You are on page 1of 25

Internal Report 00/4

On the Wall Boundary Condition for Turbulence Models


JONAS BREDBERG
Department of Thermo and Fluid Dynamics C HALMERS U NIVERSITY OF T ECHNOLOGY

( ( ( &  # "     '%$! 

On the Wall Boundary Condition for Turbulence Models


JONAS BREDBERG Department of Thermo and Fluid Dynamics Chalmers University of Technology

This report explains and discuss two main boundary conditions for turbulence models: the integrating and the wall function approach. The report thoroughly derives the law-of-thewall for both momentum and thermal eld. The deviation of these laws from DNS-data is discussed. The Wilcox turbulence model constants are derived on the basis of the lawof-the-wall and experiments. Using a simplistic treatment, the low-Reynolds number modications (damping functions) to the model are explained. Included in the report is also a guideline on the implementation of wall functions, both the standard and the ChiengLaunder two-layer variant. A new boundary condition for turbulence models combining wall function and integration approach is presented.

ABSTRACT

Keywords: boundary condition, wall function, turbulence model

ii

Contents
Nomenclature 1 Introduction 2 Near-wall Physics 2.1 Buffer Layer . . . . . . . . . . . . . . . . . . . . . 2.2 Viscous sub-layer . . . . . . . . . . . . . . . . . . 2.3 Inertial Sub-layer . . . . . . . . . . . . . . . . . . 3 Law-of-the-Wall 3.1 Momentum . . . . . . . . . . . 3.2 Energy/Temperature . . . . . 3.3 Turbulence . . . . . . . . . . . 3.4 Turbulence Model Constants 3.4.1 Coefcient, . . . . . 3.4.2 Coefcient, . . . . . . 3.4.3 Coefcients, . . . 3.4.4 Coefcient, . . . . . . 3.5 HRN Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv 1 1 1 1 2 3 3 4 6 7 7 7 7 7 8 8 8 9 9 10 10 11 11 11 12 13 14 14 15 15 16 16 18 18 19 19

4 Modelling Near-wall Turbulence 4.1 LRN-models: Integration Method . . . . . . . . . 4.1.1 The damping functions . . . . . . . . . . . modication, . . . . . . . . . . . . . 4.1.2 4.1.3 Production-to-dissipation rate . . . . . . . 4.1.4 Rationale behind the and coefcients 4.2 HRN-models: Wall Functions . . . . . . . . . . . 4.2.1 Standard wall functions . . . . . . . . . . 4.2.2 Launder-Spalding methodology . . . . . . 4.2.3 Improvements of the near-wall representation, Chieng-Launder model . . . . . . . 4.2.4 Modication to the Chieng-Launder model 4.2.5 Extensions to the Chieng-Launder model 4.3 New Model: Blending Integrating and Wall Function Boundary Conditions . . . . . . . . . . 4.3.1 Blending function . . . . . . . . . . . . . . 4.3.2 HRN part: Simplied Chieng-Launder model applied to a -type turbulence model . . . . . . . . . . . . . . . . . . . . . 4.3.3 LRN part: Standard . . . . . . . . . 4.3.4 The blending turbulence model . . . 5 Modelling Near-wall Heat Transfer 5.1 Integration Method . . . . . . . . . . . . . . . . . 5.2 Wall Function Method . . . . . . . . . . . . . . . 5.3 New Blending Model . . . . . . . . . . . . . . . .

iii

 

Nomenclature
4 1 GFED C ' 1 4 1 4 1 2 1 
iv Turbulence model constant Specic heat Various constants Length-scale constant Turbulence model coefcient Skin friction coefcient, Hydrualic diameter Rib-size Turbulence model constant Damping function Functions Channel half-height Heat transfer coefcient Channel height Tensor indices, streamwise: 1 (U) wall normal: 2 (V) spanwise: 3 (W) Turbulent kinetic energy Length scale Nusselt number Static pressure Rib-pitch Turbulent production,

Superscripts

Normalized value using

Subscripts

4 1 4 1 R F 1 ' S'"R $q 1 FR G 1 FR G 1

% ) V jhg i $F e # f f !f

Velocity Fluctuating velocity Reynolds stresses Friction velocity, Normalized value: Streamwise coordinate Wall normal coordinate

U r( (GH |

U (r |

G     

Temperature Fluctuating temperature Friction temperature,

Non-modied value First interior node Bulk value Centerline value HRN (logarithmic) part Laminar quantity LRN (viscous) part North face value Node value Turbulent quantity Thermal quantity Viscous sub-layer value Wall value Quantity based on the friction velocity

S'q1ED C

9 R

Prandtl number Functional Prandtl number Heat ux Turbulence model constants Reynolds number, Turbulent Reynolds number, , , Viscous sub-layer Reynolds number, based Reynolds number, Strain-rate tensor,

Effective value Normalized value using

' 1 Xq$ ' 1 XqE UU m R X R a E& 1 ' 1 X$ER U ' FR " 1  1 q$D C a ' m"R " 1 4 1 4 1 ' 1 GFED C

i (W q

 yw & AX x z yw y v A  vx ' V b bw V '  V w ' V y w V y v ' y w V V V u%

' V #

   

Latin Symbols

Greek Symbols

Thermal diffusivity Turbulence model coefcients Turbulence model coefcients Turbulence model coefcient Characteristic length scale Cell size Dissipation rate Normalized wall distance Van Karman constant Thermal conductivity, Kinematic viscosity Dynamic viscosity Density Specic dissipation rate Turbulence model coefcients Time Shear stress Wall shear,

w s cV s s s s s % s s s cr

!

l o p w 3

k n q #  

S"R $ 1 a ' 4 1 ' 1 X$ U 2 1 ' ' S"R $ 1

') % # ! (&$" 

w H %

cg f % g f cg % g y h d d h

w V V w ' w yw $ & &y w x w

! p V ! #

h g bf V bd f cg d % FcecV

x"3 r 3 r uvr Y(r tr s dp  (W i i (W

B @ 8  A 9 7 B

h V V f bdV bddV bd %V

$3 r

fd $3 r

 

  d 6 6

V U

P Q I

 0

Y `W

3 5 H W W

1 Introduction
One of the most common engineering problems is computing turbulent ows that are inuenced by an adjacent wall. Examples of this are ows in turbomachinery, around vehicles, and in pipes. The main two effects of a wall are:

2 Near-wall Physics
The near-wall region may be sub-divided into three different areas, Tennekes and Lumley [31]: viscous sub-layer buffer layer

Increasing the production of turbulence through the shearing mechanism in the ow.

The wall gives rise to a boundary layer, where the velocity changes from the no-slip condition at the wall to its free stream value. The variation is usually largest in the near-wall region, and hence the strongest gradients are found here. Similarly, for heat transfer applications, 1.5 there exists a thermal boundary layer with equally large Total shear Turbulent shear gradients. Because both heat transfer and friction are Laminar shear computed using gradients of the dependent variable, it is very important to accurately capture this near-wall variation. 1 The standard method is to apply a very ne mesh close to the wall, to resolve the ow. This method is called the integration method, which necessitates an LRN (lowReynolds-number) type of turbulence model. At higher Reynolds numbers, the region under the wall inuence 0.5 diminishes. However, because it is equally important to PSfrag replacements accurately capture the near-wall gradients, a large number of nodes are then necessary. From an engineering point of view, this becomes inconceivable and a function that bridge the near-wall region is instead introduced, 0 0 10 20 30 40 50 thereby signicantly reducing the computational requirements. The anticipation is that this can be done with only a small deterioration in the results. This latter ap- Figure 1: Laminar and turbulent shear in the near wall proach is denoted the wall function method, which uses region. Channel ow, DNS-data, [25]. an HRN (high-Reynolds-number) type of turbulence model.

The turbulence is negligible in the viscous sub-layer, while the viscous effects are small in the inertial sublayer. In the buffer layer, however, both turbulent and viscous effects are of importance, see Sahay and Sreenivasan [29] and Fig. 1.

2.1

Buffer Layer

The maximum turbulent production occurs in the buffer layer at roughly , slightly dependent on the Reynolds number. Due to large variations in the different turbulence source terms, see Figs. 2 and 3, the modelling becomes very difcult. Today there exists no general method for applying a turbulence model, with the rst computational interior node located in the buffer layer. Instead of trying to model the behaviour in the buffer layer, the common practise is to place the rst near-wall node in either the viscous sub-layer (LRN-models), or in the inertial sub-layer (HRN-models). A viscous approach is valid in the former case, while a turbulent approach is more correct in the latter, for the rst interior computational cell.

2.2

Viscous sub-layer

In the viscous sub-layer the following asymptotic relation for the velocity, temperature, turbulence kinetic 1

  t 3 r

 t s

Damping the wall normal components, making the turbulent ow anisotropic.

inertial sub-layer

f f  s h f f h s s f

0.25 0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2 0.25 0 10 20 30 40 50
Production Dissipation

10 9 8 7

PSfrag replacements

5 4 3 2 1 0 0

acements

Figure 4: Near-wall variation in DNS-data, [25].


20

0.25 0.2 0.15 0.1 0.05 0 0.05


Viscous diffusion Turbulent transport Pressure-diffusion Sum of diffusion terms

18 16 14 12

10 8 6 4

acements

0.1

PSfrag replacements
0 10 20 30 40 50

0 10 20 30 40 50 60 70 80 90 100

Figure 5: Variation in Channel ow, DNS-data, Fig. 4

the dissipation rate, while the turbulent kinetic energy (1) should be solved and not set a priori. (2) (3) 2.3 Inertial Sub-layer (4) The inertial sub-layer is the region observed from where the variables are normalised according to the around and outwards, where the assumed vanomenclature. riations are, see e.g. Wilcox [36]: The constants used in the relation for the turbulence quantities in Fig. 4 are and , which gi(5) ves close agreement in the wall vicinity. As can be seen from the gure, the relations for the velocity, tempera(6) ture and dissipation rate are a fair approximation, even up to , although the model for turbulent kinetic (7) energy yields a strong over-estimation for larger values. In light of this, a plausible turbulence model would (8) be to assume a variation in both the velocity prole and 2

energy and dissipation rate is valid:

  t 3 r %

Figure 3: Viscous diffusion, turbulent transport and pressure diffusion terms in the -equation. Channel ow, DNS-data, [25].

in the inertial sub-layer. [25]. Labels as in

% ) ) s 2'10o '% (&o s

  t $3 r

Figure 2: Production and dissipation in the -equation. Channel ow, DNS-data, [25].

 e

V o w

#  !"! #  $"!          
2

, DNS , Model , DNS , Model , DNS , Model , DNS , Model

f h t s

 f t ' 

  t 3 r

f t 

 t E3 r 

' ' s s i(W s

f " t s

s s

10

. Channel ow,

3.1

Momentum

Assuming that a Couette-like ow prevails, the following simplications can be made to the momentum equations:

1D ow, with variation only in the wall-normal direction: . Fully developed ow with zero gradients in the streamwise direction; , apart from a pressure gradient: . Negligible convection.

Because Eq. 10 is now an ordinate differential, it can easily be integrated to yield:

Inserting this into Eq. 9, assuming that the integration yields:

At the centreline of a channel the shear stress, , and the velocity gradient, , are zero and the pressure gradient can consequently be established as:

The shear stress variation can thus be written as:

laminar

turbulent

h f t cg b bw w # g

g u% g

The streamwise and the wall-normal tum equation then reduces to:

momen-

where , and are justied by DNS-data. For air, Kays and Crawford [20] give the following values: and in the relation for the temperature. The inertial sub-layer is also denoted the logarithmic (log) region, due to the above characteristic proles, which feature a logarithmic behaviour. The models assumed for the inertial sub-layer are compared with DNS-data in Fig. 5. There is excellent agreement for both the velocity and temperature relations beyond . The turbulent quantities are however less accurate, and the assumed constant values for the turbulence kinetic energy are not correct. At there is a descrepancy of around 25% when using either Eq. 7 or 8 as compared with DNS-data.

3 Law-of-the-Wall
The variation in the different parameters in the logarithmic region, postulated in the previous section, is based on the law-of-the-wall and its consequences. Below, the law-of-the-wall for both momentum and temperature are derived. The latter is an extension of the former, based on the modied Reynolds analogy. The assumed variation in the turbulence quantities can only be proven, see section 3.3 and 3.4, in conjunction with a specic turbulence model. In this paper the Wilcox turbulence model is used. The constants included are derived in section 3.4.

! F # b bw V z % q bw V # % q 

!

 #  g  b w b w W g h cg bw V # % q g W

&A h ~ h cg g

% &B

h ! W b bw w # t &7 h W

B t e

# $t

&e 6 qr t

t h

! "   b wb V #

h ! t f W

! g B t ! h W

h ! W

b wb V #

 

!   % q

t f

  t ) 

  t 

 f t o x i(W t )o   f

 f  f t

 f t o

ff hg t s

(9) (10)

(11) ,

(12)

(13)

(14)

Note that is negative and hence the left-handside (LHS) of Eq. 14 i.e. the total shear stress linearly decreases from the wall to the centre of the channel. This is of course in excellent agreement with DNS-data for fully developed ow, see Fig. 1. If Eq. 14 is normalised with the friction velocity, , and a length-scale based on wall quantities: (15)

The relation becomes:

PSfrag replacements (16)

Figure 6: U-velocity prole with linear and log-law ap(17) proximation. Channel ow, DNS-data, [25]. (18) where is an integration constant, which is given a value of based on experimental data. The logarithmic law or log-law for the velocity is thus:

The velocity prole using this relation and the linear relation, , is compared with DNS-data in Fig. 6. , The intercept between the two laws is found at located in the buffer layer, which is, as is obvious from (20) the gure, not accurately modelled by either of the equations. For large the rst term on the LHS can be neglected and thus the relation reduces to:
0.01

(19)

(21)
Prt=t/t

Note however that this does not give a relation for itself. On the basis of purely dimensional grounds, it is possible to establish a relation for the velocity prole, similar to Eq. 17, although as a function of the wall distance in outer variables only, van Karman [33]: (22)

0.005

t t

where is the centreline velocity and is the defect velocity. This relation is called the velocity-defect law, Tennekes and Lumley [31] 0 0 Somewhere in the ow it is plausible to assume that PSfrag replacements 0 100 200 300 400 both relations for the -velocity hold. Differentiating Eqs. 17 and 22 and adding a proportional constant as done by Clauser [9] gives: Figure 7: Diffusivities and turbulent Prandtl number for air. Channel ow, DNS-data [25], [18]. (23) where is the von Karman constant, equal to Integrating the latter relation gives: .

3.2

Energy/Temperature

(24) 4

An inspection of the structure of the governing equation for energy and momentum reveals that the near-wall

h t s

and the friction velocity, the following relation is given:

 t 3 r 

' (%   s e  f t s % 

s t s %

These relations are called the law-of-the-wall expressed in inner variables, as proposed by Prandtl [27]. If the inertial sub-layer is instead approached from the core of the ow, and the equation is normalised with the mean ow length scale:

 t 3 r 

e % s

where the turbulent Reynolds number is dened as: . For a large ratio of , one can assume that the total shear is constant. Hence the velocity and shear stress can be written as a function of only:

s %

 f f  

s $ E3 r

$3 r n t s cV n % 3 r w s

$3 r sw s s t V s %

o o s n n t 6 s t 8

n 8 t

% 2'&o s t 6

ed t s V w s e 6 t % s s

n @ t s V w

b bw V

25
DNS Linear Logarithmic

20

15

w V

% DV

w B V t $3 r

s
%

10

10

10

10

(25)

0.5

mechanisms would be similar only if the Prandtl numbers (both the molecular and turbulent) are unity. The denition of the turbulent Prandtl number is:

Air: Water: Table 1: Critical

0.7 5.9

13.2 7.55

(27) (28)

and re-arranging, the equation yields:

This gives the variation of as a function of , although in an integrated form, which is not preferable when making comparisons with the law-of-the-wall for the momentum equation. The formulation could however be simplied if it were possible to divide the nearwall region into a laminar and a turbulent layer. In the case of the momentum equation, the velocity prole follows approximate a linear relation in the viscous sublayer (laminar) region and a logarithmic relation in the inertial sub-layer (turbulent) region. If the thermal eld behaved similarly, the above integral can be divided (29) and integrated successfully. The location of this partition depends on the molecular Prandtl number, which where is the specic heat of the uid and is the for a value of unity would be identical to that of the mo(viscous) thermal diffusivity. mentum equation, i.e. . For a lower value of the Note that the pressure gradient, contrary to the momen- Prandtl number, e.g. air ( ), this critical value tum equation, does not appear in this equation. If a non- will be higher, with the opposite being true for a high negligible pressure gradient were present, much of the Prandtl number uid, e.g. water ( ). Examples similarity between the momentum and energy equation of these cross-over locations, or critical values, are giwould be lost. ven in table 1, see also Kays and Crawford [20]. Taking Assuming that the turbulent heat ux, , can be the value for air, the two integrals would be: described with the turbulent thermal diffusivity and the turbulent Prandtl number, the relation can be re-written (37) as: (30) where it was assumed that the turbulent part could be neglected in the rst integral and the laminar part in the second integral. The rst integral is easily integrated to yield the linear-law for the temperature as:

When integrating this equation from the wall to the centre of the channel, the heat ux is found as:

wxw

x (W i

wxw

where the integration constant, heat ux applied at the wall:

, is a function of the

However the second part can only be integrated if the variation in and is known. In the inertial sublayer the turbulent Prandtl number can with acceptable (32) , or the accuracy be assumed constant. The ratio

' 7 a i s  s (W t

(31)

  t (W i

' a  x (W 9&w x Q  i w  s

f t (W i t s

s x  i (W Ayw x w (W   i s

!p V # ! w V

 s (W ' a i 

t s

t s

Reynolds, in an attempt to describe the heat exchange process, proposed a model which is based on similarities between the momentum and energy transfer. Reynolds and analogy presupposes that the two diffusivities, , are equal and hence, from Eq. 26, that the turbulent Prandtl number is unity, . For a uid with a unit molecular Prandtl number, , this would result in identical temperature and velocity proles. The validity of the Reynolds analogy can be observed in Fig. 7. As can only be acnoted the assumption of a constant ceptable in the inertial sub-layer, and with a value less than unity, . Although the validity of the Reynolds analogy is limited, there still exists a strong similarity between the velocity eld and the temperature eld, which implies that there could exist a corresponding law-of-the-wall for the thermal eld. With the same assumptions as for the momentum equation, i.e. 1D fully developed ow eld and thermal eld, the energy (temperature) equation in the wallnormal direction can be simplied to:

On introducing normalised values:

)  x (W x q (W q   i i ! p t ) x (W i  i (W t ! q x q p

The turbulent Prandtl number and the turbulent momentum diffusivity, , and turbulent thermal diffusivity, , are plotted in Fig. 7. The denitions of and are:

Re-arranging and integrating once more gives:

(26)

values, air and water

x d s

i (W

x 

xw

xw


b b w

x (W i

t (W i t x (W i

b w 

% b b w b bw V

x i (W (W iq t f x q

xw

x (W (W i iq x q

i f x (W

t p

b b w u% b wb V

!p t

x (W i t x  t f x  x  xw

(33)

(34) (35)

(36)

(38)

turbulent Reynolds number, can be approximated using the mixing-length theory, which gives:

model (see section 3.5) reduces in the inertial sub-layer to:

Adopting this and evaluating the integral yields: (40)

(41)

20 18 16 14 12
DNS Linear Logarithmic

10 8 6 4 2 0 10
0

These relations can be proven by re-inserting them into the above transport equations. Using Eqs. 49 and 46 in Eq. 42 yields:

acements

10

10

There is a number of different approaches to the turbulent law-of-the-wall, or log-laws, where the derivations below mirror that of Wilcox [36]. If 1D fully developed ow is assumed, and the pressure inuence is neglected 1 (even in the streamwise direction) the Wilcox HRN
1 Note that the log-law may be derived for any turbulence model. The Wilcox is used however, because it will later be used in the description of blending turbulence model.

The turbulent kinetic energy equation is also fullled. Turning to the dissipation rate equation, Eq. 44, and 6

3.3 Turbulence

t ' o ' V V

o o o f t a V a V t f a V V w 2'% V#

% %

o ' V o V t o V

o V

Figure 8: Temperature prole with linear and log-law approximation. Channel ow, DNS-data, [18].

Thus these wall functions satisfy the momentum equation. Eqs. 46, 47, 48 and 49 are substituted into Eq. 43 which gives:

f t  ' V w V

o o t V V t 2'% o V # o V t f

w V

o o V V ' V '2% o V

V t f o

t xw

where the constants are found by using properties for air ( and ) and the van Karman constant, . The two formulations, the linear-law and the log-law, are compared with DNS-data in Fig. 8. Similar to the momentum (velocity) model, the agreement is less accurate in the buffer layer where neither equation gives acceptable performance. For air, the cross-over between , and hence the lamithe two laws is found at nar thermal layer extends further out than the laminar momentum layer.

where the laminar viscosity in the diffusion terms is neglected. Note that the neglected pressure gradient in the momentum equation (Eq. 42) is a severe limitation in the following derivation. This makes the wall function approach even more questionable for the turbulent quantities than for the velocity eld. The above equations can be satised in the inertial sub-layer with the following set of wall functions, see Wilcox [36]: (46) (47) (48) (49)

t xw

Thus the law-of-the-wall for the thermal eld, or the temperature log-law, is, adding Eq. 38 and 40

w ' ' % x w xw ' % x % xw

t t t f f f

 ' (% ' (% o   s  f   t s x (W (W   t s  i  i

 t 3 r 

 s

 '  7 a ' 7 a ' (% o so  x i (W t  s x (W t s i

  t s 

w s o t xw s

 f t x (W i

o V t xw f t o  f t (W  i s

(39)

(42) (43) (44) (45)

(50)

(51)

using the wall functions, Eqs. 46, 47, 48 and 49, gives:

from Eq. 47 and the measured relation of above, Eq. 53, we can write the coefcient as:

By studying how grid-turbulence dies out, it is possible to determine the coefcient, . In decaying turbulence, the turbulence equations simplify to:

3.4 Turbulence Model Constants


0.3

Measurements [32] indicate that the turbulence kinetic energy decays as:

0.2

0.15

0.1

The turbulence model by Wilcox [34] gives this coefcient the value of .

0.05

acements

There is no consensus within the turbulence modelling community about the values for these coefcients. Diffe0 0 20 40 60 80 100 rent Schmidt numbers are arrived at depending on the set of experimental data used. In recent years there has Figure 9: Shear stress to turbulent kinetic energy. also been a greater tendency to introduce modications for these coefcients, and the values used in turbulence Channel ow, DNS-data, [25]. models thus differs. The two Schmidt numbers for the model [34] are optimised to:

Experiments have shown that the shear stress and the turbulent kinetic energy is related as

where the friction velocity is dened as . In the Wilcox . Now, using the relation for the turbulent kinetic energy

model

is given the value of

t 

(54)

   f t f f  f f  f ff f t ' o ' 

 f t o

in the logarithmic region. The validity of this approximation is shown in Fig. 9. If the shear stress can be assumed to be constant and the laminar part is neglected, then:

The nal coefcient, , is set using the deduced relation from the specic dissipation rate equation, Eq. 52. With and the above constants, is calculated as: (60)

  f

# ! $F e

' V #

3.4.1 Coefcient,

(53)

3.4.4 Coefcient,

3.4.3 Coefcients,

  f f t f t

and hence the ratio from above, the correct value for
 

. With would be:

  f f t

0.25

 g f  f   t f D   t } f 

 

f  f f  f t 

x 9  s s

To satisfy the last relation, it is necessary to establish an identity between coefcients , , , and .

Solving this equation system gives: (58)


'

  t t

(52)

3.4.2 Coefcient,

 f  f t # # ' # ' V  f t  f t V

t t

' t ' o t V o ' V t o

'

  t E3 r

t ' o f t 'o t ' o t ' V V ' ' h V ' o o V o

'

% %

w V

! F t x  !    

o V ' (% o V

#  f

 

t f

(55)

(56) (57)

(59)

3.5 HRN Turbulence Model


On the basis of the above derived turbulence constants it is possible to write the Wilcox model as follows:

4 Modelling lence

Near-wall

Turbu-

with the constants as derived earlier:

The Wilcox HRN

model was given in the previous

 t t f t hg  t  t  ff

 f cg h h f cg g x w w g ' Y W t 00

f cg  h h w f cg g x w Q g Y W t 00

With respect to wall treatment, there are two different categories of turbulence models. The rst are the LRNmodels, which use a rened mesh close to wall in order to resolve all the important physics. The second met(62) hod employs the HRN-models, which bridge the nearwall region using wall functions. The latter approach is of course less demanding of computer resources, but a signicant amount of information is lost. (63) The wall function method is based on the previously mentioned law-of-the-wall and is strictly valid only in the inertial sub-layer (log-layer), 1D fully developed ow, where the pressure gradient can be neglected. The situations in which all these simplications are fullled are relatively few, and the predictions made using wall function or HRN turbulence models are thus generally less accurate than those that apply the integrated method of LRN turbulence models. (61) The reason why there is still an interest in wall functions is the need to reduce computational requirements, as well as the better numerical stability and convergence speed. If it were possible to improve the predictions using wall functions, more problems of engineering interest could be solved with academic quality, resulting in fewer safety margins and less need of experimental validation when designing new products. The previous section derived the commonly used wall functions based on the law-of-the-wall, while this section will focus on alternatives to these wall functions. However, before embarking on strategies in constructing wall functions, it is benecial to study the near-wall physics from a modelling point of view. Attention is rst directed towards the LRN-models and their damping functions.

t xw

4.1

LRN-models: Integration Method

It is generally implied that a turbulence model that can integrated toward the wall is denoted a LRN turbulence model. This is usually the same as including damping functions for certain terms in the turbulence equations. The damping functions are introduced to represent the viscous effects near a wall. Successfully devised damping functions should reproduce the correct asymptotic behaviour in the limit of a wall. The exception to the rule that damping function LRN is the HRN turbulence model by Wilcox (1988) [34]. This model can, like the LRN version [35], be integrated to the wall but without the need to employ damping functions. In this paper the identity of damping function and LRN is used for these models as well. This nomenclature thus helps to distinguish them, with the 88 model somewhat erroneously denoted the HRN and the 93 model the LRN model.

Thus, in the near-wall region, the production terms (rst terms on the RHS of Eqs. 66 and 68) of the LRN (68) model are reduced as compared with the HRN model, through the reduced eddy-viscosity, , via . In (69) addition, the dissipation term (the second term on the RHS) of the -equation, Eq. 66, is also reduced, through . where the turbulent Reynolds number is dened as and the turbulent production as 0.1 . The coefcients involved are:
0.09 0.08 0.07 0.06
DNS

3 2.5

DNS:

DNS:

1.5 1

4.1.1 The damping functions The near-wall behaviour of the three important constants and are shown in Fig. 10. The introduction of the wall viscous effects, via the damping functions in LRN-model can observed in the gure. The results are based on both a priori DNS-data, [25], and also on a computation made using the model. As can be seen in the gure the damping functions are most effective in the viscous sub-layer and in the inner

Measurements and DNS show that the assumed relation of is valid only in the inertial sub-layer. The ratio is reduced in the near-wall region, as shown in Fig. 9. For a model the effective is computed as:

Using a formulation, both and in the relation must be substituted with the appropriate denitions. In turbulence model, is equal to the case of the LRN 9

xw

' b bw V

 t  6 t 

 !

f t b bw V 

'  x w t 

  t 3 r

Figure 10: Damping functions in Wilcox Channel ow, DNS-data, [25].

models.

xw

0 0

10

15

20

25

30

The most important damping function is the damping of the turbulence kinetic viscosity, . This is done by adding a damping function to , which reduces to an effective according to:

0.5

4.1.2

modication,



Figure 11: [25].

-coefcient, LRN

and DNS

  t E3 r

acements

3.5

                     

DNS:

PSfrag replacements



This model reduces to the HRN , see section , if the damping functions, Eqs. 65, 67 and 69, are set equal to , and , respectively.

0.05 0.04 0.03 0.02 0.01 0 0 20 40 60 80 100

xw

x f t $3 r

                

t Y W

    t utr f g  t Y r t tr s  t  t f t ` t   g t f h f g % w d  Qy

u x tr g3 r   u x vr g3 r   f g h h w f cg g x w Q g ' `W  Y

v ur g3 r  s x h h f cg w f cg g x w Q g `W Y

x v s 3 r "3 r hg ff

section and the LRN

model [35] is:

part of the buffer layer, while the HRN and LRN formulations become similar in the inertial sub-layer. In the limit of , i.e. at the wall, the coefcients (64) for the LRN and HRN formulation give: (65) LRN HRN LRN/HRN 0.025 1 0.025 0.025 0.09 0.28 (66) 2.22 0.55 4 0.075 0.075 1 0.055 0.55 0.1 (67)

Y r x 3 r  Y r x 3 r  


ff g 

t xw t


 00

00

w g f d x h t x3 r

  

(70)

(71)

1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 5 10 15 20

(72)

4.1.3 Production-to-dissipation rate

Figure 12: Production-to-dissipation ratio, HRN and Another important parameter in a turbulence model is LRN formulation of the . Channel ow, DNS-data the correct modelling of the production-to-dissipation at . rate, . By comparing the HRN and LRN formulation model, it is possible to show the effect these of the damping functions have on the production-to-dissipation Decrease the coefcient for the dissipation term in ratio. the -equation. From Eq. 66, and assuming 1D ow, the productionIncrease the coefcient for the production term in to-dissipation rate is: the secondary equation ( or ). Decrease the coefcient for the destruction term in the secondary equation.

(73) The last method is commonly employed in the models. By properly balancing the dissipation rate equation, a desirable level of is predicted that matches the computed . Because the production of dissipation is automatically decreased through the reduction of , the where is the value of in the logarithmic region, simplest correction is to reduce the destruction term in . The log-laws, Eqs. 46 and 48, were used in the the dissipation rate equation. This is usually done in a model by introducing a damping function , see relations for and , respectively. The approximate Patel et al. [26]. identity is correct only in the inertial sub-layer for which There is more to the damping functions than balanthe log-laws are valid. In the near-wall region, cing the equations, however. As previously discussed, it , the above relation may only be used qualitatively. The a priori computed production-to-dissipation ratios is also important to predict an accurate production-tofor the HRN and LRN formulation are compared with dissipation ratio, which puts an additional demand on the actual values of and . In a moDNS-data in Fig. 12. del, where the dissipation term is a function of both The LRN formulation, similar to the DNS-data, increases the production in the buffer layer, which results in and , it is unrealistic to believe that merely modifying , as well as a a higher peak of for the LRN-model as compared the -equation would yield a correct correct level of . Hence, apart from modifying the speciwith the HRN-model. Slightly dependent on the Reynolds number, DNS-data give the peak of turbulent ki- c dissipation rate equation, the dissipation term in the -equation is also modied. In the LRN model these netic energy at around , close to the maxima of two aspects are accomplished through both an increase the -ratio given in Fig. 12. of production of (increasing ), and a reduction of the dissipation term in the -equation (decreasing ). Note 4.1.4 Rationale behind the and coefcients however that, contrary to the model, the destruction is unchanged. , term in the secondary equation As was previously shown, the damping function, was introduced to reduce the turbulence viscosity in the Considering Patel et al. [26], it seems likely that the near-wall region. This also effectively reduces the pro- minimum number of damping functions required for an duction terms in the - and -equations. In order to pre- Eddy-Viscosity Model (EVM) to correctly predict the turdict the same level of turbulence, the dissipation term bulent kinetic energy, , in the near-wall region is three. in the -equation must also be reduced. This can be ar- In that paper the only model that accurately reproduces model, ranged in three different ways: the prole of is the Lam-Bremhorst [21] 10

'

xw

Y $ W

 t

 t $3 r 

  

and the second part (non ) of perform where the role of the damping function, , found in turbulence models. Any deviation from the correct is primarily modelled with the damping function, . Note however that the turbulence equations are strongly coupled, and thus a change to the damping functions in either or would also require a modication of . PSfrag replacements The variation in the computed using the LRN is compared with DNS-data in Fig. 11. The agreement is good, especially as compared to other models, see e.g. Patel et al. [26].

    
25 30

, while the turbulent viscosity is dened as in Eq. 64. The effective thus becomes:

2
DNS LRN HRN

x x v sur g3 r Y(r "3 r x Y x v s 3 r "3 r (r "3 r   t o  e o  ' V

 f  f
t '




t s

'



u%

 t Y W

" Y W

   f  f t

" Y W



 f


f "

where

For the specic dissipation rate, , the friction velocity, , is substituted by the relation for in the log-law as:

Set at the rst node iteratively with the use of the law-of-the-wall.

4.2.2 Launder-Spalding methodology The standard wall function method is very limited in its usage. This is especially true for re-circulating ows, where the turbulent kinetic energy becomes zero in separating and re-attachment points, where, by denition, is zero. This singular behaviour severely deteriorates the predictions of, for instance, heat transfer in ribroughened channels. Launder-Spalding [24] proposed a modication to the standard wall function method that involves the following steps:

(74)

(i) an added source term simulating the correct wall friction or (ii) a modied viscosity, an effective viscosity, , that ensures the correct friction even though the velocity gradient is erroneous. Through the law-of-the-wall:

The same reason and method of specifying the wall viscosity as for the standard wall functions apply in this case. However, instead of using in the log-law, the identity is utilised:

(75)

the wall friction is computed as

(76)

In the -equation, the integrated production term is dewith . ned from (1D): For a wall function based model this can either be set directly using a source term, , or via a modied (83) effective wall viscosity [10]:

11



(77)

In the inertial sub-layer the laminar shear stress, , is negligible and, by assuming that the shear

where

is dened with

as: (82)

% g x  g

w A  t vx ('%  q 5 x o e v 7  # t A  t V vx

Y W#

g t% g q

where the subscript is used for the rst interior node. The necessary modication could either be made through

In a turbulent boundary layer, the strongest velocity gradient is found close to the wall. With a wallfunction based turbulence model, which utilises a relatively coarse mesh, it is impossible to resolve these nearwall gradients. The predicted wall friction would thus be largely in error if a modication is not introduced:

s o

2'%  5 q V# t

F !

~ % q t % l q l

s o %

' (% 5 ! F t $ %  q V# t

5 2'% o t % V s

' V # !  s 5 2'% F ! o % V# t

! g % g q

Set

with .

Solve the momentum equation with a modied wall viscosity. Solve the turbulent kinetic energy, with modied integrated production and dissipation terms. Set using the predicted .

Solve the momentum equation with a modied wall friction, either through an added source term or via a modied effective viscosity.

E $ x o v AX E e t $

E o e AX t vx

In its simplest form, the logarithmic laws, Eqs. 46, 47, 48 are directly applied to the rst interior node. These wall functions are here referred to as the standard wall function method. The three steps involved in a CFDcode would then be:

' V

4.2.1 Standard wall functions

The new value is then used to compute and re-fed into the expression. This process is repeated until convergence, with the turbulent kinetic energy nally set as:

w V $ t s

2'% s 5% t V o

The rationale behind wall function is the reduced computational requirement and the increase numerical stability and convergence speed. By adopting a mesh, where the rst interior node is located in the inertial sub-layer, it is possible to use the law-of-the-wall to specify the boundary condition for the dependent variables , and .

(78)

(79)

(80)

(81)

x e v AX t V

4.2 HRN-models: Wall Functions

which is one of only two models tested employing three damping functions.

The turbulent kinetic energy is set in the rst node by iteratively computing the friction velocity from the lawof-the-wall. This iterative process is needed because appears implicitly in the log-law. Initially is set from Eq. 47, as . Using this relation, a new is found from the log-law:

t $

! F

stress is constant, the turbulent shear stress, , is equal north to the wall shear stress: PSfrag replacements (84)

wall

(85)

Eq. 81

The integration of is equal to since is constant. To set the value for the dissipation rate, production is assumed to be equal to dissipation:

Figure 13: Assumed variation of variables in the rst computational node, Chieng-Launder model. vs=edge of , viscous sub-layer. Comparing the two formulations, Eqs. 88 and 91, gives:

(86)

(92)

The production can be re-written as:

Eq. 88

Eq. 91

which is a direct consequence of substituting the friction velocity with the log-law for turbulent kinetic energy, , into Eq. 86.

4.2.3 Improvements of the near-wall representation, Chieng-Launder model (89) It is evident from the above derivation that there is room for a certain value of . Approximating the velocity for improvement in the Launder-Spalding model, espegradient with discrete values, , and, assuming cially in its representation of the near-wall region. Chithat this value could be established in the inertial sub- eng and Launder [7] enhanced the model by introdulayer, the velocity gradient could be re-written with the cing a more complete model of the near-wall turbulence. help of the law-of-the-wall: They argued that there exists two distinct regions, the viscous sub-layer and the inertial sub-layer, which have (90) different turbulent structure. The partition line between these regions is dened by a viscous sub-layer Reynolds number: Hence the integrated dissipation rate can also be written

Eq. 87

12

Eq. 90

The near-wall variation in the variables assumed in the Chieng and Launder model is shown in Fig. 13. The turbulent shear stress is zero for and is (91) linearly interpolated between the wall friction and the

w t 3 r

as:

o o ' 9xXa v Xa  t a V t x

7  t V vx

tion ( ) necessitates re-consideration. The shortcoming of the above derivation is the assumption that the law-of-the-wall is valid all the way from the wall, which is erroneous. However, by choosing a different relation for the integrated , other approximations could be found. The integrated velocity gradient in the production term can, according to the mean value theorem, be re-written as:

The latter formulation is used in the Launder-Spalding (87) model, with the integrated dissipation rate in the equation set as: where the constant shear stress approximation, Eq. 54, , and the law-of-the-wall, (93) , were used. Integrating the dissipation rate between the wall and the edge of the rst cell, , yiwhere in Eq. 91 has been substituted by . elds: Note that in [24], is used rather than , although (88) this is most probably a typographical error. The dissipation rate equation is not solved for the rst interior node, and is instead xed according to: The mathematical singularity at the wall of this equa-

(94)

(95)

x Xa  'xa x 99 9a  v v

! F

a V o ('% 5 99a Xa  t 'x vx

 ! "5 2'%  t ! f ('%   & 2'%  l 

The accuracy of this assumption is shown in Fig. 1. The production can be set using the law-of-the-wall:

vs

g t % g b bw

l t g u% g

% ! & g u% g !  l ' (% % & 5  # t l o % x v A g !  ! t % l % q t % l  t % g 

x

2'% o 2'% o t 5 a V t 5 l V l ' V t      t `W  V   t Y t

o  ' (% o 2'% f & l a V t a V 

u% l l

 ! t   g t 

g 2'% o 5 lV t % ll % g

g t% g g g  % g l t % g 

o g ' V V t % g b bw V t Y W

 o &  V ' V # t b w b V # t 

! F t x  x  t  t
 t 

t Y W# f 2'%

t Y W

!

  

where the new constants, and , are related to the standard values, through the identities: and . For local equilibrium ow, (96) , and with the above identities, the modied law-of-the-wall reduces to the standard formulation. The integrated production following the same grounds where subscripts and are used for the value at the as for the Launder-Spalding model yields (1D): north edge of the cell and at the wall, respectively. The turbulent kinetic energy is approximated as:

The dissipation rate is approximated in the region below by the viscous sub-layer approximation used in numerous turbulence models, e.g. Jones and Launder [15], while the log-law is used for :

The rst part is identical zero, since the turbulent shear stress is assumed to be zero in the viscous sub-layer. The second part is split into two, where the rst part is easily integrated:

(98)

The wall friction, , used in these relations is deduced from the above modied law-of-the-wall, Eq. 100. The integration of the dissipation term is given as:

(99)

with

4.2.4 Modication to the Chieng-Launder model The assumption of is not valid for all types of ows. Using DNS-data on fully developed channel ows, is equivalent to , which is the location of the cross-over of the linear law, , and the . Thus, for equilibrium ows, log-law,

13

s 5 ('% o t s %

(100)

s t s %

' 

 t s

' AXa & 4p t 'x  F 'x t % 99a 4p 9 

f t 3 r

$  z $ $ t 

f t 3 r

However, since is unknown, the problem is not closed. is approximated by extrapolating the slope of from the computed s in the two rst near wall nodes. It should be noted however that this methodology is not numerically satisfying. It can be shown using DNS-data that is found at . The prole of around this value can be seen in Figs. 4 and 5. The maxima of is found at . Thus the computed slope of using nodes 1 and 2, located at say and becomes negative, which may be of concern in terms of numerical stability. In the Chieng-Launder model, the common practise of estimating the friction velocity with the turbulent kinetic energy is taken one step further, with the law-of-thewall modied as:

where . The constant, , depends on the turbulent kinetic energy via variable :

   t  p   'x 'x 'x 'x p 9A 97   9Xa AXa  " 9X3x a r  t '

! # z  e  F o F !   ! o  !  F    # t $ !  g  t % g ! 

 

The turbulent kinetic energy gradient in the rst relation is approximated using the previously assumed variation in in the viscous sub-layer, i.e. . Because the nite volume method commonly applied, and the strong variation in these quantities in the nearwall region, it is advantageous to use integrated relations in the rst computational cell. The resulting terms are then incorporated into the -equation, which is subsequently solved. In additional and identical to the Launder-Spalding model, the wall friction is corrected and the dissipation rate set. However, before continuing with the integration of and , it is proper to dene the location of the viscous sub-layer. In the Chieng-Launder model, is deduced from the assumption that the sublayer Reynolds number is constant and equal to 20, and hence:

In the second part, the velocity gradient is re-written using the law-of-the-wall:

 % !  t % g !   g

(97)

% g  F  F !  ! g

x v 7

5 t 5

! g "    % g x    t t Y W# 

 

 e ' x  V t v A o t o

f t s

'

shear stress at the edge of the cell for

f t 5

 f t o

t s  h t s

w 5

$ o 'x x 99a v Xa  t ' g h w t w h t ' g

 }   ! z  !  t x  f t x

z  

'(% o

'

f t 3 r

wf t

# $ !  t %

t t

(101)

ff " t s

(102)

(103)

(104)

(105)

(106)

viscous sub-layer in order to accurately predict the wall friction and heat transfer. To be able to signicantly improve the results, it is necessary to mimic the LRN method of resolving the viscous sub-layer, especially in nonequilibrium regions. Thus the wall function approach of bridging the viscous sub-layer seems unattainable, apart from regions in which the ow is in equilibrium. Hence the only possible path to reducing the computational demands and maintain accuracy would be a turbulence model that adopts itself to the ow, employing (107) a LRN type model where necessary and switching to a HRN type model otherwise. While such articial intelwhere , and is the linear extrapolated ligence may sound incredible it is possible to achieve it turbulent kinetic energy at the wall. through a smartly devised model. Ciofalo and Collins [8] used an approach however where they proposed to let the sub-layer thickness vary, The second problem is associated with the underlying dependent on the turbulence intensity. physics of the wall functions, i.e. the law-of-the-wall. When employing wall function, it is necessary to apply the rst computational node in the inertial sub-layer. 4.2.5 Extensions to the Chieng-Launder model This fact is less of a problem in equilibrium ow but it Amano et al. ([2], [3]) devised two different models ba- is accentuated in regions of ow separation 2 , where the sed on the same zonal principle as the Chieng-Launder wall shear stress and hence decreases. For these remodel. In both models, the dissipation rate equation is gions, the size of the rst cell must be enlarged to ensure solved using the layered approach rather than allowing that the cell node is kept within the logarithmic region. it to be set as in the Chieng-Launder model. The results Hence the mesh generation becomes a delicate business were only marginally improved however, and hence the using wall functions for re-circulating ows. two-layered approach was extended in the second model Both of the above explained problems are solved with to a three-layered version. Here the parameters, the two-layer Chieng-Launder model. The CL-model, were allowed to vary in accordance with measured data however gives rise to additional complexity. Through for the viscous sub-layer, the buffer layer and the inerprolonged mathematical operations the viscous subtial sub-layer. layer and the inertial sub-layer are modelled simultaneously, although the author encountered numerical dif4.3 New Model: Blending Integrating culties using this approach. If the two regionscould and Wall Function Boundary Condi- be separate mathematically, a numerically simpler and more robust wall function could be constructed. tions Such a turbulence model will be presented here. A less The physical treatment improved, w hen progressing complete derivation is given in Bredberg et al. [6], where from the Launder-Spalding model to the more advanced predictions of heat transfer in rib-roughened channels models, although at the expense of increased numerical are included. The reader is referred to the paper for a complexity. The modications by Johnson and Launder discussion of the accuracy and for comparisons between [14] and Ciofalo and Collins [8] showed that the impro- the model and other turbulence models. vements with the zonal approach were not particularly Dependent on some ow quantity, the model will eitgreat if a variable sub-layer thickness was not included. her employ a wall function or use the integrating method The further renement by Amano et al. enhanced the as a wall boundary condition, and hence greatly simppredictions only marginally, with a signicantly increalify the numerical scheme as compared to the Chiengsed complexity. Launder model. The new model will, like the wall funThus, here, instead of further optimising the layered ction, approach reducing the computational demand models, a new strategy will be adopted. by coarsening the mesh, when possible, and still be able Two main problems with wall-functions must be addto produce accurate results through the use of the LRN ressed to make such an effort worthwhile: part of the model, when necessary. The physical conseThe negligence of the physics in the viscous sub- quence of separating and re-circulating regions can aclayer, tually be utilised with such an adopting turbulence model. In non-equilibrium nodes where the physics reduces The necessity of locating the rst computational the value, the hybrid model automatically switches to node in the inertial sub-layer the LRN mode, producing accurate results even for a reThe rst problem has been discussed in numerous pa- lative coarse mesh. Thus even a uniform mesh could be pers, see e.g. Launder [22]. The conclusion made in sufcient for complex re-circulating ow, and, as a rethese studies is that the standard wall function formu- sult, reduces the workload of mesh-generation substanlation (HRN-models) are indisputably inferior to the in- tially. tegrating formulation (LRN-models). This is particu2 Although the log-law in itself is rather questionable in such cirlarly true for non-equilibrium ows, where it is of critical importance to include the variation found in the cumstances. 14

is a good partition value, since the two formulations become a laminar and a turbulent approximation, respectively. However, in a subsequently paper by Johnson and Launder [14], it was found that the heat transfer prediction could be improved by letting the viscous sub-layer thickness, i.e. the sub-layer Reynolds number, vary according to the turbulence level:

! h    f t  ! $3 r  $3 r t $3 r

f t 3 r

4.3.1 Blending function The critical part of a hybrid turbulence model is the construction of the function that effectively changes the model into either LRN or HRN mode. It is numerically undesirable to adopt a function that abruptly switches between the two different modes. It is also physically questionable, because the ow is only strictly laminar in the immediate near-wall region enabling an LRN model to be used while only in the inertial sub-layer can the ow be treated as fully turbulent. The latter is of course a requirement for using a wall function-based HRN-model, see Fig. 1. In the buffer layer it is thus most appropriate to employ a smoothing function that blends the two formulation together. One of the simplest mathematical functions that accomplishes this is the exponential function, which can be designed to operate from to dependent on some ow quantity. On the grounds of the above discussion, it would be most benecial to base the exponential function on . However owing to numerical issues, it is disadvantageous to include any wall distance, and hence the blending function is instead constructed using the turbulent Reynolds number, as:

improves predictions for non-equilibrium ows. This is utilised by the Launder-Spalding model and subsequent models. In addition the HRN part of the new model assumes a variation of the near-wall turbulence similar to what is done in the Chieng and Launder model. However, since the viscous sub-layer is treated separately from the HRN part in the new model, a simplied model is sufcient here. The model is also adopted to a type turbulence model. The assumed variation of the dependent variables for this model are shown in Fig. 14. Both the turbulent kinetic energy and the shear stress have lost their linear variation in the inertial sub-layer, as compared to the Chieng-Launder model. This of course simplies the treatment immensely. The specic dissipation is assumed to vary either according to the log-law, , or the linear-law, . As was stated above, only the region above the viscous sub-layer is of interest in this section. The integrated production of turbulent kinetic energy in the HRN region, i.e. from , is found as:

(108)

acements


1.2
, new model , LRN

0.8

Figure 14: Assumed variation of variables rst interior node, new model.

0.6

0.4

PSfrag replacements 4.3.2 HRN part: Simplied Chieng-Launder mo0.2 del applied to a -type turbulence model The logarithmic equations based on , Eqs. 25 and 41 are unsuitable in separated ows. This applies especially for heat transfer since the predicted Nusselt number becomes zero in separation and re-attachment points using these equations. In contrast, experiments indicate a maximum level of heat transfer at or in the vicinity of with these points (Launder [23]). Substituting

0 0

10

20

30

40

the turbulent production is negligible in the near-wall

15

! x  $

Figure 15: The accuracy of the assumption, Channel ow.

where is some turbulent quantity. The LRN and HRN parts needed in this model are treated separately below.

where and are set to and , respectively, in order to comply with the equilibrium values of the turbulence model. The a priori result using this equation is compared with the turbulent shear stress predicted using the LRN model in Fig. 15. Note that
   

  

h  f

(109)

5 &w $ % &#

5 o

' (%

! t F

where is a tunable constant. The model is then constructed to linearly combine the LRN and HRN formulations using this blending func, and hence at the wall, is tion. Since multiplied with the LRN part and is multiplied with the HRN part:

The last identity can be used only if it can be assumed that the shear stress is constant and that the laminar shear stress is negligible in the inertial sub-layer, see Fig. 1 and the discussion in connection with Eq. 84. The wall shear stress in the new model is set similar to the Chieng-Launder model (see Eq. 100) as: (111)

  % !  % g b w b V #  t Y W #  g


' &$

6 t 6

r 6

' 7 

f t x3 r

Qx 

r r 6 t ) ) r

~h3 t 6 w x w t "3 r x

(110)

50

9 8 7

6 5 4 3 2 1 0 0 20 40 60 80 100

(112)

1 0 1 2 3

Figure 17: Integrated dissipation term, -equation.

(115)

The latter relation is valid only for computational nodes within (Wilcox [36]). To establish a grid independent solution, 7-10 nodes are required in this range. Both requirements can however be eased, especially the latter, with negligible deterioration of the predictions, see e.g. Bredberg et al. [5], [4].

where is the contribution of the non-primary shear stress components ( ). The production term and the dissipation term are shown in Fig. 16 and Fig. 17, respectively.

16

u%  t b W 2'%  o E ' w 6 l   'x xa 9Xa v XA 6   t  

(114)

# % F 6 ' !

The asymptotic boundary conditions applicable for the turbulent quantities are:

b W %  Y t `W 

with the production of turbulence computed as . The model coefcients are:

The new blending turbulence model is summarised below and its performance is found in Bredberg et al. [6]. The production and dissipation terms in the turbulent kinetic energy equation are:

4.3.4 The blending

turbulence model

The LRN part of the model is identical to the Wilcox [34] (HRN) turbulence model. The model was shown in Eqs. 61- 63, and is shown again here for clarity:

g % g xw 6 l 

Y t W

 t t  t f t f f t   

f cg h

g xw f cg h g x

f hg d % g  h w f cg ' Y `W t g  h cg w w f g Y W t t xw

4.3.3 LRN part: Standard

(113) PSfrag replacements

5 6 7 8 9 10 0 20 40 60 80 100

where was assumed constant. Contrary to the Chieng-Launder model, the dissipation rate is not set using the log-law. Rather, the mean integrated value from the log-law is used:

DNS LRN New, sum New, LRN part New, HRN part

Figure 16: Integrated production term, [25]. Channel ow, DNS-data,

 t 3 r 

region; hence any discrepancy comparing to the ChiengLaunder model is a consequence of the simplistically assumed variation in the turbulent shear stress. At this moment it may perhaps be of interest to compare the turbulent shear stress given by DNS-data, Fig. 1, with the assumed variations of the two models. The assumption of a constant turbulent shear stress beyond (new model) is perhaps not any worse than the linear variation from the wall shear to edge of the cell in the ChiengPSfrag replacements Launder model. Note especially in Fig. 1 that the (laminar) wall shear is higher than any given turbulent shear, and hence the linear increase in the turbulent shear from the wall shear is not physically founded. The integrated dissipation rate of the turbulent kinetic energy is based on the logarithmic law, Eq. 48:

10
DNS LRN New, sum New, LRN part New, HRN part

('% 'x 9Xa   x o v 7X   t t

 o $   ('% x o v 7X t e  

v XA t x a
o w




  s ' !

x v A9 e t

  t


-equation.

0 0

w f d x h

(116)

(117)

3 2.5 2

1.5

1
  

1.5 1 0.5 0 0.5 1 0

0.5

acements

PSfrag replacements

0.5

0.5

0.5

1 1.5 2 2.5 3 0 20 40 60 80 100

To close the model, the blending function, (Eq. 108), and the viscous sub-layer values, and , must be dened. The viscous sub-layer thickness, , is calculated using the denition of the viscous sub-layer Reynolds number and the assumed variation of the turbulent kinetic energy:

Figure 19: Sum of integrated source terms in -equation. The third source term on the RHS of the turbulent kinetic energy equation, Eq. 66, is the diffusion term, which is not modied in the present model, with the standard relation used. The accuracy using this simplistic model is shown in Fig. 18. It is interesting to combine the three terms on the RHS of the turbulent kinetic energy equation in order to estimate its balance. This is done in Fig. 19. Both the DNS-data and the LRN mode yield perfect balance. The new model is not very well balanced, yielding a large negative source in the buffer layer and slowly increasing the value in the logarithmic region which yields a large positive source in the off-wall region. This latter fact may be the reason why the new model gives a fairly large discrepancy of the predicted in the centre of a channel. The specic dissipation in the rst near-wall node is set as:

The constants, (in the blending function), and are optimised using DNS-data [25]. The combination of and gave the overall best results, as compared with DNS-data. The effect of these values is a rather narrow blending region, see Fig. 21. The calculated and using LRN , [35] data (a priori) and the selected constants are shown in Fig. 22. The new model is dened by Eqs. 116, 117, 118 119 and the blending function, :

(122)

(118)

17

and the above routines for the specication of

and

3 r

acements

' w & 5 2% w 6 w 6 t !w m o

Ey 3 r w yw $3 r

 t

 SQ x

w 3 r w

$3 r

   

t 3 r

t 6

z  e o v 7X ' x ! $  e ('% e 6 w  6 t

1.5

The wall viscosity is set as:

DNS LRN New

A comparison with DNS-data and the LRN is shown in Fig. 20.

Figure 18: Integrated diffusion term, -equation.

Figure 20:

1 0

20

40

60

80

100

20

40

60

80

variation.


100

DNS LRN New

DNS LRN New

model

(119)

(120) (121)

1 0.9 0.8 0.7 0.6

5 Modelling Transfer

Near-wall

Heat

0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 20

acements

8 7 6 5

4 3 2 1 0 0

acements

Figure 22: Comparison of the new models variation of viscous sub-layer quantities, with those given at (constant lines).

5.1

Integration Method

The integrated or LRN method implies that the heat transfer is governed purely by viscous effects in the limit of the wall. The heat transfer is thus estimated using the Fourier law, such as:

where is the thermal conductivity given as . The in the heat ux indicates that it is the laminar or the molecular part. Using the simple gradient diffusion hypothesis (SGDH) for the turbulent region, the turbulent heat ux is computed as:

The total heat transfer is then calculated by adding the . two parts together, 18

i (W q t

x Fp  p t x  gp x

T p (W  p qi E t p t

x '

10

20

30

40

i x (W x mp x q t

t 3 r

Figure 21: Blending function, , using LRN

 

    
data
50

In contrast to previous sections, this one is written in a brief and and abbreviated style. This is not to imply that the thermal eld is any easier to predict or that the modelling is any simpler, but is merely a consequence of the lesser efforts made to improve the thermal models. To exemplify this, it can be noted that, in some predictions, an RSM turbulence model which involves seven turbulent transport equations is combined with a constant turbulent Prandtl number heat transfer model, i.e. a zero-equation isotropic model. The turbulent Prandtl number heat transfer model using the simple gradient diffusion hypothesis, or SGDH, relies on the similarity of the turbulence and heat transfer elds. The SGDH assumes that they are identical apart from a proportionality factor, the turbulent Prandtl number. The validity of this approximation in fully developed channel ow can be seen in Fig. 7. Using the SGDH approximation, it is also implied that the thermal eld is isotropic. With an RSM the individual Reynolds stresses are known however, and can be used to predict the turbulent anisotropic level in the ow. If it is assumed that the anisotropicity of the heat transfer is identical to that of the turbulence, a GGDH, or Generalized Gradient Diffusion Hypothesis, model can be employed for the heat transfer prediction. Although a GGDH would be more appropriate than a SGDH, the fundamental approximation of an identity between the ow eld and thermal eld is still used, and should be questioned. Similar to the turbulence eld, it is also possible to solve a set of equations for thermal quantities. Examples of this approach are the two, of the Nagano group, Abe et al. equation model, [1] and EAHF (Explicit Algebraic Heat Flux) models of So and Sommer [30] and Dol et al. [11]. Here, however, only the simple model based on the turbulent Prandtl number is used.

  

Rb r

(123)

(124)

0.5 0.45 0.4 0.35 0.3

5.3

New Blending Model

0.2

0.1 0.05 0 0 20 40 60 80 100

Figure 23: Coefcient, in log-law for temperature. Re-arranging, the linear law for temperature is given Model Eq. 132 and DNS-data, , [18]. (see also Eq. 38): Using a constant Prandtl number model, it is understood that all wall effects are embedded in the turbulence model, and hence any LRN modications to the heat transfer model are normally excluded. However, as shown in Fig. 7, the turbulent Prandtl number is also affected by the wall, and there thus exists a number of heat transfer models that include some wall modications, see e.g. Reynolds [28] and Kays [19].

The non-dimensionalised temperature in the blending model thus becomes:

see Fig. 23. The expression gives in the logaritmic layer, as indicated by experimental data [12, ) indicate 16, 20]. Note that the DNS-data ( a slightly lower value of . There seems to be a Reynolds number dependency for this coefcient, however, as another DNS [17] ( ) yields an even lower value, . Results obtained using this model are presented in Bredberg et al. [6].

(126)

where , and , [16]. The constant, , depends on the Prandtl number, which according to Jayatillaka [13] is:

Acknowledgments
Funding for the present work has been provided by STEM, the Volvo Aero Corporation and ALSTOM Power via the Swedish Gas Turbine Center.

(127)

19

  t 3 r

 f

f t 3 r  f t o ) 

 

) f t o

and for air ( ), see with experimental data in Kader and Yaglom [16]. The reader is also referred to Huang and Bradshaw [12] and Kays and Crawford [20] for a recent discussions on the topic. Kawamura et al. [18] provide valuable DNS-data for the temperature equation in fully developed channel ow. The above relation for the wall temperature is not commonly used, however, and is instead expressed with the normalised velocity, , as:

t ) o

  

(125)

f "3 r  x

) o

 f t ) o 

) o

The law-of-the-wall for the temperature was derived in section 3.2. From that the temperature prole in the inertial sub-layer is approximated with a logarithmic equation as given by Eq. 40:

) o

5.2 Wall Function Method

where is dened as for the turbulence eld, see Eq. 108. To achieve accurate results with this heat transfer model in the case of a rst node location in the buffer and must be modied. This layer, the values of is not unreasonable since constant values are valid only in the logarithmic layer. In the new model, we chose to modify , which is reduced in the buffer layer, as found from DNS computations [18] and [17]. Reasonable agreement with the predicted by DNS was found for the relation: (132)

acements

has not yet been given explicitly for the LRN region. Introducing normalised variables, and , Eq. 123 can be re-written as:

6 ) e } T ) o 6 (W 6 t i

w y v AX t x

x x v AX v 7X # (W w F !p w i # t ! p

! p y v AX x #

(W t i

0.15

' ) e (% ) o t

0.25

DNS New model

During the numerical optimisation of the new model it was found that the most appropriate blending parameter for the thermal eld was the normalised temperature, . The non-dimensionalised wall distance in Eq. 125, is however based on instead of the commonly used although less satisfying :

s w AX $ e $ vx

! ) ) s 2'% 0o t p ! V#

 f t (W i

) o o t x (W i

  t 3 r

V u% t s V s

i (W f  x (W  f f i

x (W i  x i v 9a (W 

) f t 0o

) o

 i W x (W ) t  (W i i  (W V x (W t i s i s

  

 f

  t ) 

   t  (W i

) o

(128)

(129)

(130)

(131)

References
[1] K. Abe, T. Kondoh, and Y. Nagano. A new turbulence model for predicting uid ow and heat transfer in separating and reattaching ows - I. ow eld calculations. Int. J. Heat and Mass Transfer, 37:139151, 1994. [2] R.S. Amano. Development of a turbulence nearwall model and its application to separated and reattached ows. Numerical Heat Transfer, 7:5975, 1984. [3] R.S. Amano, M.K. Jensen, and P. Goel. A numerical and experimental investigation of turbulent heat transport downstream from an abrupt pipe expansion. J. Heat Transfer, 105:862869, 1983. [4] J. Bredberg. Prediction of ow and heat transfer inside turbine blades using EARSM, and turbulence models. Thesis for the Degree of Licentiate of Engineering, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1999. Also available at www.tfd.chalmers.se/bredberg. [5] J. Bredberg and L. Davidson. Prediction of ow and heat transfer in a stationary 2-D rib roughened passage using low-Re turbulent models. In 3:rd European Conference on Turbomachinery, pages 963 972, London, 1999. IMechE. [6] J. Bredberg, S-H. Peng, and L. Davidson. On the wall boundary condition for computing heat transmodels. In J.H. Kim, editor, HTDfer with Vol. 366-5, ASME Heat Transfer Divison - 2000, volume 5, pages 243250, Orlando, 2000. The American Society of Mechanical Engineers. [7] C.C. Chieng and B.E. Launder. On the calculation of turbulent heat transfer transport downstream an abrupt pipe expansion. Numerical Heat Transfer, 3:189207, 1980. [8] M. Ciofalo and M. W. Collins. predictions of heat transfer in turbulent recirculating ows using an improved wall treatment. Numerical Heat Transfer, Part B, 15:2147, 1989. [9] F.H. Clauser. The turbulent boundary layer. Adv. Appl. Mech, 4:151, 1954. [10] L. Davidson and B. Farhanieh. CALC-BFC. Report 95/11, Dept. of Thermo and Fluid Dynamics, Chalmers University of Technology, Gothenburg, 1995. [11] H.S. Dol, K. Hanjali , and S. Kenjeres. A comparac tive assessment of the second-moment differential and algebraic models in turbulent natural convection. Int. J. Heat Fluid Flow, 18:414, 1997. [12] P.G. Huang and P. Bradshaw. Law of the wall for turbulent ows in pressure gradients. AIAA Journal, 33:624632, 1995.

[13] C.L.V. Jayatillaka. The inuence of Prandtl number and surface roughness on the resistance of the laminar sublayer to momentum and heat transfer. Progress in Heat and Mass Transfer, 1:193, 1969. [14] R.W. Johnson and B.E. Launder. Discussion of: On the calculation of turbulent heat transfer transport downstream an abrupt pipe expansion. Numerical Heat Transfer, 5:493496, 1982. [15] W.P. Jones and B.E. Launder. The prediction of laminarization with a two-equation model of turbulence. Int. J. Heat and Mass Transfer, 15:301314, 1972. [16] B.A. Kader and A.M. Yaglom. Heat and mass transfer laws for fully turbulent wall ows. Int. J. Heat and Mass Transfer, 15:23292351, 1972. [17] N. Kasagi, Y. Tomita, and A. Kuroda. Direct numerical simulation of passive scalar in a turbulent channel ow. J. Heat Transfer, 114:598606, 1992. [18] H. Kawamura, H. Abe, and Y. Matsuo. DNS of turbulent heat transfer in channel ow with respect to Reynolds and Prandtl number effects. Int. J. Heat and Fluid Flow, 20:196207, 1999. [19] W.M. Kays. Turbulent Prandtl number - where are we? J. Heat Transfer, 116:284295, 1994. [20] W.M. Kays and M.E. Crawford. Convective Heat and Mass Transfer. McGraw-Hill, Inc, 1993. [21] C.K.G. Lam and K.A. Bremhorst. Modied form of -model for predicting wall turbulence. J. the Fluids Engineering, 103:456460, 1981. [22] B.E. Launder. Numerical computation of convective heat transfer in complex turbulent ows: Time to abandon wall functions? Int. J. Heat and Mass Transfer, 27:14851491, 1984. [23] B.E. Launder. On the computation of convective heat transfer in complex turbulent ows. J. Heat Transfer, 110:11121128, 1988. [24] B.E. Launder and D.B. Spalding. The numerical computation of turbulent ows. Computational Methods Appl. Mech. Eng., 3:269289, 1974. [25] R.D. Moser, J. Kim, and N.N. Mansour. Direct numerical simulation of turbulent channel ow up to Re=590. Physics of Fluids, 11:943945, 1999. Data available at www.tam.uiuc.edu/Faculty/Moser/. [26] V.C. Patel, W. Rodi, and G. Scheuerer. Turbulence models for near-wall and low Reynolds number ows: A review. AIAA Journal, 23:13081319, 1985. [27] L. Prandtl. Zur turbulenten stromung in roh ren und langs platten. Ergeb. Aerod. Versuch, G ttingen, IV Lieferung, 18, 1932. o [28] A.J. Reynolds. The prediction of turbulent Prandtl and Schmidt numbers. Int. J. Heat and Mass Transfer, 18:10551069, 1975.

20

[29] A. Sahay and K.R. Sreenivasan. The wall-normal position in pipe and channel ows at which viscous and turbulent shear stresses are equal. Physics of Fluids, 11:31863188, 1999. [30] R.M.C So and T.P. Sommer. An explicit algebraic heat-ux model for the temperature eld. Int. J. Heat and Mass Transfer, 39:455465, 1996. [31] H. Tennekes and J.L. Lumley. A First Course in Turbulence. Massachusetts Institute of Technology, Cambridge, 1972. [32] A.A. Townsend. The structure of turbulent shear ow. Cambridge University Press, Cambridge, 1976. [33] T. von Karman. Mechanishe ahnlichkeit un turbulenz. Nachr. Ges. Wiss., G ttingen, pp 68, 1930. o [34] D.C. Wilcox. Reassessment of the scaledetermining equation for advanced turbulence models. AIAA Journal, 26:12991310, 1988. [35] D.C. Wilcox. Comparison of two-equation turbulence models for boundary layers with pressure gradient. AIAA Journal, 31:14141421, 1993. [36] D.C. Wilcox. Turbulence Modeling for CFD. DCW Industries, Inc., 1993.

21

You might also like