Lecture Notes in Turbulence
Lecture Notes in Turbulence
Steve Berg
1 Conservation Equations 2
1.1 Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Conservation of Moment . . . . . . . . . . . . . . . . . . . . . . 4
2 Turbulent Theory 9
2.1 What is Turbulence ? . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Turbulence Spectrum . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Correlation in Turbulence . . . . . . . . . . . . . . . . . . . . . 16
2.4 Turbulence Equations . . . . . . . . . . . . . . . . . . . . . . . 23
3 Turbulence Models 26
3.1 Zero Equation Model . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 One Equation Model . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Two Equation Model . . . . . . . . . . . . . . . . . . . . . . . . 29
3.4 Reynolds Stress Model (RSM) . . . . . . . . . . . . . . . . . . . 31
4 Summary 36
A Turbulence Pictures 38
B Examination Questions 41
1
Chapter 1
Conservation Equations
Transport equations are the basics for all CFD simulations. Before we discuss
turbulence models we will derive these equations for laminar flow. Turbulence
will be added in chapter 2.4. This chapter is based on lectures notes from
Mathiesen (2000) and as an alternative you can also read chapter 6 in Munson
et al. (2002).
kg
mx = ρux A [ ]
s
This equation is used for all directions for calculation of mass flow. In our
case we will only derive the equation of mass for two direction (x and y). It is
easier to understand the procedures for the derivation of the equations when
there are a few variables as possible. The third flow direction part (z-direction)
can easily be added to the two direction equation.
2
(ρuy )|y+∆y ∆x∆z
6
y + ∆y
y
6
x x + ∆x
(ρuy )|y ∆x∆z
∂ρ ∂
∆V = (ρux )∆x∆y∆z
∂t ∂t
Mass flow in and out of the control volume is then given by:
(ρux ) |x ∆y∆z +(ρuy ) |y ∆x∆z −(ρux ) |x+∆x ∆y∆z −(ρuy ) |y+∆y ∆x∆z
| {z } | {z } | {z } | {z }
inf low−x inf low−y outf low−x outf low−y
We then put the inflow and outflow for x directed flow in the same equation,
and do the same for the y directed flow:
3
∂ρ {(ρux ) |x+∆x −(ρux ) |x } {(ρuy ) |y+∆y −(ρuy ) |y }
∆x∆y∆z = − ∆y∆x∆z− ∆y∆x∆
∂t ∆x ∆y
∂ρ ∂ ∂
= − (ρux ) − (ρuy )
∂t ∂x ∂y
If we re-arrange this equation and add the last term for the flow in z-
direction, we will end up with the final expression for mass conservation in 3
dimensions.
∂ρ ∂ ∂ ∂
+ (ρux ) + (ρuy ) + (ρuz ) = 0
∂t
|{z} |∂x {z } ∂y |∂z {z }
| {z }
Accumulation convection−x convection−y convection−z
y + ∆y
y
6
x x + ∆x
(ρux )uy |y ∆x∆z
4
Accumulations of Incoming impuls Outgoing impuls Sum of all forces
impuls per = per time unit - per time unit + on the system
time unit
∂
(ρu~x )4V
∂t
4V is the size of the control volume 4x4y4z. For the convective trans-
port we can derive equations in the same manner as in section 1.1.
ρu u |
x y y+∆y − ρux uy |y
ρu u |
x x x+∆x − ρux ux |x
− ∆x∆y∆z − ∆x∆y∆z
| ∆x {z } | ∆y
{z }
changes of x-impuls in x-direction changes of x-impuls in y-direction
Sum of all forces on the system is a function of pressure, viscous forces and
gravity.
Pressure is an isotropic variable, ie it has the same value in all directions.
The conservation equation for pressure can be written as:
P |y+∆y ∆x∆z
?
y + ∆y
y
6
x x + ∆x
P |y ∆x∆z
P|
x+∆x − P |x
(P |x − P |x+∆x )∆x∆z = − ∆V
∆x
5
? τxy |y+∆y ∆x∆z
y + ∆y
6
?
y -
6 τxy |y ∆x∆z
x x + ∆x
Shear stress is a deformation force that alters the control volume shape
while the normal stresses changes the volume.
The viscous forces are given by:
τ |
xy y+∆y − τxy |y
τ |
xx x+∆x − τxx |x
− ∆x∆y∆z − ∆x∆y∆z
∆x ∆x
-
gx
?
gy
ρu u |
x y y+∆y − ρux uy |y
ρu u |
∂ x x x+∆x − ρux ux |x
∆V (ρux ) = − ∆V − ∆V
∂t ∆x ∆y
τ |
P |x+∆x − P |x xy y+∆y − τy |y
τ |
xx x+∆x − τx |x
− ∆V − ∆V − ∆V + ρgx ∆V
∆x ∆x ∆y
∆V → 0 (∆x → 0, ∆y → 0)
The conservation of impulse in 2 dimensions in x-direction is:
6
∂ ∂ ∂ ∂P ∂τxx ∂τxy
(ρux ) + (ρux ux ) + (ρux uy ) = − − − + ρgx
∂t ∂x ∂y ∂x ∂x ∂y
∂ ∂ ∂ ∂P ∂τyy ∂τxy
(ρuy ) + (ρuy ux ) + (ρuy uy ) = − − − + ρgy
∂t ∂x ∂y ∂y ∂y ∂x
∂ ∂ ∂P ∂τij
(ρui ) + (ρui uj ) = − − + ρgj
∂t ∂x ∂xj ∂xj
Duj ∂P ∂τij
ρ =− − + ρgi
Dt ∂xj ∂xi
The stress tensor, τij , is fluid depended. We have already mentioned that
viscosity is divided in shear- and normal stresses. For a Newtonian fluid we
can write:
∂ui ∂uj 2 ∂uk
τij = −µ + − δij (1.1)
∂xj ∂xi 3 ∂x
| {z } | {z k}
shear stresses normal stresses
The term k is the sum of x, y and z, and is given by equation 1.2. δij is the
Kroenicker delta and is equal to one when i = j, and zero when i 6= j. That
means, when Kroenicker delta is equal to zero, normal stresses are absent, and
the flow is incompressible (ρ is constant). The volume will not change, so we
can ignore the last term in equation 1.1. If we include combustion, there will
7
be large temperature gradients, and we can no longer assume that the density
is constant.
The stresses in equation 1.1 can be simplified to:
∂u
i ∂uj
τij = −µ +
∂xj ∂xi
∂τij ∂2u ∂ 2 ui
i
= −µ +
∂xi ∂xi xj ∂xi xi
∂τij ∂ ∂u
i ∂ 2 ui
= −µ +
∂xi ∂xj ∂xi ∂xi xi
∂ui
=0
∂xi
∂τij ∂2u
i
= −µ
∂xi ∂xi xj
From this we write the two general equation which are generally used
in CFD: Navier-Stokes equation (1.3) and Euler equation (1.4). For Navier-
Stokes we assume that both ρ and µ are constant. The Euler equation is often
used in aero dynamics, since the viscosity for air is negligible (µ ∼ 0).
∂ ∂ ∂P ∂ 2 ui
(ρui ) + (ρui uj ) = − +µ + ρgi (1.3)
∂t ∂xi ∂xi ∂xi ∂xj
∂ ∂ ∂P
(ρui ) + (ρui uj ) = − + ρgi (1.4)
∂t ∂xi ∂xi
8
Chapter 2
Turbulent Theory
In this chapter we will explain what turbulence really is, how it is created, and
how it is transferred from the mean flow to the smaller whirls. We will describe
the differences between isotropic-, homogeneous- and stationary turbulence.
The turbulent flow can be divided in spectrum, and we will derive equations
for the smallest scales. Correlations are important in turbulence. From these
numbers we can derive energy spectrum for all scales, and derive the Taylor
micro scales and the Reynolds number, Reλ , associated with it. At the end of
this chapter we will derive the basic Reynolds Averaged Navier-Stokes (RANS)
equation which is used in a CFD code for simulation of turbulent flow.
Re = UD/v
4000
2000
Laminar
t, sec
9
U
Moving plate
u
h
Wall
a wide range of length and time scales of the turbulent motion. The size
of the largest whirls are determined by the geometry, and the size of the
smallest whirls are determined by Kolmogorov scales. The flow is almost
always 3 dimensional. Dissipation by viscous forces dampen turbulent unless
new energy is constantly supplied. Turbulence relies on extracting energy from
the mean flow, eg a shear layer. Vortex stretching is an important mechanism
feeding energy into vortices.
Boussinesq assumed as early as 1877 that transport by turbulence is an
extra stress, or an extra viscosity.
du
τ =µ (2.2)
dy
10
y y
A A A A
Figure 2.3: Turbulence balls is assumed to behave in the same way as molecules
in laminar flow
migrating from low velocity layer to a high velocity layer. This momentum
transfer will result in a shear layer, where the layer near the wall has nearly zero
velocity, and the shear layers further up have an gradual increasing velocity.
This idea about momentum transfer, caused by molecular movement between
velocity gradients, is not restricted to boundary layers only. It also takes place
in other types of flow like jets and turbulent mixing layers.
Prandtl used Boussinesq’s idea in equation 2.1 to develop a simple way
to calculate turbulence (algebraic or zero equation model). Prandtl used the
wall mixing model where he assumed that the eddies behaves in the same way
as molecules in laminar flow. Fluid lumps moving toward the low velocity
regions, across line A-A, have some of their excess moment removed by the
lower velocity fluid. Conversely, lumps moving away from the lower velocity
region gain momentum from their new high velocity surrounding fluid. This
is illustrated in figure 2.3. Lower velocity regions acts as a momentum sink.
The eddies transport mass from one place to another. Flux of momentum
is called stress, or said in another way, turbulent stresses is the transport of
momentum (ρu0 ) with the υ fluctuation.
This way of representing stress, where we assume that the eddies in tur-
bulent flow behave in the same manner as molecules in laminar flow, is called
the “eddy-viscosity approximation”. We use this approximation in zero-, one-
and two equations model, see section 2.4 and chapter 3. To verify that the
this flux of momentum really is a stress, we can compare the units for both
molecular viscosity in equation 2.2 and turbulent viscosity in equation 2.3.
du kg m/s kg N
τ =µ · = = 2
dy m · s m m · s2 m
0 0 kg m m kg N
τturb = ρu v 3
· · = 2
= 2
m s s m·s m
11
Turbulent Transport Larger whirls increases the transport across the flow.
They transport fluids at distances equal to largest whirls, le . Smaller whirls
does the same, but over a shorter distance. They are less important for trans-
port, but are important for chemical reactions 1 . Larger whirls are more
important for prediction of flow than smaller whirls.
Big whirls have little whirls that feed on their velocity, and little
whirls have lesser whirls and so on to viscosity - in the
molecular sense
Figure 2.4: Breakdown of whirls from larger to smaller whirls (Ertesvåg, 2000)
ρvD vD
Repipe = =
µ υ
v is the velocity for the fluid along the pipe and D is the pipe diameter. µ
is dynamic viscosity and υ is kinematic viscosity. For a vortex, the diameter
is defined as `, and the vortex rotating velocity is u. Notice the difference
between velocity in a pipe and in a vortex; axial velocity for a pipe and radial
velocity for a vortex.
1
We will come back to this in Combustion Technology, FACE 9
2
From the English scientist Lewis Fry Richardson, 1922
12
Instabilities in flow causes a degredation u00 `00
of larger whirls to smaller ones with
Reu00 = ν
1
0
`
characteristic length `00 , and velocity u00
u0
`00
The larger whirls are defined by the
characteristic length `0 ,
u00
and velocity u0 .
u0 `0
Reu0 = ν
1
η
This continues until the Reynolds number υ
are equal to 1. Characteristic
length is η, and velocity is v.
υη
Reη = ν
=1
Figure 2.5: Energy is transferred from the larger to the smaller whirls
(Ertesvåg, 2000)
The main flow will transfer energy to the larger whirls. These whirls are
determined by the the outer dimensions of the flow, eg the pipe diameter.
Time scales for the small whirls are short. They are in close equilibrium with
13
Universal Equilibrium Range
Energy Containing
Range
Dissipation Range Inertial Subrange
η `e
Figure 2.6: Turbulent spectrum with the larger energy containing eddies to
right and dissipation eddies to left
the local properties. They are called universal equilibrium range. The term
universal is used because this assumption goes for all high Reynolds number
flows. The different scales in a turbulence spectrum are illustrated in figure
2.6. The spectrum is divided in 2- or 3 parts. High Reynolds number flow
has 3 parts while low Reynolds number flow has only two parts. The inertial
subrange vanish for lower Reynolds number. The universal equilibrium range
can be divided in two separate ranges, inertial sub-range and dissipation range.
Inertial subrange In the inertial region the turbulence scales are indepen-
dent of both the large scales and the small scales if the Reynolds number is
large. For turbulent flow with low Reynolds number this region is “small”, ie
only dissipation- and energy containing range are present.
This region is characterised by the amount of energy transported through
the spectrum per time (ie ε) and the sizes of the eddy (ie 1/κ). The energy in
the region can be estimated as:
E ≈ ε a κb
a b
L2 L2a−b
a b 1
E(κ) = (ε κ ) = =
T3 L T 3a
2
which gives: a = 3 and b = − 53
2 5
E ∼ CK · ε 3 κ− 3
14
E(κ)
slope = − 53
κ1
Figure 2.7: - 5/3 low for inertial subrange. Both the wavenumber κ- and the
energy E(κ) axis are given in logarithmic function
the whirls will dissipate into a molecular movement. These whirls are named
after the person who first came up with this ”idea”, Kolmogorov. Kolmogorov
also assumed that:
As illustrated in figure 2.5, Reynolds number for the smallest whirls are
equal to 1, ie inertia forces equals viscous forces (or there is a balance between
momentum forces and diffusion).
ηυ
Reη = =1
ν
The scales of the smallest eddies are determined by viscosity, ν (m2 /s), and
dissipation (increase of thermal energy), ε (energy/time=m2 /s3 ). The length
scale, η, for the smallest eddies can be expressed as:
η = ν a εb
m : 1 = 2a + 2b s : 0 = −a − 3b
3
which gives a = 4 and b = − 14 . The Kolmogorov length scale is then given
by:
15
14
ν3
η=
ε
Writing the velocity scales in the same way gives the Kolmogorov velocity
scale υ and the Kolmogorov time scale τ as:
1
υ = (νε) 4
ν 1
2
τ=
ε
We have seen above that there can be derived expressions for inertial- and
dissipation range. But these expressions, derived by Kolmogorov, are derived
from dimensional analysis, and not from experimental measurements. There
are actually smaller scales than the Kolmogorov scales. We will come back to
this later in this chapter.
Auto correlation and cross correlation are two important statistical factors in
turbulent analysis. These correlations are meant for isotropic turbulence but
we also use them as approximations to real turbulence. First we need a basic
explanation about what correlation really is.
Correlation is a measure of linear relationship, or the study the strength
of relationship between two random variables. One simple kind of association
between the variables x and y produces pairs of values or, graphically, points
that scatters around a straight line. A numerical measure of this relationship
is called the sample correlation coefficient (SCC), and is given in equation
2.4. Small amount of scatter around a line indicates strong association and
large amount of scatter is a result of weak association. Examples are given in
figure 2.8. High values for SCC indicates a strong relationship, as shown in
(a). High negative values in (b) indicates that there is an anti-correlation. If
values for SCC approximates 0.5, as in (c), there is not any clear relationship.
A further description about correlation can be found in any basic book about
basic statistics, such as Bhattacharyya and Johnson (1977).
Pn
i=1 (xi− x̄)(yi − ȳ)
SCC = q P (2.4)
n 2
Pn
2
i=1 (xi − x̄) i=1 (yi − ȳ) )
16
(a) SCC=0.915 (b) SCC=-0.912 (c) SCC=0.492
For standardisation we often define this function with help of the root-
square-mean (σ) values at the same points. The result is:
Qij
Rij =
σi σj
17
f (r) g(r)
1 1
0 0 r
r
0 0
6
6
u0n u0n (r)
r
-
If two points are at the same spot, we are “looking” at the same fluctuation.
The correlation is then 1.
f (0) = g(0) = 1
When distances increases, the correlation between them will gradually be-
come smaller. If correlation is equal to zero, they are uncorrelated.
f (∞) = g(∞) = 0
The value for f (r) can be negative for a part of of r. The value for g(r)
must be negative for a part of r. A fluctuation across the line must result in
fluctuation in opposite direction (continuity). This is illustrated in figure 2.9
We can define a length scale Lf where the area f (0) is equal to the area
under the curve f (r) ≥ 0, as indicated in figure 2.10. Since f (0) = 1, we can
write the longitudinal macroscale:
18
f (r) f (r)
1 1
0 0
r r
0 Lf 0 λf
Z ∞
Lf = f (r)dr
0
We can define the lateral correlation function, g(r), in the same way:
Z ∞
Lg = g(r)dr
0
It can be shown that Lf = 2Lg . These two length scales tells us the
distance fluctuations will influence each other.
Taylor series are polynomials that can be used to approximate other func-
tions. The general formula for Taylor series are given in equation 2.6.3
f 00 (a) f ka f na
Pn (x) = f (a)+f 0 (a)(x−a)+ (x−a)2 . . .+ (x−a)k +. . .+ (x−a)n
2! k! n!
(2.6)
Taylor series for f (r) around r = 0 gives:
1 00
f (r) = f (0) + f 0 (0)r + · f (0) · r2 + . . .
2
We know that f (0) = 1 and f 0 (0) = 0. If we round off the Taylor series
after the r2 link, the result will be a parable which crosses the r-axis. We
define a length scale λf as a values instead of r where the parable crosses
the r-axis (f (r = λf ) = 0). This length scale is called the longitudinal micro
scales.
2
λ2f = −
f 00 (0)
From the lateral correlation function g(r) we can define a lateral microscale
in the same manner as for the longitudinal micro scales.
3
A more detailed discussion about this topic can be found in Kreyszig (1993)
19
2
λ2g = −
g 00 (0)
√
It can be shown that (Hinze, 1975) λf = 2λg. Dissipation of turbulence
energy is:
∂u0i ∂u0i
ε=v
∂xj ∂xj
2 2 2
∂u01 ∂u02 ∂u03
= =
∂x1 ∂x2 ∂x3
2 2 2 2 2 2
∂u01 ∂u01 ∂u02 ∂u02 ∂u03 ∂u03
= = = = =
∂x2 ∂x3 ∂x1 ∂x3 ∂x1 ∂x2
0 2 0 2
∂u1 ∂u1
ε=v 3 +6
∂x1 ∂x2
2
∂u01 2u02
= −u02 · f 00 (0) =
∂x1 λ2f
2 0 2
∂u01 4u02 2u02
00 00 ∂u1
= −u02 02
· 2f (0) = 2 = 2 = −u · g (0) = 2
∂x2 λf λg ∂x1
2
∂u01 u02 u02
ε = 15v = 30v 2 = 15v 2
∂x1 λf λg
u02 k 2
ε = 15v 2
= 10v 2 , u02 = k
λ λ 3
We only have one length scalar, as in isotropic turbulence. λ is equal to
λg , and with this length scale we can define a turbulence number:
u0 λ
Reλ =
v
We know that the larger whirls are determined by the outer dimensions,
and the smaller whirls by viscous forces. A turbulence Reynolds number, such
as Reλ , gives a measure for the degree of turbulence. It also gives the difference
20
between the smaller (η) and the larger (`e ) whirls. If Reλ is high, there will
be a large difference between small and large whirls. It can also be shown that
for isotropic turbulence (Hinze, 1975):
`e 3
∼ 0.1Reλ2
η
What are Typical Values for Reλ ? Ertesvåg (2000) discusses different
values for Reλ . The highest Reynolds numbers are found in the atmosphere
and oceans. The outer dimensions for these flows are huge, and typical
Reynolds number are 104 . Typical flows in laboratories are much lower. A
Reynolds number (Reλ ) at 150-200 is considered high. In special cases a
Reynolds number up to 103 is achievable.
k2
ReT =
vε
1 1
0 2 2 10vk 2
u = k λ=
3 ε
1
u0 λ 20 k 2
2
Reλ = =
v 3 vε
2
3k · 10vk 20 k 2 20
Re2λ = = = ReT
v2ε 3 vε 3
F
Rij (~r, t) → Eij (~κ, t)
Eij is the energy spectrum tensor and ~κ is the wave number vector (1/l).
The energy spectrum E(κ, t) are independent of directions, and is defined as:
Z ∞
1
E(κ, t)dκ = u0i u0i = k(t)
0 2
21
E(κ)
D(κ)
`0 κe, `e κd, `d η −1
Figure 2.11: Energy and dissipation spectrum. The x-axis is logarithmic and
y-axis is linear (Ertesvåg, 2000)
This energy spectrum is the area under the E(κ) curve in figure 2.11.
Dissipation spectrum D(κ, t) is the area under the other curve in the same
figure, and is defined as:
Z ∞
D(κ, t)dκ = ε(t)
0
Both curves in figure 2.11 have a maxima at different locations from the
smallest (η) and largest whirls (`0 ). The largest whirls in a flow have the same
diameter as the outer dimensions, which is `0 . For example, the largest whirls
in a pipe have the same length as the pipe diameter. There are not many of
these larger whirls; most of the energy containing whirls are smaller. They
are indicated in figure 2.11 by the letters `e . The larger whirls, `0 , have more
energy than the smaller whirls, `e . But `e are more numerous, so most of
the energy is located around this scale. We learned earlier that the smallest
whirls are the Kolmogorov scales. This scales are derived from dimensional
analysis, and experiments has shown that there are actually smaller scales in
a flow. Most of the dissipation, however, takes place in larger whirls (`d ).
Hinze (1975) is referring to some measurements which indicates that `d is
typically 10 times larger than the Kolmogorov scales for isotropic flow. Pope
(2000) is referring to other measurements for real flows (non-isotropic) where
`d can be up to 60 times larger. Taylor scales are larger than the Kolmogorov
scales, and some people often relate this scales to `d , ie that the Taylor scales
represents the eddies sizes in which most of the dissipation occurs. There are
some disagreement about this statement. According to Tennekes and Lumley
(1972) the Taylor scales does not represent any group of eddy sizes in which
dissipative effects are strong. It is not a dissipation scale, because it is defined
with the assistance of a velocity scale which is not relevant for the dissipating
eddies.
22
E(κ)
their “strength”, the particles will slip from each other and the whirls becomes
larger. The area under the energy spectrum is the turbulence energy k; this
area is decreasing as well. The length scales `e is increasing, and the top of
the curve will move towards left (lower wave number and higher/larger length
scales). Kolmogorov length scales are also increasing. The relationship `e /η
(is a measure for the extent of the spectrum) is decreasing. All this are shown
in figure 2.12.
The energy spectrum is decreasing and it shifts toward the longer length
scales when turbulence decreases.
∂ρ ∂
+ =0 (2.7)
∂t ∂xj
∂ ∂ ∂p τij
(ρui ) + (ρui uj ) = − + + ρSi (2.8)
∂t ∂xj ∂xi ∂xj
ρ = const.
∂ui ∂uj 2 ∂uk
>
τij = µ + − µ δij
∂xj ∂xi 3
∂xk
23
u0
u
ū
ui = ui + u0i (2.9)
∂ ∂ ∂p τij
(ρ(ui + u0i )) + (ρ(ui + u0i )(uj + u0j )) = − + + ρSi
∂t ∂xj ∂xi ∂xj
∂ ∂ ∂p τij
(ρ(ui + u0i )) + (ρ(ui uj + ui u0j + u0i uj + u0i u0j )) = − + + ρSi
∂t ∂xj ∂xi ∂xj
We need to average each term in the equation. The mean value for a term
with only one fluctuation is equal to zero. We need three laws for averaging:
φ=φ
φ+ψ =φ+ψ
ψφ0 = ψ · φ0
0 0 0
∂ ∂ ∂p τ ij
0 0 + u0 u0i u0j )) = −
>
(ρ(ui + u )) + (ρ(u u + u u u + + + ρS i
>
i i j i j i j
∂t ∂xj ∂xi ∂xj
The final result is Reynolds averaged equation (we assume that ρ = con-
stant) are given in equations 2.10 and 2.11.
∂uj ∂u0j
=0 and =0 (2.10)
∂xj ∂xj
24
∂ ∂ ∂p ∂
(ρui ) + (ρui uj ) = − + (τ ij − ρu0i u0j ) + ρS i (2.11)
∂t ∂xj ∂xi ∂xj
ρu0i u0j is the Reynold stresses, and is a new unknown variable that needs to
modelled. That is the purpose for a turbulence model. There are two different
approaches to model the Reynold stresses, either the eddy-viscosity approach
introduced in chapter 2.1, or we can model all six equations for turbulence
stresses. The eddy viscosity approach is often modelled by equation 2.12.
The six equation for turbulent stresses, represented by equation 2.13, can be
derived from Navier-Stokes equations in the same way as we derived the ex-
pressions for the turbulent equation of momentum (equation 2.11). Definition
of the different terms in equation 2.13 are discussed in chapter 3.4.
ρ = const.
∂ui ∂uj 2 ∂u
l
7
−ρui uj = µt + − ρk + µt δij (2.12)
∂xj ∂xi 3 ∂xl
∂u0i u0j
∂ 0 0 ∂ 0 0 0 0 ∂uj 0 0 ∂ui ∂
ρui uj + ρui uj uk = − ρui uk + ρuj uk + µ
∂t ∂xk ∂xk ∂xk ∂xk ∂x
| {z } | {z } | {z k }
ρCij ρPij ρDij,v
(2.13)
25
Chapter 3
Turbulence Models
This chapter is about the different turbulence models. This are zero equation
model, one equation model and the two equation models. Zero- and one
equation models are (there are exceptions) incomplete, ie before we start a
simulation we need to know the length scales (or time scales). Two equation
models was derived in the beginning of the 70’, and was a break through in
turbulence modelling. The most popular two equation model, the standard
k-ε model, is the most (some would probably say - “only”) used turbulence
model in industry.
Zero equation model are the simplest of all turbulence model. It was developed
by Prandtl in 1925. He visualised a simplified model for turbulent motion in
a boundary layer where the turbulent balls moves around as molecules. He
showed that the two velocity fluctuations can be given by:
du du
u0 = ` v0 = `
dy dy
2
0 0 2 du
τturb ∼ υ (ρu ) ∼ ρ`
dy
26
3.2 One Equation Model
In one equation models we are modelling the turbulence energy, and solving
dissipation (ε) analytical. They are incomplete models since the length scales
have to be related to some typical flow dimensions. This model was a fore-
runner for the two equation model. There are (lately) developed complete one
equation models, eg Spalart-Allmaras turbulence model.
We need to develop an equation for turbulence energy - k. Kinetic energy
per mass unit in a flow is 12 ui ui . If we subtract the mean values from this,
we end up with the kinetic energy for the turbulent fluctuations 12 u0i u0i . The
mean value of this expression is the “mean kinetic turbulence energy” or just
“turbulence energy”
1 1
k = u0i u0i = (u02 + u02 02
2 + u3 )
2 2 1
We will not derive the equation for k, but we will use a method from
Ertesvåg (2000) and show how it can be done in 5 steps.
2. Reynolds averaging: ui = ui + ui
3. Subtract the last equation from the first. The result is an equation for
the fluctuation u0i = ui − ui .
∂ ∂ ∂
(ρu0i ) + . . . = . . . =
(ρui ) + . . . = . . . − (ρūi ) + . . . = . . .
∂t ∂t ∂t
4. multiply this equation with fluctuation u0i ; the result is an equation for
1 0 0
2 ui ui (fluctuating turbulence energy)
∂ ∂ 1 0 0 averaging ∂ 1 0 0
u0i (ρu0i )+. . . = . . . =
(ρ ui ui )+. . . = . . . −→ (ρ ui ui )+. . . = . . .
∂t ∂t 2 ∂t 2
∂u0 ∂u0i
∂ ∂ ∂ ūi ∂ ∂k ∂ 1
0 0 − ρu0i u0i u0j − pu0j − µ i
(ρk) + (ρkūj ) = −ρui uj + µ +
∂t ∂xj ∂xj ∂xj ∂xj ∂xj 2 ∂xj ∂xj
| {z } | {z } | {z } | {z } | {z }
ρCk ρPk ρDk,v ρDk,t ρε
This equation is exact, ie it has been derived from the basic Navier-stokes
equation and we have not made any assumption, simplifications or trying to
model any terms.
27
Pk production term
ε dissipation
We cannot solve the exact equation for k. Some of these terms are unknown
and have to modelled (the equations is not exact any more). The triple corre-
lation u0i u0i u0j and pressure-velocity correlation p0 u0j are unknown, and Dk,t has
to be modelled. The viscous gradient term, Dk,v , can be calculated directly,
and does not need any modelling. In high Reynolds number flow, this term is
small compared to the turbulent viscosity, Dk,t . Pk can be calculated directly.
Dissipation term, ε, has to be modelled.
Turbulent diffusion, Dk,t , is modelled as a gradient model:
µt ∂k
(3.1)
σk ∂xj
u03 k 3/2
ρε ∼ ρ → ε ∼ CD
L L
28
√ √
vt ∼ u0 `0 u0 ∼ 2k ∼ k `0 = L
√
vt = CL kL (CL = 0.09)
∂ ∂ ∂ µt ∂k
(ρk) + (ρkuj ) = µ+ + ρPk − ρε
∂t ∂xj ∂xj σk ∂xj
∂ui ∂uj ∂ui
ρPk = µt +
∂xj ∂xi ∂xj
√
µt = ρvt = ρ kL
k 3/2
ε = CD
L
L is an algebraic expression and must be determined before the simulation
starts.
As opposed to zero (analytical)- and one equation models, two equations mod-
els are complete. They can be used to predict properties of a given turbulent
flow with no prior knowledge of the turbulence structures. We are using the
same k - equation as for the “one equation model”, and in addition we need
to find a transport equation for the length- or time scale. There are several
different approaches to find this extra scales, and the result is a variety of two
equation models. Some of them are well documented, whilst other are not
documented at all. The most popular model is the “Standard k − ε model”.
An exact equation for ε can be derived from the Navier-Stokes equations. The
final result will be:
∂ε ∂ε ∂ ūi ∂ 2 ūi
= −2v u0i,k u0j,k + u0k,i u0k,j − 2v u0k u0i,j − 2v u0i,k u0i,m u0k,m
+ ūj
∂t ∂xj ∂xj ∂xk ∂xj
2 0 0 ∂ ∂ε 0 0 0 v 0 0
−2v ui,km ui,km + v − v uj ui,m ui,m − 2 pm uj,m
∂xj ∂xj ρ
29
∂ ∂ ∂ µt ∂k
(ρk) + (ρkūj ) = µ+ + ρPk − ρε (3.2)
∂t ∂xj ∂xj σk ∂xj
∂ ∂ ∂ µt ∂k
(ρε) + (ρεūj ) = µ+ + ρPε − ρQε (3.3)
∂t ∂xj ∂xj σk ∂xj
ε ε
Pε = Cε Pk Qε = Cε2 ε
k k
The idea behind this relationship is that when turbulence energy increases,
the disintegration has to increase as well. Or said in another way: when Pk
increases, ε should also increase, if it does not, k can get unphysical high
values. The best way to increase ε is by increasing Pε - and that is the reason
why Pε is dependent on Pk . When turbulence energy decreases, disintegration
must also decrease. The complete set of equation and its constants for the
standard k − ε turbulence model are (Ertesvåg, 2000):
k2
µt = ρvt = Cµ ρ
ε
∂ ūi ∂ ūj 2
−ρu0i u0j = µt + − ρkδij
∂xj ∂xi 3
∂ ∂ ∂ µt ∂k
(ρk) + (ρkūj ) = µ+ + ρPk − ρε
∂t ∂xj ∂xj σk ∂xj
∂ ∂ ∂ µt ∂ε ε ε
(ρε) + (ρεūj ) = µ+ + Cε1 ρPk − Cε2 ρε
∂t ∂xj ∂xj σε ∂xj k k
∂ ūi ∂ ūj ∂ ūi
ρPk = µt +
∂xj ∂xi ∂xj
30
What is ε ? For high Reynolds number, most of the dissipation takes place
in the smaller whirls. The mechanical energy is transferred from the mean
flow to the larger whirls, and thereafter to the smaller whirls.
The exact equation for ε represents the dissipating eddies, or the length
scale for these. Since this equation is hard to model, we are using a simplified
model where ε is given by the larger whirls. ε in the k − ε model is not the
dissipation, but it represents the energy transfer from larger to smaller whirls.
In stationary flow the energy transfer from the larger whirls are equal to the
dissipation in the smaller whirls.
Since the ε equation we use in k−ε models (and RSM) differs from the exact
equation derived from the Navier -Stokes equation, this equation is often seen
as the weak part in the turbulence model. In standard k − ε model (and other
two equation models) we only have one equation for the Reynolds stresses.
Advantages with the k − ε model are:
• stable
Disadvantages:
The Reynold stress model (RSM) does not use the eddy viscosity approach
described in chapter 2.1. Instead we are deriving 6 individual equations for the
stresses. This equation is already given in 2.13, but for simplicity we repeat
it:
∂u0i u0j
∂ 0 0 ∂ 0 0 0 0 ∂uj 0 0 ∂ui ∂
ρui uj + ρui uj uk = − ρui uk + ρuj uk + µ
∂t ∂xk ∂xk ∂xk ∂xk ∂x
| {z } | {z } | {z k }
ρCij ρPij ρDij,v
31
Pij production, or energy transferred from the mean flow to the Reynolds
stresses
The convection term Cij , production term Pij , and viscous diffusion term
Dij,v can be solved directly. The viscous diffusion term is often negligible
compared to the turbulent diffusion term, but it remains important where
there are significant gradients. The other terms like turbulence diffusion Dij,t ,
re-distribution term Φij , and dissipation εij needs to be modelled.
The turbulent diffusion model is modelled as:
k 0 0 ∂u0i u0j
∂
Dij,t = Cs uk u` (3.4)
∂xk ε ∂x`
Cs is often set to 0.22. k/ε is used as a time scale. Dij,t can be interpreted
in two different ways (in the same way as in the k in equation 3.1):
2. as a model for the triple correlation gradient, and the pressure term is
ignored
∂u0i u0j
∂ 0 k
Dij,t = Cµ k (3.5)
∂xk ε ∂xk
This model (equation 3.5) is easier to program than the other model in
equation 3.4 , and in some cases its proved to be more stable. Fluent uses a
similar simplified approach.
Dissipation tensor, εij , is assumed to be isotropic, ie dissipation is equal
for all normal stresses.
2
εij = εδij
3
Dissipation takes place in the smaller whirls. These whirls are indepen-
dent of the larger whirls and the main flow. For the small whirls, the flow
probability for the different directions are all equal. The dissipation must be
equal distributed between the three energy components u02 02 02
1 , u2 and u3 . This
gives: ε11 = ε22 = ε33 = 2/3 ε
32
The pressure strain term is the most uncertain terms we are modelling in
the Reynolds stress equation. The classical approach to modelling Φij uses
the following decomposition:
Pressure strain term is also called the re-distribution term. It takes from
the terms who has a lot, and gives it away the the terms who has less (“Robin
Hood term”). Or said in another way: it shares the energy between the
components, or it makes the turbulence isotropic.
The slow pressure strain, Φij,1 , is modelled as:
ε 0 0 2
Φij,1 = −C1 ρ ui uj − δij k
k 3
33
k 3/2
ε 3 3
Φij,1w = C1w u0k u0m nk nm δij − u0i u0k nj nk − u0j u0k ni nk
k 2 2 C` εd
k 3/2
3 3
Φij,2w = C2w φkm,2 nk nm δij − φik,2 nj nk − φjk,2 ni nk
2 2 C` εd
where C1w = 0.5, C2w = 0.3. nk is the xk component of the unit normal
3/4
to the wall, d is the normal distance to the wall, and C` = Cµ /κ, where
Cµ = 0.09 and κ is the von Karman constant (= 0.4187).
The wall reflection term dampens the normal fluctuations near the solid
surfaces. But this term is not only a near-wall term, it also influences the
flow outside the boundary layer. For flow in complex geometries, it may
be difficult to define the wall distance. Developers of boundary layers and
numerical techniques makes the necessary fine mesh close to the wall, and the
results from the RSM are often satisfying. In industry, there are often limited
time or resources to use fine grids. According to Ertesvåg (2000), there are
indications that a corse grid near the wall can worsen the result compared
to a k-ε model. A typical simulation which require the solution of Reynolds
stresses due to swirl, rotations, etc could be solved more correctly with a k-ε
model if we cannot afford a fine mesh.
u2
ū
u01 /ū1
u02 /ū1
u03 /ū1
u3
u1
Figure 3.1: Turbulence variations for the 3 velocity components close to the
wall
The advantages with RSM is that it reproduces (or trying to reproduce) the
dynamics of each stress component which enables better modelling of stresses.
This will lead to better capturing of (Hanjalic, 2004):
• swirling flows
34
• secondary motions (pressure induced and stress induced)
35
Chapter 4
Summary
When we use the theory for isotropic turbulence, and measure the correlation,
in our case the autocorrelation, we can find the macro scales and the Taylor
scales (or Taylor micro scales). The last is interesting because it gives us Reλ .
This number gives us the difference between the larger and smaller whirls,
and this number must be higher than 150-200 for the existence of the inertial
sub-range.
The theory for the two most interesting turbulence models, standards k −ε
and RSM, are discussed. The former is the most popular turbulence model
and is widely used in industry. There are many other turbulence models that
resemble the standard k − ε model, and they might perform better in some
types of flow.
The problem in turbulence modelling is to know when to use the different
turbulent models. In most cases the two equation models will be adequate for
many tasks. If the flow has sudden changes in flow directions, the Reynolds
stress model should be used. But there are back sides with the RSM. We have
already mentioned that the boundary layer should be better resolved when
we are using RSM. If we cannot resolve the boundary layer, we may have to
stick with two equation models, even if the physics of the flow indicates that
the RSM is preferred. Another problem with the RSM are difficulties with
convergence - it is harder to make a simulation converge when using RSM
compared to two equation models. This may be one of the reasons why many
CFD users prefer the two equations models.
36
Bibliography
37
Appendix A
Turbulence Pictures
38
Figure A.1: Flow at low Reynolds number
Figure A.2: A von Kárman street is created below the bolts at low Reynolds
number
Figure A.3: Flow velocity is increased and some parts in the flow becomes
turbulent
39
Figure A.4: Further increase in Reynolds number will make the flow turbulent,
and there is a “stretching” of vortices
40
Appendix B
Examination Questions
• Explain the relationship between the mean flow Reynolds number and
the range of scales
• Define the regions of confined turbulent flow, and discuss physical aspects
of these
41
Question 6: Reynolds averaging and closure
• There are different turbulence models to solve the new terms, −u0i u0j , to
close the RANS equations. What is the major difference between 0-, 1-,
2-equation models and Reynolds stress models?
• Dimension analysis show that the new terms −ρu0i u0j are also stresses,
same as the viscous stresses, τij . For incompressible Newtonian fluids,
the viscous stresses can be expressed as:
∂ui ∂uj
τij = τji = µ +
∂xj ∂xi
The last term seems beyond the direct analogy to the viscous stresses.
Why does the hypothesis include this term?
42