You are on page 1of 48

Publication 97/2

An Introduction to Turbulence Models


Lars Davidson, http://www.tfd.chalmers.se/lada
Department of Thermo and Fluid Dynamics C HALMERS U NIVERSITY OF T ECHNOLOGY G oteborg, Sweden, November 2003

  ! "


Nomenclature 1 Turbulence 1.1 Introduction . . . . . . . . . . . . . . . . 1.2 Turbulent Scales . . . . . . . . . . . . . . 1.3 Vorticity/Velocity Gradient Interaction . 1.4 Energy spectrum . . . . . . . . . . . . . Turbulence Models 2.1 Introduction . . . . . . . . . . 2.2 Boussinesq Assumption . . . 2.3 Algebraic Models . . . . . . . 2.4 Equations for Kinetic Energy 2.4.1 The Exact Equation . 2.4.2 The Equation for 2.4.3 The Equation for 2.5 The Modelled Equation . . 2.6 One Equation Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3 5

# $ % &

12

'213(547698@6A8CB '213(54FG E6 8FA E6 8CB

')( ')# ')$ ')$ ')$ ')D ')& (3H (I' (Q( (Q( (Q# (Q$ (W% (Q& YQ( YQY Ycb YQ# YQ$ YQ$

Two-Equation Turbulence Models 3.1 The Modelled Equation . . . 3.2 Wall Functions . . . . . . . . . 3.3 The Model . . . . . . . . 3.4 The Model . . . . . . . 3.5 The Model . . . . . . . .

0RSP 0RUT 0RSV

. . . . .

22

Low-Re Number Turbulence Models 4.1 Low-Re Models . . . . . . . . . . . . . . . . . 4.2 The Launder-Sharma Low-Re Models . . . . 4.3 Boundary Condition for and . . . . . . . . . . . 4.4 The Two-Layer Model . . . . . . . . . . . . . 4.5 The low-Re Model . . . . . . . . . . . . . . . 4.5.1 The low-Re Model of Peng et al. . . . 4.5.2 The low-Re Model of Bredberg et al. .

0XRSP

28

0 RaP 0RUT 0RUT 0RUT

0RSP P`

Reynolds Stress Models 5.1 Reynolds Stress Models . . . . . . . . . . . . . . . 5.2 Reynolds Stress Models vs. Eddy Viscosity Models 5.3 Curvature Effects . . . . . . . . . . . . . . . . . . . 5.4 Acceleration and Retardation . . . . . . . . . . . .

YQ& bWH bF' bQb

38

  " !%$'&)( 10%2  ! # 3547698A@CBD'EGFIHQPAR

STVUWSYX`UWSba SdcT UWSdcX Sde T UWSfe X SfghTVUWSfgiX Sdp q r


0 6 68 E6 A E6 8 U

tv tw tX t t8 t 8 tyx E vX v vi w E w X w

constants in the Reynolds stress model constants in the Reynolds stress model constants in the modelled equation constants in the modelled equation constant in turbulence model energy (see Eq. 1.8); constant in wall functions (see Eq. 3.4) damping function in pressure strain tensor turbulent kinetic energy ( ) instantaneous (or laminar) velocity in -direction instantaneous (or laminar) velocity in -direction time-averaged velocity in -direction time-averaged velocity in -direction shear stresses uctuating velocity in -direction normal stress in the -direction uctuating velocity in -direction Reynolds stress tensor instantaneous (or laminar) velocity in -direction time-averaged velocity in -direction uctuating (or laminar) velocity in -direction normal stress in the -direction shear stress instantaneous (or laminar) velocity in -direction time-averaged velocity in -direction uctuating velocity in -direction normal stress in the -direction

T s X t 8t 8 u u8

u8

u8

h BD'EGFIHPR
P
boundary layer thickness; half channel height dissipation wave number; von Karman constant ( ) dynamic viscosity dynamic turbulent viscosity kinematic viscosity kinematic turbulent viscosity instantaneous (or laminar) vorticity component in

 d8

Hy bF'

u 8 -direction



dE 8
TG 8 V   V  V 

time-averaged vorticity component in -direction specic dissipation ( ) uctuating vorticity component in -direction turbulent Prandtl number for variable laminar shear stress total shear stress turbulent shear stress

Pc130

u8

u8

B!'F R#" 8%$'6 w


&
centerline wall

0%2 0 & !%$  @6 H "68 H@

Almost all uid ow which we encounter in daily life is turbulent. Typical examples are ow around (as well as in) cars, aeroplanes and buildings. The boundary layers and the wakes around and after bluff bodies such as cars, aeroplanes and buildings are turbulent. Also the ow and combustion in engines, both in piston engines and gas turbines and combustors, are highly turbulent. Air movements in rooms are also turbulent, at least along the walls where wall-jets are formed. Hence, when we compute uid ow it will most likely be turbulent. In turbulent ow we usually divide the variables in one time-averaged part , which is independent of time (when the mean ow is steady), and one uctuating part so that  . There is no denition on turbulent ow, but it has a number of characteristic features (see Tennekes & Lumley [41]) such as:  !   . Turbulent ow is irregular, random and chaotic. The ow consists of a spectrum of different scales (eddy sizes) where largest eddies are of the order of the ow geometry (i.e. boundary layer thickness, jet width, etc). At the other end of the spectra we have the smallest eddies which are by viscous forces (stresses) dissipated into internal energy. Even though turbulence is chaotic it is deterministic and is described by the Navier-Stokes equations. "# %$ '&")(' 0 . In turbulent ow the diffusivity increases. This means that the spreading rate of boundary layers, jets, etc. increases as the ow becomes turbulent. The turbulence increases the exchange of momentum in e.g. boundary layers and reduces or delays thereby separation at bluff bodies such as cylinders, airfoils and cars. The increased diffusivity also increases the resistance (wall friction) in internal ows such as in channels and pipes. @9 "& "132 " 54 3 768& . Turbulent ow occurs at high Reynolds number. For example, the transition to turbulent ow in pipes occurs that A5B"CED A5B"FGD layers at . 3S # in  boundary '&T IHP RQ , and . Turbulent ow is always three-dimensional. However, when the equations are time averaged we can treat the ow as two-dimensional. HPR# %&T&T7U 1 . Turbulent ow is dissipative, which means that kinetic energy in the small (dissipative) eddies are transformed into internal energy. The small eddies receive the kinetic energy from slightly larger eddies. The slightly larger eddies receive their energy from even larger eddies and so on. The largest eddies extract their energy from the mean ow. This process of transferred energy from the largest turbulent scales (eddies) to the smallest is called cascade process.

E6

6 E6

 7h 

h Q  ' y 7Q

(Q Y3HQ H 

' HQHQHQHQH

 7



Q19 y7 1&

. Even though we have small turbulent scales in the ow they are much larger than the molecular scale and we can treat the ow as a continuum.

H 

71 

F P @6B!"y41P R

As mentioned above there are a wide range of scales in turbulent ow. The larger scales are of the order of the ow geometry, for example the boundary layer thickness, with length scale and velocity scale . These scales extract kinetic energy from the mean ow which has a time scale comparable to the large scales, i.e.

E6



B'

4 1 2B
 

The kinetic energy of the large scales is lost to slightly smaller scales with which the large scales interact. Through the cascade process the kinetic energy is in this way transferred from the largest scale to smaller scales. At the smallest scales the frictional forces (viscous stresses) become too large and the kinetic energy is transformed (dissipated) into internal energy. The dissipation is denoted by which is energy per unit time and unit mass (   "!$# ). The dissipation is proportional to the kinematic viscosity times the uctuating velocity gradient up to the power of two (see Section 2.4.1). The friction forces exist of course at all scales, but they are larger the smaller the eddies. Thus it is not quite true that eddies which receive their kinetic energy from slightly larger scales give away all of that the slightly smaller scales but a small fraction is dissipated. However it is assumed that most of the energy (say 90 %) that goes into the large scales is nally dissipated at the smallest (dissipative) scales. The smallest scales where dissipation occurs are called the Kolmogorov scales: the velocity scale % , the length scale & and the time scale . We assume that these scales are determined by viscosity and dissipation . Since the kinetic energy is destroyed by viscous forces it is natural to assume that viscosity plays a part in determining these scales; the larger viscosity, the larger scales. The amount of energy that is to be dissipated is . The more energy that is to be transformed from kinetic energy to internal energy, the larger the velocity gradients must be. Having assumed that the dissipative scales are determined by viscosity and dissipation, we can express % , & and in and using dimensional analysis. We write

% 

'#



'#( 

XP
1

"!)#

4@' '2B
0

where below each variable its dimensions are given. The dimensions of the left-hand and the right-hand side must be the same. We get two equations,

"3 0 H y  0  68y7  7 h"

one for meters

' (5 ( U



4@' (WB
0  #

and one for seconds

which gives and so that


%

#"

 '21 b . In the same way we obtain the expressions for V T    T  X  T    U U 4 PcB P V   P! 4@' b B D % $ PH "8 6VD 4 8 @Q6 @6 h 4 "68 HQ@ $ H 68 "8 6V&
& & ! 0

R ' R RSY U

4@' YWB
0

The interaction between vorticity and velocity gradients is an essential ingredient to create and maintain turbulence. Disturbances are amplied the actual process depending on type of ow and these disturbances, which still are laminar and organized and well dened in terms of physical orientation and frequency are turned into chaotic, three-dimensional, random uctuations, i.e. turbulent ow by interaction between the vorticity vector and the velocity gradients. Two idealized phenomena in this interaction process can be identied: vortex stretching and vortex tilting. In order to gain some insight in vortex shedding we will study an idealized, inviscid (viscosity equals to zero) case. The equation for instanta ) reads [41, 31, 44] neous vorticity (

6 x d d 8(' x 8x

d 8 d E 8 TG8 d x 6A8(' x  d 8(' x x 0) 8 x21 6 1 ' x

4@' #WB
0

where ) 21 is the Levi-Civita tensor (it is  if 3 , 4 are in cyclic order, if 3 , 4 are in anti-cyclic order, and if any two of 3 , 4 are equal), and where ' denotes derivation with respect to . We see that this equation is not an ordinary convection-diffusion equation but is has an additional term on the right-hand side which represents amplication and rotation/tilting of the vorticity lines. If we write it term-by-term it reads

4 B x

ux

X'

R '

d T 6 T 'T  d X 6 T 'X  d 6 T ' d T 6 X 'T  d X 6 X 'X  d 6 X ' d T 6 'T  d X 6 'X  d 6 '


! ! ! ! !

! !

4@' $WB
0

! !

The diagonal terms in this matrix represent vortex stretching. Imagine a slender, cylindrical uid element which vorticity 7 . We introduce a cylindrical coordinate system with the -axis as the cylinder axis and as the

uT

uX

 65 &b 1 Q8 '

"3 0 H y  0  68y7  7 h"

d T


d T
' ' !#"$%'&(0)
0 0

radial coordinate (see Fig. 1.1) so that 7 stretch the cylinder. From the continuity equation

TUH UH B. 4d b

A positive

6 T 'T

will

6 T 'T

'  1 41

6 X B 'X H

we nd that the radial derivative of the radial velocity must be negative, i.e. the radius of the cylinder will decrease. We have neglected the viscosity since viscous diffusion at high Reynolds number is much smaller than the turbulent one and since viscous dissipation occurs at small scales (see p. 6). Thus there are no viscous stresses acting on the cylindrical uid element surface which means that the rotation momentum

6 X

Xd

4@' %QB
0

remains constant as the radius of the uid element decreases (note that also 1 is constant). Equation 1.7 shows that the vorticthe circulation 2 ity increases as the radius decreases. We see that a stretching/compressing will decrease/increase the radius of a slender uid element and increase/decrease its vorticity component aligned with the element. This process will affect the vorticity components in the other two coordinate directions. 65 H The off-diagonal terms in Eq. 1.6 represent vortex tilting. Again, take a " " ' slender uid element with its axis aligned with the axis, Fig. 1.2. The velocity gradient ' will tilt the uid element so that it rotates in clock-wise direction. The second term ' in line one in Eq. 1.6 gives a contribution to . Vortex stretching and vortex tilting thus qualitatively explains how interaction between vorticity and velocity gradient create vorticity in all three coordinate directions from a disturbance which initially was well dened in one coordinate direction. Once this process has started it continues, because vorticity generated by vortex stretching and vortex tilting interacts with the velocity eld and creates further vorticity and so on. The vorticity and velocity eld becomes chaotic and random: turbulence has been created. The turbulence is also maintained by these processes. From the discussion above we can now understand why turbulence always must be three-dimensional (Item IV on p. 5). If the instantaneous ow is two-dimensional we nd that all interaction terms between vorticity and velocity gradients in Eq. 1.6 vanish. For example if and all ! derivatives with respect to are zero. If the third line in Eq. 1.6 ! ! vanishes, and if (' the third column in Eq. 1.6 disappears. Finally, the

d X

d T

6 TX

d X6 TX

uX

  

6 8 sH
!

6 s H

6 s H

Q1

& U'9

`

dX

6 T 4u X B

dX


' (
0



!  )
0

remaining terms in Eq. 1.6 will also be zero since

d T 6 'X R 6 X '  H d X 6 T ' R 6 ' T s s Hy


! ! ! !

The interaction between vorticity and velocity gradients will, on average, create smaller and smaller scales. Whereas the large scales which interact with the mean ow have an orientation imposed by the mean ow the small scales will not remember the structure and orientation of the large scales. Thus the small scales will be isotropic, i.e independent of direction.

@ h D R#$ " 6 E

q 4 B 4@' DWB where Eq. 1.8 expresses the contribution from the scales with wave number to the turbulent kinetic energy 0 . The dimension between and  of wave number is one over length; thus we can think of wave number as proportional to the inverse of an eddys radius, i.e '21 1 . The total

The turbulent scales are distributed over a range of scales which extends from the largest scales which interact with the mean ow to the smallest scales where dissipation occurs. In wave number space the energy of eddies from to  can be expressed as

turbulent kinetic energy is obtained by integrating over the whole wave number space i.e.

 

q 4 B

4@' &WB
0

&

Q1

& U'9

`

q 4 B

(  %!' )    &(   )$  % )  0)(0)( !( " ' 0 Y   0 0      " & )$ "#( )  $ " )( & "$ '  0"  0% "%  "    0   %      0
0

The kinetic energy is the sum of the kinetic energy of the three uctuating velocity components, i.e.
0

X X X 0 ('  t  v  w (' t 8 t 8 4@' ' H B q The spectrum of is shown in Fig. 1.3. We nd region I, II and III which

correspond to: I. In the region we have the large eddies which carry most of the energy. These eddies interact with the mean ow and extract energy from the mean ow. Their energy is past on to slightly smaller scales. The eddies velocity and length scales are and , respectively. III. Dissipation range. The eddies are small and isotropic and it is here that the dissipation occurs. The scales of the eddies are described by the Kolmogorov scales (see Eq. 1.4) II. Inertial subrange. The existence of this region requires that the Reynolds number is high (fully turbulent ow). The eddies in this region represent the mid-region. This region is a transport region in the cascade process. Energy per time unit ( ) is coming from the large eddies at the lower part of this range and is given off to the dissipation range at the higher part. The eddies in this region are independent of both the large, energy containing eddies and the eddies in the dissipation range. One can argue that the eddies in this region should be characterized by the ow of energy ( ) and the size of the eddies . Dimensional reasoning gives

'21
0

4@' 'Q'2B This is a very important law (Kolmogorov spectrum law or the R#W13Y
 

q 4 B S)(10 32f P 54

law) which states that, if the ow is fully turbulent (high Reynolds

'H

Q1

& U'9

`
R #W13Y

number), the energy spectra should exhibit a -decay. This often used in experiment and Large Eddy Simulations (LES) and Direct Numerical Simulations (DNS) to verify that the ow is fully turbulent. As explained on p. 6 (cascade process) the energy dissipated at the small scales can be estimated using the large scales and . The energy at the large scales lose their energy during a time proportional to ) , which gives

21

21

4@' ')(WB
0

'Q'

0%2 0 & !%$   & " @6 H "68 H@

When the ow is turbulent it is preferable to decompose the instantaneous variables (for example velocity components and pressure) into a mean value and a uctuating value, i.e.

6G E6 8  t 8 8Q A E Q

4C( '2B
0

One reason why we decompose the variables is that when we measure ow quantities we are usually interested in the mean values rather that the time histories. Another reason is that when we want to solve the Navier-Stokes equation numerically it would require a very ne grid to resolve all turbulent scales and it would also require a ne resolution in time (turbulent ow is always unsteady). The continuity equation and the Navier-Stokes equation read

 6G8 2  4 6 8 B ' 8  4 6G876 x B ' x


where ' denotes derivation with respect to . Since we are dealing with incompressible ow (i.e low Mach number) the dilatation term on the righthand side of Eq. 2.3 is neglected so that

4 B x

H R '8





6G8(' x


ux

6 x ' 8 R (Y 8 x 6

1 '1 

 ' x

4C( W (B C4 ( W YB
0 0

 6G8  4 6G876 x B ' x

R ' 8  74 6G8(' x  6 x ' 8 B ' x


 #

4C( b B
0

Note that we here use the term incompressible in the sense that density is ) , but it does not mean that density is independent of pressure ( constant; it can be dependent on for example temperature or concentration. Inserting Eq. 2.1 into the continuity equation (2.2) and the Navier-Stokes equation (2.4) we obtain the time averaged continuity equation and NavierStokes equation

1 

  4 A E6 8 B ' 8 2  G E6 8  G E6 8 6 E x ' x 2


A new term appears on the right-hand side of Eq. 2.6 which is called the Reynolds stress tensor. The tensor is symmetric (for example ). It represents correlations between uctuating velocities. It is an additional stress term due to turbulence (uctuating velocities) and it is unknown. We need a model for to close the equation system in Eq. 2.6.

t Xt T

t 8 tx

H R E '8


 4 GE6 8(' x 

E6 x ' 8 B9R t 8 tx  ' x

4C( #WB
0

4C( $WB
0

t Tt X

t 8t x

')(

7  6  9"7

V 

V V

V   



( '
0

'H

& "  "" )


 

 

This is called the closure problem: the number of unknowns (ten: three velocity components, pressure, six stresses) is larger than the number of equations (four: the continuity equation and three components of the NavierStokes equations). For steady, two-dimensional boundary-layer type of ow (i.e. boundary layers along a at plate, channel ow, pipe ow, jet and wake ow, etc.) where

 &  U 9 

E E6 U u

Eq. 2.6 reads


E6 E6

 E

4C( %QB
0

E6

R u

E


E6 R tv  


4C( DWB
0

denotes streamwise coordinate, and coordinate normal to the ow. Often the pressure gradient is zero. To the viscous shear stress on the right-hand side of Eq. 2.8 appears an additional turbulent one, a turbulent shear stress. The total shear stress is thus

u uT

E6 1

E 1

uX

&TQih &  &T& b

tv V  E6 R

In the wall region (the viscous sublayer, the buffert layer and the logarithmic layer) the total shear stress is approximately constant and equal to the wall shear stress  , see Fig. 2.1. Note that the total shear stress is constant only close to the wall; further away from the wall it decreases (in fully developed channel ow it decreases linearly by the distance form the wall). At the wall the turbulent shear stress vanishes as , and the viscous  shear stress attains its wall-stress value  . As we go away from the D wall the viscous stress decreases and turbulent one increases and at they are approximately equal. In the logarithmic layer the viscous stress is negligible compared to the turbulent stress. In boundary-layer type of ow the turbulent shear stress and the ve locity gradient have nearly always opposite sign (for a wall jet this

t Xt

v H

'H

E6 1

')Y

7  6  9"7

X T


v H

v H

6X4FB

( (
0

) 

&

$#  )  " & " " "




tv R X
6

)


() $


$
0

is not the case close to the wall). To get a physical picture of this let us study the ow in a boundary layer, see Fig. 2.2. A uid particle is moving to with (the turbudownwards (particle drawn with solid line) from lent uctuating) velocity . At its new location the velocity is in average smaller than at its old, i.e. . This means that when the par) comes down to (where ticle at (which has streamwise velocity the streamwise velocity is ) is has an excess of streamwise velocity compared to its new environment at . Thus the streamwise uctuation is positive, i.e. and the correlation between and is negative ( ). If we look at the other particle (dashed line in Fig. 2.2) we reach the same conclusion. The particle is moving upwards ( ), and it is bringing a decit in so that . Thus, again, . If we study this ow for a long time and average over time we get . If we change the sign we will nd that the sign of of the velocity gradient so that also changes. Above we have used physical reasoning to show the the signs of and are opposite. This can also be found by looking at the production term in the transport equation of the Reynolds stresses (see Section 5). In cases where the shear stress and the velocity gradient have the same sign (for example, in a wall jet) this means that there are other terms in the transport equation which are more important than the production term. There are different levels of approximations involved when closing the equation system in Eq. 2.6.

t H

X E6 4 6X4

TB X E6 4 X B X 6 4 B X TB T t
E6 1 H

t H

tv H tv H

v H

tv H

E6 1

tv tv

 9  %

An algebraic equation is used to compute a turbulent viscosity, often called eddy viscosity. The Reynolds stress tensor is then computed using an assumption which relates the Reynolds stress tensor to the velocity gradients and the turbulent viscosity. This assumption is called the Boussinesq assumption. Models which are based on a turbulent (eddy) viscosity are called eddy viscosity mod-

  6 y7& 

'b

  %
1
els.

& &T &

&&

 U "

1"

In these models a transport equation is solved for a turbulent quantity (usually the turbulent kinetic energy) and a second turbulent quantity (usually a turbulent length scale) is obtained from an algebraic expression. The turbulent viscosity is calculated from Boussinesq assumption. These models fall into the class of eddy viscosity models. Two transport equations are derived which describe transport of two scalars, for example the turbulent kinetic energy and its dissipation . The Reynolds stress tensor is then computed using an assumption which relates the Reynolds stress tensor to the velocity gradients and an eddy viscosity. The latter is obtained from the two transported scalars.

Q3S) "7 6 y %& 

% ) S " 6 y7&  P

Q76 & &b &T& IH@ 4 

determining the length scale of the turbulence. Usually an equation for the dissipation is used.

6 y%&  Here a transport equation is derived for the Reynolds tensor t 8 tx . One transport equation has to be added for
P

Above the different types of turbulence models have been listed in increasing order of complexity, ability to model the turbulence, and cost in terms of computational work (CPU time).

H R`R`8@ RR`R# E ' $ 68 HQ@

In eddy viscosity turbulence models the Reynolds stresses are linked to the velocity gradients via the turbulent viscosity: this relation is called the Boussinesq assumption, where the Reynolds stress tensor in the time averaged Navier-Stokes equation is replaced by the turbulent viscosity multiplied by the velocity gradients. To show this we introduce this assumption for the diffusion term at the right-hand side of Eq. 2.6 and make an identication

 4IAE6 8(' x

6 E x ' 8CBAR t 8 tyx  ' x5  4




B 4 G E6 8(' x

E6 x ' 8 B  ' x

which gives

t 8 tyx
0

4A R E6 8#' x

E6 x ' C8 Bf

If we in Eq. 2.9 do a contraction (i.e. setting indices side gives

3  4

4C( &WB
0

) the right-hand

t 8t Q 8 s (Q0

where is the turbulent kinetic energy (see Eq. 1.10). On the other hand the continuity equation (Eq. 2.5) gives that the right-hand side of Eq. 2.9

')#




9    6 7& 7

is equal to zero. In order to make Eq. 2.9 valid upon contraction we add to the left-hand side of Eq. 2.9 so that

(W13Y y 8 x 0

t 8 tyx

4A R E6 8#' x


Note that contraction of

#"

88 T T  X X  '  '   ' Y P F 48 " H P R


! !

8x

E6 x ' 8CB (Y 8 x 07
gives

4C( ' H B
0

In eddy viscosity models we want an expression for the turbulent viscosity " . The dimension of is  # (same as ). A turbulent velocity scale multiplied with a turbulent length scale gives the correct dimension, i.e.

6 6 ('7&V &" 0 6 y

4C( 'Q'2B
0

Above we have used and which are characteristic for the large turbulent scales. This is reasonable, because it is these scales which are responsible for most of the transport by turbulent diffusion. In an algebraic turbulence model the velocity gradient is used as a velocity scale and some physical length is used as the length scale. In boundary layer-type of ow (see Eq. 2.7) we obtain

is the mixing length, and the model is called the mixing length model. It is an old model F is and is hardly used any more. One problem with the model is that  unknown and must be determined. More modern algebraic models are the Baldwin-Lomax model [2] and the Cebeci-Smith [6] model which are frequently used in aerodynamics when computing the ow around airfoils, aeroplanes, etc. For a presentation and discussion of algebraic turbulence models the interested reader is referred to Wilcox [46].

 X 8 F 6 F where is the coordinate normal to the wall, and where  8


4C( ')(WB
0

 

D 8 @ 698 " @ h QQ 5 h9 0 "


'4768 H@ R H

The equation for turbulent kinetic energy is derived from the Navier-Stokes equation which reads assuming steady, incompressible, constant viscosity (cf. Eq. 2.4)

T 0 X t 8t 8

4 6A876 x B ' x R ' 8  A 6 8#' x x

4C( ')YWB
0

')$

Q  h" & Pi" Q"

The time averaged Navier-Stokes equation reads (cf. Eq. 2.6) Subtract Eq. 2.14 from Eq. 2.13, multiply by obtain

t x B 'x 4 A E6 8IE6 x B ' x R E ' 8  A E6 8#' x x R 4 t 8

 R E  '8 t 8 

 6 8 6 x R E6 8 E6 x  ' x t 8  6A8 R AE6 8  ' x x t 8  4 t 8 tx B ' xVt 8

t8

4C( ' b B
0

and time average and we

4C( ')#WB
0 0

 4 E6 8  t 8 B 4 E6 x  t x B9R E6 8 6 E x  ' x t 8  E6 8 t x  t 8 E6 x  t 8 t x  ' x t 8 4C( Using the continuity equation 4 t x B ' x H , the rst term is rewritten as  E6 8 t x ' x t 8 t 8 t x E6 8(' x 4C( We obtain the second term (using 4 E6 x B ' x H ) from  E6 x 0  ' x# 6 E x  ' t 8 t 8 (  'x 4C( ' 6 E x t 8 t 8#' x  t 8 t 8(' x t 8  6 E xYt 8  x ' (
The left-hand side can be rewritten as
0 0

')$WB '2%QB ')DWB

The third term in Eq. 2.16 can be written as (using the same technique as in Eq. 2.18)

The rst term on the right-hand side of Eq. 2.15 has the form

' 4 t x t 8t 8B x ' (

4C( ')&WB
0

R ' 8 t 8Q R4 t 8CB ' 8

4C( (3H B
0

The second term on the right-hand side of Eq. 2.15 reads

t (8 ' x xYt 8 4 t 8#' b x t 8 B ' x R t #8 ' xbt 8(' x 4 t (8 ' x t 8 B ' x ' 4 t 8 t 8 B ' x x 0 ' x x ( x E6 8(' x R tyx E6 x 0FB ' x R t 8ty G
 

4C( (I'2B
0

For the rst term we use the same trick as in Eq. 2.18 so that

4C( (Q(WB
0

The last term on the right-hand side of Eq. 2.15 is zero. Now we can assemble the transport equation for the turbulent kinetic energy. Equations 2.17, 2.18, 2.20, 2.21,2.22 give

' t xbt t x t ' xY t 8(' x ( 8 8 R 0 '  ' x R (8   

4C( (QYWB
0

The terms in Eq. 2.23 have the following meaning.

'2%

 

Q  h" & Pi" Q"

 9" 1  6
1"

7 ( i"

. The large turbulent scales extract energy from the mean ow. This term (including the minus sign) is almost always positive. The two rst terms represent by pressure-velocity uctuations, and velocity uctuations, respectively. The last term is viscous diffusion. . This term is responsible for transformation of kinetic energy at small scales to internal energy. The term (including the minus sign) is always negative.

Q19  7

6%$ '  &" 7

IH@ # 7& &T7Uh"

In boundary-layer ow the exact


E6 0

E 0

tv

E6 R

read  equation v  (' vt 8 t 8 R

Note that the dissipation includes all derivatives. This is because the dissipation term is at its largest for small, isotropic scales where the usual boundary-layer approximation that is not valid.

0 R t (8 ' xYt 8(4C' x ( (cb B 


0

 

The equation for the instantaneous kinetic energy is derived from the Navier-Stokes equation. We assume steady, incompressible ow with constant viscosity, see Eq. 2.13. Multiply Eq. 2.13 by so that

Q Q Q  h"  '213(5476G8@6A8CB

t 81 u

t 81

T XG 6 876G8
6 8

6G8 4 6A876 x B ' x R6G8 ' 8  A 6 8 6A8(' xAx

4C( (Q#WB
0

The term on the left-hand side can be rewritten as

4 6 8 6 8 6 x B ' x R 6 8 6 x 6 8(' x 6 x 476 8 6 8 B ' x R (' 6 x 476 8 6 8 B ' x (' 6 x 476G876G8 B ' x 4 6 x U B 'x where '213(5476A8@6A8CB . R6 8 ' 8 R 476 8 B ' 8

4C( (Q$WB
0

The rst term on the right-hand side of Eq. 2.25 can be written as

4C( (W%QB
0

The viscous term in Eq. 2.25 is rewritten in the same way as the viscous term in Section 2.4.1, see Eqs. 2.21 and 2.22, i.e. Now we can assemble the transport equation for 2.27 and 2.28 into Eq. 2.25

6G876G8(' x x ' x x R A 6 8(' x 6G8(' x

4C( (QDWB
0

by inserting Eqs. 2.26,

x x R 476 8 B ' 8 R 6 8(' x 6 8#' x 4 6 x S B 'x 'A


')D

4C( (Q&WB
0

Q  h" & Pi" Q"

We recognize the usual transport terms due to convection and viscous diffusion. The second term on the right-hand side is responsible for transport of by pressure-velocity interaction. The last term is the dissipation term which transforms kinetic energy into internal energy. It is interesting to compare this term to the dissipation term in Eq. 2.23. Insert the Reynolds decomposition so that

6G8(' x 6A8#' x G 6 E 8(' x 6A E 8(' x  t 8(' xYt 8(' x 6G8(' x A 6 8#' x D t (8 ' xYt 8(' x
E6

As the scales of

is much larger than those of

t 8 , i.e. t 8#' x 6G8(' x we get


0

4C( Y3H B
0

4C( YI'2B

This shows that the dissipation taking place in the scales larger than the smallest ones is negligible (see further discussion at the end of Sub-section 2.4.3).

Assume steady, incompressible ow with constant viscosity and multiply the time-averaged Navier-Stokes equations (Eq. 2.14) so that

Q Q Q  h"  2 ' 13(54IE6 8 6 E 8 B The equation for '213(54FA E6 85G E6 8 B is derived in the same way as that for '213(5476 8@6A8 B . 
 

G 6 E 8 6A E6 8 4 A E6 8 E6 x B ' x R G E6 8 E ' 8  A E 8(' xAx R A E6 8 4 t 8 tx B ' x

4C( YQ(WB
0

The term on the left-hand side and the two rst terms on the right-hand side are treated in the same way as in Section 2.4.2, and we can write

4 E6 x S E B ' x E ' xAx R 4 G E6 8 E B ' 8 R G E6 8(' x A E6 8#' x R G E6 8 4 t 8 tyx B ' x U where E '213(54 G E6 8 6A E 8CB . The last term is rewritten as R G E6 8 t 8 tyx B ' x  4 t 8 tyx B G E6 8(' x E6 8 4 t 8 tx B ' x R4 G
Inserted in Eq. 2.33 gives

4C( YQYWB
0

4C( Ycb B
0

4 E6 x S E B ' x E ' xAx R 4 G E6 8 E B ' 8 R G E6 8(' x A E6 8#' x R 4IA E6 8 t 8 tx B ' x  t 8 tx A E6 8#' x 0

4C( YQ#WB
0

On the left-hand side we have the usual convective term. On the right-hand side we nd: transport by viscous diffusion, transport by pressure-velocity interaction, viscous dissipation, transport by velocity-stress interaction and loss of energy to the uctuating velocity eld, i.e. to . Note that the last term in Eq. 2.35 is the same as the last term in Eq. 2.23 but with opposite sign: here we clearly can see that the main source term in the equation (the production term) appears as a sink term in the equation. It is interesting to compare the source terms in the equation for (2.23), (Eq. 2.29) and (Eq. 2.35). In the equation, the dissipation term is

R 6 8(' x 6 8(' x R E6 8#' x E6 8(' x R t #8 ' x t 8(' x ')&

4C( YQ$WB
0

Q Q 6 y  6  "7 0  E 0

In the equation the dissipation term and the negative production term (representing loss of kinetic energy to the eld) read

R 6G C4 ( W Y %QB E 8(' x A E6 8(' x  t 8 tyx A E6 8#' x U and in the 0 equation the production and the dissipation terms read R t 8 tyx 6A E #8 ' x R t (8 ' xYt #8 ' x C4 ( Q Y DWB
0 0

The dissipation terms in Eq. 2.36 appear in Eqs. 2.37 and 2.38. The dissi' is distributed into pation of the instantaneous velocity eld the time-averaged eld and the uctuating eld. However, as mentioned above, the dissipation at the uctuating level is much larger than at the time-averaged level (see Eqs. 2.30 and 2.31).

6 8# G E6 8 t 8

H PP

47698HQ@
U

In Eq. 2.23 a number of terms are unknown, namely the production term, the turbulent diffusion term and the dissipation term. In the production term it is the stress tensor which is unknown. Since we have an expression for this which is used in the Navier-Stokes equation we use the same expression in the production term. Equation 2.10 inserted in the production term (term II) in Eq. 2.23 gives

 6 9"   

G R t 8 tx G 6 E 8(' x E6 8(' x
0

E6 x ' 8  A E6 8#' x R (Y 0 A E6 8(' 8

4C( YQ&WB
0

Note that the last term in Eq. 2.39 is zero for incompressible ow due to continuity. The triple correlations in term III in Eq. 2.23 is modeled using a gradient law where we assume that is diffused down the gradient, i.e from region of high to regions of small (cf. Fouriers law for heat ux: heat is diffused from hot to cold regions). We get

"9   y7 68%$  &T  

where 1 is the turbulent Prandtl number for . There is no model for the pressure diffusion term in Eq. 2.23. It is small (see Figs. 4.1 and 4.3) and thus it is simply neglected. The dissipation term in Eq. 2.23 is basically estimated as in Eq. 1.12. The velocity scale is now

x ' t x t 8t 8 R 1 0 ' (

4C( bWH B
0

6%&T&T7U1 7

 

0
5 4

4C( bF'2B
0

so that

x t 8#' x 0 P s t (8 ' Y

4C( b (WB
0

(3H

Q Q  h" 6 7& 
The modelled

equation can now be assembled and we get

E6 x 0FB ' x



 1 

0 'x  'x


1 R 0 54

4C( b YWB
0

We have one constant in the turbulent diffusion term and it will be determined later. The dissipation term contains another unknown quantity, the turbulent length scale. An additional transport will be derived from which we can compute . In the model, where is obtained from its 5 4 own transport, the dissipation term .  in Eq. 2.43 is simply For boundary-layer ow Eq. 2.43 has the form

0XRS P
1 

 E6 0  E 0

 '4768 H@  H P R



0 1




E6

0 54

4C( bQb B
0

In one equation models a transport equation is often solved for the turbulent kinetic energy. The unknown turbulent length scale must be given, and often an algebraic expression is used [4, 48]. This length scale is, for example, taken as proportional to the thickness of the boundary layer, the width of a jet or a wake. The main disadvantage of this type of model is that it is not applicable to general ows since it is not possible to nd a general expression for an algebraic length scale. However, some proposals have been made where the turbulent length scale is computed in a more general way [14, 30]. In [30] a transport equation for turbulent viscosity is used.

(I'



"

 0%(   H PP 

0%2 0 & !%$  '4768 HQ@

  & "

An exact equation for the dissipation can be derived from the Navier-Stokes equation (see, for instance, Wilcox [46]). However, the number of unknown terms is very large and they involve double correlations of uctuating velocities, and gradients of uctuating velocities and pressure. It is better to derive an equation based on physical reasoning. In the exact equation for the production term includes, as in the equation, turbulent quantities and and velocity gradients. If we choose to include and (' in the production term and only turbulent quantities in the dissipation term, we take, glancing at the equation (Eq. 2.43)

t 8 tx

A E6 8 x

(0"" $"

R S e T 0P  G E6 8(' x X P S X e R
0

E6 x ' 8  G E6 8(' x

4CY '2B
0

Note that for the production term we have write the transport equation for the dissipation as

E6 x PcB ' x



S e T 4 P3130FB d
1

. Now we can

 e 

P 'x  'x


P 4 S e T 1 R S e X PcB 0

4CY (WB
0

For boundary-layer ow Eq. 3.2 reads

  E6 P E P

"

 4PP @ "698HQ@ R u




P e 




S e T P
0

E6

X
R

SeX PX

0 4CY YWB
0

The natural way to treat wall boundaries is to make the grid sufciently ne so that the sharp gradients prevailing there are resolved. Often, when computing complex three-dimensional ow, that requires too much computer resources. An alternative is to assume that the ow near the wall behaves like a fully developed turbulent boundary layer and prescribe boundary conditions employing wall functions. The assumption that the ow near the wall has the characteristics of a that in a boundary layer if often not true at all. However, given a maximum number of nodes that we can afford to use in a computation, it is often preferable to use wall functions which allows us to use ne grid in other regions where the gradients of the ow variables are large. In a fully turbulent boundary layer the production term and the dissi pation term in the log-law region ( ) are much larger than the

Y3H ' HQH (Q(



  Q91 7'& 


20 10

10

20 0 200 400

(
A5B D

WV

Y ' P  () bQbWHQH t  1 6


0


$
D

Hy H3b Y

)

0

RQ    TS )$ 



  (546"4 !$()79#$8@%'&& ()()1100 23 23 D 1(BAC0F#$EH4GI(5%'1& 0 ()10


800 1000 1200
0

600

 

)(%! 0)

IU

  )  


bY

# 0

other terms, see Fig. 3.1. The log-law we use can be written as

q & H


'

0)

q t 


4CY b B C4 Y W #B
0 0

Comparing this with the standard form of the log-law

6 t  YX

0)

t 

a`

4CY $WB
0

we see that

'

4  by . In the log-law where we have replaced the dissipation term region the shear stress is equal to the wall shear stress  , see Fig. 2.1. The Boussinesq assumption for the shear stress reads (see Eq. 2.10)

' )Iq In the log-layer we can write the modelled 0 X  E6 H  R P


4CY %QB
0

equation (see Eq. 2.44) as

tv R

4CY DWB
0

t X , and inserting Eqs. 3.9, Using the denition of the wall shear stress V  3.15 into Eq. 3.8 we get  t X X Sdp  4CY ' H B 0 
0

V R

tv

E6

4CY &WB
0

(QY



  Q91 7'&
 Hy Y

From experiments we have that in the log-law region of a boundary layer D  so that . & When we are using wall functions and are not solved at the nodes adjacent to the walls. Instead they are xed according to the theory pre9  sented above. The turbulent kinetic energy is set from Eq. 3.10, i.e.

t X 130

SYp Hy HW&

TX X  0 S p t 


b 7  

S

where the friction velocity  is obtained, iteratively, from the log-law (Eq. 3.4). Index denotes the rst interior node (adjacent to the wall). The dissipation is obtained from observing that production and dissipation are in balance (see Eq. 3.8). The dissipation can thus be written 9 as

4CY 'Q'2B
0

t 
!

 

A

where the velocity gradient in the production term computed from the log-law in Eq. 3.4, i.e.

R tv 6 1

4CY ')(WB
0

has been

6 t

4CY ')YWB
0

For the velocity component parallel to the wall the wall shear stress is used as a ux boundary condition (cf. prescribing heat ux in the temperature equation). When the wall is not parallel to any velocity component, it is more convenient to prescribe the turbulent viscosity. The wall shear stress  is obtained by calculating the viscosity at the node adjacent to the wall from the log-law. The viscosity used in momentum equations is prescribed at the nodes adjacent to the wall (index P) as follows. The shear stress at the wall can be expressed as

9    ( y3 !

V 6E V

' E6
&

E6  ' ' 
&

where  ' denotes the velocity parallel to the wall and & is the normal distance to the wall. Using the denition of the friction velocity 

X t
&

we obtain

' 6 ' t X ' t  t  6 ' Substituting t  1  E6  ' with the log-law (Eq. 3.4) we nally can write ' 0) t q  4 B where t  F1 .
&

&

&

&

&

(cb

QQ X  0 RSP 6 

" #"

 H P

In the model the modelled transport equations for and (Eqs. 2.43, 3.2) are solved. The turbulent length scale is obtained from (see Eq. 1.12,2.42)

0XRSP

0P
!

X

4CY ' b B
0

The turbulent viscosity is computed from (see Eqs. 2.11, 2.41, 1.12)

1 and , which we hope We have ve unknown constants should be universal i.e same for all types of ows. Simple ows are chosen where the equation can be simplied and where experimental data are used to determine the constants. The constant was determined above (Subsection 3.2). The equation in the logarithmic part of a turbulent boundary layer was studied where the convection and the diffusion term could be neglected. constant . We look at the In a similar way we can nd a value for the & equation for the logarithmic part of a turbulent boundary layer, where the convection term is negligible, and utilizing that production and dissipation are in balance 1 , we can write Eq. 3.3 as

X S p 0 T  X 5 S p 0 P
0

SVpUWS e TbUWS e X`U

4CY ')#WB
0

Sbp

SeT

Ch e


0
!

P
P

S b 7

 

4 S e T R S e X B P0

4CY ')$WB
0

The dissipation and production term can be estimated as (see Sub-section 3.2)

X

t  U

!  !

4CY '2%QB
0

which together with

using Eqs. 3.17, 3.18, 3.15 as


!

7 4CY ')DWB In the logarithmic layer we have that 0F1 H , but from Eqs. 3.17, 3.18 we nd that Pc1 H . Instead the diffusion term in Eq. 3.16 can be rewritten
0

hS p



1 P gives

X e 0  0 e hS p    e
 !

XX X S pT X

4CY ')&WB
0

(Q#

QQ X 6 y 0 RUT 

Inserting Eq. 3.19 and Eq. 3.17 into Eq. 3.16 gives

X S e T S e X R T X S p he

4CY (3H B
0

The ow behind a turbulence generating grid is a simple ow which allows us to determine the constant. Far behind the grid the velocity & gradients are very small which means that 1 . Furthermore and the diffusion terms are negligible so that the modelled and equations (Eqs. 2.43, 3.2) read

SeX

H

H

b 7

S

Assuming that the decay of is exponential   . Insert this in Eq. 3.21, derivate to nd   Eq. 3.22 yields

E6 u P R Sde X P 0

E6 u0 R P

X
0 0

4CY (I'2B
0

R u

u


P31 u

4CY Q ( (WB , Eq. 3.21 gives 0


0

and insert it into

SeX





'

Experimental data give  [43], and is chosen. We have found three relations (Eqs. 3.10, 3.20, 3.23) to determine three of the ve unknown constants. The last two constants, 1 and , are optimized by applying the model to various fundamental ows such as ow in channel, pipes, jets, wakes, etc. The ve constants are given the following values: , , , 1 , .

'` (Q# y H HW$

SVe X '` &Q(

4CY (QYWB
0

"

Sdp  Hy HW& S e T ` ' bQb S e X '` &Q( H P

'` H he '` YI'


0

The model is gaining in popularity. The model was proposed by Wilcox [45, 46, 36]. In this model the standard equation is solved, but as a length determining equation is used. This quantity is often called specic dissipation from its denition . The modelled and equation read

0R T

4 E6 x 0FB ' x 4



E6 x T B ' x



 1g  

T T P3130


0 'x  x '

R  ! T 0

4CY (cb B
0

x T g  T '  ' x  0 4 SfghT 1 R SfgiX 0 T B 0 U P  T!07 T


(Q$

4CY (Q#WB
0

QQ X 6 y 0 RSV 

The constants are determined as in Sub-section 3.3: , , , 1 and . When wall functions are used and are prescribed as (cf. Sub-section 3.2):

SfgiX YW1 bWH

g (


Hy HW& S ghT #W13&


0

T X X 0  7 4  B t  U T


 T X t  4 B


4CY (Q$WB P P P 130

In regions of low turbulence when both and go to zero, large numerical problems for the model appear in the equation as becomes zero. The destruction term in the equation includes , and this causes even if also goes to zero; they must both go to zero problems as  at a correct rate to avoid problems, and this is often not the case. On the contrary, no such problems appear in the equation. If  in the equation in Eq. 3.24, the turbulent diffusion term simply goes to zero. Note that the production term in the equation does not include since

0RaP

T f S ghT 1 T SfghTW E6 x 8  E6 x E6 x 8 SfghT  E6 x 8  E6 x E6 x 8 0 0 u u8 u u u8 u In Ref. [35] the 0 R T model was used to predict transitional, recirculating


ow.

"

H P

One of the most recent proposals is the model of Speziale [39] where the transport equation for the turbulent time scale is derived. The exact equation for is derived from the exact and equations. The modelled and equations read

0 R V

E6 x 0FB ' x E6 x V B ' x



V 0F1cP

V @ S e T B 1  4 S e X R '2B 0 4CY Q   X V 'x x  4 ' R   ' 0 V  ( DWB  x x (  x x (  0  T 0' V' R V  X V'V' Sfp 0IV U P 0F1cV  The constants are: SYp , S e T and S e X are taken from the 0XRaP model, and 1  T  X '` YQ$ .



 1 

0 'x  x '


1R 0 
V

4CY (W%QB
0 0

(W%

 ! 0 "

'2

0%2 0 & !%$ 

  &"

In the previous section we discussed wall functions which are used in order to reduce the number of cells. However, we must be aware that this is an approximation which, if the ow near the boundary is important, can be rather crude. In many internal ows where all boundaries are either walls, symmetry planes, inlet or outlets the boundary layer may not be that important, as the ow eld is often pressure-determined. For external ows (for example ow around cars, ships, aeroplanes etc.), however, the ow conditions in the boundaries are almost invariably important. When we are predicting heat transfer it is in general no good idea to use wall functions, because the heat transfer at the walls are very important for the temperature eld in the whole domain. When we chose not to use wall functions we thus insert sufciently many grid lines near solid boundaries so that the boundary layer can be adequately resolved. However, when the wall is approached the viscous the ow is viscous domeffects become more important and for inating, i.e. the viscous diffusion is much larger that the turbulent one (see Fig. 4.1). Thus, the turbulence models presented so far may not be correct since fully turbulent conditions have been assumed; this type of A5B models are often referred to as highA5B number models. In this section we will discuss modications of highnumber models so that they can be used all the way down to the wall. These modied models are termed low Reynolds number models. Please note that high Reynolds number and low Reynolds number do not refer to the global Reynolds number A5B A5B F A5B F , , etc.) but here we are talking about the local (for example A5 B turbulent Reynolds number formed by a turbulent uctuation and turbulent length scale. This Reynolds number varies throughout the computational domain and is proportional to the ratio of the turbulent and A5B physical viscosity , i.e. . This ratio is of the order of 100 or larger in fully turbulent ow and it goes to zero when a wall is approached. We start by studying how various quantities behave close to the wall when  . Taylor expansion of the uctuating velocities (also valid for the mean velocities ) gives

21  1

9 1

H
 

X t 0  T  X X bb v  T 5 X X bb 4 b '2B S S T Y S X w  5 bb where bb SYX are functions of space and time. At the wall we have noslip, i.e. t v w H which gives S . Furthermore, at the wall t 1 u w 1 H , and the continuity equation gives v 1 H so that


G E6 8

t8

(QD

S4

0XRSP 6 7&

0.3 0.2

0.1 0



1


0.1 0.2

P
5 10

( # $ $ )     )(" b 0 ' A5B  %3    (I ' # 0 6 1 DQ" &3 H (%! ) 1  (0" 0'   %  )  0   ) (  " )  ""  $ " &   #  $  ) "%  ! &  B

IU

"!

t 1

'

t  )



15

20

25

30

 %!#) $ 0% "  $ #  )(%! 0 ) 1 6  $# Hy  HW) #3 H 0 ( ()(   " ) #  %$  P   ) 0" %  (" (  (" 4 ) 0 



 "

0

$#



&%



(%

$#

0)

T H . Equation 4.1 can now be written t T  X X X bb v X X bb  S T Y S w  X bb


From Eq. 4.2 we immediately get

4 b (WB
0

X X X T  bb v X XX X bb w S T T X bb tv X b XT b X S T 0 4  BA bb E6 1 T bb t XX
!
214365

     

4  B 4 X B 4 B 4 2 XB 4 B 4 B


X
!

4 b YWB
0

In Fig. 4.2 DNS-data for the fully developed ow in a channel is presented.

35H

 H PR

There exist a number of Low-Re number models [32, 7, 10, 1, 27]. When deriving low-Re models it is common to study the behavior of the terms when  in the exact equations and require that the corresponding terms in the modelled equations behave in the same way. Let us study

0XRSP

(Q&

2
3

S4

0XRSP 6 7&

2.5 2 1.5 1 0.5 0 0

t c1 t


0.14 0.12 0.1

w c1 t  v c1 t 
20 40

0.08 0.06 0.04 0.02 60 80 100

w c 1 E6

t c 1 E6
 

v c 1 E6


(  )(" b 0 (  5 A B    (I' # 0 % )( ) " t c   8

# $


0 0

0.2

    6 X 1 3 % DQ&3H t  1 6 Hy HW#3H t 8
0 0


$ ) 

 "

F1  %!#) $ 0%
0.4
0

0.6

&%

0.8

0 (%!  )


"  %$   



the exact

equation near the wall (see Eq. 2.24).

E6 0


 E 0

R


tv

X 4 0X

The pressure diffusion term is usually neglected, partly because it is not measurable, and partly because close to the wall it is not important, see Fig. 4.3 (see also [28]). The modelled equation reads

v1


E6 R v R  '  vt t  ( 8 8   B 4 B t 8 ' x R t #8 ' xb ( 4 B


!


4b b B
0

E6 0


 E 0

X 4 B 0 X R  P 4 B



E6

  

0 1  4 B




4 b #WB
0

When arriving at that the production term is

 d Sp 0 P

4 B 4 B


4 B

4 B

we have used

4 b $WB
0

Comparing Eqs. 4.4 and 4.5 we nd that the dissipation term in the modelled equation behaves in the same way as in the exact equation when

Y3H

S4

0XRSP 6 7&

0.25 0.2

0.15 0.1 0.05


0 0.05 0.1 0 5 10

(  )(" b 0 Y      (I) ' # 0  "  ) 0 #  " % !  &

IU

$#

    1  t  1 6 6 3 % Q D 3 & H $#  )  ( (" ) #  %$0    Hy % HW#3H   ""    ) 0"% (" ( (" )     t  1


A5B 

$ $ ) 
 $# 

15

20

25

30

 "

0

$#

 %!#) $ 0% " )$ #  ) %!#0)  0   0  )("  $#  )    $" " &  # $ )



 . However, both the modelled production and the diffusion term  are of whereas the exact terms are of ! . This inconsistency of the modelled terms can be removed by replacing the constant by where is a damping function so that when  and   when . Please note that the term damping term in this case is not correct since actually is augmenting when  rather than damping. However, it is common to call all low-Re number functions for damping functions. Instead of introducing a damping function , we can choose to solve for a modied dissipation which is denoted , see Ref. [25] and Section 4.2. It is possible to proceed in the same way when deriving damping functions for the equation [39]. An alternative way is to study the modelled equation near the wall and keep only the terms which do not tend to zero. From Eq. 3.3 we get

rp

'

rp

4 B

#3H r p

rp

rp

4 2B

4 B

ST p

SYp r p

P`

rp

 E6 P E P

4 B

uT

 P P S T e 0 1  he  T T   T 4 B 4 B X X 4 B P P  X R SeX 0   X 4 B 4 B



4 b %QB
0

where it has been assumed that the production term 1 has been suitable modied so that 1 ( ! . We nd that the only term which do not vanish at the wall are the viscous diffusion term and the dissipation term

4 B

YI'



Q Q 2  6 TS  Q h`  2 H PX R S e X P 0

3

S4

0RUP 6 y %&


so that close to the wall the dissipation equation reads

4 b DWB
0

The equation needs to be modied since the diffusion term cannot balance the destruction term when  .

354 @ h B 4E 4 3 H

365

 H PAR

There are at least a dozen different low Re models presented in the literature. Most of them can be cast in the form [32] (in boundary-layer form, for convenience)

0RSP

 E6 0   E 0 u  E6 P ` E P `

r r r Different models use different damping functions ( p , T and X ) and q different extra terms ( and ). Many models solve for P ` rather than for P where is equal to the wall value of P which gives an easy boundary condition P ` H (see Sub-section 4.3). Other models which solve for P use no extra source in the 0 equation, i.e. H .
0

Sprp 0 P` P P` 

P`  r P`  he  S T e T X   0 r R S e X X P `0  q X





 1 

E6

 

R P

E6

4 b &WB
0

4b ' H B
0

4b Q ' '2B 4 b ')(WB


0

Below we give some details for one of the most popular low-Re models, the Launder-Sharma model [25] which is based on the model of Jones & Launder [20]. The model is given by Eqs. 4.9, 4.10, 4.11 and 4.12 where

0XRSP

 Q h` 

 6  S

rp rT rX

!

'S R Hy Y ( q (  X 0 P`



'

4@'


R A Y b


!

 X X E6 X

X R A X 


13#3H B

X


4 b ')YWB
0

YQ(

% '6 h"
q rX

7'6 "  P '6 P ` 0

The term was added to match the experimental peak in around  [20]. The term is introduced to mimic the nal stage of decay of turbu lence behind a turbulence generating grid when the exponent in  changes from  to  .

(3H

#"

` ' (Q# ( # H @ 4 D H@ 8 68 HQ@ H  4@ 


0 RUP P` 0

In many low-Re models is the dependent variable rather than . The main reason is that the boundary condition for is rather complicated. The largest term in the equation (see Eq. 4.4) close to the wall, are the  dissipation term and the viscous diffusion term which both are of so that

X H 0X R  P

4 B
0

4b ' b B P 4 b ')#WB
0

From this equation we get immediately a boundary condition for as

P  7 0 X

From Eq. 4.14 we can derive alternative boundary conditions. The exact form of the dissipation term close to the wall reads (see Eq. 2.24)

 t X 

 D

where Taylor expansion in Eq. 4.1 gives

1 u

X w  1 and t

4 b ')$WB
0

have been assumed. Using

P 0 ('
so that

 T 

X S XT

bb

4 b '2%QB
0

In the same way we get an expression for the turbulent kinetic energy

 T 

X S XT X bb ('

4 b ')DWB
0

 T 

X S XT b b X

4 b ')&WB
0

Comparing Eqs. 4.17 and 4.19 we nd

P  7 (

0  YQY

4 b (3H B
0

Q Q % S 2    0XRSP 6 y

In the Sharma-Launder model this is exactly the expression for in Eqs. 4.12 . and 4.13, which means that the boundary condition for is zero, i.e. In the model of Chien [8], the following boundary condition is used

P`

P ` H
0

P  7 ( 0 X
This is obtained by assuming

P ( X X T 0 T

T  ST

4 b (I'2B
in Eqs. 4.17 and 4.18 so that

4 b (Q(WB
0
1

which gives Eq. 4.21.

 H P H Y3 4hD h
3

Near the walls the one-equation model by Wolfshtein [48], modied by Chen and Patel [7], is used. In this model the standard equation is solved; the diffusion term in the -equation is modelled using the eddy viscosity assumption. The turbulent length scales are prescribed as [15, 11]

p S0 ' R
 !

!

(0 is the normal distance from the wall) so that the dissipation term in the -equation and the turbulent viscosity are obtained as:

4@R A 1 X p B U e S 0 ' R
# 

!

4@R A 1 X e B

Sdp P 0 e U

X

The Reynolds number

0 p

4 b (QYWB
0  !

and the constants are dened as

0 0 U Sdp Hy HW& U S h Sp h 0XRSP



U X p %cH U X e ( S
A

4@R A 1 X p B takes, e.g., the value y H &Q# . The matching of the one-equation model and the 0X RaP model does not pose any problems but gives a smooth distribution of and P across the matching line. ' R
!

The one-equation model is used near the walls (for ), and the A5B standard highin the remaining part of the ow. The matching line could either be chosen along a pre-selected grid line, or it could be dened as the cell where the damping function

(Q#3H

Ycb

Q Q 

S4
21

0XRUT 6 y

3 5

PH

HP
0R T

A model which is being used more and more is the model of Wilcox [45]. The standard model can actually be used all the way to the wall without any modications [45, 29, 34]. One problem is the boundary condition for at walls since (see Eq. 3.25)

0 RT

P 0 T T

4 B


4 b (cb B
0

tends to innity. In Sub-section 4.3 we derived boundary conditions for by studying the equation close to the wall. In the same way we can here use the equation (Eq. 3.25) close to the wall to derive a boundary condition for . The largest terms in Eq. 3.25 are the viscous diffusion term and the destruction term, i.e.

X X H T X R S giX T

4 b (Q#WB
0

The solution to this equation is

The equation is normally not solved close to the wall but for is computed from Eq. 4.26, and thus no boundary condition actually needed. This works well in nite volume methods but when nite element methods are used is needed at the wall. A slightly different approach must then be used [16]. Wilcox has also proposed a model [47] which is modied for viscous effects, i.e. a true low-Re model with damping function. He demonstrates that this model can predict transition and claims that it can be used for taking the effect of surface roughness into account which later has been conrmed [33]. A modication of this model has been proposed in [36].

$ X T f S giX T

4 b (Q$WB
0

( # T

0XRaT

YQ#

  
The

Q Q  0XRUT

S4

Q Q 


0XRUT 6 y

S4

0XRUT 6  y 

0 T
2 2

 x 4 F 0 B  E6 1  ux u x   x 4 T B  E6 g  ux ux  r p 0 T r 1 ' RSHy %3(Q( ! R


model of Peng et al. reads [36]



ux ux
 A

1


T 4f S ghT r g 0


R S 1 r 1 T!0

S g  R SfgiX 0 T B f 0


ux ux

r p Hy HW(Q#

rg S1 Sfg
 
A new

 X ! Hy &W%3#  Hy A HQH5 ' R (3HQH    A T X ! U rg '   '  b Y R '` #  b Y  Hy HW& U SfghT Hy b ( U SfgiX Hy H %3# Hy %3# U 1 Hy D U g '` YQ#
 A 

' R

!

'H R

 A


'H



!

 A

'` #

T X


Q Q  09R T 0 

S4

0XRUT 6  %  6 9 i 1 ghT T 1 R & giX T X  0 0 T g    0 0 ux ux


&

4 b (W%QB
0

model was recently proposed by Bredberg et al. [5] which reads

T
2

x u x 4 E6 0FB' x u x 4 E6 T B'
&

& 1 0 TP

ux




 1 

ux  g


ux



ux

4 b (QDWB
0


4 b (Q&WB
0

The turbulent viscosity is given by

 & p r p 0 T  r p Hy HW&  y ' H &I' A  ' R


!

!

 A

(Q#

X 


YQ$

Q Q 

S4

0XRUT 6 y

with the turbulent Reynolds number dened as in the model are given as

0F1I4 T B . The constants


4 b Y3H B
0

& 1 &

y H HW& U & p ' U & g '` ' U giX  Hy H %3( U 1 ' U g '` D

&

ghT Hy b & U

YW%

 X#& " 72  " "

 

&"

In Reynolds Stress Models the Boussinesq assumption (Eq. 2.10) is not used but a partial differential equation (transport equation) for the stress tensor is derived from the Navier-Stokes equation. This is done in the same way as for the equation (Eq. 2.23). Take the Navier-Stokes equation for the instantaneous velocity (Eq. 2.4). Subtract the momentum equation for the time averaged velocity ((Eq. 2.6) and multiply by . Derive the same equation with indices 3 and 4 interchanged. Add the two equations and time average. The resulting equation reads

tx

E6 8

6 8

E6 1 t 8 t x B ' 1 R t 8 t 1 6 E x ' 1  R t x t 1 E6 (8 ' 1 & x 8x 8

R t 8 txVt

1 

tx 8 1

Ra ( t 8 ' 1 t x ' 1 ( P)8 x


where

8x

4 t 8(' x  t x ' 8 B  x 8 t 8 x 1 R 4 t 8 tyx B ' 1 




' 1

4C# '2B
0

8x
! !

is the pressure-strain term, which promotes isotropy of the turbulence; is the dissipation (i.e. transformation of mechanical energy into heat in the small-scale turbulence) of ;

8x

is the production of )];

t 8 tx

[note that

88

T X
1

88 T T  X X


P 8x
&

8x

and

8x

t 8 tyx

are the convection and diffusion, respectively, of

t 8 tx .

Note that if we take the trace of Eq. 5.1 and divide by two we get the equation for the turbulent kinetic energy (Eq. 2.23). When taking the trace the pressure-strain term vanishes since

8 8 ( t 8#' 8 H

4C# (WB
0

due to continuity. Thus the pressure-strain term in the Reynolds stress equation does not add or destruct any turbulent kinetic energy it merely redistributes the energy between the normal components ( , and ). Furthermore, it can be shown using physical reasoning [19] that acts to reduce the large normal stress component(s) and distributes this energy to the other normal component(s).

tX vX

8 x

w X

YQD

4 3 Q768&
5

1 

&&

6 7&

D'@ HQP R B6 R`R H PAR t 8t x t 8x


4 t x 8 Pc8 x

We nd that there are terms which are unknown in Eq. 5.1, such as the 1 , the pressure diffusion 1   21 triple correlations and the pressure strain , and the dissipation tensor . From Navier-Stokes equation we could derive transport equations for this unknown quantities but this would add further unknowns to the equation system (the closure problem, see p. 12). Instead we supply models for the unknown terms. The pressure strain term, which is an important term since its contribution is signicant, is modelled as [24, 17]

t 8x B1

8x 'T 8x 'X

8x

The object of the wall correction terms ' and ' is to take the effect of 0 coordinate system, the wall into account. Here we have introduced a  0 with  along the wall and normal to the wall. Near a wall (the term near may well extend to ) the normal stress normal to the wall is damped (for a wall located at, for example, this mean that the normal stress is damped), and the other two are augmented (see Fig. 4.2). In the literature there are many proposals for better (and more complicated) pressure strain models [40, 23]. The triple correlation in the diffusion term is often modelled as [9]

c ' T ca a ' T 0 5 r 04 ( #Q#u P

c8 x ' T  c8 x ' X R ST 0P t 8 tx R (Y 8 x 0   Y S X R 8 x R (Y 8 x 1  X R ( S cT 0P t r S c T P t X r

8x 'T

8x 'X

4C# YWB
0

c T

ca a T
R

vX

(3HQH

u H

8x

SYa t 1 t  0 4 t 8 tx B ' 1  P
Pc8 x

'

17U   S   4C# b B "


0

The pressure diffusion term is for two reasons commonly neglected. First, it is not possible to measure this term and before DNS-data (Direct Numerical Simulations) were available it was thus not possible to model this term. Second, from DNS-data is has indeed been found to be small (see Fig. 4.3). The dissipation tensor is assumed to be isotropic, i.e.

P)8 x Y( 8 x P

From the denition of (see 5.1) we nd that the assumption in Eq. 5.5 is equivalent to assuming that for small scales (where dissipation occurs) the

P 8x

4C# #WB
0

YQ&



4 3 Q768&

1 

&&

6 7& tx t8 tyx

6 6

H %&

&T 0  6 y7&

two derivative (' 1 and ' 1 are not correlated for 3 4 . This is the same as 4 which is assuming that for small scales and are not correlated for 3 a good approximation since the turbulence at these small scale is isotropic, see Section 1.4. We have given models for all unknown term in Eq. 5.1 and the modelled Reynolds equation reads [24, 17]

t8

Stress Models), see [18, 22, 26, 38].

t xYt 1 E6 8#' 1 4 E6 1 t 8 tyx B ' 1 R t 8 t 1 E6 x ' 1 R S a t 1 t  0 4 t 8 tx B '   8 x R ( 8 x P 4C# $WB  4 t 8 tyx B ' 1  Y P Y  '1 where 8 x should be taken from Eq. 5.3. For a review on RSMs (Reynolds
0

D'@ HQP R B6 R`R H PAR

'D $

8AR#"yHR8 6VD H PAR


0XRSP

Whenever non-isotropic effects are important we should consider using RSMs. Note that in a turbulent boundary layer the turbulence is always non-isotropic, but isotropic eddy viscosity models handle this type of ow excellent as far as mean ow quantities are concerned. Of course a model give very poor representation of the normal stresses. Examples where non-isotropic effects often are important are ows with strong curvature, swirling ows, ows with strong acceleration/retardation. Below we present list some advantages and disadvantages with RSMs and eddy viscosity models. Advantages with eddy viscosity models: i) simple due to the use of an isotropic eddy (turbulent) viscosity; ii) stable via stability-promoting second-order gradients in the meanow equations; iii) work reasonably well for a large number of engineering ows. Disadvantages with eddy viscosity models:

i) isotropic, and thus not good in predicting normal stresses ( ); ii) as a consequence of i) it is unable to account for curvature effects; iii) as a consequence of i) it is unable to account for irrotational strains. Advantages with RSMs: i) the production terms need not to be modelled; ii) thanks to i) it can selectively augment or damp the stresses due to curvature effects, acceleration/retardation, swirling ow, buoyancy etc. Disadvantages with RSMs:

t XUv XUw X

bWH

T(Q $ 9T&

A 0 B r B 0

U (r)

! # () " ( # 0 '   6Q 4 1 B U 6  H 0


%

$

Q(


) 1

S)(10


32 0 2)0

i) complex and difcult to implement; ii) numerically unstable because small stabilizing second-order derivatives in the momentum equations (only laminar diffusion); iii) CPU consuming.

#"

476

" 6R

Curvature effects, related either to curvature of the wall or streamline curvature, are known to have signicant effects on the turbulence [3]. Both types of curvature are present in attached ows on curved surfaces, and in separation regions. The entire Reynolds stress tensor is active in the interaction process between shear stresses, normal stresses and mean velocity strains. When predicting ows where curvature effects are important, it is thus necessary to use turbulence models that accurately predict all Reynolds stresses, not only the shear stresses. For a discussion of curvature effects, see Refs. [12, 13]. When the streamlines in boundary layer type of ow have a convex (concave) curvature, the turbulence is stabilized (destabilized), which dampens (augments) the turbulence [3, 37], especially the shear stress and the Reynolds stress normal to the wall. Thus convex streamline curvature decreases the stress levels. It can be shown that it is the exact modelling of the production terms in the RSM which allows the RSM to respond correctly to this effect. The model, in contrast, is not able to respond to streamline curvature. A The ratio of boundary layer thickness to curvature radius is a common parameter for quantifying the curvature effects on the turbulence. The work reviewed by Bradshaw demonstrates that even such small amounts A of convex curvature as can have a signicant effect on the turbulence. Thompson and Whitelaw [42] carried out an experimental inves-

0RSP

1 Hy H5'

bF'

T(Q $ 9T&

y x

streamline

( &( " 0) " &(0% & 0 )   # () $   $ "   ) # 0(           &( "  "( )    %!! 0 " % )(%  $ "   4   %   B  )    0 0 ) u  %' ( & 0) "  ) $ )           0
2

tigation on a conguration simulating the ow near a trailing edge of an A D . They reported a 50 percent deairfoil, where they measured crease of (Reynolds stress in the normal direction to the wall) owing to and was also substantial. In addition curvature. The reduction of they reported signicant damping of the turbulence in the shear layer in the outer part of the separation region. An illustrative model case is curved boundary layer ow. A polar coor(see Fig. 5.1)) with locally aligned with the streamline dinate system 1 1 (with 1 and  ), the radial is introduced. As inviscid momentum equation degenerates to

vX

1 tX

tv R

Hy HWY

6 X
1

6 6 34 B

6 21

6 H

H

4C# %QB
0

Here the variables are instantaneous or laminar. The centrifugal force exerts a force in the normal direction (outward) on a uid following the streamline, which is balanced by the pressure gradient. If the uid is displaced by some disturbance (e.g. turbulent uctuation) outwards to level A, it encounters a pressure gradient larger than that to which it was accustomed 1 1 1  , as  .  , which from Eq. 5.7 gives at 1 1 1  Hence the uid is forced back to . Similarly, if the uid is displaced 1  and inwards to level B, the pressure gradient is smaller here than at 1 cannot keep the uid at level B. Instead the centrifugal force drives it back to its original level. It is clear from the model problem above that convex curvature, when , has a stabilizing effect on (turbulent) uctuations, at least in 1 the radial direction. It is discussed below how the Reynolds stress model responds to streamline curvature. Assume that there is a at-plate boundary layer ow. The ratio of the normal stresses and is typically 5. At one station, the ow is deected upwards, see Fig. 5.2. How will this affect the turbulence? Let us study the effect of concave streamline curvature. The production terms owing to rotational strains can be written as

4@6 )B 4 6 B

4 1 B 4 1 B

6 )1

tX

vX

8x

b(

T(Q $ 9T&

7 (6 5 T  (Q ("(yhQ

 #

21  21  " # 0 0)  " # 0)  "  # 0) " # 0 0)  



 2 ' $  # '   ! # "%$ & (' 0) '  1$ ' (
0

  

X 43 U t R
43 U tv R X 43 U v R

IU IU IU
    

TT TX XX 1
E6 1

R R R

t X E R v X u E ( tv

( tv

E6

4C# DWB
0

E6

4C# &WB
0

0XRSP E6 1 X E 1 u H v

E6

E X u 

4C# ' H B
0

4C# 'Q'2B
0

As long as the streamlines in Fig. 5.2 are parallel to the wall, all pro duction is a result of . However as soon as the streamlines are de ected, there are more terms resulting from . Even if is much smaller that it will still contribute non-negligibly to as is much larger than . Thus the magnitude of will increase ( is nega tive) as . An increase in the magnitude of will increase , which in turn will increase and . This means that and will be larger and the magnitude of will be further increased, and so on. It is seen that there is a positive feedback, which continuously increases the Reynolds stresses. It can be said that the turbulence is 57608@9BA4C!DEDFG6@5 owing to concave curvature of the streamlines. model is not very sensitive to streamline curvature However, the (neither convex nor concave), as the two rotational strains are multiplied by the same coefcient (the turbulent viscosity). If the ow (concave curvature) in Fig. 5.2 is a wall jet ow where , the situation will be reversed: the turbulence will be 8H9BA4CIDEDFG6@5 . If the streamline (and the wall) in Fig. 5.2 is deected downwards, the situation will be as follows: the turbulence is stabilizing when , and desta bilizing for . The stabilizing or destabilizing effect of streamline curvature is thus dependent on the type of curvature (convex or concave), and whether there is an increase or decrease in momentum in the tangential direction with radial distance from its origin (i.e. the sign of 1 ). For convenience, these cases are summarized in Table 5.1.

TT

TX

XX

E 1 u TX

TX

tX

E 1 u T X TX

tX

v XR

tv

0RSP

E6 1

E6 1 H

E6 1 H

E6 21

bY

yi" 

6 4

i h16h"
x2 y x1 x x2 x1


" "

P h 4768 H@ 4@

694 468HQ@ tv 61 u R

# Y
0

 )   )     (
0

When the ow accelerates and/or decelerate the irrotational strains ( , and ) become important. In boundary layer ow, the only term which contributes to the pro duction term in the equation is ( denotes streamwise direction). Thompson and Whitelaw [42] found that, near the separation point as well as in the separation zone, the production term is of equal importance. This was conrmed in prediction of separated ow using RSM [12, 13]. In pure boundary layer ow the only term which contributes to the production term in the and -equations is . Thompson and Whitelaw [42] found that near the separation point, as well as in the sepa is of equal importance. ration zone, the production term As the exact form of the production terms are used in second-moment closures, the production due to irrotational strains is correctly accounted for. D In the case of stagnation-like ow (see Fig. 5.3), where the production due to normal stresses is zero, which is also the results given by second-moment closure, whereas models give a large production. In order to illustrate this, let us write the production due to the irrotational strains and for RSM and :

E 1

E 1

E6 1 u

X X R 4t R v B 6 1 u

t v E6 1 R X X E6 1 u R 4t R v B
0XRaP

tX

vX

The production term 1 in be 8 of the two strains.

E 1 0XRSP X X A  y H # 4 T T  X X ' B R t u E6 R v  E6 X  E X 0RSP  1 ( u    X D v X we get T T  X X D H since E6 1 u R If t


E6 1 u

0RSP

due to continuity. model, however, will be large, since it will

E 1

bQb

3 4 S  )'2 !%$  "


[1]

A new turbulence model for predicting uid ow and heat transfer in separating and reattaching ows - 1. Flow eld calculations. Int. J. Heat Mass Transfer 37 (1994), 139151.

 !"#%$&#('#0)1

[2]

23#456879 @2@#BACD#EGFH Thin-layer approximation and algebraic model for separated turbulent ows. AIAA 78-257, Huntsville, AL, 1978.
Effects of streamline curvature on turbulent ow. Agardograph progress no. 169, AGARD, 1973. Calculation of boundary-layer development using the turbulent energy equation. Journal of Fluid Mechanics 28 (1967), 593616.

[3] [4]

2PIQ#RS#56TVU&

2PIQ#RS#56TWU&9YXG`I(IQ79RaRbdcef#gihp6q`4r4sf$H

[5]

2PI(s`I'(tUD`'vuvxwyFq(#Yc#579Ra Ai An improved turbulence model applied to recirculating ows. International Journal of Heat and Fluid Flow 23, 6 (2002), 731743.
Academic Press, 1974.

0R T

[6] [7] [8]

P`pps7!"&#uQC79he P` FH#U#`hQ`4pV P(7`1

Analysis of Turbulent Boundary Layers.

Near-wall turbulence models for complex ows including separation. AIAA Journal 26 (1988), 641648. Predictions of channel and boundary layer ows with a low-reynolds-number turbulence model. AIAA Journal 20 (1982), 33 38.

[9] [10]

c#4`2#F&#I(4r6XP Transport equation in turbulence. Physics of Fluids 13 (1970), 26342649. c#57RaiAi Calculation of the turbulent buoyancy-driven ow in a rectangular cavity using an efcient solver and two different low turbulence models. Numer. Heat Transfer 18 reynolds number (1990), 129147.

0XRSP

[11]

Reynolds stress transport modelling of c#57Ra A shock/boundary-layer interaction. 24th AIAA Fluid Dynamics Conference, AIAA 93-3099, Orlando, 1993.

[12]

c#57RaA Prediction of the ow around an airfoil using a Reynolds stress transport model. ASME: Journal of Fluids Engineering 117 (1995), 5057.

b#

3 4 S 

[13] [14]

c#57RaAi Reynolds stress transport modelling of shock-induced separated ow. Computers & Fluids 24 (1995), 253268.
Calculation of some parabolic and elliptic ows using a new one-equation turbulence model. In 5th Int. Conf. on Numerical Methods in Laminar and Turbulent Flow (1987), C. Taylor, W. Habashi, and M. Hafez, Eds., Pineridge Press, pp. 411422. Navier-Stokes stall predictions using an algebraic stress model. J. Spacecraft and Rockets 29 (1992), 794800.

c#57RaAi #

4 R Ra &

[15] [16]

c#57RaiAi #7 7ie

DCi7 U& The Full Multigrid Method Applied to Turbulent Flow in Ventilated Enclosures Using Structured and Unstructured Grids. PhD thesis, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Goteborg, 1997. 7QR    #A #sI0 2 Ground effects on pressure uctuations in the atmospheric boundary layer. Journal of Fluid Mechanics 86 (1978), 491511.
F&# #4b7v  G Advanced turbulence closure models: A view of current status and future prospects. Int. J. Heat and Fluid Flow 15 (1994), 178203. F&79 t Turbulence, second ed. McGraw-Hill, New York, 1975. tpsRb   (#YA0#`I02 The prediction of laminarization with "7C vt

[17]

[18]

[19] [20]

a two-equation model of turbulence. Int. J. Heat Mass Transfer 15 (1972), 301314. [21]

The collaborative testing of turbulence models (organized by ). Data Disk No. 4 (also available at Ercoftacs wwwP. Bradshaw server http://fluindigo.mech.surrey.ac.uk/database), 1990.

[22] [23]

A #sI0@2 Second-moment closure: Present ... and future? Int. J. Heat and Fluid Flow 10 (1989), 282300. A #sI0v2 #A07@uvxw U On the elimination of wall-topography parameters from second-moment closures. Physics of Fluids A 6 (1994), 9991006. A #sI02`ppe@# 7!  Progress in the development of a Reynolds-stress turbulence closure. Journal of Fluid Mechanics 68, 3 (1975), 537566. A #sI02 #Wu#I(C#2 Application of the energy dissipation model of turbulence to the calculation of ow near a spinning disc. Lett. Heat and Mass Transfer 1 (1974), 131138.

[24]

[25]

b$

3 4 S 

[26] [27]

AsR  7`I0   Modelling engineering ows with Reynolds stress turbulence closure. J. Wind Engng. and Ind. Aerodyn. 35 (1991), 2147. A 7` XPB# AsR  7`I0   Low-Reynolds-number eddyviscosity modelling based on non-linear stress-strain/vorticity relations. In Engineering Turbulence Modelling and Experiments 3 (1996), W. Rodi and G. Bergeles, Eds., Elsevier, pp. 91100.

[28]

 #R  I0B$H "7C t #  7 U Reynolds-stress and dissipation-rate budgets in a turbulent channel ow. Journal of Fluid Mechanics 194 (1988), 1544.
Two-equation eddy-viscosity turbulence models for engineering applications. AIAA Journal 32 (1994), 15981605. On the connection between one- and two-equation models of turbulence. In Engineering Turbulence Modelling and Experiments 3 (1996), W. Rodi and G. Bergeles, Eds., Elsevier, pp. 131140.

[29] [30]

shQ`I0 XP shQ`I0XP

[31] [32]

Uv#hQ 8

Incompressible Flow. John Wiley & Sons, New York, 1984.

Turbulence models for near-wall and low Reynolds number ows: A review. AIAA Journal 23 (1985), 13081319.

Uv#`hQ`4p V 7  # uQ sI(`I0!e

[33]

Uv#`hQ`4p0V9# )8 t Application of turbulence models to separated ows over rough surfaces. ASME: Journal of Fluids Engineering 117 (1995), 234241. UD`' uvxwyFH9Yc #r7vR  fAif#gF45C3p`IQ'Yuv Performance of two-equation turbulence models for numerical simulation of ventilation ows. In 5th Int. Conf. on Air Distributions in Rooms, ROOMVENT96 (Yokohama, Japan, 1996), S. Murakami, Ed., vol. 2, pp. 153160.
The two-equations turbulence model applied to recirculating ventilation ows. Tech. Rep. 96/13, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1996.

[34]

[35]

UD`'vuvxwyFH9c #r7vR  Ai(# F45C3psI'uv

0 RT

[36]

UD`'3uvxwyFH9ic#57RaAi # F4rCD`IQ'Duv A modied lowReynolds-number model for recirculating ows. ASME: Journal of Fluids Engineering 119 (1997), 867875.

0XRST

[37]

Calculation of curved shear layers with two-equation turbulence models. Physics of Fluids 26 (1983), 14221435.

7  # u(p `I(`I0 e

b%

3 4 S 
u(qA #7 )19#

[38]

(#'vFH Second-order near-wall turbulence closures: a review. AIAA Journal 29 (1991), 18191835.

[39] u` 7#45  s798@#%`IR  ! Critical evaluation of two-equation models for near-wall turbulence. AIAA Journal 30 (1992), 324331. [40] u` 7#45 up#I#I0uv# #`hQR7(!" Modelling the pressurestrain correlation of turbulence: an invariant dynamical system approach. Journal of Fluid Mechanics 227 (1991), 245272. [41] [42]

!`vsR53FHP# ACD4r t A First Course in Turbulence. The MIT Press, Cambridge, Massachusetts, 1972. !CsRa H28#  (79hs4b#r6Tqt Characteristics of a trailingedge ow with turbulent boundary-layer separation. Journal of Fluid Mechanics 157 (1985), 305326.
The Structure of Turbulent Shear Flow, second ed. Cambridge University Press, New York, 1976.

[43] [44] [45] [46] [47] [48]

!&6qR `(e

 (7hQvXP

Fluid Mechanics. McGraw-Hill, Inc., New York, 1994.

 74rQQEce Reassessment of the scale-determining equation. AIAA Journal 26, 11 (1988), 12991310.  74rQQEvce Turbulence Modeling for CFD. DCW Industries, Inc., 5354 Palm Drive, La Canada, California 91011, 1993.  74rQQEvce Simulation of transition with a two-equation turbulence model. AIAA Journal 32 (1994), 247255.  4 RShQ 79   The velocity and temperature distribution in onedimensional ow with turbulence augmentation and pressure gradient. Int. J. Mass Heat Transfer 12 (1969), 301318.

bD

You might also like