Anda di halaman 1dari 20

The k −  Turbulence Model

ME 448/548 Lecture Notes

Gerald Recktenwald∗
August 3, 2009

1 Overview
This article gives a very brief overview of the basic k −  turbulence model used in CFD codes. The
major topics discussed are
• The Reynolds Averaged Equations
• The modified Boussinesq Eddy-Viscosity Concept
• Prandtl’s Mixing Length Hypothesis
• The k and  equations
• Boundary Conditions and Wall Functions
Good background reading for this material can be found in the books by Ferziger and Perić [1,
Chapter 9] Panton [4, Chapter 23], Pope [5, Chapter 10], and Tannehill et al. [7, § 5.2, § 5.4].

1.1 What Is Turbulence?


Turbulence is a fluid flow phenomenon characterized by
• unsteadiness,
• fluid motions that appear irregular or topologically complex,
• fluid motions occurring on a wide range of physical scales,
• rapid mixing of passive contaminants (smoke, heat, concentrations of chemical species)

Unsteadiness Imagine using a probe to measure the fluid velocity at a point in a turbulent flow.
To be specific, consider inserting a probe through a small hole in a duct so that the sensing element
of the probe is located near the centerline. Assume that the probe has a high frequency response,
meaning that it is capable of detecting any time variation in the velocity signal. Figure 1 shows the
output of the probe as it would be displayed on an oscilloscope. The velocity signal is characterized
by a nominal average U , and a superimposed fluctuation, u0 . The average value is representative
of the mass flow rate through the duct. The fluctuating component is an essential feature of the
turbulence.
∗ Associate Professor, Mechanical and Materials Engineering Department Portland State University, Portland,

Oregon, gerry@me.pdx.edu
1.1 What Is Turbulence? 2

1.8

1.6

1.4

1.2

1
u

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2
t

Figure 1: Turbulent velocity signal at a point. The Reynolds decomposition separates the signal
into an average, U , and a fluctuating component, u0 .

Flow Irregularity Let us continue with our imagined experiment with duct flow. Suppose there
was some way to make the flow patterns inside the duct visible. First of all, assume that the duct
walls were transparent. Then suppose that we were to inject smoke from a series of small pipes
having outlets at a small number of vertical locations on the center plane of the duct. Finally, in
order to keep the smoke from obscuring all the details, illuminate the center plane of the duct with
a sheet of laser light. The light has the effect of showing smoke that is only in the plane. All other
smoke in the duct is not illuminated and, therefore, invisible. Figure 2 is an example of such a flow
visualization.
The first thing apparent from the smoke visualization is that the flow is extremely complicated.
The flow swirls and fluctuates with no apparent pattern. It would be difficult, for example, to trace
out the trajectory of a particular puff of smoke.
Further complicating an identification of flow patterns is the realization that our visualization
system is illuminating the flow in a single plane. The smoke that we see is not confined to move in
the plane. In fact, any smoke that we see at one instant is likely to be swept out of the plane in
the next instant. Thus, although our view is two-dimensional, the flow features are fundamentally
three-dimensional as well as being unsteady.

Multiple Scales After staring at the smoke in the duct, we may begin to adjust our perspective.
In particular, if we allow our eyes to follow a billow of smoke as it is carried downstream, we will
see small whorls of smoke turning and tumbling. If we were to take photographs of the smoke, we
would be able to identify whorls of many different sizes. Fluid dynamicists call the whorls “eddies”.
An exact definition of a turbulent eddy is somewhat elusive. Pope [5, p. 183] describes an eddy
as “a turbulent motion, localized within a region of size `, that is at least moderately coherent over
this region”. Although the shape of an eddy is not precise, we often think of it as being roughly
circular in cross-section. Because an eddy is embedded in an unsteady, three-dimensional flow, it
is in constant motion. Furthermore, the eddy will be stretched and twisted until it breaks up into
other smaller eddies. The notion that the eddy is coherent means that it exists for a period of time,
1.2 Why is Modeling Necessary? 3

Figure 2: Turbulent wake behind a cylinder at Re = 1770. White areas are oil fog illuminated by a
sheet of laser light. Image is from the compilation of Van Dyke [9].

and that while it exists, the fluid that constitutes the eddy, has a temporary degree of organization.
For a high Reynolds number turbulent flow, there will be eddies of many different sizes1 . In fact
a large scale eddy will likely be composed of many smaller scale eddies. All of these eddies, these
somewhat ill-defined, transient structures in the flow, are in constant motion.
In our imaginary duct flow, the largest eddy will be no larger than D/2, where D is the char-
acteristic size of the duct cross-section. The size of the smallest eddy depends on the Reynolds
number of the flow, but in general it will be much smaller than D/2. The size of the smallest eddy
is determined by the mechanism of viscous dissipation.
The largest eddies tumble and interact largely under the influence of inertia, and to some extent
pressure. The largest eddies are virtually immune to viscous effects. As the large eddies break up
into smaller eddies, the role of inertia continues to dominate until the eddies become small enough
that viscous resistance converts their kinetic energy (of translation and rotation) to heat.

Rapid Mixing To complete our visualization of the turbulent duct flow, we turn off the laser
light sheet, and the streams of smoke. We illuminate the duct with a broad beam of light, and
introduce a small puff of smoke in the center of the duct. The puff of smoke is quickly dispersed as
it moves downstream. For a high Reynolds number flow, the dispersal happens so quickly that the
puff might seem to disappear almost as soon as it is introduced. The smoke disappears because it
has been diluted by rapid mixing with the rest of the air in the duct. The smoke particles remain,
but at such a low concentration that they are no longer visible.
The rapid mixing of the turbulent flows is due to bulk transport of fluid. A blob of fluid is
stretched and reshaped by the eddy motion. This mixing is order of magnitudes more effective
than the Brownian motion that is responsible for the transport properties that we call viscosity and
thermal conductivity.

1.2 Why is Modeling Necessary?


Suppose that to simulate a turbulent flow, we attempt to obtain a numerical solution to the unsteady
Navier-Stokes equations. This is the approach of a strategy called Direct Numerical Simulation
(DNS) of turbulence. DNS is only applicable as a research tool at relatively low Reynolds numbers.
The difficulty with DNS is in the spatial grid and time scale requirements to model all the important
motions in a turbulent flow. Furthermore, any direct simulation of turbulence yields an unsteady
flow field. Extraction of mean quantities (say overall drag, or local heat transfer rate) requires the
time (or ensemble) averaging of a vast amount of data.
1 The meaning of “high” is somewhat problem-dependent. Let us just assume that the velocity and length scales
of the problem are sufficient that the flow is very far from being laminar.
2 REYNOLDS AVERAGED EQUATIONS 4

Table 1: Computational cost for direct numerical simulation of isotropic turbulence. Adapted from
Pope [5, Table 9.2, p. 349]. In November 2005, the Blue Gene/L computer was the fastest computer
in the world. Blue Gene/L has 131072 processors and achieved a sustained 280 TFlops on the
LINPACK benchmark. 1 Tflop = 1012 floating point operations per second. A high end personal
computer has a theoretical performance of between 1 and 10 Gflop. 1 Gflop = 109 floating point
operations per second.
CPU Time
Supercomputer Blue Gene/L
ReL N N3 PC at 1 Gflop
at 1 Tflop at 280 Tflop
6
100 52 1.1 × 10 3 minutes 1.94 µsec 0.2 µsec
1000 291 1.0 × 107 47 hours 0.05 sec 0.2 msec
10,000 1640 1.2 × 108 5.3 years 47 sec 0.2 sec
9
100,000 9200 2.0 × 10 5300 years 5.3 years 6.9 days
10
1,000,000 52000 3.8 × 10 53,000 centuries 5300 years 19 years

Pope [5, § 9.1.2] presents the following estimates of the computational cost of using DNS to
simulate a homogeneous turbulent flow. Homogeneous turbulence is an idealization of turbulent
flow far from any bounding walls, and free of any shear or other flow structure. For a homogeneous
turbulent flow in a region of size L, and having maximum mean velocity U , the Reynolds number
is Re = U L/ν. Let N be the number of points necessary to resolve all relevant flow scales in any
one coordinate direction. The total number of mesh points necessary to simulate the flow is
9/4
N 3 ∼ 4.4ReL
If the simulation is performed on a computer with a sustained throughput of one Gigaflop (109
floating point operations per second) the time to complete the simulate in days is
 3
ReL
TG ∼
800
Evaluating these formulas for a range of ReL gives the results in Table 1. Clearly, DNS is not
applicable for engineering design work.

2 Reynolds Averaged Equations


The governing equations for flow in Cartesian coordinates are the continuity equation
∂ρ ∂ ∂ ∂
+ (ρu) + (ρv) + (ρw) = 0 (1)
∂t ∂x ∂y ∂z
and the momentum equations (in conservative form, no body forces)
∂ ∂ ∂ ∂ ∂p ∂τxx ∂τxy ∂τxz
(ρu) + (ρuu) + (ρvu) + (ρwu) = − + + + (2)
∂t ∂x ∂y ∂z ∂x ∂x ∂y ∂z
∂ ∂ ∂ ∂ ∂p ∂τyx ∂τyy ∂τyz
(ρv) + (ρuv) + (ρvv) + (ρwv) = − + + + (3)
∂t ∂x ∂y ∂z ∂y ∂x ∂y ∂z
∂ ∂ ∂ ∂ ∂p ∂τzx ∂τzy ∂τzz
(ρw) + (ρuw) + (ρvw) + (ρww) = − + + + (4)
∂t ∂x ∂y ∂z ∂z ∂x ∂y ∂z
2.1 Reynolds Decomposition 5

where the velocity vector is u = êx u + êy v + êz w, and the shear stresses are
     
∂u 2 ∂v 2 ∂w 2
τxx = µ 2 − (∇ · u) τyy = µ 2 − (∇ · u) τzz = µ 2 − (∇ · u)
∂x 3 ∂y 3 ∂z 3
     
∂u ∂v ∂u ∂w ∂v ∂w
τxy = τyx = µ + τxz = τzx = µ + τyz = τzy = µ +
∂y ∂x ∂z ∂x ∂z ∂y

2.1 Reynolds Decomposition


Figure 1 shows a short sample of a turbulent signal. The Reynolds decomposition identifies an
average value U and a fluctuating component u0 such that the instantaneous signal is

u = U + u0 (5)

The average of a signal is also designated with an overbar, i.e.

ū = U

By definition of the average


u0 = 0
The governing equations are averaged with the Reynolds averaging rules which apply to any two
functions f and g
1. f + g = f¯ + ḡ
2. af = af¯ (a is a constant)

∂f ∂ f¯
3. =
∂s ∂s
4. f¯g = f¯ḡ

It is possible to define averages that do and do not satisfy these rules.


A common form of averaging is time-averaging, which is defined by
Z t0 +T
U = lim (1/T ) u(t) dt (6)
T →∞ t0

In general both the average and fluctuating component are functions of space. For time averaging,
as defined by Equation (6) the average velocity will not be a function of time, i.e., the average flow
is steady. Some turbulent flows are unsteady in the mean, for example the flow inside the cylinder
of an IC engine. In these cases the ensemble average (see below) is used.
The averaging process has the important property that although u0 = 0,

u0 u0 6= 0

Terms like u0i u0j are called Reynolds stresses, and arise from the to process of Reynolds averaging of
the momentum equations.
2.2 Ensemble Averaging 6

6
Sample 1
4

2
0 1 2 3 4 5 6 7 8
6
Sample 2
4

2
0 1 2 3 4 5 6 7 8
6
Sample 3
4

2
0 1 2 3 4 5 6 7 8
6
Ensemble
4

2
0 1 2 3 4 5 6 7 8

Figure 3: Ensemble averaging of a velocity at a point in a turbulent flow with a periodic forcing
function.

2.2 Ensemble Averaging


Consider an unsteady flow with a periodic forcing function (e.g. sinusoidally varying inlet pressure
or inlet velocity, or a boundary that moves with simple harmonic motion). Figure 3 depicts three
samples (top three plots) and the ensemble average of the velocity measured at a point. If the
velocity vector is u = u(x, t), then the ensemble averaged velocity is
n
1X
U(x, t) = lim ui (x, t) (7)
n→∞ n
i=1

where n is the total number of samples and i is the index of the sample. U(x, t) is the ensemble
average of u at a fixed location, x.
Ensemble averaging satisfies the Reynolds averaging rules. By applying the Reynolds averag-
ing rules to the governing equations, a new set of equations is obtained. These equations have
ensemble averaged velocities (U, V, W ) and second order turbulence correlations as dependent vari-
ables. Note that if ensemble averaging is used, the average velocities will be functions of time, e.g.,
U = U (x, y, z, t). The ensemble averaging has achieved a separation of the fluctuations from the
average quantities.

2.2.1 Example of Reynolds Averaging


Each component of the velocity field is assumed to fluctuate so that we can identify three pairs of
average velocity (components) and fluctuating components

u = U + u0 v = V + v0 w = W + w0
2.2 Ensemble Averaging 7

Apply the Reynolds averaging rules to the left side of Equation (2). For simplicity, assume that
there are no fluctuations of pressure or density, i.e., p0 = 0 and ρ0 = 0. Consider each term

∂ ∂
(ρu) = (ρu) rule 3
∂t ∂t

= (ρū) rule 2, with ρ = constant
∂t

= (ρU ) by definition of U
∂t

∂ ∂ h i
(ρuu) = ρ(U + u0 )(U + u0 ) rule 3
∂x ∂x
∂  
= ρ U U + 2U u0 + u0 u0 rule 2, with ρ = constant
∂x
∂  
= ρU U + ρu0 2 by definition of U , and since U u0 = U u0 = 0
∂x
Similarly,
∂ ∂ 
(ρvu) = ρV U + ρu0 v 0
∂y ∂y
∂ ∂ 
(ρwu) = ρW U + ρu0 w0
∂z ∂z
Thus,

∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
(ρu) + (ρuu) + (ρvu) + (ρwu) = (ρU ) + (ρU U ) + (ρV U ) + (ρW U )
∂t ∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂  ∂  ∂ 
+ ρu0 u0 + ρu0 v 0 + ρu0 w0
∂x ∂y ∂z
The second line of the preceding equation consists of second order correlations that are a direct result
of the averaging process. These second order correlations of the fluctuating velocity components are
called the Reynolds stresses.
Applying the Reynolds averaging rules to the right hand side of Equation (2) gives

∂ p̄ ∂τxx ∂τxy ∂τxz ∂p ∂ τ̄xx ∂ τ̄xy ∂ τ̄xz


− + + + =− + + +
∂x ∂x ∂y ∂z ∂x ∂x ∂y ∂z
Notice that no extra correlation terms appears due to averaging of the right hand side.
Putting the averaged x-direction momentum equation back together and moving the Reynolds
stresses to the right hand side gives.
∂ ∂ ∂ ∂
(ρU ) + (ρU U ) + (ρV U ) + (ρW U ) (8)
∂t ∂x ∂y ∂z
∂p ∂   ∂   ∂  
=− + τ̄xx − ρu0 u0 + τ̄xy − ρu0 v 0 + τ̄xz − ρu0 w0
∂x ∂x ∂y ∂z

Equation (8) suggests how the Reynolds stresses get their name. Averaging the governing equations
(to simplify the analysis) results in additional terms that appear as stresses.
2.3 Flow That Is “Steady in the Mean” 8

Repeating the averaging process for the continuity and the y and z momentum equations gives
∂ρ ∂ ∂ ∂
+ (ρU ) + (ρV ) + (ρW ) = 0 (9)
∂t ∂x ∂y ∂z
∂ ∂ ∂ ∂
(ρV ) + (ρU V ) + (ρV V ) + (ρW V ) (10)
∂t ∂x ∂y ∂z
∂p ∂   ∂   ∂  
=− + τ̄yx − ρv 0 u0 + τ̄yy − ρv 0 v 0 + τ̄yz − ρu0 w0
∂y ∂x ∂y ∂z
∂ ∂ ∂ ∂
(ρW ) + (ρU W ) + (ρV W ) + (ρW W ) (11)
∂t ∂x ∂y ∂z
∂p ∂   ∂   ∂  
=− + τ̄zx − ρw0 u0 + τ̄zy − ρw0 v 0 + τ̄zz − ρw0 w0
∂z ∂x ∂y ∂z

2.3 Flow That Is “Steady in the Mean”


In the preceding section the governing equations were transformed by performing an ensemble
average. Now consider a flow that lacks any periodic forcing function and that does not exhibit
large scale periodic motion of any kind. This is the case in many practical situations, for example,
in turbulent flow in ducts. One expects that while the unsteady turbulent fluctuations exist, the
average velocity field appears to be steady. Such is the case depicted by the velocity signal in
Figure 1.
When time averaging of the velocity field at each point in the flow yields constant value of U ,
V , W and p, then the flow is said to be steady in the mean. For such flows it is appropriate to use
Equation (6) to perform Reynolds averaging of the governing equations. The result is the same as
Equations (8) through (11) with the exception of the time derivatives, which vanish.
If the flow is steady in the mean then the governing equations are simplified. Although the
Reynolds stresses are still present, at least the average velocity field is steady. If appropriate models
for the Reynolds stresses can be introduced, a steady numerical solution can be obtained even
though the flow is characterized by a high degree of fluctuation in the velocity signal. This is the
heart of turbulence models used in CFD codes.

2.4 The Closure Problem


The Reynolds averaging process has created new dependent variables — the Reynolds stresses —
without adding to the set of governing equations. This is called the turbulence closure problem. For
three-dimensional, incompressible flow there are six Reynolds stresses, which can be represented by
the following matrix.  0 0 
u u u0 v 0 u0 w0
 u0 v 0 v 0 v 0 v 0 w0 
u0 w0 v 0 w0 w0 w0
The matrix is symmetric because u0 v 0 = v 0 u0 and v 0 w0 = w0 v 0 . The Reynolds stresses are field
variables, i.e., each term in the matrix is a function of space in the flow.

3 Basic Turbulence Modeling


Consult Tannehill [7, § 5.4] and Wilcox [11] for overviews of turbulence modeling applied to CFD
codes. Pope [5, Chapter 10] explains the fundamental deficiencies of this modeling approach.
3.1 Eddy-Viscosity Concept 9

3.1 Eddy-Viscosity Concept


The viscous stress tensor is
  
∂Uj ∂Ui 2 ∂Uk
τij = µ + − δij (12)
∂xi ∂xi 3 ∂xk

By direct analogy we propose (guess, assume, hope) that the Reynolds stresses have the same
form as the viscous stresses. This model introduces a turbulence viscosity µt or eddy viscosity such
that [2, 5, 7]   
0 0 ∂Uj ∂Ui 2 ∂Uk 2
−ρ ui uj = µt + − δij − ρ δij k (13)
∂xi ∂xi 3 ∂xk 3
where k is called the turbulence kinetic energy
1 0 0 1 0 0 
k = uk uk = u u + v 0 v 0 + w0 w0 (14)
2 2
Equation (13) effectively replaces the Reynolds stresses by two new unknowns, k and µt . Both
µt and k vary from point to point in the flow. It is important to realize that Equation (13) is used
because it is convenient, not because it is an accurate model of the Reynolds stresses.
Substituting Equation (12) and Equation (13) into the Reynolds averaged momentum equations,
and simplifying for steady, incompressible flow gives
∂ ∂ ∂
(ρU U ) + (ρV U ) + (ρW U )
∂x ∂y ∂z
     
∂ p̂ ∂ ∂U ∂ ∂U ∂ ∂U
=− + µeff + µeff + µeff + SU (15)
∂x ∂x ∂x ∂y ∂y ∂z ∂z

∂ ∂ ∂
(ρU V ) + (ρV V ) + (ρW V )
∂x ∂y ∂z
     
∂ p̂ ∂ ∂V ∂ ∂V ∂ ∂V
=− + µeff + µeff + µeff + SV (16)
∂y ∂x ∂x ∂y ∂y ∂z ∂z

∂ ∂ ∂
(ρU W ) + (ρV W ) + (ρW W )
∂x ∂y ∂z
     
∂ p̂ ∂ ∂W ∂ ∂W ∂ ∂W
=− + µeff + µeff + µeff + SW (17)
∂z ∂x ∂x ∂y ∂y ∂z ∂z

where
µeff = µ + µt (18)
1 2
p̂ = p − µeff ∇ · u + ρk (19)
3 3
and SU , SV , and SW are additional source terms due to the non-uniform viscosity. For example,
∂µeff ∂V ∂V ∂µeff ∂µeff ∂W ∂W ∂µeff
SU = − + − .
∂y ∂x ∂y ∂x ∂z ∂z ∂z ∂x
Use of the eddy viscosity concept in a CFD code involves replacing the true viscosity, µ, by µeff .
In general µeff  µ. We now turn to the job of computing µt .
3.2 Prandtl Mixing Length Hypothesis 10

Table 2: Comparison of the molecular model of viscosity from the kinetic theory of gases with
Prandtl’s mixing length hypothesis.

molecular viscosity of gases turbulence viscosity


1
µ = ρ`f Vm µt = ρ`m Vt
3
`f = mean free path `m = mixing length

Vm = velocity of molecules Vt = velocity scale of turbulence

3.2 Prandtl Mixing Length Hypothesis


The Prandtl mixing length hypothesis is based on a direct analogy with the molecular theory of
viscosity from the kinetic theory of gases. In general the (true) viscosity is a macroscopic manifes-
tation of the microscopic behavior of a fluid. For a gas, the viscosity arises because momentum is
transferred across velocity gradient by the random motion of gas molecules (see, e.g., Panton [4,
Chapter 6] and Tennekes and Lumley [8]). Prandtl hypothesized that the turbulent eddies enhance
the exchange of momentum (and passive scalars) by augmenting the molecular diffusion process.
The relationship between Prandtl’s mixing length model and the viscosity model from the kinetic
theory of gases is summarized in Table 2.
The mixing length hypothesis has two parts:
1. A model for the turbulence viscosity
ρu0i u0j
µt =
∂U/∂y
where ∂U/∂y is the mean velocity gradient. (The mixing length was first developed for bound-
ary layer flows where the only important velocity gradient is ∂U/∂y.)
2. A model for the turbulent velocity fluctuations

∂U
Vt = `m
∂y

One problem with the mixing length hypothesis is that µt is not defined wherever ∂U/∂y = 0,
e.g. at the centerline of a pipe. Another problem is that the mixing length hypothesis does not
include the influence of upstream events, i.e. turbulence cannot be convected downstream. Yet
another problem is that the mixing length model requires specification of the mixing length. For
some flows, e.g. boundary layers, and the far field of isolated turbulent jets and wakes, experimental
data is available so that specification of a mixing length is possible. For a general flow, especially
one involving strong recirculation, identification of a single mixing length is not possible. The idea
of the mixing length is still useful as a conceptual model, and as a point of reference for more
sophisticated models.

3.3 Constant Eddy Viscosity Model


A simplistic model of turbulence involves asserting that µeff is constant throughout the flow. The
trick, of course, is to specify a value of µeff that is meaningful. The constant eddy viscosity model
3.3 Constant Eddy Viscosity Model 11

will not allow accurate prediction of the smaller scale features of the flows. An estimate of µeff is
obtained by assuming `m is uniform throughout the flow.

3.3.1 Example: Flow in a Pipe


We can easily estimate the magnitude of the eddy viscosity in a pipe. Prandtl’s model is
µt ∼ ρVt `m
In general, the magnitude of the turbulent fluctuations are small compared to the mean velocity in
the pipe.
u0
0.01 ≤ ≤ 0.15
U
To estimate µt , take u0 /U ∼ 0.1 and Vt ∼ u0 so that
Vt ∼ 0.1U
The flow in the pipe contains eddies ranging from half the diameter of the pipe down to the scale
at which kinetic energy is dissipated by viscosity. To make a somewhat arbitrary choice of a single
length scale that characterizes the turbulent mixing, take
D
`m ∼
4
Combining the preceding expressions gives the following estimate of turbulence viscosity
µt = ρ(0.1U )(0.25D) = 0.025ρU D
From the estimate of µt we can compute an effective Reynolds number
ρU D ρU D
Reeff = = = 40
µt 0.025ρU D
Therefore, the constant viscosity model has the effect of modeling the fluid as if it were very viscous.
At first this may seem counter-intuitive, but deeper examination shows that the large apparent
viscosity is consistent with Prandtl’s mixing length hypothesis.
Recall that the mixing length hypothesis is based on an analogy with the model of viscosity
obtained from the kinetic theory of gases. There, it is assumed that the small scale random motion
of individual molecules gives rise to a transport of momentum whenever there is a velocity gradient.
In the mixing length model, fluid blobs transported by turbulent fluctuations take the role of the
individual molecules in the kinetic theory. Thus, turbulent fluctuations enhance the transport of
momentum. The increase in momentum transport is reflected by the increase in the apparent
viscosity.
Though this model is intuitively compelling, the reality of turbulent mixing is not so simple.
Pope [5, § 10.1] uses experimental data and scaling arguments to show that the mechanics of molec-
ular diffusion and turbulent mixing happen on vastly different time scales. Molecular mixing happens
very quickly because the velocity of molecules on the microscopic scale is very high2 . Turbulent
eddies rotate on a time scale comparable to the mean flow velocities. As a consequence, models of
turbulent mixing that rely on mean velocity gradients only work when the mean flow field changes
relatively slowly in the flow direction.
p
2 From the kinetic theory of gases (see, e.g. Wark [10]), the RMS velocity of a molecule of gas is vrms = 3kT /m,
where k = 1.38 × 10 −23 J/K/molecule is Boltzmann’s constant, T is the absolute temperature of the gas, m = M/NA
is the mass per molecule, M is the molar mass of the gas, and NA = 6.024 × 1026 atoms/k-mol is Avogadro’s number.
For Nitrogen (M = 28 kg/k-mol) at 294 K, vrms = 511 m/s.
4 THE K −  MODEL 12

3.3.2 Constant Turbulence Viscosity Models in CFD Codes


The constant viscosity model attempts to simulate the effect of turbulent fluctuations by increasing
the apparent viscosity of the fluid. This is a crude interpretation of the mixing length model, for it
assumes that the mixing length is uniform throughout the flow. In fact, turbulent transport occurs
due to eddies of many different sizes. Near walls, the eddies are smaller. Downstream protruding
objects such as cylinders or other bluff bodies, the eddy size is strongly influenced by the size of the
object. Shear layers, e.g. where two streams of different velocity mix, create high levels of turbulent
fluctuations. All of these flow features are not accounted for by a model that assumes a uniform
turbulence viscosity.
Though the constant turbulence viscosity model cannot be expected to yield accurate simulation
of many flow features, it is still useful in CFD codes. The constant viscosity model is especially help-
ful for preliminary analysis, and for developing an initial guess at the flow field. More sophisticated
turbulence models require significantly greater computational effort. In addition, more complex
models may also introduce convergence problems for the calculations. Therefore, when beginning
a new CFD simulation, it is helpful to use the constant turbulence viscosity model to generate a
first guess at the flow field. The computations can then be restarted with a more sophisticated
turbulence model after the large scale features of the flow have been established.

4 The k −  Model
The k −  model was first proposed by Jones and Launder [3]. It is now consider the standard
turbulence model for engineering simulation of flows.
The modified Boussinesq eddy viscosity model overcomes the first problem of the mixing length
hypothesis, viz. that µt is not defined in regions of zero shear (∂U/∂y = 0). To relate µt to the
Reynolds stresses, and assume that √
Vt ∝ k
so that
µt = C ρ`m k 1/2 (20)
where C is a constant. Using this model µt is nonzero everywhere in the flow that k is nonzero.
The new independent variables of the turbulence model are `m and k.

4.1 The k Equation


An exact equation for k is obtained by taking the inner product of the velocity vector and the
momentum equation (in vector form). The result after some algebra is a conservation equation for
k (see [2]).
"  #
Dk ∂ p ∂Uj
=− u0 +k − u0i u0j
Dt ∂xi i ρ ∂x
| {z } | {z i}
production by deformation
convective diffusion
"  #
0 ∂uj ∂u0j
 0 
∂ 0 ∂u i ∂uj ∂ui
+ ν u + −ν +
∂xi j ∂xj ∂xi ∂xj ∂xi ∂xi
| {z } | {z }
production by viscous shear dissipation
4.2 The  Equation 13

Like the Reynolds averaged equations this equation also has higher order correlations. The solution
is to model these correlations. The standard form of the model is (see [7])

ρ k 3/2
     
Dk ∂ µeff ∂k ∂Ui ∂Uj 2 ∂Uj
= + µt + − ρ δij k − cD (21)
Dt ∂xi σk ∂xi ∂xj ∂xi 3 ∂xi `m
| {z } | {z } | {z }
diffusion production dissipation

Note that the model equation for k includes the mixing length, `m in the dissipation term.
Equation (21) is a conservation equation for k. The left hand side has terms for storage and
convection. These terms are balanced on the right hand side by diffusion, and source terms due to
turbulence production (source) and turbulence dissipation (sink).
If `m were known, then we could obtain a numerical solution to Equation (21) using the same
approximations applied to the momentum equations. The result would be a spatial distribution of
k on the computational mesh. From the discrete k(x, y, z) field, Equation (20) would be used to
compute the local µt , which in turn would be used in the discrete form of the momentum equations.
The net effect is that solution of the momentum equations would require simultaneous solution of
the k equation. This would be an improvement over the constant viscosity model.
The question remains, “How can `m be specified?”

4.2 The  Equation


In a turbulent flow, some mechanism or device is responsible for adding energy to the fluid. In a duct
flow, a fan converts electrical energy to work, and imparts momentum to the fluid. In atmospheric
flows, e.g. the weather in the troposphere, solar heating causes large columns of air to rise, while
precipitation induces downdrafts. These external energy inputs cause large scale motions in the
fluid. Motions on a large scale correspond to large Reynolds numbers. The organization of these
large scale motions is not stable, i.e. a large swirling vortex tends to break up into smaller vortices.
High Reynolds number turbulent flows have motions that range over many length scales. The
largest scales correspond to the scale of the energy input mechanism or to the size of the bounding
walls. At the large scales, viscosity is unimportant. In other words, the large vortex structures exist
and breakup into smaller scale vortices under the influence of pressure and inertia alone. Viscous
effects are important predominantly at the smallest scales of motion. At the smallest scales the
velocity gradients are smoothed out by the dissipative effect of viscosity.
In a typical turbulent flow, the kinetic energy of the largest scale motions is transferred to
successively smaller scale motions without loss. It is only at the small scales that energy dissipation
occurs.
We can use a simple scaling argument to estimate the dissipation rate (see Panton [4, § 22.10]).
Consider a flow where u0 is linear velocity associated with the largest eddy. The kinetic energy
of this eddy is proportional to u20 . If the diameter of the eddy is L, then the eddy completes a
revolution in time L/u0 . Let  be the rate at which energy is dissipated. An upper estimate of  is

kinetic energy of an eddy u20 u3


∼ = = 0 (22)
time for one rotation L/u0 L

Ultimately all of this energy is dissipated. In a steady flow the amount of energy dissipated is equal
to the amount of energy that must be continually supplied to the flow.
In a turbulent flow the energy is dissipated at the smallest scales. Thus, although Equation (22)
provides an estimate for the magnitude of the dissipation rate, it does not correspond to the actual
dissipation mechanism. The turbulence kinetic energy, k is a measure of the energy in the velocity
4.3 Calculation of Effective Viscosity 14

fluctuations. If the turbulence is isotropic, i.e. if |u0 | = |v 0 | = |w0 | then3


3 0 0
k= uu isotropic turbulence
2

and we can estimate the velocity of the fluctuations with |u0 | ∼ k. Furthermore if we define ` as
the dissipation length scale, i.e. the length scale on which the viscous dissipation mechanism occurs,
then we can write the dissipation rate as

k 3/2
 ∼ CD (23)
`
where CD is to be a constant that is assumed adjust the magnitude of the right hand side to the true
magnitude of . Equation 23 is a scaling relationship, not the equation that allows us to compute
the distribution of .
We can develop a relationship between µt and  as follows. Solve Equation (23) for ` .

k 3/2
` ∼ CD (24)

Use ` as an indication of the scale for `m . We do not assert that ` and `m are equal. Rather
assume that their ratio is constant. Thus, if `m = C 0 ` , where C 0 is a constant, then we use the
following estimates in Prandtl’s mixing length model
µt ∼ ρVt `m Prandtl

Vt ∼ k isotropic turbulence
k 3/2
` ∼ CD scale estimate

to get
k 3/2  1/2 
 
µt = Cρ CD k

or
k2
µt = Cµ ρ (25)

The standard form of the modeled conservation equation for  is (check this?)

ρ 2
     
D ∂ µeff ∂ ∂Ui ∂Uj 2 ∂Uj
= + c,1 µt + − ρ δij k − c,1 (26)
Dt ∂xi σ ∂xi ∂xj ∂xi 3 ∂xi k
| {z } | {z } | {z }
diffusion production dissipation

4.3 Calculation of Effective Viscosity


A typical turbulent flow would involve the following iterative loop:
1. Set-up and solve equations for the velocities and pressure, using the current guess at the
effective viscosity.
2. Solve the k and  equations
3 |u0 | = |v 0 | = |w0 | is a consequence of isotropy, not the definition of it.
4.4 Wall Functions 15

ue

overlap layer
(a.k.a. log region,
inerial sublayer)
<
δ ~ 0.1δ 30 ~ y+

buffer layer
8 <~ y+ <~ 30

viscous sublayer
~ 0.1δ 0 ² y+ <~8

Figure 4: Velocity profile in a turbulent boundary layer.

3. At each point in the flow compute


k2
µt = Cµ ρ

4. Use µt as appropriate to compute the effective diffusion coefficient for each dependent variable.
5. Return to step 1 until convergence.

4.4 Wall Functions


In the basic CFD model of turbulent flow, the boundary conditions for the dependent variables (u, v,
w, T , k, ) are implemented with wall functions. Wall functions are derived from a semi-empirical
model of turbulent boundary layer flow called the law-of-the-wall. More sophisticated near-wall
treatments are available.
As depicted in Figure 5 the k −  model applies to the central part of the calculation domain.
Because velocity and temperature gradients are very steep near the wall, it is often impractical to
resolve all the details of the flow in the near-wall region. Wall functions are an economical (in terms
of mesh points and CPU effort) way to bridge the region between the true wall boundary values and
the turbulent core flow. Before discussing the implementation of wall functions we will first review
the features of turbulent boundary layers and introduce the law-of-the-wall.

4.4.1 Law of the Wall


See Panton [4] for a clear and concise introduction to the scaling laws for turbulent boundary layers.
Panton also gives a nice introduction to turbulent flow in general. More detailed discussions are
provided by Hinze [2] and Pope [5].
Figure 4 depicts the profile of the mean velocity in a turbulent boundary layer. The boundary
layer is a region of thickness δ over which the velocity varies from 0 to ue , where ue is the local
velocity of the external flow. The external velocity is assumed to vary slowly with position, i.e.
ue is not expected to change much over distances of the order of, say, 10δ. When this is true,
the boundary layer will have a universal structure that is the same in many different flows. This
assumption of universal structure is at the heart of the near wall model of turbulence used in the
common CFD models of turbulent flow.
4.4 Wall Functions 16

In the left half of Figure 4 the velocity profile is divided into two regions: an inner region between
the wall and roughly 0.1δ, and an outer region between approximately 0.1δ and the free stream.
The inner region is further divided into three layers as shown in the right half of Figure 4. Wall
functions are computational models used to bridge the numerical approximation in the turbulent
core flow (the edge of which corresponds to the outer region), and the wall without resolving the
details of the inner region. In other words, wall functions are models of the turbulent momentum
transport in the region 0 ≤ y ≤ 0.1δ.
The extent of the sublayers of the inner region are characterized by a dimensionless variable
yu∗
y+ = (27)
ν
where y is the distance normal to the wall, u∗ is called the friction velocity and ν is the kinematic
viscosity of the fluid. Note that y + has the same form as a Reynolds number where the characteristic
velocity is u∗ instead of ue .
The delineation of an inner region (y < 0.1δ) and outer region (y > 0.1δ) reflects a fundamental
truth about turbulent boundary layers. The outer region is characterized by the flow outside the
boundary layer, and the inner region is characterized by the conditions at the wall. Near the wall
the velocity profile has the functional form (see, e.g., [4])

u = fin (y, ρ, ν, τw , w ) (28)

where τw is the wall shear stress, and w is a length scale characterizing the wall roughness4 . In the
outer region the velocity profile has the functional form
 
dp
u − ue = fout y, δ, ρ, , τw (29)
dx

Comparing Equation (28) and (29) we see that the near wall region is insensitive to the external
pressure gradient, the boundary layer thickness, δ, and the external velocity, ue . Also note that the
wall shear stress, τw , affects both the inner and outer profiles.
The dimensionless form of the inner part of the profile (Equation (28)) is

u+ = Fin (y + , +
w) (30)

where
u
u+ = (31)
u∗
and the friction velocity is defined by r
τw
u∗ = (32)
ρ
Note that τw has the units of pressure (stress), and (by definition of u∗ ) τw = ρu2∗ . The dimensionless
wall roughness is
w u∗
+
w = (33)
ν
Equation (30) is called the law-of-the-wall. This so-called law is a semi-empirical model which
describes the near-wall behavior of boundary layers and flow in ducts. The law-of-the-wall covers
the three sublayers depicted in the right half of Figure 4.
4 Do not confuse the roughness scale, w , with the turbulence dissipation rate, , used in the k −  model.
4.4 Wall Functions 17

Wall functions are used to compute μeff near solid surfaces.

k-ε model is used to compute μeff


in the central region of the flow

Figure 5: Wall functions apply to thin layers between solid walls and the central part of the flow
domain.

In the viscous sublayer the shear stress is dominated by viscous forces5 . The velocity profile in
the viscous sublayer is
u+ = y + or u=y (34)
i.e., the mean velocity profile in the viscous sublayer is nominally linear. Equation (34) is hard to
verify experimentally because the region y + < 5 is very close to the wall. It is difficult to build a
probe that is small enough to accurately measure the velocity in this region. We do know, however,
that all velocity fluctuations are zero at the wall so the importance of the Reynolds stresses vanish
as y → 0.
The middle part of the inner layer is called the buffer layer. It has no simple correlating equation.
The outer part of the inner layer is called the overlap layer or the log-law layer. The log-law layer
is correlated by an equation of the form

u+ = c1 ln y + + c2 (35)

where c1 and c2 are constants.


The dimensionless form of the outer part of the profile (Equation (29)) is
 
u − ue y δ dp
= Fout , 2 (36)
u∗ δ ρu∗ dx

In a CFD code the outer part of the velocity profile is not used.

4.4.2 Wall Function Boundary Conditions


Wall functions provide a computational glue layer that allow the turbulent core flow to be linked
to the physical boundary conditions at the wall of the domain. Wall functions are used because it
is often not economical to resolve the structure of the turbulent boundary layer, especially in the
sublayers close to the wall. Simulation of the full turbulent boundary layer would require many
5 Some older texts refer to this as the “laminar” sublayer. It is misleading to consider this a laminar layer because

the flow in the sublayer is characterized by slower motions driven by the turbulent eddies outside of the viscous
sublayer. Although viscous forces are dominant in this sublayer, the flow in the sublayer is not necessarily smooth
and sheet-like — the defining characteristic of laminar flow.
4.4 Wall Functions 18

I
yI
ures
B

Figure 6: Wall functions model the variation of dependent variables between nodes adjacent to a
solid boundaries. Node B is on the boundary. Node I is the nearest interior node close to node B.

nodes (cells) to be placed very close to the wall. This consumes memory (to store element and node
data) and increases the CPU time to achieve a converged solution6 . Instead of resolving the details
of the wall boundary layers, wall functions attempt to impose the boundary condition information
on the core flow in such a way that the correct flux of momentum (drag) and heat at the wall is
obtained. One way to look at wall functions is to consider them as a computational device for
specifying µeff for nodes on and near the wall. The µeff near the wall needs to be modified by the
presence of the wall, and wall functions achieve approximate values of µeff from the law-of-the-wall.
Figure 5 depicts the flow very close to a solid wall. The circles labeled “B” and “I” represent
a node on the boundary, and the nearest internal node on the computational mesh. In the wall
function model, node I is assumed to lie outside of the log-law region, i.e. yI+ > 30. The purpose
of the wall function is to model the wall shear stress from the value of the mean velocity at node I.
The model enables a prediction of the wall shear stress (and wall heat flux if there energy transport
is being simulated) without resolving the detailed structure of the boundary layer with many nodes
between node B and node I.
The log-law part (cf. Equation (35)) of the law-of-the-wall can be written (see e.g. Rodi [6])
ures 1
= ln(y + E) (37)
u∗ κ
where ures is the resultant velocity parallel to the wall, κ = 0.4 is von Kármán’s constant, and E
is the roughness parameter (E = 9.0 for smooth walls). Use a finite-difference model to relate the
wall shear stress to the velocity gradient in the y direction

∂u uI − uB
τw = µeff ≈ µeff (38)
∂y w yI
where µeff is the effective viscosity. Note that the velocity profile between node B and node I is
assumed to look like the right half of Figure 4. The role of the wall function is to provide the estimate
of µeff that compensates for the crude estimate of the velocity gradient given by Equation (38).
Since the no-slip condition requires uB = 0, Equation (38) can be rearranged as
yI
µeff = τw
uI
Using the definitions of y + , u+ , and u∗ the preceding equation becomes
(yI+ ν/u∗ ) τw y + y+
µeff = τw + = 2 + = µ I+ (39)
uI u∗ u∗ u uI
6 Because the solution is a superlinear function of the total number of cells in the mesh.
5 SUMMARY 19

where yI+ and u+ I are evaluated at the interior point, “I”.


Evaluation of µeff from Equation (39) depends on the value of yI+ . Ideally the first interior point
is far enough from the wall so that yI+ > 30. In that case, Equation (37) is used to relate yI+ to u+ I ,
i.e.
1 yI+ κyI+
u+
I = ln(y +
I E) =⇒ =
κ u+I ln(yI+ E)
If yI+ < 30 the wall function model breaks down because there is no universal functional relationship
for the buffer layer (8 ≤ y + ≤ 30). When yI+ < 30 most CFD codes print a warning message. Some
codes apply ad-hoc analytical corrections to the wall function boundary condition. If yI+ < 5
Equation (34) may be used, i.e,
µeff = µ if yI+ < 5
Otherwise
κµy +
µeff = if yI+ > 30
ln(Ey + )

5 Summary
1. Reynolds averaging leads to equations governing the mean or ensemble averaged flow field
2. Reynolds averaging also introduces the Reynolds stresses and creates the closure problem

3. Closure in the k −  model is achieved by


a. relating −ρ u0i u0j to µt
b. using model equations to evaluate µt
For the constant eddy-viscosity model, µt is assumed to be uniform throughout the flow field.
For the k −  model µt is a local function of the k and  fields. Wall functions are used to
compute the value of µt near solid boundaries for the k −  model.
4. Eddy-viscosity models are not accurate for flows with anisotropic Reynolds stresses, and flows
for which the local production and dissipation of turbulence kinetic energy is not in balance.

5. k is the turbulence kinetic energy


• k has a direct physical significance
• there is an exact equation for k, but it has higher order correlations that cannot be
directly computed
• In the k −  model, the “k” is an approximation to the true k
. turbulence is assumed to be isotropic
. k represents energy contained in eddies of many sizes
. k is governed by a transport equation with terms that are models of the exact terms
6.  is the turbulence dissipation rate

•  represents the action of the small eddies that are responsible for dissipating the kinetic
energy of turbulence into heat.
• an exact differential equation for  can be derived, but it contains higher order correlations
that cannot be directly computed
6 ADVICE 20

• In the k −  model, the “” equation is an approximation to the true  equation


7. Boundary conditions in a k −  model are often implemented by wall functions
• The turbulent boundary layer model is used to avoid resolving steep gradients near walls

6 Advice
Turbulence modeling is an inexact art. Most commercial codes offer a variety of turbulence models.
Here we have described the standard k −  model. We have barely scratched the surface of this
important topic.
It is important that you do additional research on your application to determine which turbulence
model should be used. You should not expect the k −  model to give good results in flow with
rapid streamwise changes in the mean flow variables, or in flows with strong swirling components.
Other models to investigate are the k − ω model and Reynolds stress models. Many commercial
CFD codes include these models as options. The k − ω model has some advantage over the standard
k −  model for simulating near-wall flows and flows with strong streamwise pressure gradients. The
Reynolds stress models are theoretically more sound because they do not rely on the eddy viscosity
model to relate the Reynolds stresses to the mean flow gradients.
Research in this area is continuing, and you should expect to be faced with more choices in
turbulence modeling in the future. Above all, remember this: just because you have obtained a
simulation of a turbulent flow with a turbulence model, in no way guarantees that this simulation
is accurate.

References
[1] Joel H. Ferziger and Milovan Perić. Computational Methods for Fluid Dynamics. Springer-
Verlag, Berlin, third edition, 2001.
[2] J.O. Hinze. Turbulence. McGraw-Hill, New York, second edition, 1975.
[3] W.P. Jones and B. E. Launder. The prediction of laminarization with a two-equation model of
turbulence. International Journal of Heat and Mass Transfer, 15:301–314, 1972.
[4] Ronald L. Panton. Incompressible Flow. Wiley, New York, second edition, 1996.
[5] Stephen B. Pope. Turbulent Flows. Cambridge University Press, Cambridge, UK, 2000.
[6] Wolfgang Rodi. Turbulence Models and Their Application in Hydraulics. International Associ-
ation for Hydraulic Research, Delft, Netherlands, second edition, 1984.
[7] John C. Tannehill, Dale A. Anderson, and Richard A. Pletcher. Computational Fluid Mechanics
and Heat Transfer. Taylor and Francis, Washington, D.C., second edition, 1997.
[8] H. Tennekes and J.L. Lumley. A First Course in Turbulence. MIT Press, Cambridge, MA,
1972.
[9] Milton Van Dyke, editor. An Album of Fluid Motion. The Parabolic Press, Stanford, CA, 1982.
[10] Kenneth Wark. Thermodynamics. McGraw-Hill, New York, third edition, 1977.
[11] David C Wilcox. Turbulence Modeling for CFD. DCW Industries, Inc., La Cañada, CA, 1993.

Anda mungkin juga menyukai