Sie sind auf Seite 1von 17

5.

TURBULENCE MODELLING SPRING 2008 (Revised)

5.1 Notation
5.2 Effect of turbulence on the mean flow
5.3 Eddy-viscosity models
5.4 Reynolds-stress transport models (second-order closure)
5.5 Non-linear eddy-viscosity models
5.6 References
Appendix: Derivation of the Reynolds-stress transport equations
Examples

5.1 Notation

Index notation will be used where appropriate:


x ≡ ( xi ) ≡ ( x1 , x 2 , x3 ) ≡ ( x, y, z )
u = (u i ) ≡ (u1 , u 2 , u 3 ) ≡ (u , v, w)

The Einstein summation convention (implicit summation over a repeated index unless told
otherwise) is widely used e.g.
∂U i ∂U 1 ∂U 2 ∂U 3
means + +
∂xi ∂x1 ∂x 2 ∂x3

Turbulent quantities are split into “instantaneous = mean + fluctuation” denoted by either:
u = u + u′ (“overbar/prime” notation)
or
u * = U + u (“upper-case/lower-case” notation)

The following basic rules of averaging apply:


φ′ = 0
φ 2 = φ 2 + φ′ 2 (variance)
φ =φ + φ′ ′ (covariance)

5.2 Effect of Turbulence on the Mean Flow

The Navier-Stokes equations for the mean flow include the net transport of momentum by
turbulent fluctuations.

For an incompressible fluid, the momentum balance (per unit volume) gives
DU i ∂P ∂ ij
= − + + Fi
Dt ∂xi ∂x j (1)
mass × acceleration forces
Fi means “any other forces” (e.g. buoyancy forces; Coriolis forces). There is an implied
summation over a repeated subscript (here, j).

Turbulent Boundary Layers 5-1 David Apsley


τ22
τ12
The stress component ij is the effective force per unit area
• in the xi direction y τ21
• on an area normal to the xj direction τ11 x τ 11
• exerted by fluid in xj+ on that in xj– τ 21

The various components acting on a rectangular fluid element in τ12


the x-y plane are shown right.
τ 22
The stress tensor is composed of two parts – viscous forces due to molecular viscosity and net
transport of momentum due to random turbulent fluctuations:
∂U ∂U j
= ( i + ) − ui u j
∂x j ∂xi
ij
(2)
viscous turbulent

In the boundary-layer context (∂/∂x ∂/∂y) the only dynamically-significant component is


∂U
≡ 12 = − uv (3)
∂y
because this is the only component of stress in the area A
dominant x1 direction whose magnitude differs on the
opposite sides of a fluid element and, therefore, gives rise
τ
y+ δ y1
2

to a net force. The net force in the x1 direction due to this


component acting on top and bottom of a block, area A and
height y is (x,y)
δy

A − y− 1 y A = A y
y + 12 y 2 ∂y
Dividing by the volume A y, the net force per unit volume
is

.
τ
y- δ y 1
2

∂y

Objective in turbulence modelling:


model the Reynolds stresses − u i u j in order to close the mean-flow equations.

One route to a turbulence model:


(1) Deduce − uv from theory for an equilibrium boundary layer.
(2) Then make the model tensorially-invariant to apply it to an arbitrary flow.

Second route to a turbulence model:


(1) Concoct a tensorially-correct formulation by inspired physics.
(2) Calibrate a few parameters by looking at the special case of a boundary layer.

This course will consider examples of both approaches.

Turbulent Boundary Layers 5-2 David Apsley


5.3 Eddy-Viscosity Models

5.3.1 Basis

In the overlap region of a turbulent boundary layer the direct effects of viscosity are
unimportant and the following approximations are valid:
∂U u
dimensional analysis log law : = ∂U
∂y y = u y
∂y
zero pressure gradient constant stress : = u2
Hence,
∂U
= t (stress ∝ rate of strain) (4)
∂y
where
t = t
and, in the log-law region,
t = u y (5)

(4) can be made tensorially-invariant with an appropriate trace:


∂U i ∂U j 2
( turb )
= ( + ) − k ij (6)
∂x j ∂xi
ij t
3

This is called an eddy viscosity model. There is an obvious analogy with viscous stresses. t
and t are, respectively, the dynamic and kinematic eddy viscosities.

The total stress is the sum of viscous and turbulent parts. So,
2 ∂U ∂U j
ij + k ij = eff ( i + )
3 ∂x j ∂xi
where the effective viscosity is the sum of molecular and turbulent parts:
eff = + t
At high Reynolds numbers, t throughout most of the boundary layer.

5.3.2 Assessment

Advantages
• Easy to implement in existing solvers – simply use a variable viscosity.
• Additional viscosity helps to stabilise the solution algorithm.
• Theoretical foundation in simple, but common, shear flows.

Disadvantages

• Merely a model – its basis is questionable in complex flows.


• Reduces turbulent transport to the specification of a single scalar t. Hence, at most
one Reynolds stress ( − uv ) can be represented accurately.

Turbulent Boundary Layers 5-3 David Apsley


5.3.3 Gradient Transport For Other Properties

Besides momentum, other physical quantities are advected (i.e. transported with the fluid
flow). These include enthalpy, dissolved salts and pollutants.
A
The instantantaneous rate of transport by advection across an area A of any ρ uA θ
quantity with concentration per unit mass is
( u i Ai )
Averaging gives the net rate of transport per unit area:
u i = u i + u i′ ′ (7)

Turbulent fluctuations contribute a net flux u i′ ′ in the same manner as for momentum and
it is appropriate, therefore, to make Reynolds’ analogy and apply a gradient-transport model:

− u i′ ′ = t (8)
∂xi
The eddy diffusivity t is assumed to be proportional to the eddy viscosity:

t = t
(9)


is a called a turbulent Prandtl number and its value is often taken as approximately 1.0.

5.3.4 Mixing-Length Models

Original Idea (Prandtl, 1925)


∂U ∂U
u ~ lx , v ~ ly
∂y ∂y
where lx and ly are typical sizes of turbulent eddy. Then
2
∂U
− uv = l 2
(10)
∂y
m

where the mixing length lm is the geometric mean (l x l y )1 / 2 , absorbing any constants of
proportionality. This corresponds to an eddy-viscosity relationship with
2 ∂U
t = lm (11)
∂y

Alternative Viewpoint y
U(y)
The kinematic eddy viscosity has dimensions of (velocity)×(length),
suggesting that it be modelled as
t = u0lm
where lm is a typical size of turbulent eddy and u0 is a turbulent velocity
lm
scale. The mixing-length hypothesis assumes
∂U lm dU
u0 = lm (12) dy
∂y
which is the difference in mean velocity over a vertical displacement lm.

Turbulent Boundary Layers 5-4 David Apsley


Log-Law Region

In the log-law region,


∂U u
= u y and =


∂y
t
y
Eliminating uτ gives
2 ∂U
t = ( y) (13)
∂y
and hence, comparison with (11) gives
lm = y (14)
i.e. the mixing length is proportional to distance from the wall.

Near-Wall Behaviour

Near a solid boundary, lm must vanish faster than y. A popular modification is van Driest’s
(1956) damping factor:
y+
l m = y[1 − exp(− )], ( A = 26) (15)
A
(This gives l m ∝ y 2 as y → 0).

Outer Layer

The mixing length cannot grow indefinitely since turbulent eddies are clearly smaller than the
boundary-layer depth . Common practice is to apply an upper bound which is a fraction of ;
e.g.
l m = min( y,0.09 ) (16)

5.3.5 Two-Equation Turbulence Models

The mixing-length formulation is geometry-dependent and unsatisfactory in complex flows.


It is desirable to base the expression for t on scales of turbulent motion which are free of
geometric constraints. On dimensional grounds,
[ t] = [velocity] × [length] or [velocity]2 × [time]

An obvious candidate for the velocity scale is k , where k = 12 (u 2 + v 2 + w 2 ) is the turbulent


kinetic energy (per unit mass).

An independent dimensional scale with physical meaning is the rate of dissipation of


turbulent kinetic energy, – but there are plenty of alternatives. The quantity k/ is a measure
of eddy turnover time (indicating how fast turbulence is dissipated by viscosity) and is called
the turbulence timescale.

Since [k] = [velocity]2 and [k/ ] = [time], any kinematic eddy viscosity constructed from k
and must be of the form

Turbulent Boundary Layers 5-5 David Apsley


k2
t = constant ×
The constant is derived from equilibrium boundary-layer properties. Values of k and are
determined by solving modelled transport equations.

Derivation of the k- Eddy-Viscosity Formula

In the overlap region of a turbulent boundary layer:


∂U u
(A) Log law (overlap region): =


∂y y

(B) High-Re; constant stress: − uv = w


= u2


(C) Local equilibrium: P (k ) = (i.e. production = dissipation).

(D) Structure functions (i.e. ratios of any pair of stresses) are constant; in particular,
(−uv)
= constant (17)
k

The rate of production of turbulent kinetic energy (per unit mass) for a simple shear flow is
(see the Appendix for a derivation):
∂U
P ( k ) = −uv (18)
∂y
Substitution for ∂U/∂y and − uv from assumptions (A) and (B) respectively gives
u3
= P (k ) = (19)


y
Assumptions (A) and (B) lead to an eddy viscosity
− uv
t = = u y
∂U/∂y


Eliminating the geometric height y gives


u 4 (−uv) 2
t = =


Finally, using the assumption (D) (consistent with the constant-stress assumption) that, in this
layer, the structure parameter
(−uv)
= constant , C 1 / 2 say
k
with a measured value of ∼0.3. The eddy-viscosity relation may then be written in terms of k
rather than − uv ; i.e.
k2
t =C (20)

where
C = 0.09
 (21)

Turbulent Boundary Layers 5-6 David Apsley


Modelled transport equations

An exact equation for k can be derived (see the Appendix). Once the eddy-viscosity formula
has been applied, only the turbulent diffusion term needs modelling, and this is usually done
in the form of a gradient-transport hypothesis (diffusion ∝ mean-flow gradient). An
equation can be formally written down, but most terms require modelling. The “standard” k
and transport equations are of the form (in the absence of body forces):
Dk ∂ ∂k
= [( + ( kt ) ) ] + P (k ) − (22)
Dt ∂xi ∂xi
D ∂ ∂
= [( + t
) ] + (C 1 P ( k ) − C ) (23)
Dt ∂xi ∂xi
( ) 2
k
where
∂U i
P ( k ) = −u i u j (rate of production of k) (24)
∂x j

Constants vary somewhat between authors but a moderately standard set is


C = 0.09, C 1 = 1.44, C 2 = 1.92, (k) = 1, ( ) = 1.3 (25)



 

To be consistent with the log law region ( u 2 /k = C 1 / 2 , P ( k ) = = u 3 / y ) these should (but


  

don’t quite) satisfy


(C 2 − C 1 ) ( ) C = 2


 

(26)

In a slowly-developing boundary layer, the LHS of both (22) and (23) is essentially zero,
whilst for any simple shear flow the production term simplifies considerably to
∂U ∂U 2
P ( k ) = −uv = t( ) (27)
∂y ∂y

5.3.6 Boundary Conditions

Near a solid boundary:


• gradients in the direction normal to the boundary are very large (numerical schemes
need many nodes to resolve the flow);
• molecular viscosity becomes important (log-law assumption inaccurate);
• different stress components u i u j behave very differently.

Numerical schemes adopt two contrasting approaches.

High-Reynolds-Number Turbulence Models

• Bridge the near-wall region with wall functions; i.e. assume profiles (based on
equilibrium boundary-layer theory) between near-wall node and boundary.
• OK if the equilibrium assumption is reasonable (e.g. slowly-developing boundary
layers); dodgy in highly non-equilibrium regions (particularly near separation points).
• The near-wall node should optimally be placed in the region 30 < y+ < 50
(pragmatism usually expands this to 15 –150). This means that numerical meshes

Turbulent Boundary Layers 5-7 David Apsley


cannot be arbitrarily refined close to solid boundaries.

Low-Reynolds-Number Turbulence Models


• Resolve the flow right up to the boundary.
• Include effects of molecular viscosity in the coefficients of the turbulence model (C , 

C 1, C 2 and possibly extra terms on the RHS of the equation).


 

• Try to ensure the correct near-wall behaviour:


2 k
k ∝ y2, ~ 2 ~ constant , t ∝ y
3
( y → 0) (28)
y
• Correct resolution of the flow requires a numerical simulation to place a near-wall
node at y+ ≤ 1. This can be very computationally intensive, particularly for high-speed
flows.

5.3.7 Choice of Turbulence Scalars

The turbulent kinetic energy k is a very common choice of turbulence scale, but there are
many possibilities for the “length-scale-determining” second variable:
k2
k- t ∝ Launder and Spalding, 1974; Launder and Sharma, 1974
k
k- t ∝ Wilcox, 1988, 1998; Menter, 1994
k- t ∝k
t ∝ k
1/ 2
k-l l
and many variants of models with nominally the same transport variables.

5.4 Reynolds-Stress Transport Models (Second-Order Closure)

By subtracting the mean from the instantaneous momentum equations the differential
equations governing the behaviour of the individual stresses u i u j can be derived (see the
Appendix). Solution of (modelled forms of) these is called second-order closure (SOC) or
Reynolds-stress transport modelling (RSTM) or differential stress modelling (DSM).

These equations each take the form (in the absence of body forces):
D ∂
(u i u j ) = d ijk + Pij + ij − ij (29)
Dt ∂x k
where
D ∂ ∂
(u i u j ) = ( + U k )(u i u j ) = advection (rate of change following mean flow)
Dt ∂t ∂x k
∂U j ∂U i
Pij = − u i u k + u j uk = production (by the mean flow)
∂x k ∂x k
ij = redistribution by pressure fluctuations
ij = dissipation by viscous action
dijk = diffusion by molecular viscosity and turbulence

Turbulent Boundary Layers 5-8 David Apsley


For the precise forms of these last three terms see the Appendix.

Advantages

• The important production and advection terms are exact and do not require modelling.

Disadvantages

• The redistribution, diffusion and dissipation terms are not exact and require extensive
modelling.
• Computational expense – 6 additional stress-transport equations in 3 dimensions.
• Numerical instability – absence of an extra stabilising diffusive term.

In simple shear flows (∂U/∂y the only non-zero velocity derivative) the production terms for
the normal stresses ( u 2 , v 2 and w 2 ) reduce to
∂U Turbulent Stresses
P11 = −2uv , P22 = P33 = 0 y+
∂y 400

Since − uv is usually positive in this type of flow, energy is 300

preferentially put into the u 2 component. Add to this the


200
preferential damping of the wall-normal component v 2 by
the presence of a solid boundary and it is easy to see why
100
uu+
the normal stresses are anisotropic with u 2 > w 2 > v 2 for a uv+ vv+
ww+
boundary layer along y = 0. 0
-2 0 2 4 6 8

Standard Model

Accepting the notion that the purpose of the pressure-strain correlation is to reduce
anisotropy, that dissipation takes place at small scales (where velocity gradients are large and
the eddies isotropic) and that diffusion is a response to mean gradients, a typical model for
these terms is of the form:
ui u j 2
ij = −C1 ( − 3 ij ) − C 2 ( Pij − 13 Pkk ij ) + wall − reflection terms (30)
k
The first two terms are, respectively, the return-to-isotropy and isotropisation
of production (IP) terms. The wall-reflection terms are necessary to ensure
that the correct levels of anisotropy are retained at solid boundaries.
ij = 3
2
ij (31)
u k u l k ∂ (u i u j )
d ijk = + Cs (32)
∂xl
kl

The turbulent part is called the generalised gradient diffusion hypothesis (GGDH).

Turbulent Boundary Layers 5-9 David Apsley


5.5 Non-Linear Eddy-Viscosity Models

These represent a “half-way house” between linear eddy-viscosity models and second-order
closure which can accommodate anisotropy but through a formulation based on mean-flow
gradients and not having to solve extra transport equations. They mimic the response of
turbulence to complex strains by using a non-linear constitutive relation between turbulent
stress and mean rate of strain: i.e. roughly of form:
stress = C1 (rate of strain) + C 2 (rate of strain) 2 + C 3 (rate of strain) 3 + ...

Notation

The mean-flow gradient ∂U i /∂x j is usually split into symmetric and antisymmetric parts
called the mean strain and vorticity tensors, respectively:
1 ∂U ∂U j 1 ∂U i ∂U j
S ij = ( i + ), ij = ( − ) (33)
2 ∂x j ∂xi 2 ∂x j ∂xi
The magnitudes of these tensors are represented by
S = (2S ij S ij )1 / 2 , = (2 ij ij )1 / 2 (34)
and in a simple shear flow these reduce to
∂U
S= = (35)
∂y

Turbulent stress is conveniently re-expressed in terms of the dimensionless anisotropy tensor:


ui u j 2
aij = − ij (36)
k 3

Since velocity gradients have dimensions 1/[time], a general relationship between stress and
mean strain takes the form
k k
a = f ( S, ) (37)

If the RHS is to be a symmetric and traceless tensor function of S and , then, in general, it
can be written as a linear combination of 10 bases (Pope, 1975). However, since the
symmetric, traceless 3×3 anisotropy tensor has only 5 independent components, a smaller
subset is sufficient and experience at this university is that a cubic eddy-viscosity model
allows successful inclusion of normal-stress anisotropy and streamline curvature effects on
turbulence. One (but far from the only!) canonical way of writing this is
k
a = − ( )2C S 

k
[
+ ( )2 1 (S 2 − 13 {S 2 }I) + 2 ( S−S )+ 3 ( 2
− 13 { 2
}I) ]
−( ) [ ]
k 3
1 {S 2 }S + 2 { 2
}S + 3 ( 2
S+S 2
−{ 2
}S − 23 { S }I) + 4 ( S2 − S2 )
(38)
where successive lines give linear, quadratic, cubic terms; matrix notation is used to simplify
writing down contracted tensor products, {} denotes a trace (sum of diagonal elements) and I
is the identity matrix.

Turbulent Boundary Layers 5 - 10 David Apsley


The coefficients C , i and i vary significantly between models (see, e.g., Apsley et al.,


1997) but important general points arise from this form of constitutive relation.

(1) The linear term corresponds to the usual linear eddy-viscosity model:
k k 2 ∂U i ∂U j
a = −2C S ⇔ − u i u j = (C )( + ) − 23 ij k
∂x j ∂xi

(2) The quadratic terms are associated with normal-stress anisotropy. e.g. in a simple shear
flow (∂U/∂y the only non-zero velocity gradient) the expression gives
u2 2 2
= +( 1 +6 2 − 3)
k 3 12
2 2
v 2
= +( 1 −6 2 − 3) (39)
k 3 12
w2 2 2
= −( 1 − 3)
k 3 6
where
k ∂U
= (40)
∂y
2
whereas a linear model would set all normal stresses equal to 3 k in this type of flow.

(3) In a 2-dimensional incompressible flow the 3- and 4-related


terms vanish and the
∂U i
quadratic terms do not contribute to the production of turbulence energy ( P ( k ) = −u i u j ).
∂x j

(4) The 1- and 2-related terms evoke a sensitivity to curvature since, in a curved shear flow
∂U ∂U s ∂V U
(with = , = − s , Rc the local radius of curvature),
∂y ∂R ∂x Rc
∂U s U s
{S 2 } + { 2 } ≡ 12 ( S 2 − 2 ) = −2 (41)
∂R Rc
and the curvature is known to be stabilising
(reducing turbulence) if Us increases in the
direction away from the centre of curvature
(∂Us/∂R > 0) and destabilising (increasing
turbulence) if Us decreases in the direction away
from the centre of curvature (∂Us/∂R < 0). In the
form (38) the response is correct if 1 and 2 are stable unstable
both positive.

Turbulent Boundary Layers 5 - 11 David Apsley


5.6 References

Apsley, D.D., Chen, W.-L., Leschziner, M.A. and Lien, F.-S., 1997, Non-linear eddy-
viscosity modelling of separated flows, J. Hydraulic Res., 35, 723-748.
Craft, T.J., 1998, Developments in a low-Reynolds-number second-moment closure and its
application to separating and reattaching flows, Internat. J. Heat Fluid Flow, 19, 541-
548.
Craft, T.J., Launder, B.E. and Suga, K., 1996, Development and application of a cubic eddy-
viscosity model of turbulence, Int. J. Heat Fluid Flow, 17, 108-115.
Gibson, M.M. and Launder, B.E., 1978, Ground effects on pressure fluctuations in the
atmospheric boundary layer, J. Fluid Mech., 86, 491-511
Hanjalic, K., 1999, Second-moment turbulence closures for CFD: needs and prospects,
International Journal of Computational Fluid Dynamics, 12, 67-97
Hanjalic, K., Jakirlic, S. and Hadzic, I., 1997, Expanding the limits of "equilibrium" second-
moment turbulence closures, Fluid Dynamics Res., 20, 25-41
Launder, B.E., Reece, G.J. and Rodi, W., 1975, Progress in the development of a Reynolds-
stress turbulence closure, J. Fluid Mech., 68, 537-566
Launder, B.E. and Sharma, B.I., 1974, Application of the energy-dissipation model of
turbulence to the calculation of flow near a spinning disc, Letters in Heat and Mass
Transfer, 1, 131-138
Launder, B.E. and Spalding, D.B., 1974, The numerical computation of turbulent flows,
Computer Meth. Appl. Mech. Eng., 3, 269-289
Menter, F.R., 1994, Two-Equation Eddy-Viscosity Turbulence Models for Engineering
Applications, AIAA J., 32, 1598-1605
Pope, S.B., 1975, A more general effective-viscosity hypothesis, J. Fluid Mech., 72, 331-340
Wilcox, D.C., 1988, Reassessment of the scale-determining equation for advanced turbulence
models, AIAA J., 26, 1299-1310
Wilcox, D.C., 1998, Turbulence modelling for CFD, 2nd Edition, DCW Industries

Turbulent Boundary Layers 5 - 12 David Apsley


Appendix: Derivation of the Reynolds-Stress Transport Equations
(For Reference and Hangover-Cure Only!)

Restrict to constant-density fluids. fi are the components of problem-dependent body forces


(buoyancy, Coriolis forces, ...). Use prime/overbar notation (e.g. u = u + u ′ ) and summation
convention throughout.

Continuity
∂u k
Instantaneous: =0 (A1)
∂x k
∂u k
Average: =0 (A2)
∂x k
∂u k′
Subtract: =0 (A3)
∂x k

Result (I): Both mean and fluctuating quantities satisfy the incompressibility equation.

Momentum

∂u i ∂u 1 ∂p ∂ 2ui
Instantaneous: + uk i = − + + fi (A4)
∂t ∂x k ∂xi ∂x k ∂x k
∂u i ∂u ∂u ′ 1 ∂p ∂ 2ui
Average: + u k i + u ′k i = − + + fi (A5)
∂t ∂x k ∂x k ∂xi ∂x k ∂x k
Du i 1 ∂p ∂ ∂u i
Rearrange: =− + ( − u i′u ′k ) + f i (A6)
Dt ∂xi ∂x k ∂x k
(Note: u ′k can be “taken through” the ∂/∂xk derivative whenever required, due to the
incompressibility condition)

Result (II): Mean flow equation is the same as the instantaneous equation except for
additional apparent stresses − u i′u ′j . Note: this is actually the net transport of momentum by
turbulent fluctuations, not a force.

∂u i′ ∂u i′ ∂u i ∂u i′ ∂u i′ 1 ∂p ′ ∂ 2 u i′
Subtract: + uk ′
+ uk ′
+ uk ′
− uk =− + + f i′ (A7)i
∂t ∂x k ∂x k ∂x k ∂x k ∂xi ∂x k ∂x k
∂u ′j ∂u ′j ∂u j ∂u ′j ∂u ′j 1 ∂p ′ ∂ 2 u ′j
Similarly j: + uk + u ′k + u ′k − u ′k =− + + f j′ (A7)j
∂t ∂x k ∂x k ∂x k ∂x k ∂x j ∂x k ∂x k

Form u i′ (A7) j + u ′j (A7) i :

Turbulent Boundary Layers 5 - 13 David Apsley


∂ ∂ ∂u j ∂u ∂
(u i′u ′j ) + u k (u i′u ′j ) + u i′u k′ + u ′j u ′k i + (u i′u ′j u ′k )
∂t ∂x k ∂x k ∂x k ∂x k
∂p ′1 ∂p ′ ∂ 2 u ′j ∂ 2 u i′
= − (u i′ + u ′j ) + (u i′ + u ′j ) + (u i′ f j′ + u ′j f i′)
∂x j ∂xi ∂x k ∂x k ∂x k ∂x k
Rewrite the pressure terms and rearrange:
∂ ∂ ∂ ∂ 1
(u i′u ′j ) + u k (u i′u ′j ) = [ (u i′u ′j ) − p ′(u i′ jk + u ′j ik ) − u i′u ′j u ′k ]
∂t ∂x k ∂x k ∂x k
∂u j ∂u i p ′ ∂u ′ ∂u ′j ∂u i′ ∂u ′j
− (u i′u ′k + u ′j u k′ ) + (u i′ f j′ + u ′j f i′) + ( i + )−
∂x k ∂x k ∂x j ∂xi ∂x k ∂x k

Result (III): Reynolds-stress transport equation:


D ∂d ijk
(u i′u ′j ) = + Pij + Fij + ij − ij (A8)
Dt ∂x k
where:
D ∂ ∂
(u i′u ′j ) = (u i′u ′j ) + u k (u i′u ′j ) advection (by mean flow)
Dt ∂t ∂x k
∂u j ∂u
Pij = −(u i′u k′ + u ′j u k′ i ) production (by mean strain)
∂x k ∂x k
Fij = u i′ f j′ + u ′j f i′ production (by body forces)
p ′ ∂u i′ ∂u ′j
= ( + ) pressure-strain correlation
∂x j ∂xi
ij

∂u i′ ∂u ′j
= dissipation
∂x k ∂x k
ij

∂ (u i′u ′j ) 1
d ijk = − p ′(u i′ + u ′j ) − u i′u ′j u k′ diffusion
∂x k
jk ik

Contract (A8), then divide by 2. Change subscript k to an i to minimise confusion.

Result (IV): turbulent kinetic energy equation:


Dk ∂d i( k )
= + P (k ) + F (k ) − (A9)
Dt ∂xi
where
∂u
P ( k ) = −u i′u ′j i production (by mean strain)
∂x j
F ( k ) = u i′ f i′ production (by body forces)
∂u i′ 2
= ( ) dissipation
∂x j
∂k 1
d i( k ) = − p ′u i′ − 12 u ′j u ′j u i′ diffusion
∂xi

Turbulent Boundary Layers 5 - 14 David Apsley


Examples

Question 1. (Summation convention)


∂U i ∂U 1 ∂U 2 ∂U 3 ∂U ∂V ∂W
is shorthand for + + or + + . Expand the following in
∂xi ∂x1 ∂x 2 ∂x3 ∂x ∂y ∂z
similar fashion.
DU i ∂U i ∂U i
(a) (= +Uk ) when i = 1 and when i = 2.
Dt ∂t ∂x k
∂U i
(b) P ( k ) = −u i u j
∂x j
∂U j∂U i
(c) Pij = −(u i u k + u j uk
) when i = 1, j = 1 and when i = 1, j = 2.
∂x k ∂x k
(d) For a general matrix M, what quantities are represented by Mii, MijMji and MijMjkMki?

Question 2.
What do the following quantities reduce to in a simple shear flow (a flow in which ∂U/∂y is
the only non-zero velocity component)?
(a) P(k) (use the definition in Question 1 above)
(b) Pij (for each combination of i and j; use the definition in Question 1 above)
(c) S = (2S ij S ij )1 / 2
(d) = (2 ij ij )1 / 2

Question 3.
The rate of production of turbulent kinetic energy per unit mass is given by
∂U i
P ( k ) = −u i u j
∂x j
Show that, in incompressible flow,
(a) only the symmetric part Sij of the mean velocity gradient affects production;
(b) only the anisotropic part of the turbulent stress tensor kaij = u i u j − 23 k ij affects
production;
(c) with the eddy-viscosity hypothesis, P(k) is greater than or equal to zero.

Question 4.
(a) If the mean velocity profile in the fully-turbulent region is
1
U + = ln y + + B
and the shear stress is assumed to be constant, deduce an expression for the kinematic
eddy viscosity t.
(b) If the velocity profile is as above, but the shear stress varies as
= w (1 − y/ )
deduce an expression for t.

Turbulent Boundary Layers 5 - 15 David Apsley


Question 5.
Prove that for the high-Re k and transport equations to be consistent with mean and
turbulent profile in the log law region the constants should satisfy
(C 2 − C 1 )C 1 / 2 ( ) = 2

Question 6.
Show that, near the wall, the shear stress in a slowly-developing boundary layer varies with y
according to
dP
= w + e y + ...
dx

Question 7.
For fully-developed channel flow (depth 2 ), show that the Reynolds shear stress can be
written
y ∂U
− uv = w (1 − ) − S (where S = )
∂y
and hence that the rate of production of turbulent kinetic energy is
y
P ( k ) = w (1 − ) S − S 2

Question 8.
The shear-stress distribution in Couette flow (flow between parallel walls, one stationary and
the other moving with speed U0) is given by

=0
∂y

If molecular viscosity can be neglected and the shear stress is given by a mixing-length eddy-
viscosity model with
y
l m = y (1 − )
h
find the mean-velocity profile as a function of u and y+. (You may assume that the profile


approaches the standard log law as either wall is approached.)

Question 9.
By expanding the fluctuating velocities in the form
u = a1 + b1 y + c1 y 2 +
v = a 2 + b2 y + c 2 y 2 +
w = a 3 + b3 y + c3 y 2 +
show that u 2 = b12 y 2 + and derive similar expressions for v 2 , w 2 , uv , k and t.

Turbulent Boundary Layers 5 - 16 David Apsley


Question 10.
Use the turbulent kinetic energy equation and the near-wall behaviour of k from above to
show that the near-wall behaviour of is
2 k
~ 2 ~ constant ( y → 0)
y
(Note that this implies that some modification is required in the equation as y → 0).

Question 11. (Non-linear eddy-viscosity models)


Prove that, with the non-linear constitutive relation given in Section 5.6,
(a) the 3- and 4-related terms vanish in 2-d incompressible flow;
(b) the quadratic terms do not contribute to turbulence production in 2-d incompressible
flow;
(c) the normal stresses in a simple shear flow are given by
u2 2 2
= +( 1 +6 2 − 3)
k 3 12
2 2
v 2
= +( 1 −6 2 − 3)
k 3 12
2 2
w 2
= −( 1 − 3 )
k 3 6
k ∂U
where = .
∂y

Question 12.
Show that at each point in a curved shear flow there is a streamwise-oriented coordinate
system with
∂U ∂U s ∂V U
= , =− s
∂y ∂R ∂x Rc
where Us is the velocity magnitude and Rc the radius of curvature and hence show that
∂U s U s
{S 2 } + { 2 } = 12 ( S 2 − 2 ) = −2
∂R Rc
Note that {} denotes the trace (sum of diagonal elements) of a matrix.

Turbulent Boundary Layers 5 - 17 David Apsley

Das könnte Ihnen auch gefallen