w   friction velocity, y   
1/2

u  
If node P is in the fully turbulent region use
u 1  u y 
 ln E    , 30  y   500, E  9.793
u    
If node P is within the buffer region the two above layers are blended
using a function suggested by Kader (1981)
u
u   eulam

 e1/ uturb

, u 
where u
a( y  )4
 , a  0.01, b  5
1  by 
ME555 : Computational Fluid Dynamics 166 I. Sezai – Eastern Mediterranean University

83
Similarly, the general equation for the derivative du+/dy+ is
 
du   dulam 1/  duturb
 e  e
dy  dy  dy 

This approach allows the fully turbulent law to be easily modified and
extended to take into account other effects such as pressure gradients
or variable properties.
This formula guarantees reasonable representation of velocity profiles
in the cases where y+ falls inside the wall buffer region (3 < y+ < 10 ).
Friction velocity can be calculated from u  k P C
1/ 2 1/4

However, in most of the LES models k equation is not solved and kP is


not available. Since uτ appears on both sides of the non-linear
logarithmic law equation, then an iterative method should be used to
solve for uτ.
ME555 : Computational Fluid Dynamics 167 I. Sezai – Eastern Mediterranean University

2) The Werner-Wengle model


2) The Werner-Wengle model (Werner,1993)
To avoid the iterative solution procedure required for determining uτ,
Werner and Wengle proposed analytical integration of power-law
near-wall velocity distribution resulting in the following expressions
for the wall shear stress τw:
 2 uP 
 for uP  A2/(1 B )
 z 2 z
w   2

 1  B 11 BB     1 B
1 B B
1 B    
  A   +   uP  for uP  A2/(1 B )
 
2  z  A  z   2 z

uP = velocity parallel to wall, A = 8.3, B = 1/7, Δz = near-wall control


volume length scale (height if CV is rectangular).
Then uτ is calculated from u   w  
1/ 2

ME555 : Computational Fluid Dynamics 168 I. Sezai – Eastern Mediterranean University

84
3) Spalding’s law of the wall method
The universal velocity profile proposed by Spalding is a fit of the
laminar, buffer and logarithmic regions of the equilibrium boundary
layer into a single equation:
1  u 1 1 3
   
2
y  u   e 1  u  u  u 
E 2 6 
where κ = 0.42 and E = 9.1.
If y+ = yPρuτ/μ and u+ = uP/uτ are inserted into the above equation, an
equation of only one unknown uτ is obtained. This non-linear equation
can be solved using the Newton-Raphson method (Villiers, 2006):
u  un 1  f / f 
where (n–1) → previous iteration value
1  1 1 3
f  u   y   e u  1   u    u     u   
2

E 2 6 
f u 
y 1   u u  u

1 
1 
   
2 3
f     e   u  u 
u u u E  u u u 2u 
ME555 : Computational Fluid Dynamics 169 I. Sezai – Eastern Mediterranean University

The Newton-Raphson method converges rapidly to a tight tolerance


when applied in the above procedure.
By using initial values from the previous timestep, only 1 or 2
iterations are sufficient.
The predicted wall shear is then calculated from
 w   u2

ME555 : Computational Fluid Dynamics 170 I. Sezai – Eastern Mediterranean University

85
Implementation of LES
For the most general case incompressible fluids where µ is not
constant, LES momentum Eqns. (3.89) can be written as
   ui      ui u j    p
t

x j
  ui u j         
x j   x j xi   x j
 ij  
xi (3.105)
where  ij   ui u j   ui u j
The SGS turbulence model for τij is
1  u u  1
 ij  2  SGS Sij   kk  ij    SGS  i  j    kk  ij
 x  (3.94)
3  j xi  3
Then, the second term on the RHS of Eqn. (3.105) becomes

    u u j     1 

x j
 ij     SGS  i 
x j 
     kk  (3.106)
 x j xi   xi  3 

ME555 : Computational Fluid Dynamics 171 I. Sezai – Eastern Mediterranean University

Substituting (3.106) into (3.105) gives


   ui   p*    u u j  
t

x j
  ui u j     (    SGS )  i 
xi x j 
 
(3.106)
 x j xi  
where 1
p*  p   kk , (k  1,3)
3
Note that Eqn. (3.106) is the same as that of the Navier-Stokes Eqns.
given by (2.32a,b,c) with
μ replaced by (μ + μSGS) and
p replaced by p*
After the solution, τ11, τ22, τ33 can be obtained by solving the system of
equations (3.107), which are obtained from Eqn. (3.94):

 2 / 3 1/ 3 1/ 3  11   2  S 11 
 1/ 3 2 / 3 1/ 3     2 S  (3.107)
   22   22 
 
 1/ 3 1/ 3 2 / 3   33   2  S 33 
ME555 : Computational Fluid Dynamics 172 I. Sezai – Eastern Mediterranean University

86
Smagorinsky-Lilly SGS Model Equations in Generic Form
The Smagorinsky-Lilly SGS model equations can be written in generic form as
 (  )
 div(  U )  div( grad  )  s
t

Equation   s
Continuity 1 0 0
  u    v    w  p *
x-Momentum u eff  eff    eff    eff 
x  x  y  x  z  x  x
  u    v    w  p *
y -Momentum v eff  eff    eff    eff 
x  y  y  y  z  y  y
  u    v    w  p *
z -Momentum w eff  eff    eff    eff 
x  z  y  z  z  z  z

1  ui u j 
eff     SGS ,  SGS   (CSGS )2 S   (CSGS ) 2 2Sij Sij , Sij     ,
 2  x j xi 
2 2 2 2 2 2
 u   v   w  1  u v  1  w u  1  v w 
Sij Sij                      ,
 x   y   z  2  y x  2  x z  2  z y 
1
p*  p   kk (k  1,3), CSGS  0.065  0.25,   (Vcell )1/3
3
ME555 : Computational Fluid Dynamics 173 I. Sezai – Eastern Mediterranean University

References (in addition to that given in textbook)


G. Medic and P. A. Durbin (2002) . Towards improved prediction of heat transfer
on turbine blades, Journal of turbomachinery, 124:187-192.
P. A. Durbin(1996). On the k-ε stagnation point anomaly. International journal of
heat and fluid flow, 17(1):89–90.
T. Jongen and Y. P. Marx (1997). Design of an unconditionally stable, positive
scheme for the k-e and two-layer turbulence models, Computers and Fluids, vol.
26, No 5, pp. 469-487.
Y. Zang, R.L., Street and J.R. Hoseff, 1993 . A dynamic subgrid-scale eddy
viscosity model, Physics of fluids A, 5:3186-3196.
B. Kader, Temperature and Concentration Profiles in Fully Turbulent Boundary
Layers. Int. J. Heat Mass Transfer, 24(9):1541-1544, 1981
H. Werner and H. Wengle. Large-Eddy Simulation of Turbulent Flow Over and
Around a Cube in a Plate Channel. In Eighth Symposium on Turbulent Shear
Flows, Munich, Germany, 1991.
D.B. Spalding (1961). A single formula for the law of the wall. Journal of Applied
Mechanics, Trans ASME, Series E, Vol. 28, pp. 455-458.
E. Villiers (2006). The Potential of Large Eddy Simulation for the Modeling of
Wall Bounded Flows. PhD thesis, Imperial College of Science, Technology and
Medicine, Department of Mechanical Engineering.
ME555 : Computational Fluid Dynamics 174 I. Sezai – Eastern Mediterranean University

87
P. A. Durbin (1991). Near-wall turbulence closure modeling without damping
functions, Theoretical and Computational Fluid Dynamics, vol. 3, No. 1, pp. 1-13.
P. A. Durbin (1993). Application of a near-wall turbulence model to boundary layers
and heat transfer, Int. J. Heat and Fluid Flow, vol. 14, no. 4, pp. 316-323.
P. A. Durbin (1995). Separated flow computations with the k-ε-v2 model, AIAA
Journal, vol. 33, no. 4, pp.659-664.
K. Hanjalic , M. Popovac and M. Hadziabdic (2004). A robust near-wall elliptic-
relaxation eddy-viscosity turbulence model for CFD, Int. J. Heat and Fluid Flow,
vol. 25, pp. 1047-1051.
P. Moin and J. Kim (1982). Numerical investigation of turbulent channel flow.
Journal of Fluid Mechanics, vol. 118, pp.341–377.
W.P. Jones and M. Wille (1995). Large eddy simulation of a jet in a cross-flow. In
10th Symp. on Turbulent Shear Flows, pages 4:1 – 4:6, The Pennsylvania State
University.

ME555 : Computational Fluid Dynamics 175 I. Sezai – Eastern Mediterranean University

88

You might also like