Download as pdf or txt
Download as pdf or txt
You are on page 1of 75

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/334130163

Turbulence: Class notes

Book · July 2019

CITATIONS READS
0 77

1 author:

Francisco J. Valdes-Parada
Metropolitan Autonomous University
133 PUBLICATIONS   1,183 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Porous Media View project

All content following this page was uploaded by Francisco J. Valdes-Parada on 01 July 2019.

The user has requested enhancement of the downloaded file.


Turbulence:
Class notes

Francisco J. Valdés Parada


iqfv@xanum.uam.mx;
Last update: October 23, 2017
2
Index

Página

1 Introduction 5
1.1 The experiments of Reynolds, Taylor and Bénard . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Reynolds experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Taylor experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.3 Bénard experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Scales of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Richardson’s energy cascade . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Kolmogorov microscale and integral macroscale . . . . . . . . . . . . . . . . . 9
1.3 A glimpse of the closure problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2 The fluid mechanics equations 13


2.1 Mass and momentum equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Mechanical energy equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 The rate of dissipation of energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Vorticity equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4.1 Q and R invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Helmholtz decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Concluding remarks so far . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3 Statistical nature of turbulence 21


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Freely-evolving turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3 Turbulence and memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 A statistical approach for turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Structure functions and energy spectrum . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.6 Another look at Kolmogorov’s theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.7 Probability distribution of the velocity field . . . . . . . . . . . . . . . . . . . . . . . 34

4 Turbulent shear-flows 39
4.1 Reynolds stress and the closure problem . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.2 Closure problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Turbulent flow in a pipe and between two parallel plates . . . . . . . . . . . . . . . . 43
4.4 The eddy-viscosity theories of Boussinesq and Prandtl . . . . . . . . . . . . . . . . . 48
4.5 Energy transfer in turbulent flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3
4.6 Current models of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6.1 The k − ε model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6.2 The k − ω and SST models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6.3 Spalart-Allmaras model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6.4 Reynolds stress model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6.5 DNS and Large-eddy simulation . . . . . . . . . . . . . . . . . . . . . . . . . 54

5 The phenomenology of Taylor-Richardson and Kolmogorov 57


5.1 Richardson’s theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.2 Kolmogorov’s theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3 Turbulent mass transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4 Random walk and Taylor diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.5 Richardson’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

References 71

4
1 | Introduction

Contents
1.1 The experiments of Reynolds, Taylor and Bénard . . . . . . . . . . 5
1.1.1 Reynolds experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Taylor experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.3 Bénard experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Scales of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Richardson’s energy cascade . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Kolmogorov microscale and integral macroscale . . . . . . . . . . . . . 9
1.3 A glimpse of the closure problem . . . . . . . . . . . . . . . . . . . . 11

Turbulence is the last unsolved problem of classical physics and at the moment there is no
satisfactory universal theory to address this problem. Despite this tremendous drawback, there
is need to study problems of flow and transport taking place under turbulent conditions. This
has motivated the emergence of many approximate theories that are currently used and need to
be carefully understood. In this first chapter we briefly describe some classical experiments and
applications of turbulence and make special emphasis on the scales of turbulence and the closure
problem for turbulence.

1.1 The experiments of Reynolds, Taylor and Bénard


1.1.1 Reynolds experiment
Our final classical example of turbulence is the classical experiment performed by Reynolds in 1883.
Reynolds wanted to determine which conditions were sufficient to distinguish between laminar and
turbulent flow. With this in mind he devised an experimental setup in which he made a Newtonian
and incompressible fluid to flow through a pipe under several flow rates. This devise allowed him
to propose his now famous dimensionless number. He observed that the critical value of Re at
which turbulence first appeared is quite sensitive to perturbations at the inlet. For example, he
found that if no particular effort was made to minimize the perturbations, turbulence may appear
for Re ∼ 2000. Nowadays, it is possible to minimize the inlet perturbances and still have laminar
flow Re = 90,000.
Currently, it is common agreement that for Re > 104 , turbulence tends to appear in the annular
boundary layer near the inlet. These initial spots of turbulence merge later on in the pipe unit

5
transition to turbulence. There are flows in which turbulent motion
appears first in small patches. Provided Re is large enough, the

> I
he inlet '
Initiation Formation of turbulent slug Laminar region Turbulent slug

Figure 1.1: Turbulent slugs and laminar region in Reynolds experiment.

occupying a complete cross-section over a determined distance (this portion of the system is known
as turbulent slug as shown in Figure 1.1). After this slug is formed, a laminar region is found and
then another slug is formed near the exit. If Re > 2000, the inlet and outlet slugs tend to grow in
length eventually eliminating the laminar zone and the flow becomes completely turbulent. When
this is not the case, the length of the slugs is quite small and they quickly decay.

1.1.2 Taylor experiment


Consider the Newtonian and incompressible flow of a fluid in the annular gap between two concentric
cylinders as shown in Figure 1.2. The inner circle, of radius R, rotates at an angular velocity Ω.
The gap has a width d and the whole system has a length L. This problem is typically studied
under creeping conditions in the introductory courses of fluid mechanics as Couette flow. If the
speed of rotation is sufficiently low, the fluid within the gap rotates because of the drag from the
inner cylinder. However, at a certain critical speed, toroidal (or Taylor ) vortices appear due to the
centrifugal force. Under creeping flow condition, the only relevant component of the velocity vector
is its angular component, which is obviously a function of r. This is because the centrifugal force is
balanced with the radial pressure gradient so that there is no net radial movement. The fluid has
a natural tendency to move upward, however the viscous forces compensate this tendency.
At the critical rotation rate the viscous forces are no longer capable to contain radial
disturbances and the flow becomes unstable to the slightest perturbation. Naturally, all fluid
cannot go upward because the outer fluid is on its way. Consequently, the flow is separated in
bands (or cells). This behavior is known as a Rayleigh instability. In this regime the flow would
appear to form rings as in a screw.
If the rotation velocity is further increased (say 25% more), Taylor vortices become unstable
and give rise to wavy Taylor vortices. These vortices look like non-axisymmetric waves around the
cylinder. Note that, even at this flow rate the flow is still considered laminar.
Interestingly, when the rotation speed is increased even more and flow becomes fully turbulent,
the flow pattern once again resembles the original Taylor vortices observed at the critical speed.
However, there is now a crucial difference, which is the appearance of chaotic components of motion,
similar to Brownian particles motion.
Certainly, not only the rotation speed, Ω is the only parameter that is crucial for the development
of the vortices. The fluid kinematic viscosity, ν, the angular gap, d, the radius R and length L
of the inner cylinder are also crucial. These factors can be grouped in the following dimensionless

6
an instability of the basic rotary flow. The net motion of a fluid
particle is now helical, confined to a toroidal surface.

(a)

tric
he inner
mes

Taylor vortices Wavy Taylor vortices Turbulent Taylor vortices

Figure 1.2: Development of vorices in the Taylor experiment: a) Critical Taylor vortices, b) Wavy
Taylor vortices and c) Turbulent Taylor vortices.

numbers  2
Ωd d L
Ta = Rd; , (1.1)
ν R R
where T a is the Taylor number. For conditions in which d  R  L, the only parameter that
matters is the Taylor number. The critical value of this number is 1.7 × 103 .

1.1.3 Bénard experiment


Consider now the fluid between two flat parallel plates, which are separated by a distance, d as
shown in Figure 1.3. The lower plate is maintained at a temperature T = T0 + ∆T and the
upper plate is fixed at a temperature T = T0 . This is the typical setup for determining thermal
conductivities of fluids. At sufficiently low values of ∆T , the fluid is stagnant and heat is transferred
from one plate to the other only through conduction. There is a natural tendency of the fluid to
move upwards due to the hot plate heating. However, for sufficiently low values of ∆T , this force
is balanced by the hydrostatic fluid pressure and therefore the fluid remains immobile.
As ∆T is increased, heat is transferred by both diffusion and convection and the hot fluid starts
to rise and the cold fluid tends to go down. The phenomenon gives rise to regular (Bénard ) cells.
Under steady conditions, the rate of the buoyancy force is balanced by viscous dissipation within
the fluid.
If we increase the temperature difference, the potential energy gained by the fluid when it rises
is not exactly compensated by viscous forces, therefore the fluid will accelerate and instabilities are
produced. In this way, the system will be stable if ∆T is sufficiently small and the viscous forces

7
say x0 in a frame of reference travelling with the cylinder. Despite the
nominally identical conditions we find that the function u(x0, t) is quite
different on each occasion. This is because there will always be some
ulence minute difference in the way the experiment is carried out and it is in
the nature of turbulence that these differences are amplified. It is
striking that, although the governing equations of incompressible flow
are quite simple (essentially Newton's second law applied to a con-
tinuum), the exact(b) details of u(x, t) appear to be, to all intents and
' Cold (T= TO)
f \ \ X X X X N X purposes,
\ \ \ irandom and unpredictable.
Now suppose that we do something different. We measure u(x0, t)
for some considerable period of time and then calculate the time-
two flat, No motionaverage of the signal, u(x0). We do this a second time and then a
d
heated. I i
third. Each time we obtain the same value for u(x 0 ), as indicated in
Figure 1.8. The same thing happens if we calculate u2. Evidently,
iescent. As although u(x, t) appears to be random and unpredictable, its statistical
'^^Hot(T = T0 + ATr
sets in, properties are not. This is the first hint that any theory of turbulence
ion cells Low AT has to be a statistical one,Bénard cells
and we shall Turbulent
return to this point time and convection
flow.
again in the chapters which follow. (High AT)
In summary, then, the statistical properties of a turbulent flow are
uniquely determined by the boundary conditions and the initial condi-
Figure 1.3: Bénard experiment consisting
tions. If a sequence of two
of nominally parallel
identical platesareheld
experiments atout,
carried different temperatures.
The first of these groups, Ta, is called the Taylor number. (Different
authors use slightly different definitions of Ta.) When the apparatus is
very long, L ;$> R, and the gap very narrow, d <C R, it happens to be
the value of Ta, and only Ta, which determines the onset of Taylor
vortices. In fact, the critical value of Ta turns out to be 1.70 X 10 and
the axial wave number of the vortices isfe=Realization 27l/l1 = 3.12/a, 1 being
the wavelength.
Figure 1.8 A cylinder is towed through a
quiescent fluid and u isLet
x measuredus now consider a quite different experiment, often called
at location
x . The records of u (t) in two nominally
0 x

quite different. Bénard convection or Rayleigh-Bénard convection. Suppose that a


identical realizations of the experiment are

fluid is held between two large, flat, parallel plates, as shown in


Figure 1.4: Sketch of the velocity profiles near a cylinder that is moved through quiescent fluid.
Figure 1.3(a). The lower plate is maintained at temperature
areTlarge
= TOand+ will ATbeand the asupper
unstable plate oratviscous
∆T increases temperature
forces decrease.T0. At
The low values
instability of
criterion
is dictated by the Rayleigh number:
AT the fluid remains stagnant and heat passes from one plate to the
other by molecular conduction. Ra =
gβ∆TOfd3 course 3there is an upward buoy-
≥ 1.7 × 10 (1.2)
να
ancy force which is greatest near the lower plate and tends to drive the
here β is the thermal expansion coefficient of the fluid and α is the thermal diffusivity of the fluid.
fluid
As in the upward. However,
Taylor experiment, as theattemperature
low values of isAT
difference this above
increased forceitsiscritical
exactly
value,
balancedappear
instabilities by aand vertical pressure
turbulent convective gradient.
vortices are We now inslowly
observed increase
the system. UnderAT.these
conditions we find small portions of fluid moving randomly as well.
AtBeforea critical
moving on, value of AT the
it is convenient fluidabout
to think suddenly starts
the analogies to convert
of these as shown
two experiments. In both
in Figure 1.3(b).
cases, the system is stable for high fluid viscosities. As the viscosity is reduced, the system becomes
unstable to infinitesimal perturbations and the system bifurcates (changes) to a more complicated
state,The whichflow is still consists
laminar, and ofregular
hot rising fluid
flow cells appear.andAs cold fallingis regions.
the viscosity further and This
further
decreased, the flow becomes fully turbulent consisting of a mean component plus random chaotic
takes the form of regular cells, called Bénard cells, which are remin-
motions.
iscent of Taylor vortices.2 The rising fluid near the top of the layer
8
loses its heat to the upper plate by thermal conduction, whereupon it
starts to fall. As this cold fluid approaches the lower plate it starts to
heat up and sooner or later it reverses direction and starts to rise again.
To gain more understanding about the random character of turbulent motion consider the
following experiment: suppose that there is a pool with a Newtonian and incompressible fluid that
contains a vertical cylinder. Initially, the entire system was at rest and then we decide to create
turbulence by moving the cylinder at a fixed speed in order to create a turbulent wake. If we repeat
this experiment 100 times and we measure the velocity as a function of time at fixed a point x0 in a
frame of reference traveling with the cylinder in each experiment we would obtain different results
in each experiment as shown in Figure 1.4. This is because when each experiment is performed, we
can expect minor differences to appear between one realization and another. These tiny differences
are amplified by turbulence. Interestingly, if in the same experiment we plot the time average of
the velocity, say u(x0 ), we would obtain the same value in each experiment. This would also be
the case if we record u2 . In other words, even though u(x, t) seems to become a random variable
under turbulent conditions, its statistical properties are not.
Thus far we can identify the following properties of turbulence:

1. The velocity field fluctuates randomly in time and it is highly disordered in space, exhibiting
a wide range of length scales.

2. Small changes at the initial conditions lead to large changes in the subsequent motion.

3. The statistical properties of the velocity are deterministic fields.

Homework

1. Verify at home at least two of the experiments described in class.

2. Investigate the cause of turbulence of the following phenomena: a solar flare, Jupiter’s
storms, turbulent wakes in buildings and drag effects in planes and cars.

1.2 Scales of turbulence


1.2.1 Richardson’s energy cascade
Consider a process that produces large turbulent eddies. The largest eddies have the largest Re
values and are created by instabilities of the mean flow. They are subject to inertial instabilities
and can break up into smaller vortices that are, most likely, to be also driven by inertial forces and
are also unstable. However, their Reynolds number may be expected to be smaller than the one
of the previous eddies. This cascade of breaking down eddies is mostly driven by inertial forces
because the Reynolds number, although decreasing in each offspring, is considerably large. This
cascade is finite and it stops when the eddy size becomes so small that the Reynolds number is
on the order of unity. At this point, viscous forces are quite important and so is viscous energy
dissipation as illustrated in Figure 1.5.

1.2.2 Kolmogorov microscale and integral macroscale


Let u, ` and v and η represent the velocities and characteristic lengths associated to the largest
and smallest eddies, respectively. The lifespan of an eddy is called the turn-over time and it is on
the order of magnitude of `/u. Since the larger eddies move due to inertial forces, their energy is

9
Different scales in a turbulent flow

© o
First instability

Second instability

0000
Energy flux Third instability

A schematic representation of
scade (after Frisch 1995). See I
's sketch—Plate 3. Viscosity

is destroyed
Figure 1.5:only in theoffinal
Sketch the stages
eddiesof this process
cascade when the structures
in Richardson’s theory
are so fine that Re, based on the small-scale structures, is of order
best describedunity. In this
in terms of thesense viscosity
kinetic energy. plays a rather
Hence, passive
the rate role, the
at which mopping
larger up
eddies pass down
whatever energy cascades down from
their kinetic energy per unit mass can be estimated to be: above.
Let us now see if we can determine the smallest scale  in2 aturbulent
 3
u u
flow.
Rate of Let H and
energy v represent
of the typicalper
largest eddies velocities
unit massassociated
=O with =theO lar- (1.3)
`/u `
gest and smallest eddies respectively. Also, let ! and t] be the length
At the smallestscales of the
scale, largestestablished
we have and smallest structures.
that energy isNow we know
viscously that most
dissipated, since the rate of
viscous dissipation T
∇v ), weof
eddiesis break-up
given by ν(∇von a :timescale have theturn-over
their following time
estimate:
(all of the
experimental evidence confirms that this is so), and so the  rate
νv 2 at

Rateenergy
which of energy
(per of themass)
unit smallest eddiesdown
is passed per unit mass =cascade
the energy O (1.4)
η 2from
the largest eddies is,
Since the energy from the largest eddies is ultimately transferred to the smallest ones, we have:
n ~ u 2 /(i/tt) = tt3/i. u3 νv2
= 2 (1.5)
When conditions are statistically ` steady
η this must match exactly the
In addition, werate
knowof that
dissipation of energy
the Reynolds at theofsmallest
number scales.
the smallest If it is
eddies didabout
not, the
thenunity, therefore,
there would be an accumulation vη of energy at some intermediate scale,
= O(1)we want the statistical struc-
and we exclude this possibilityν because (1.6)
ture of the turbulence to be the same from one moment to the next.
Combining these equations gives rise to the following estimates:
The rate of dissipation of energy at the smallest scales is,
!
νRe3/4
     
` ν u
η=O ; v=O =O =O (1.7)
Re3/4 η ` Re1/4
where Sy is the rate of strain associated with the smallest eddies,
where
y//?. This yields u`
Re = (1.8)
ν

10
Since the dissipation of turbulent energy, e, must match the rate at
which energy enters the cascade, H, we have,

H 3 /l 1.1
For example, in some wind tunnels, we may have Re = 103 and ` = 1cm, it thus follows that
η = 0.056 mm. Indeed, as the Reynolds number is increased, the characteristic size of the smallest
eddies is also decreased. The scales of v and ν are called the Kolmogorov microscales whereas u
and ` are the integral macroscales.

Homework Denoting the rate of viscous energy dissipation as

v2 m2
ε=ν [=]
η2 s3
Demonstrate that the following identities apply:
1/4
ν3

η= ; v = (νε)1/4
ε

1.3 A glimpse of the closure problem


Let us consider the unsteady turbulent flow of a Newtonian and incompressible fluid. The governing
equations for the mass and momentum transport, even under turbulent conditions are,

∇·v=0 (1.9a)

∂v p
+ v · ∇v = −∇ + ν∇2 v (1.9b)
∂t ρ
The Navier-Stokes (NS) equation may be simply recast in the following form:

∂v
= F1 (v, p) (1.10)
∂t
To eliminate the pressure dependence, we may take the divergence of the NS equation, taking into
account the continuity equation, to obtain:
p
∇2 = −∇ · (v · ∇v) (1.11)
ρ
Assuming that this equation has a unique solution, we have

p = g(v) (1.12)

Consequently,
∂v
= F2 (v) (1.13)
∂t
This equation is, in principle, solvable and it should yield the velocity fields at each instant and
at each position, even under turbulent conditions. However, as stressed above, comparison with
experiments is quite complicated because even the smallest variation in initial conditions yields to
quite different velocity profiles. Nevertheless, it appears that the time average of each experiment
is the same. This motivates the derivation of theories that describe the statistical measures of
turbulence. Unfortunately, due to the nonlinear nature of the NS equations, there is no closed

11
theory for average quantities. In order to obtain a glimpse of this problem, let us assume for the
moment that the form of Eq. (1.13) for one component of the NS is:

∂v
= γv 2 (1.14)
∂t
Averaging over several realizations yields:

∂hvi
= γhv 2 i (1.15)
∂t
However, to solve this equation we need to know about hv 2 i. This may be solved my multiplying
Eq. (1.14) by v:
∂v 2
= 2γv 3 (1.16)
∂t
and then averaging
∂hv 2 i
= 2γhv 3 i (1.17)
∂t
However, this equation now depends of hv 3 i, which the would also require a new equation and so
on. This is the reason why the problem for turbulence is unclosed. Many attempts have been made
to produce a closure problem but none have been generally applicable. Researchers are currently
working on producing powerful numerical codes to solve for v without needing to average for every
realization. However, current codes are not able to solve macroscale models for each time and
position for any value of the Reynolds number.

12
2 | The fluid mechanics equations

Contents
2.1 Mass and momentum equations . . . . . . . . . . . . . . . . . . . . . 13
2.2 Mechanical energy equation . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 The rate of dissipation of energy . . . . . . . . . . . . . . . . . . . . 15
2.4 Vorticity equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4.1 Q and R invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Helmholtz decomposition . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 Concluding remarks so far . . . . . . . . . . . . . . . . . . . . . . . . 19

2.1 Mass and momentum equations


The equation of conservation of mass in a fluid is
∂ρ
+ ∇ · ρv = 0 (2.1a)
∂t
which, for incompressible flows reduces to

∇·v=0 (2.1b)

The movement of a fluid is governed by Newton’s second law, or Euler’s first equation, which in a
differential form is known as Cauchy’s equation of motion and it is given by:
Dv
ρ = −∇p + ρg + ∇ · τ (2.1c)
Dt
For incompressible flow, Newton’s viscosity law is

τ = 2µS (2.2)

where S is the strain tensor and it is defined as


1
∇v + ∇vT

S= (2.3)
2
Clearly, this tensor is the symmetric part of the tensor ∇v, which can be decomposed as

∇v = S + Ω (2.4)

13
where we introduced the vorticity tensor:

1
∇v − ∇vT

Ω= (2.5)
2
The strain tensor is physically related to the rate of deformation of a fluid whereas the vorticity
tensor is related to its rate of rotation.
In this way, after taking the divergence of the stress tensor and substituting the result into
Cauchy’s equation gives rise to the well-known Navier-Stokes equations:

∂v
ρ + ρv · ∇v = −∇p + ρg + µ∇2 v (2.6)
∂t
At this point, it is worth noting that, if we take the divergence of this equation, we obtain the
following Poisson equation:
∇2 p = −ρ∇ · (v · ∇v) (2.7)
Let us assume for the moment that the pressure is subject to homogeneous boundary conditions,
using Green’s functions, the solution of the problem is:
Z
p(x) = − G(x − x0 )ρ∇ · (v · ∇v)dV (x0 ) (2.8)
V

If this result is substituted back into the NS equation, it results an integro-differential equation for
the velocity. This indicates that the velocity fields are non-local.

2.2 Mechanical energy equation


Taking the dot product of this equation with the velocity yields

Dv
ρv · = −v · ∇p + ρv · g + µv · ∇2 v (2.9)
Dt
Or, using the identities
Dv 1 Dv 2
v· = (2.10a)
Dt 2 Dt
v · ∇p = ∇ · (pv) (2.10b)
1 2 2
∇ v = v · ∇2 v + ∇v : (∇v)T (2.10c)
2
we obtain the mechanical energy equation:

ρ Dv 2 µ
= −∇ · (pv) + ρv · g + ∇2 v 2 − µ∇v : (∇v)T (2.11)
2 Dt 2
Notice that, on the basis of the continuity equation, we have:

Dv 2 ∂v 2 ∂v 2
= + v · ∇v 2 = + ∇ · (v 2 v) (2.12)
Dt ∂t ∂t

14
In this way, the mechanical energy equation may be written as:

1 ∂v 2
 
1 p
=− ∇ · (v 2 v) − ∇· v + v·g
|2 {z
∂t} 2
| {z } ρ
| {z }
|{z}
rate of volumetric work
rate of accumulation of KE convective transport of KE rate of pressure work
ν 2 2
+ ∇ v − ν∇v : (∇v)T (2.13)
|2 {z } | {z }
rate of viscous energy dissipation
rate of viscous work

2.3 The rate of dissipation of energy


The viscous force of a fluid acting over a surface is:
Z Z
f = n · τ dA = ∇ · τ dV (2.14)
A V

From which, we obtain the local force per unit volume to be:

f̂ = ∇ · τ (2.15)

Hence, the rate of viscous power is


Z Z
W = n · τ · vdA = ∇ · (τ · v)dV (2.16)
A V

From here we obtain the local power per unit volume as:

Ŵ = ∇ · (τ · v) = (∇ · τ ) · v + τ : ∇v = f · v + τ : ∇v (2.17)

In addition, taking into account the symmetry of the stress tensor, we have:

1 1 1
τ : ∇v = (τ + τ T ) : ∇v = (τ : ∇v + τ T : ∇v) = (τ : ∇v + τ : (∇v)T ) = τ : S (2.18)
2 2 2
In this way,
Ŵ = f · v + τ : S (2.19)

The first term on the rhs of this equation is the rate of change of the mechanical energy of the fluid
and the second one is the rate of change of internal energy per unit volume of the fluid. From here,
we can deduce the rate of change of internal energy per unit mass:

1
ε= τ :S (2.20)
ρ
R R
In this way, ρεdV = τ : SdV represents the rate of loss of mechanical energy due to viscous
V V
dissipation.

15
localized blobs of vortici
analogous to the blobs of
street behind a cylinder. T
and spread by diffusion,
remains the same.
Since vorticity cannot
might ask where the vorti
fluid particles upstream
Figure 2.10 Karman street behind a cylinder.
Figure 2.1: Sketch of the streamlines of flow around a sphere at sufficiently large Reynolds number
values that ensure turbulence.
momentum (vorticity) yet
heat is useful. The hot blo
2.4 Vorticity equation
Let us commence our derivations by first noting that
1
v · ∇v = ∇v 2 − v × w (2.21)
2
where w is the vorticity vector, which is defined as:

w=∇×v (2.22)

the vorticity indicates the tendency of a fluid particle to rotate as would be seen by an observer
traveling with the flow. Actually Stokes defined vorticity to be twice the angular velocity of a
fluid particle. It is crucial to understand that the vorticity gives no information about the global
rotation of the fluid. Rather, it only gives information about the rotation of fluid particles.
Using Eq. (2.21), the NS equation takes the form
 2 
∂v v p
+∇ + − v × w = g + ν∇2 v (2.23)
∂t 2 ρ
v2
here 2 + ρp is Bernoulli’s function. Taking the curl (∇×) on both sides of the above equation yields

∂w
= ∇ × (v × w) + ν∇2 w (2.24)
∂t
This result can be further modified by using the identity:

∇ × (v × w) = w · ∇v − v · ∇w (2.25)

Consequently,
Dw
= w · ∇v + ν∇2 w (2.26)
Dt
This equation is clearly simpler than the NS equation because it is clearly local and it does not
depend of the pressure. Hence, when we talk about an eddy on turbulence, we mean a blob of
vorticity and its motion. This motion is governed by the moment of inertia and the viscous torque
of a fluid particle as clearly indicated by the terms on the right-hand side of the above equation.
Before moving on, it is convenient to recall the benchmark example of flow around a sphere.
For sufficiently large Reynolds numbers, we should expect to observe the streamlines depicted in
Figure 2.1. Clearly vortices are generated after the encounter of the fluid with the obstacle. In
other words, vorticity is originated at the surface of the obstacle. This means that the boundary

16
layer is a space in which vortices can be transported by diffusive mechanisms. This observation
follows from Eq. (2.26) by noting that:
 wv 
w · ∇v = O(v · ∇w) = O (2.27a)
L 
 νw
ν∇2 w = O (2.27b)
L2
We know that within the boundary layer viscous forces predominate, and therefore, it is reasonable
to assume that
vL
 1, in the boundary layer (2.28)
ν
Consequently, Eq. (2.26) is reduced to
∂w
= ν∇2 w (2.29)
∂t
indicating that the vortices should be transported mainly by diffusive conditions. In fact, it is
interesting to notice that the above equation has the same structure as the heat transfer equation:
∂T
= α∇2 T (2.30)
∂t
The message to be kept here is that vorticity is intense near the boundary layers where large velocity
gradients are present and decreases outside from the boundary layer where the velocity profiles are
flat. Certainly, for non-viscous flow, the vortices would remain frozen in space and travel across
the system by inertia.
Furthermore, the enstrophy is defined as

w2 = w · w = (∇ × v)2 (2.31)

Taking the dot product of Eq. (2.26) with w leads to

1 Dw2 ν
= w · (w · ∇v) + ∇2 w2 − ν∇w : (∇w)T (2.32)
2 Dt 2
If we compare this equation with the mechanical energy equation, we can conceive the last two
terms are related to the viscous transport of estrophy and by the lost of estrophy due to mechanical
energy. As a final note, it may be convenient to remember the following: enstrophy is to vorticity
what kinetic energy is to velocity.

2.4.1 Q and R invariants


A fundamental question in turbulence is the following: do the local velocity gradients contribute
more to the rate of strain or to the rate of rotation? To answer this question, let us define the
invariant:
1
Q = − ∇v : ∇v (2.33)
2
Since S = 12 (∇v + ∇vT ), we have:
1 1
S : ST = ∇v : ∇v + ∇v : ∇vT (2.34)
2 2

17
In this way:
1 1
∇v : ∇v = S : ST − ∇v : ∇vT (2.35)
2 2
Furthermore, from the enstrophy definition we have:

w2 = ∇v : ∇vT − ∇v : ∇v (2.36)

To obtain this result, we used the identity εijk εlmk = δil δjm − δim δjp . Substitution of the above
result in Eq. (2.35) yields
1
∇v : ∇v = S : ST − w2 (2.37)
2
Consequently:  
1 T 1 2
Q=− S:S − w (2.38)
2 2
this result is often written in a normalized form:
S : ST − 12 w2
Λ= (2.39)
S : ST + 21 w2

In this way, positive values of Λ indicate that flow is dominated by strain and negative values of Λ
indicate that the flow is governed by rotation (or vorticity).
Another invariant that is also frequently used is:
 
1 3 T
R= (S · S) : S + ww : S (2.40)
3 4

Homework

• Demonstrate that the vorticity is a solenoidal field.

• Demonstrate the identities given in eqs. (2.34) and (2.36).

2.5 Helmholtz decomposition


In the previous paragraphs we have highlighted the relevance of vorticity for the study of turbulent
transport, however one should not forget about the velocity fields which is also of primal importance.
To clearly localize the importance of vorticity in the prediction of velocity, it is relevant to use
Helmholtz decomposition, which is stated as follows:

Let a be a vector field, which is twice continuously differentiable. Let a be defined


everywhere in space and vanishing at infinity together with its first derivatives. Then,
a can be decomposed into a rotation-free component and a divergence-free component:

a = ∇φ + b (2.41)

where φ is a potential field that is obviously irrotational, therefore, b is the rotational


part, which is solenoidal.

18
In the case of velocity, we have thus:

v = vp + vw (2.42)

where vp = ∇φ and
∇ × vw = w (2.43)
Since both v and w are solenoidal, it follows that vw is also solenoidal and:

∇2 φ = 0 (2.44)

This is consisting the Helmholtz decomposition, requiring that vw is divergence-free.


Taking the curl of Eq. (2.42) yields

∇ × v = ∇ × ∇φ + ∇ × vw = w (2.45)

Notice that v is only equal to vp for conditions of purely potential flow, i.e., under non-viscous
momentum transport conditions. The relevance of the decomposition given in Eq. (2.42) is that
it allows understanding that the velocity field is the result of the potential flow limit and the
corrections due to vorticity. In other words, the flow is mixed with rotational and irrotational
components at each point.

Homework Use Helmholtz decomposition to show that the Navier-Stokes equation can be
expressed as:
 
∂vω ∂φ ρ
ρ +∇ ρ + |∇φ| + p + ρ∇ · [vω ∇φ + ∇φvω + vω vω ] = µ∇2 vω
2
(2.46)
∂t ∂t 2

This equation is interesting because it shows that if one solves the problem of potential flow,
one could use this result to predict the corrections of the velocity due to vorticity.

2.6 Concluding remarks so far


With the material reviewed here, we might attempt to define turbulence as follows:

Incompressible turbulence is a spatially complex distribution of vorticity, which moves


in a chaotic manner. The vorticity field is random and it exhibits a large distribution
of time and length scales.

Indeed, when we refer to a turbulent eddy, what we mean is a blob of vorticity and the associated
flow. Such an eddy evolves in the velocity field by material and diffusive movements. In other words,
turbulence is a complex mixture of vortex blobs that move according to the vorticity equation.
Hinze proposed the following definition

Turbulent fluid motion is an irregular condition of flow in which the various quantities
show a random variation with time and space coordinates, so that statistically distinct
average values can be discerned.

19
20
3 | Statistical nature of turbulence

Contents
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Freely-evolving turbulence . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3 Turbulence and memory . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 A statistical approach for turbulence . . . . . . . . . . . . . . . . . . 26
3.5 Structure functions and energy spectrum . . . . . . . . . . . . . . . 28
3.6 Another look at Kolmogorov’s theory . . . . . . . . . . . . . . . . . 32
3.7 Probability distribution of the velocity field . . . . . . . . . . . . . 34

3.1 Introduction
The turbulent velocity is a random variable in the sense that different values arise from the same
realization of experiments. Formally, a random variable is an event that is neither certain neither
impossible and therefore it does not have a unique value. A fundamental question in turbulence is
the following: if the NS equations are deterministic, why are the solutions random?
The answer to this question relies in the fact that in any turbulent flow there are unavoidably
perturbations in initial, boundary conditions or in fluid properties to which the flow is highly
sensitive. To have an idea of this situation, let us consider the following set of differential equations
that were originally studied by Lorenz (1963)
dx
= σ(y − x) (3.1a)
dt
dy
= ρx − y − xz (3.1b)
dt
dz
= −βz + xy (3.1c)
dt
Taking σ = 10, β = 8/3 and ρ = 28 and the initial condition:
x(0) = y(0) = z(0) = 0.1 (3.2)
we may numerically solve this system of equations using the Euler method for a tmax = 100 and
N t = 106 . Let us also consider a slight change in the initial conditions, say:
x(0) = 0.100001; y(0) = z(0) = 0.1 (3.3)

21
The numerical solution of the problem for both types of initial conditions is shown in Figure 3.1.
We observe that, at first, there are no differences between both solutions, but later on noticeable
differences appear between both solutions. This indicates that, beyond some time it is not possible
to obtain accurate solutions of this problem, despite the fact that these are deterministic equations.
Since the velocity arising from the NS equation is a random variable, it is not enough to have a
particular solution of this equations, since a small change in the initial and/or boundary conditions
would translate into a different solution. It is therefore convenient to characterize the velocity in
terms of its probability density function (PDF) and we shall discuss about this topic later.

Homework
The logistic function is given by

xn+1 = axn (1 − xn )

where a is a constant. Take x(1) = (a − 1)/a and make a program for the logistic equation at
different values of Nmax and make a plot of the final value of x vs a.

The type of chaos that arises from the Lorenz or the Logistic equations is known as deterministic
chaos and has the following properties:

1. Long-term behaviors are practically impossible to predict.

2. The system behavior is highly sensitive to initial conditions.

3. Small amounts of external noise (i.e., changes in the boundary conditions) rapidly grow to
control the system.

Many non-linear systems have a tendency towards chaos. Lorenz defined chaos as:

When the present determines the future, but the approximate present does not approximately
determine the future.

Landau suggested that a laminar flow goes through a succession of bifurcations (or changes) that
lead to more complicated states as the Reynolds number is increased. He actually proposed that
an infinite number of bifurcations take place as Re → ∞. Despite the appealing nature of Landau’s
vision, it turns out that it is overly simplistic because the infinite succession of bifurcations that he
suggested is just one of many pathways that lead to chaotic results.
We have pointed out that the non-linear term in the NS equations is the source of chaos. This
may lead us to think that turbulence is governed by the Euler equation:

∂v
ρ + ρv · ∇v = −∇p (3.4)
∂t
Actually, this is not the case. Imagine the following experiment: imagine that we can solve the
Euler equation from a certain initial condition. The result should be obvious: as time progresses,
fluid mixture takes place. Now imagine that at a given time we stop the simulation and take the
current solution as the initial condition and then we run the simulation backwards in time. In this
way, t becomes −t and v becomes −v, if we make these changes of variables in the Euler equation,
we actually obtain the same equation, hence as time evolves we recover the original initial condition.

22
Figure 3.1: Response of Lorenz equations under different initial conditions

23
flow and the turbulence, with the former supplying energy to the

(b) Stage (i) Stage (ii) Stage (iii)

-a -S-A-^
Voi -*
3>-^^J
^ ^3 -^ ^

on of grid Fully developed turbulence


ges of
Transition to
nce. The
greatly
fully developed Small-scale
opment of turbulence eddies decay first
irection.

Figure 3.2: a) Sketch of an experimental system to produce grid turbulence and b) Stages of
development of turbulence.

This is an unphysical result that we should not be able to obtain it in practice. This is because in
reality, even in turbulent flow there is an influence of the viscous term that can not be ignored. If
we repeat our changes of variables in the NS equations instead of the Euler equation, we observe
that it is due to the viscous term that the same equation is not obtained when we backwards in
time. In other words, it is the finite viscous contribution the one responsible for the irreversibility
of turbulent flow.

3.2 Freely-evolving turbulence


Consider the system sketched in Figure 3.2 in which a grid is inserted in a relatively quiescent
stream of gas with the intention of generating turbulence downstream. The following stages are
identified:

1. In stage (i), the motion consists of a discrete number of laminar vortices arising from the grid.

2. Stage (ii) corresponds when these laminar vortices later converge into a field of fully developed
turbulence. This stage is characterized because it contains all the length scales of turbulence,

24
of turbulence

One realization of u'x

Ensemble average of (u'x)2


ce (u'x) (t), is a
time in any one
nsemble average is

Figure 3.3: a) Sketch of the turbulent velocity fluctuations and b) Sketch of the turbulent kinetic
energy in freely evolving turbulence.

behaviour of the logistic equation.) To overcome this problem we


from the integral scale to the Kolmogorov scale. This stage is also known as the asymptotic
average
state | (u') 2 (t) over the many realizations of the experiment to give
of turbulence.
a so-called ensemble average, denoted by |((u') )(t). This tells how
3. In stage (iii) the small vortices have disappeared and only remain the large eddies, which
rapidly,
slowly onacross
rotate average, the fluid
the system. In thisloses
stage,its
thekinetic energy
Reynolds number,ascomputed
a resultfrom
of the size
ofviscous dissipation
the eddies as itofpasses
is on the order through
magnitude of one.the test section.

4. StageIt(iv)
turns out
is the that,
final one while
and it isu'called
is apparently random
the final period in any
of decay, oneinertia
in which realiza-
is negligible
2
tion, the ensemble average, ((u') ), is a smooth function of time as
and a spatially complex laminar flow is achieved.
Let illustrated in Figure
us assume now that we 3.4. That is,
are capable of although
installing aisnon-intrusive
u' chaotic,probe
the average
that measures the
rate of loss of energy is a smooth, repeatable function of time. This
fluid velocity and that this probe travels at the same velocity as the mean flow, say v.isSince the
probe measures the velocity and we also assume that we know the mean flow, we can then calculate
yet another
the turbulent manifestation of the fact that, although a turbulent
fluctuations:
velocity field appears to be quite v0 =random,
v−v and is different from one (3.5)
realization
If the experimentto the next,several
is repeated the statistical properties
times, we should of avturbulent
note that 0 is differentflow arerealization
in each
and theperfectly reproducible.
result should be chaotic along time (see Figure 3.3a). If we compute the ensemble average
hv02 i, we should obtain a smooth time-decaying function as the one sketched in Figure 3.3b). In
Note that there appears to be many 'frequencies' contributing to ii
other words, even though v0 is chaotic, it turns out that hv02 i is deterministic and smoothx and thus
in Figure 3.4. This is because the velocity at any one point is the result
reproducible.
of a multitude
In previous sessions, weof mentioned
eddies (lumps of vorticity)
that there in ofthe
is a cascade vicinity oflengths
characteristic that in fully
developed turbulence. We mentioned that the large eddies give rise to smaller eddies with less
point, each contributing to u via the Biot-Savart law. These vortices
have a wide range of spatial scales25and turn-over times, and so con-
tribute different characteristic frequencies to u'x(t).
We now go further and divide up | ((u') 2 ) according to the instant-
aneous size of the eddies surrounding the probe. That is, we calculate
2
nd nature of turbulence

Enstrophy

e distribution of energy and


ly developed turbulence. Log (wave number)

larger ofeddies:
Figure 3.4: Sketch Z/H ~ 4 X 10
the distribution s. Evidently,
of energy the small
and estrophy witheddies havetoathe
respect veryeddies’ sizes in
fully developed turbulence.
small turn-over-time, and since |to| is a measure of the rate of rotation
of an eddy, this explains why the vorticity is concentrated in the small
kinetic energy,scales.
unit the smallest eddies viscously dissipate the remaining energy. In Figure 3.4 we
sketch the distribution of energy
So, loosely (hv02 i)
speaking, weand
mayestrophy
say that(hw
2
thei)bulk
versus the energy,
of the wave number,
and k = π/r (r
being the size of the eddies). We observe that the largest eddies contain more energy than the
the bulk of the enstrophy, are held in two mutually exclusive groups
smaller eddies and these contain the largest amounts of enstrophy. this indicates that the large
eddies containofless
eddies: the energy
vorticity is smaller
than the held in the largeHowever,
eddies. eddies and theeddies
these enstrophy in
contribute very little
the small
to the net kinetic energyeddies. Actually,
because thy areasmall
moreandaccurate statement
randomly would be to say
distributed.
that the vorticity which underpins the large eddies (via the Biot-Savart
3.3 Turbulence law) is weak and and memory
dispersed, and makes little contribution to the net
enstrophy. Conversely, the small eddies are composed of intense
Let us return patches of vorticity,of and
to the experiment grid so they dominate
turbulence the enstrophy.
and repeat However,
the experiment several times using
different gridsthey
everymake
time.little
It would appear, to
contribution that
thesufficiently
net kinetic farenergy
away from the entrance,
because they the fully
developed turbulence should have no memory of the grid that was used and the statistical properties
are small and randomly orientated. This general picture is true for all
of flow should remain unchanged. This idea was proposed by Batchelor (1953); however Davidson
(2015) pointshigh-Re
out thatturbulent flows.
this is not For example,
necessarily in a According
the case. typical pipetoflow
this most of it turns out
author,
the turbulent
that there is certain energycontained
information is held inineddies of sizecondition
the initial comparablethat with the pipe
is retained throughout the
radius, yet the majority of the enstrophy is held in eddies which are
evolution of flow despite the complex nonlinear interactions. To be specific, information is retained
by turbulenceonly
as a aconsequence
fraction of of
a percent of the pipe
the conservation diameter.
of linear and angular momentum.

3.4 A statistical approach for turbulence


3.2.2 mentioned
We have previously The rate ofthat,
destruction
althoughofthe energy in fully
velocity fields are random, some average
developed
properties of the velocity, turbulence
such as the kinetic energy, should be reproducible for different realizations.
There are at least three types of averages that we should consider:
We might note that the initial process of adjustment of the flow, from
stage (i) to stage (ii), is almost completely non-dissipative. Energy is
1. Time averages
merely redistributed from one eddy size to another as the
2. Ensemble averages
eddies evolve (break-up?). The net energy loss is small because the
3. Volume Kolmogorov
averages scale is excited only once stage (ii) is reached.
In this section we are interested in the next phase: the decay of fully
26
developed turbulence. This is shown schematically in Figure 3.5(c).
There is now significant dissipation of energy and it turns out that the
smallest eddies decay fastest. This is because their turn-over-time,
which also turns out to be their break-up time, is much smaller than
that of the large eddies, /7/v<CÏ/H.
and nature of turbulence

Traces of velocity against


A and B are statically
hile A and C are not.

Figure 3.5: Sketches of traces of velocity measured downstream of flow around a cylinder in three
points.
Suppose we have a large tank of water which we vigorously stir (in
some ensemble
We have mentioned carefully averages,
prescribedwhich
manner). Theonwater
consist is then
taking left to itself
the average and
values of a number N
of experiments we wish to
or direct study the
numerical statistics of
simulations. the average
This decay process.
is defined Weas:might, for
example, have a velocity probe which measures the velocity of the
N
turbulence, u', at a particularhvx iN = locationvxnin the tank, A. A natural
1 X
(3.6)
question to ask is, 'how does theNaverage n=1 kinetic energy of the tur-
bulence decay with time?'. The problem is, of course, that we do not
In volume averaging, one would only make an experiment (or numerical simulation once) and
have a clear definition of the word 'average' in this context. The
take averages over all positions at given times. For statistically homogeneous turbulence (where
turbulence
properties do not depend ofisposition),
not statistically
both volume steady
and and so we
ensemble cannotshould
averaging interpret
match. Actually,
'average' to mean a time-average. One way out of this mess
Wood et al. (2003) have shown that this is the case in non-turbulent transport in porous is to media.
revisit
Time averages the quite
are also idea important
of ensemble and averaging, which
we shall make wethem
use of introduced in
in the following chapter.
For the moment,Section 3.2.1 of this chapter.
consider the flow around a cylinder where downstream we have placed three
measuring ports A, B and we
Suppose C asrepeat
shownthe in Figure 3.5. Clearly,
experiment the velocity
10,000 times. Each dynamics
realizationin A are quite
similar to those in B because these two points are located nearby each other. Furthermore, even
of the experiment is, as far as we are concerned, identical. However,
though there are some pointwise differences, on average the velocities are quite similar. On the one
as we have seen, minute variations in the initial conditions will
produce quite different traces of27u'(t) in each experiment. We now
tabulate (u')2 against time for each realization and then, for each
value of t, we find the average value of (u')2. This average, denoted
((u') ), is called an ensemble average. It is function of time but,
unlike (u')2(t), it is a smooth function of time (Figure 3.13). It is also
hand, if we multiply uA and uB and then we make the time average we would obtain uA uB 6= 0. On
the other hand the time-average of the multiplication between uA (or uB ) with uC we find a result
that is close to zero. We can thus say that the points A and B are statistically correlated, whereas
points A and C are uncorrelated. Certainly, correlation is established on the basis of averages. As
a final note, for now, about time averages, it is worth mentioning that time averages should match
ensemble averages in steady-on-average turbulent flows.

3.5 Structure functions and energy spectrum


In the reminder of this chapter we discuss about statistical measures that help to quantify the state
of a cloud of turbulence. These are:
1. The velocity correlation function.

2. The second-order structure function.

3. The energy spectrum.

The velocity correlation function


The velocity correlation function (also called the two-point correlation) in turbulent flow is defined
as a second-order tensor:
Q = hv0 (x)v0 (x + r)i (3.7)
However, for the developments that follow, we find it more convenient to discuss about the ij
components of this tensor:
Qij (r) = hvi0 (x)vj0 (x + r)i (3.8)
where vector x is a position vector and vector r is a vector indicating the distance at which the
second velocity fluctuation is measured with respect to the first one. Notice that Qij is defined in
terms of components of the velocity vector fluctuations. If Qij is time-independent it is said to be
statistically steady and if Qij does not depend of position, it is said to be statistically homogeneous.
In fact, notice that, for statistically homogeneous turbulence, Qij is a function of the separation
between the two points, r, and not by the location of the first point, x.
In general, Qij gives information about the degree at which the velocity components at
two points in space are correlated to each other. If the velocities fluctuations are statistically
independent, it follows that Qij = 0. This may be the case if the separation distance |r | is much
greater than the typical eddy size. The opposite situation is when |r | → 0 and thus Qij → h(vi0 )2 .
So, in general, the velocity correlation function gives information about the degree and manner at
which the velocity components at two different points are correlated to each other.
In some cases, a cloud of turbulence has statistical properties that are independent of direction.
This is known as isotropic turbulence. An example of this is fully developed turbulence in a wind
tunnel, which is approximately isotropic and homogeneous. Let u be a typical velocity fluctuations
(we will not use the prime for simplicity) of the large eddies, defined as:

u2 = Qii (0) = hu2x i = hu2y i = hu2z i (3.9)

In other words,
hu2 i = hu · ui = 3u2 (3.10)

28
assumption that, through
the mean flow is zero.)

Qyy(rix) =u2g(r).

The functions / and g a


correlation functions (or
_f(0) =g(0) = 1, and have
that / and g are not in
equation (2.3). In fact, w
The integral scale, !,
Figure 3.6:Figure
Sketch 3.15 The shape of
of the longitudinal andthe longitudinal
lateral correlation functions.
and lateral velocity correlation functions.
1= \ f(r)dr.
Then two-typical components of Qij are:

Qxx (rex ) = hux (x)ux (x + rex )i = u2 f (r) (3.11)


2 This provides a conveni
Qyy (rex ) = Qzz (rex ) = u g(r) (3.12)
which velocities are appr
where f and g are the longitudinal and lateral velocity correlation functions. By definition f (0) =
g(0) = 1 and they are decreasing dimensionless functions of r as sketched in Figure eddies.
3.6. Actually,
the integral scale of turbulence in the Richardson cascade, `, is defined as We now come to an i
Z∞
1
Z∞
1
Z∞ that there is a wide r
` = f (r)dr = 2 Qxx (rex )dr = Qxx (rex )dr (3.13)
0
u
0
Qii (0)
0
turbulent signal. This re
This provides a convenient measure of the region in which velocities are appreciablyhierarchy of tangled v
correlated,
which is associated to the size of the large eddies.
7
A less restrictive form of i
The second-order structure function
but not reflectional symmetry
It is worth noting that the correlation function does not indicate how the stay kinetic
withenergy is
the definition above
distributed across the different eddies. In the analysis of turbulent flow it is certainly desirable
to be able to measure the kinetic energy of the different eddies. To this end, theItlongitudinal
cannot be perfectly hom
velocity increment is defined as: of the tunnel. Moreover, the g
is made to avoid this then ty
∆v = ux (x + rex ) − ux (x) (3.14)
fluctuations. However, we sha
Clearly, only eddies of size r or less make a contribution to ∆v because eddies of sizes much larger
than r should have similar velocities at x and x + rex . In this way, let us define the second-order
longitudinal structure function is defined as

h[∆v]2 i =h[ux (x + rex ) − ux (x)]2 i = hux (x + rex )2 i − 2hux (x + rex )ux (x)i + hu2x (x)i
= 2Qxx (0) − 2Qxx (rex ) = 2u2 (1 − f (r)) (3.15)

29
The structure function can actually be defined, in general, at any power, p, that is: h[∆v(r)]p i.
From Figure 3.6, we have that, for sufficiently large values of r, f tends to zero, consequently,

2
h[∆v]2 i ≈ 2u2 = hu2 i (3.16)
3
Or, after little rearrangement,
3 1
h[∆v]2 i → hu2 i (3.17)
4 2
Thus it may seem that the second-order structure function gives an indication about the cumulative
energy per unit mass in eddies of sizes r or less. However, eddies larger than r make a contribution
to ∆v on the order of r × w, hence an interpretation of the second-order structure function is.

3
h[∆v]2 i = O(Energy of eddies smaller than r) + r2 O(enstrophy in eddies larger than r) (3.18)
4

The energy spectrum


An alternative way of analyzing turbulent flow is to work with wavenumbers instead of eddy sizes
and to use the Fourier transform to identify structures of different sizes. In this way, the information
of the velocity correlation function can be re-expressed in terms of the wavenumber spectrum. To this
end, let us introduce the wavenumber vector k, whose magnitude defines the wavelength |k | = 2π/r.
The velocity spectrum tensor is defined as the Fourier transform of the velocity correlation tensor:
Z Z∞ Z
1
Φ(k, t) = e−ik·r Q(r, t)dr (3.19)
(2π)3
−∞

and its inverse transform recovers the velocity correlation tensor:


Z Z∞ Z
Q(r, t) = eik·r Φ(k, t)dk (3.20)
−∞

Indeed, for r = 0, we recover:


Z Z∞ Z
Q(0, t) = Φ(k, t)dk = huui (3.21)
−∞

In this way, the velocity spectrum tensor represents the contribution to the covariance huui of
velocity modes with wavenumber k. In other words, the dependence of Q with r and of Φ with k,
give information about the directional dependence of huui. Whereas the components Qij and Φij
give information about the direction of the velocity.
The energy spectrum function is defined as
Z Z∞ Z
1
E(k, t) = Φii (k, t)δ( |k | − k)dk (3.22)
2
−∞

30
which may be viewed as Φ stripped from all directional information. The energy spectrum is a
positive definite function and its inverse Fourier transform is:
Z∞ Z∞
ikr sin(kr)
R(r, t) = E(k, t)e dk = E(k, t) dk (3.23)
kr
0 0

where R(r) = 21 hu(x) · u(x + r)i = u2 (g + f /2). Notice that, the limit as r → 0 in the above
equation, corresponds to integration of the energy spectrum over all scalar frequencies and it is
given by:
Z∞
1 1 3
R(0, t) = E(k, t)dk = Qii (0, t) = hu · ui = u2 (3.24)
2 2 2
0
Therefore, Edk represents the contribution to the turbulent kinetic energy from all modes with |k |
in the range k ≤ |k | < k + dk. In other words, the energy spectrum represents the energy in eddies
of size π/k.
The energy spectrum has another property:
Z∞
1 2
hw i = k 2 Edk (3.25)
2
0

Analogously, k 2 Edk can be interpreted as the contribution to enstrophy from the range of
wavenumbers k ≤ |k | < k + dk.
It should not be forgotten that our discussion about the second-order longitudinal structure
function and the energy spectrum function originated from our desire to gain some knowledge
about the energy of the eddies of different sizes. So far, we have seen that they are both related to
kinetic energy and enstrophy, so they should be related to each other. The demonstration is left as
a homework exercise, here we simply write the result:
Z∞
3 cos(x) sin(x)
h[∆v]2 i = E(k)H(kr)dk, H(x) = 1 + 3 2
−3 3 (3.26)
4 x x
0

Homework
The longitudinal and lateral velocity correlation functions are related by:
d 2
2rg = (r f )
dr
Show that
d 3
2r2 R = u2
(r f )
dr
Use the definitions given in eqs. (3.23) and (3.24) to obtain the following differential equation
for f :
d 3 sin(kr)
(r f ) = 3r
dr k
solve this equation subject to the boundary condition f (0) = 1 and use this result along with

31
the definition of u2 in terms of the energy spectrum given in Eq. (3.24) into Eq. (3.15) to obtain
the final result expressed in Eq. (3.26).

A reasonable estimate for H is the following:


(
(x/π)2 , x<π
H(x) ≈ (3.27)
1, x>π

Consequently,
Zπ/r Z∞
3 r2
h[∆v]2 i = 2 2
k E(k)dk + E(k)dk (3.28)
4 π
0 π/r

Or,

3 r2
h[∆v]2 i = O(Energy of eddies smaller than r) + 2 O(enstrophy in eddies larger than r) (3.29)
4 π

which is equivalent to our heuristic approximation given by Eq. (3.18).


As a matter of summary, we have two functions, the E(k) and h[∆v]2 i that serve to have an
impression about the energy across the different eddy sizes.

3.6 Another look at Kolmogorov’s theory


At this point, it is worth pondering, at least heuristically, what are the factors that determine
∆v in large and small eddies. Both of them may be assumed to be independent of the Reynolds
number and they are only functions of the velocity and the scale factor. In this way we have, from
dimensional examination:
For large eddies
h[∆v]2 i = u2 F̂ (r/`) (3.30)

For small eddies


h[∆v]2 i = v 2 F (r/η) (3.31)

Here F should be some universal function valid for all forms of isotropic turbulence. This was
proposed by Kolmogorov and it applies to a part of the spectrum of values of the vortices. This
particular part is known as the universal equilibrium range.
Of course in between the largest and smallest eddies, one may find a range of them that they are
influenced by the viscous dissipation rate, ε. This particular range is called the inertial subrange
and Kolmogorov proposed that in this range, the following relation applies:
Inertial subrange:
h[∆v]2 i = βε2/3 r2/3 , η  r  ` (3.32)

where β is a universal constant that is approximately 2 and it is known as Kolmogorov’s constant.


These ranges are identified in Figure 3.7.

32
Some elementary properties of freely evolving turbulence

(a) ¡ ([Av] 2 >

2u¿
Inertial subrange

r=t¡-
i
(b) Log(E)

Inertial subrange

Log(fc)
Figure 3.17 (a) The general shape of {[Av]2).
(b) The general shape of E(fe). Universal equilibrium range

This 2has
Figure 3.7: Sketch of h[∆v] a reassuring
i and similarity the
E including to our heuristic estimate
universal (3.18). Of and inertial subranges.
equilibrium
course (3.21b) cannot be exactly true because it is quite artificial to
categorize the influence of eddies according to whether they are
smaller or greater than r. The cut off is too sharp. Nevertheless, (3.21a)
captures the main features of the relationship between E(K) and
([Av(r)]2). The general shapes of ([Av]2) and E(fe) are shown in
Figures 3.17 (a) and (b).
Earlier
3 The westatistical
saw that description
vorticity tendsoftoturbulent
be concentrated
flows in the
smallest eddies and is rather weak in the large scales (consult
Figure 3.6). Thus, provided we are not too close to the dissipation
scales, we might simplify (3.2lb) to

j ^ ^ - all energy in eddies of size r or less

and indeed this approximation is commonly used. However, this is


often dangerous since the second term in (3.2la) can make a sig-
nificant contribution to (Av)2. For example, it is readily confirmed
from (3.21a) that, for small r, we have,

((Av)2) ~ < w > 2 +• • • = ^ 2 + • • • •


Actually it turns out that the exact relationship is (consult Chapter 6),
-, the 1.3% error having crept in because of

93
Fig. 3.3. Sketches of the sample space of U showing the regions corresponding to the
Figure
events (a) B 3.8:
(U <Examples, ( bof) Csample
v ~ )and = {V, Ispaces for two different events.
U < Vb].

Sample space
In order to be able to discuss more general events than A r { U < 10 m s-I},
we introduce an independent velocity variable V which is referred to as the
sample-space variable corresponding to33U. As illustrated in Fig. 3.3, different
events such as
3.7 Probability distribution of the velocity field
Let us introduce an independent variable, V, which is referred to as the sample-space variable,
corresponding to a component of the velocity, say u. In this way, from Figure 3.8 we may identify
events B and C such that:
B ≡ {u < Vb } (3.33a)
C ≡ {Va < u < Vb }, Va < Vb (3.33b)
The probability p of an event is a real number (0 ≤ p ≤ 1) indicating the likelihood of occurrence
of an event. For an impossible event, p is zero and for a certain event, p is one. For the two events
defined above we have,
p = P (B) = P {u < Vb } (3.34a)
P (C) = P {Va < u < Vb } = P {u < Vb } − P {u < Va } (3.34b)
The probability of any event can be determined from its cumulative distribution function (CDF),
which is defined as:
F (V ) ≡ P {u < V } (3.35)
In this way, we have
P (B) = F (Vb ) (3.36)
P (C) = F (Vb ) − F (Va ) (3.37)
The CDF has the following properties:

1. Since u < −∞ is impossible, it follows that

F (−∞) = 0 (3.38)

2. Since u < ∞ is certain,


F (∞) = 1 (3.39)

3. The CDF is a non-decreasing function, hence for Va < Vb ,

F (Vb ) ≥ F (Va ) (3.40)

The probability density function (PDF) is defined as the derivative of the cumulative distribution
function:
dF
f (V ) ≡ (3.41)
dV
Clearly, f has the units of the inverse of u. Since, F is a growing function, it follows that f is
positive as well. Also, since in the extremes F is a constant, we have that

f (−∞) = f (∞) = 0 (3.42)

and it satisfies the normalization condition:


Z∞
f (V )dV = F (∞) − F (−∞) = 1 (3.43)
−∞

34
Also, form our previous example of event C, we have
ZVb
P (C) = F (Vb ) − F (Va ) = f (V )dV (3.44)
Va

For an infinitesimal interval, we have:

F (V + dV ) − F (V ) = f (V )dV (3.45)

The PDF fully characterizes the random variable u. When two or more random variables have the
same PDF they are said to be identically distributed or statistical identical.
The mean (or expectation) of the random variable u is defined as
Z∞
µ = hui = V f (V )dV (3.46)
−∞

this is the probability-weighted-average of all possible values of u. In a more general manner, if


Q(u) is a function of u, the mean of this function is
Z∞
hQ(u)i = Q(V )f (V )dV (3.47)
−∞

Note that, while u and Q(u) are random functions, the same is not true for hui and hQ(u)i.
Consequently,
Z∞ Z∞
hhQ(u)ii = hQ(u)if (V )dV = hQ(u)i f (V )dV = hQ(u)i (3.48)
−∞ −∞
The variance is:
Z∞
σ2 = (V − µ)2 f (V )dV (3.49)
−∞
The standard deviation, σ, is the square root of the variance. In this way, the standardized random
variable û is
u − hui
û = (3.50)
σ
The skewness factor is defined as the third moment of the pdf:
Z∞
1
S= 3 (V − hV i)3 f (V )dV (3.51)
σ
−∞

This factor has to do with the symmetry of a function. For example, a Gaussian function has a
zero skewness factor.
The flatness factor or kurtosis is the fourth moment:
Z∞
1
δ= 4 (V − hV i)4 f (V )dV (3.52)
σ
−∞

35
Figure 3.9: Examples of kurtosis of a function: The curve in red has infinite kurtosis, the curve in
blue has a kurtosis of 5 and the curve in black is the Gaussian curve with kurtosis of 3.

The origins and nature of turbulence

For small r the probab


become identical. For ex

S(r 0) = S0 =
<[A

It follows that dux/dx


Figure of ~ —0.4 and a flatness
Figure 3.10:3.20 Schematic
Sketch of the pdfofofthe
the p.d.f. forgradient.
velocity duj
dx. (The departures from a Gaussian dis- dux/dx is shown schema
tribution are exaggerated.)
dux/dx are more likely
negative values ofdux/dx
this dominates the skew
36
The fact that the flatne
to extremely high values
about the spatial structur
A Gaussian function has a flatness factor of δ = 3. Kurtosis is related to the tails of a function
not to its peaks as shown in Figure 3.9.
In terms of turbulence, the PDF of velocity is usually Gaussian. It is symmetric (skewness of
zero) and it has a kurtosis of 2.9-3. One interpretation for this normal distribution is that the
velocity at any point is the consequence of a large number of randomly oriented vortical structures
so that a normal distribution arises. This is in agreement with the central limit theorem. However,
this is not the case for the pdf of the velocity gradient. In this case, it has been observed that the
distribution is similar to the one sketched in Figure 3.10, which exhibits a non-symmetrical shape
and a large kurtosis. This departure from the normal distribution is likely to be found for small
eddies, where the viscous effects are more dominant.

37
38
4 | Turbulent shear-flows

Contents
4.1 Reynolds stress and the closure problem . . . . . . . . . . . . . . . 39
4.2 Closure problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Turbulent flow in a pipe and between two parallel plates . . . . . 43
4.4 The eddy-viscosity theories of Boussinesq and Prandtl . . . . . . . 48
4.5 Energy transfer in turbulent flow . . . . . . . . . . . . . . . . . . . . 50
4.6 Current models of turbulence . . . . . . . . . . . . . . . . . . . . . . 52
4.6.1 The k − ε model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6.2 The k − ω and SST models . . . . . . . . . . . . . . . . . . . . . . . . 52
4.6.3 Spalart-Allmaras model . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6.4 Reynolds stress model . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.6.5 DNS and Large-eddy simulation . . . . . . . . . . . . . . . . . . . . . 54

4.1 Reynolds stress and the closure problem


In this chapter we study classical models for turbulence. As we discussed, there are at least three
types of averages, which under certain conditions are equivalent. We will show here the time-
average of the NS equations. Although there are other books in which the ensemble average is
presented (Pope, 2000). In the end, the resulting equation from both types of averages is the same.
Let us define the time average of a function S to be:
η=t+∆t
Z
1
S= Sdη (4.1)
2∆t
η=t−∆t

The time interval must be large enough so that the turbulent fluctuations are removed in the
averaging process. As mentioned before, for statistically steady flows, the ensemble and time-
averages are equivalent. In this case, we can use an arbitrarily large value of ∆t to smooth the
fluctuations. When this is not the case, the time-averaged functions are still functions of time. In
other words, ∆t must be large enough so that turbulent fluctuations are smoothed but also small
enough so that the time-dependence of the average variable is not lost. This is usually the case
when there is a separation of time scales. As explained before, for small eddies, the characteristic

39
time is estimated to be: η 
t∗ = O (4.2)
v
Whereas, for large eddies, the estimate is
 
∗ `
T =O (4.3)
u

Assuming that v = O(u), we have that whenever η  `, it follows that there is a separation of
time-scales: t∗  T ∗ . In other words, when the characteristic time scale associated to the velocity
fluctuations is much smaller than the time scale of the time-averaged velocity, it is possible to apply
time-averaging.
As mentioned before, the velocity can be decomposed according to:

v = v + v0 (4.4)

with v0 being the velocity fluctuations. If we time-average this equation, we obtain

v = v + v0 (4.5)

For statistically steady flows, it is evident that

v=v (4.6)

For flows that are unsteady in the time-average sense, we may assume that

v0  v (4.7)

As we know, the governing equations for turbulent flow are the continuity and the NS equations.
Let us, for the moment, direct the attention to the continuity equation and perform the time
average, so that the following expression results:
η=t+∆t
Z
1
∇ · vdη = 0 (4.8)
2∆t
η=t−∆t

Since the limits of integration are independent of position, we can interchange differentiation and
integration to obtain:
∇·v=0 (4.9)
which indicates that the time-averaged velocity is solenoidal.
Moving on to the NS equations, the result of making the time average is:
η=t+∆t
Z η=t+∆t
Z
1 ∂v 1
ρ dη + ∇ · (ρvv)dη =
2∆t ∂η 2∆t
η=t−∆t η=t−∆t
η=t+∆t
Z η=t+∆t
Z η=t+∆t
Z
1 1 1
− ∇pdη + ρgdη + µ∇2 vdη (4.10)
2∆t 2∆t 2∆t
η=t−∆t η=t−∆t η=t−∆t

40
Again, taking into account the time-independence of the integration limits and assuming that both
the density and viscosity are constant, we obtain:
η=t+∆t
Z
ρ ∂v
dη + ρ∇ · vv = −∇p + ρg + µ∇2 v (4.11)
2∆t ∂η
η=t−∆t

Where the first term is:


η=t+∆t
vt+∆t − vt−∆t
Z
1 ∂v
dη = (4.12)
2∆t ∂η 2∆t
η=t−∆t
Notice that, using Leibniz rule:
 
η=t+∆t η=t+∆t  
d(t − ∆t)
Z Z
∂v ∂  1 1 ∂v 1 d(t + ∆t)
=  vdη  = dη + vt+∆t − vt−∆t

∂t ∂t 2∆t 2∆t ∂t 2∆t dt dt
η=t−∆t η=t−∆t
vt+∆t − vt−∆t
= (4.13)
2∆t
Consequently, our average version of the NS equation is:
∂v
+ ρ∇ · vv = −∇p + ρg + µ∇2 v
ρ (4.14)
∂t
The inertial term can be further worked as follows:
∇ · vv = ∇ · (v + v0 )(v + v0 ) = ∇ · vv + ∇ · v0 v0 (4.15)
In this way, the time-averaged NS equations are:
∂v
+ ρv · ∇v = −∇p + ρg + µ∇2 v − ∇ · ρv0 v0

ρ (4.16)
∂t
This equation is similar to the NS equation except for the last term in the right-hand side, which
couples the mean flow to the turbulence. This form of the model is known as the Reynolds-averaged
NS equations or RANS. We can express the last two terms as follows:
µ∇2 v − ∇ · ρv0 v0 = ∇ · τ (T )

(4.17)
here τ (T ) is the total turbulent stress tensor and it is defined as:
τ (T ) = τ + τ (R) (4.18)
where:
τ = µ∇v (4.19a)
τ (R) = −ρv0 v0 (4.19b)
τ (R) is the turbulent (or Reynolds) stress tensor. Under these conditions the RANS equations take
the form:
∂v
ρ + ρv · ∇v = −∇p + ρg + ∇ · τ (T ) (4.20)
∂t
which looks like the NS. In fact, if there were a simple relation between τ (T ) and v, we could treat
this equation just like we would use the NS equations. Unfortunately, this is not the case. In the
following section we will show why there is no solution to this problem.

41
4.2 Closure problem
Let us now try to derive the governing equations for the velocity fluctuations. These can be obtained
by noting that v0 = v − v. In this way, it is easy to obtain that the fluctuations are solenoidal:

∇ · v0 = 0 (4.21)

Furthermore, the result of subtracting Eq. (4.16) to the NS equation is:

∂v0
+ ρv · ∇v0 + ρv0 · ∇v = −∇p0 + µ∇2 v0 + ∇ · ρv0 v0

ρ (4.22)
∂t
This non-local equation illustrates that, in order to determine the velocity fluctuations, it is
necessary to have the velocity field, its average and v0 v0 . The latter being ultimately, the quantity
that is needed to close the RANS equations. This intimate coupling between the NS, the RANS
and the velocity fluctuations has discouraged the direct solution of the closure problem. It is
computationally less expensive to solve the NS directly.
Another approach to this problem is to derive an equation for v0 v0 . This can be done by pre-
multiplying Eq. (4.22) by v0 and adding the result to the one arising from the post-multiplication
of the same equation by v0 . The result of time-averaging this sum is:

D v0 v0
+ ρ∇ · v0 v0 v0 + (∇v )T · ρv0 v0 + ρv0 v0 · ∇v = −v0 ∇p0 − (∇p0 )v0 + µv0 (∇2 v0 ) + µ(∇2 v0 )v0
Dt
(4.23)

here
Dψ ∂ψ
= + v · ∇ψ (4.24)
Dt ∂t
From Eq. (4.23) it follows that, in order to compute v0 v0 , it is necessary to have v0 v0 v0 and so on
and this is known as the hierarchical closure problem for turbulence.

Homework
1. Demonstrate the following identities:

v0 ∇p0 + (∇p0 )v0 = ∇(p0 v0 ) + ∇(p0 v0 )T − 2p0 S0 (4.25a)

v0 ∇2 v0 + (∇2 v0 )v0 = ∇2 (v0 v0 ) − 2(∇v0 )T · ∇v0 (4.25b)


and use them along with the definition of Reynold’s stress to re-write Eq. (4.23) as follows:

D v0 v0 0
ρ + ρ∇ · v0 v0 v0 = 2τ (R) · S − ∇p0 v0 − (∇p0 v0 )T + 2p0 S
Dt
+ µ∇2 v0 v0 − 2µ(∇v0 )T · ∇v0 (4.26)

Note that this equation can easily be written in terms of the stress tensor since τ (R) = −ρv0 v0 .
2. On the basis of the definition:

(v 0 )2 = vi0 vi0 (no sum convention) (4.27)

42
This empirical representation for the turbulent velocity profile can not be used to predict the shear stress at
the tube wall, because the derivative of v z with respect to r tends to infinity as r  ro . Nevertheless, it is
a reasonably accurate expression for the time averaged velocity as a function of radial position.

vz
vz,max

Figure 9-4.
Figure 4.1: Laminar andLaminar andvelocity
turbulent turbulentprofiles
velocityin profiles
a pipe of radius r0 .

In the central region of the tube, the velocity fluctuations are large enough so that the turbulent stress
Take the
far exceeds theiiviscous
component
stress,ofi.e.,
Eq. (4.26) and show that it can be expressed as follows:

ρ D(v 0 )2 τ (t )
0 τ ,1 0in2 the
 
core (9-33)
= ∇ · −p v + v · τ ii − ρ(v ) v + (τ (R) · S)ii − 2µ(S0 · S0 )ii
0 0 0 0 (4.28)
2 Dt 2
and the intense exchange of momentum creates a nearly flat velocity profile. At the tube wall, the velocity
which can
fluctuations mustbebe
interpreted
zero, as: the material rate of change of kinetic energy = transport of kinetic
energy + generation of kinetic energy - dissipation.
v  0, v  0, v v  0 , at the tube wall (9-34)

and it follows that the turbulent stress must also be zero, i.e.,
4.3 Turbulent flow in a pipe and between two parallel plates
τ (t )    v  v   0 , at the tube wall (9-35)
Consider a pipe of radius r0 with a smooth wall in which a Newtonian, steady and incompressible
flow is taking place. Under laminar flow conditions, the velocity profile is given by
We refer to the turbulent core as the region where τ (t )  τ and the viscous sublayer as the region near
the wall where τ (t )  τ . The region between the turbulent
"  core
 # and the viscous sublayer is identified as
r 2
the transition region where τ (t )  τ andvthis
z = idea
vz,max 1−
is illustrated
r0 in Figure 9-5.
(4.29)

43
Turbulent flow 313

vz

Viscous sublayer

Figure 9-5. Viscous sublayer and transition region in a tube


Figure 4.2: Sketch of the turbulent core, the transition region and the viscous sublayer in a tube.
9.3.2 Flow in a tube
For turbulent flow, thethevelocity
We can determine profile
total stress can notflow
for turbulent be computed directly.
in a tube using However,
Eq. 9-29 provided an
thatapproximate
the time-
expression that represents experimental values is
averaged flow is steady and uniform. We begin with the time averaged stress equations of motion that can
be expressed as
r 1/7
 
v z = v z,max 1 − (4.30)
v
 
 r0
  v  v    p   g    τ     v v  (9-36)
The difference between  these
t  profiles can be appreciated in Figure 4.1.
velocity
Let us now define sections of the system on the basis of the transport phenomena taking place.
Directing our core
The turbulent attention to the continuity
corresponds to the equation given by
central region of Eq.
the 9-12,
tube we makethe
where usevelocity
of the incompressible
fluctuations are
form given by Eq. b of Table 4-1 to obtain
large enough so that the turbulent stress overcomes the viscous stress:
1  1 v  vz
 r vτr  τ (R) ,  turbulent
 core
0 (4.31)
(9-37)
r r r  z
Certainly, the intense momentum exchange creates a nearly flat velocity profile.
For steady (in the time-averaged sense), uniform flow, we can follow the development in Sec. 3.3 and the
At the tube
development wall,7.2the
in Sec. non-slipthat
to conclude condition imposes that the velocity should be zero. If we take
the time average of this result, we obtain that the time-average velocity is also zero and therefore
the velocity v z is independent
1. fluctuations must of also
t (steady flow),atz (uniform
be zero i.e.,and  (axially symmetric flow).
the wall,flow)
0
2. v  0 , a plausible = 0, v =consistent
v assumption 0, v =with
0, uniform,
at theaxially
tube wall
symmetric flow. (4.32)

Consequently, the
Given these two turbulent stress
simplifications, must
we can be 9-37
use Eq. zero to
atprove
the tube r  0 and Eq. 9-36 simplifies to
that vwall:

τ (R) = −ρv0 v0 = 0,
 
at the tube wall (4.33)
0   p   g    τ     v v  (9-38)

(R)

ToLet us then
obtain identify the
the component formsviscous sublayer
for cylindrical as the portion
coordinates, we makeof use
theofsystem where
Table 5-2 τ
and Table 6-3τwhich
. Finally,
letallow
us denote the region
us to express between
the viscous stressthe turbulent
in terms core and
of the velocity the viscous
gradients and the sublayer as thein transition
turbulent stress terms of or
v  ,
inertial
r v  and v  .
region,z in which τ
(R) ≈ τ as illustrated in Figure 4.2.

44
4.32

2W

7 Flow between parallel plates.

Figure 4.3: Sketch of turbulent flow between parallel plates.

For the sake of simplicity, let us consider now the case of fully developed flow between two
parallel plates, which are separated by a distance 2W . As sketched in Figure 4.3, we find it
convenient to locate the coordinate axis at the lower wall and the maximum velocity is denoted as
v 0 . Since flow is assumed to be fully developed, the velocity component v x does not depend on x
and v y is zero. Hence, the RANS model reduces to:
 
dp ∂ ∂v x 0 0
0=− + µ − ρvy vx (4.34)
∂x ∂y ∂y
Since the pressure gradient is a constant, we make the change of variables:
dp
− = Kρ (4.35)
dx
Furthermore, let us define the total shear stress as:
∂v x
τ= µ − ρvx0 vy0 (4.36)
∂y | {z }
turbulent stress
| {z }
viscous stress

In this way, Eq. (4.34) takes the form



= −Kρ (4.37)
dy
which can be integrated subject to the symmetry boundary condition

at y = W, τ =0 (4.38)

giving rise to the following linear profile for the total shear stress
 y 
τ = ρKW 1 − (4.39)
W
Which, when evaluated at y = 0 gives the wall shear stress

τ ≡ τw = ρKW (4.40)

At this point, let us introduce the friction velocity, v∗ as


τw
v∗2 = (4.41)
ρ

45
So that the result can be rewritten in the following form

τ = ρ v∗2 − Ky

(4.42)

As indicated above, we may identify three regions in the system. Close to the wall (y  W ), we
have the viscous sublayer and Eq. (4.39) reduces to

∂v x
µ − ρvx0 vy0 = ρv∗2 , y/W  1 (4.43)
∂y

This equation can be further simplified under the assumption that in this region:

∂v x
ρvx0 vy0  µ (4.44)
∂y

Hence, taking into account the non-slip boundary condition at y = 0, we have:

v∗2
vx = y = v∗ y + , y/W  1 (4.45)
ν
where the dimensionless position y + is defined as
v∗ y
y+ = (4.46)
ν
In the turbulent core region we may assume that

∂v x
µ  ρvx0 vy0 (4.47)
∂y

The length-scale constraint behind this assumption is the following:


ν
y (4.48)
v∗

note that here we took v x = O(v y ) = O(v∗ ). Under these conditions, the solution reduces to
ν
τ = −ρvx0 vy0 = ρ v∗2 − Ky = ρv∗2 (1 − η) ,

y (4.49)
v∗
where another dimensionless length is introduced as
y
η≡ (4.50)
W
From these derivations we have the following comments:

1. The total stress is a linearly-decaying function, that has a maximum value τw at y = 0.

2. The laminar viscous stress is a rapidly decaying function that is equal to τw at y = 0.

3. The turbulent stress is a function that it is zero at y = 0 and also at y = W , hence it should
have a maximum point somewhere in this interval.

46
! Inner region T Overlap
[y«W 1 region
\ \\

(b)

Reynolds stress

-ylW

Figure 4.4: Sketch of the total, laminar and viscous stresses for turbulent flow between two parallel
Let plates.
us divide the flow into a number of regions as indicated by
Figure 4.8(a). Close to the wall, y <C W, the variation in shear stress
The above observations are sketched in Figure 4.4. The velocity profile near the wall was shown
given by
to be(4.38) is negligible
linear, however andwe we
at this point have may assume
no means that
of knowing the T is constant
velocity and core
in the turbulent
equal to
or inTthe
w. We may model the flow by
transition region. To address this issue we note that the time average velocity should not
depend of viscosity in the main turbulent core. It may depend of the fluid density and the wall
shear stress τw and position y. Since v∗ is defined in terms of both τw and ρ, we may propose that

y IWdvx«= 11 v∗ 4.39 (4.51)


dy κ y
where κ is a constant known as the Kármán constant and it has a value of 0.4. Before solving this
We shall callit this
problem regionto write
is convenient the itinner layer or inner
in a dimensionless form asregion. We retain the
viscous term in (4.39) because, adjacent ∂v x /v∗to the
1 wall, u' falls to zero and
+
= + (4.52)
so the entire shear stress is laminar. The dy
inner
κy
layer is characterized by
Integration of this equation leads to
rapid variations in T^ and T^,. Although the sum of the two stresses is
vx 1
constant, we move rapidly from av∗ situation = ln y + + Ain which T is purely vis-(4.53)
κ
cous atFrom 0 to T « data,
y =experimental T^, ait short
has beendistance
found that from
A = 5.5.the
Thewall
above(Figure
equation is4.8(b) ). of
the log-law
the wall, or the von-Kármán-Prandtl universal logarithmic velocity distribution and can be used
This suggests that we introduce a second region, the outer layer, in
to reproduce data in the turbulent core zone.
which the laminar stress is negligible,

= -u'xu'y = Vl - Ky, V*y/v > 1. (4.40)


47

Note that we have interpreted 'far from the wall' in terms of the
normalized distance V^ylv. This seems plausible since, given the
Homework
A better model for the velocity that works well in both the transition region and in the
viscous sublayer is obtained by performing the following Taylor series expansion of the velocity
about y = 0:

1 ∂ 2 v x 1 ∂ 3 v x

∂v x 2
v x (y) = v x (0) + y+ y + y3 + . . . (4.54)
∂y y=0 2! ∂y 2 y=0 3! ∂y 3 y=0

where v x (0) = 0. Starting from Equation (4.39), demonstrate that



∂v x τw
= (4.55)
∂y y=0 µ

∂ 2 v x

τw
=− (4.56)
∂y 2 y=0 µW
∂ 3 v x

=0 (4.57)
∂y 3 y=0
So that the expansion takes the form
   
vx yv∗ 1 ν yv∗ 
= 1− (4.58)
v∗ ν 2 v∗ W ν

4.4 The eddy-viscosity theories of Boussinesq and Prandtl


A first attempt to produce a turbulence model may date back to the late nineteenth century works
of Boussinesq (1877). He proposed that, by analogy with Newton’s law of viscosity, one may propose
that
 2
τ (R) = µ(t) ∇v + (∇v)T − kI (4.59)
3
In which µ(t) is the turbulent or eddy viscosity, which is presumably larger than µ and k is the
turbulent kinetic energy defined as:
1
k = v0 · v0 (4.60)
2
This model works well for free shear flows, such as axisymmetric jets and mixing layers.
Unfortunately, there is experimental evidence that the relation between D and τ (R) is much more
complex than the one given above. The obvious question about Boussinesq equation is what exactly
is the turbulent viscosity and how con we compute it? Evidently, it is a property of turbulence
and not of the fluid. Prandtl (1925), was the first one to propose a theory to compute it without
having to deal with the complexities of the closure problem. Along with the idea of a boundary
layer, Prandtl proposed the concept of a mixing length. He was inspired by the success of the gas
kinetic theory to predict the viscosity according to:
ρ
µ= lmf p vrms (4.61)
3

48
Turbulent flow 321

Figure 9-10. Mixing length process


Figure 4.5: Sketch of Prandtl’s mixing length.
and express that result in terms of the components to obtain

 vx  vy  vz


where lmf p is the mean-free path length and  v  is theroot-mean-square
0 speed of (9-60)
the gas molecules.
x rms
y z
In laminar flow, layers of fluid experience mutual drag because the molecules bounce around each
If we let the characteristic lengths in the x, y, and z directions be given by  x ,  y and  z , the order of
layer exchanging momentum (see Figure 4.5). A molecule in the slow-moving layer may move up
magnitude form of Eq. 9-60 takes the form
to the faster-moving layer making it to go slower and viceversa.
 v   vy   vz occur in turbulent flow, only instead of
Prandtl reasoned that a similar O  x phenomenon
  O    O can
   0 (9-61)
   z 
molecules one could imagine fluid particles  x   y  by turbulence.
moved Consider the mean flow shown in
Figure 4.5. The Forfluid atfluctuations,
turbulent y has a the meanchangevelocity v xwith
in the velocity (y)respect
will tohave come
position, vx ,from levels
will be only y ±of`, with ` being
the order
the measure ofthethe large
velocity eddies
itself. Becauseinof the flow.
this, we Performing
can use the estimates a Taylor series expansion of the velocity at
y + ` about y yields: vx  vx , vy  vy , vz  vz (9-62)

∂v x
v x |y+`by=Eq.v9-61
and the continuity equation represented x |y can
+ `be expressed +as . . . (4.62)
∂y y
 vx   vy   v 
  O  O
    O z  0 (9-63)
Truncating the series, we can approximate  x    y velocity
the   z  fluctuations as
 

At this point we make use of the fact that the characteristic lengths associated with the turbulent
0 order of magnitude
fluctuations are generally the same ∂v x
vx ≈ v x |y+` − v x |y ≈ ` (4.63)
∂y y
O  x   O  y   O  z  (9-64)

From the continuity


so that Eq.equation
9-63 providesfor the fluctuations, we have:
the estimates

O(vx )  O(vy )  O(vz ) (9-65)


∂vx0 ∂vy0
We can use this result with Eq. 9-59 to obtain the+ =0 (4.64)
shear stress ∂x following
|{z} ∂y representation for a component of the turbulent
|{z}
O(vx0 /`x ) O(v 0 /`y ) y

Assuming that `x = O(`y ), we obtain that vx0 = O(vy0 ). Consequently, the yx component of the
Reynolds stress tensor is:  
(R) 0 0 2 ∂v x ∂v x
τ yx = −ρvy vx = O ρ` (4.65)
∂y ∂y
(R) ∂v x
If we accept that τ yx should have the same sign as ∂y , we have
 
2 ∂v x
∂v x
τ (R)

yx = −ρvy0 vx0 = O ρ` (4.66)
∂y ∂y

49
Or, without the order of magnitude symbol:

2 ∂v x
∂v x
τ (R)

yx = −ρ` (4.67)
∂y ∂y

Since Prandtl’s mixing length is an adjustable coefficient, we may rewrite this result as:

(R) 2 ∂v x ∂v x

τ yx = ρ`m (4.68)
∂y ∂y

In addition, from Eq. (4.59), we have that the yx component of the Reynolds stress tensor is:
∂v x
τ (R)
yx = µ
(t)
(4.69)
∂y
here we have assumed that v y does not depend of x. From here, we can deduce that

(t) 2 ∂v x

µ = ρ`m (4.70)
∂y

If the mixing length can be determined experimentally and assuming that ∂v ∂y is also available,
x

one may use the above result to compute the turbulent velocity. The mixing length model can
be classified as an algebraic model and it has been severely criticized because of its inability to
represent flows different from unidirectional ones. This is because most of real flows are anisotropic
and thus the turbulent viscosity is not sufficiently represented as a scalar and there is a variety of
mixing lengths that can be considered.
As a final note, the turbulent viscosity can be written in general as

ν (t) = `∗ u∗ (4.71)

in the mixing length theory, `∗ = `m and




∂v x
u = `m
(4.72)
∂y

4.5 Energy transfer in turbulent flow


In our review of fluid mechanics we established that the viscous power of a fluid is

Ŵ ≡ ∇ · (τ · v) = (∇ · τ ) · v + τ : ∇v = f · v + τ : S (4.73)

where the first term on the rhs represents the rate of change of the mechanical energy of the fluid
and the second one represents the rate of increase of internal energy. Analogously, for turbulent
flow we have that the turbulent viscous power is:

Ŵ ≡ ∇ · (τ (R) · v) = (∇ · τ (R) ) · v + τ (R) : S (4.74)

In this case the physical interpretation is rather different, the first term represents the rate of loss
of kinetic energy from the mean flow per unit volume and the second term represents the gain of
kinetic energy by the turbulence per unit volume. This term is also called the deformation work

50
or turbulent energy generation and it represents the tendency of the mean shear to stretch and
intensify the turbulent vorticity, which ultimately translates into an increment of turbulence.
Previously, we found the following equation for the turbulent kinetic energy:
ρ D(v 0 )2
 
1
= ∇ · −p v + v · τ ii − ρ(v ) v + (τ (R) · S)ii − 2µ(S0 · S0 )ii
0 0 0 0 0 2 0 (4.75)
2 Dt 2
here we find the term (τ (R) · S)ii , which is equivalent to the gain of kinetic energy. Let us denote
this gain term as:
ρG = (τ (R) · S)ii (4.76)
The last term represents the dissipation of turbulent energy by the fluctuating viscous stresses:
ρε = 2µ(S0 · S0 )ii (4.77)
Finally, we define the transport vector of turbulent kinetic energy as follows:
1
ρT = p0 v0 − v0 · τ 0 ii + ρ(v 0 )2 v0 (4.78)
2
Consequently, the turbulent kinetic energy equation is expressed as follows:
∂k
+ v · ∇k = −∇ · T + G − ε (4.79)
∂t
where the turbulent kinetic energy is
1
k = (v 0 )2 (4.80)
2
Returning the attention to Eq. (4.71), both Kolmogorv (1942) and Prandtl (1945) suggested
that
ν (t) = ck 1/2 `m (4.81)
i.e., u∗ = ck 1/2 , with c ≈ 0.55. In this way, one may solve Eq. (4.79) to obtain k and, given that the
mixing length value is available, one may use the above equation to predict the turbulent viscosity.
This type of modeling approach is called the one equation model. However, in its current form this
model is unclosed because the viscous dissipation term, ε is unavailable. To circumvent this issue,
it has been proposed that
c3 k 3/2
ε= (4.82)
`m
In this way, the turbulent viscosity becomes:
c4 k 2
ν (t) = (4.83)
ε
where c4 ≈ 0.09. This approximation has been verified to be acceptable except near boundaries of
the flow. Finally, the energy flux vector is defined in terms of the velocity and pressure fluctuations,
which are unavailable. To tackle this problem, it has been suggested that
T = −ν (t) ∇k (4.84)
In this way, the turbulent kinetic energy equation is
∂k
+ v · ∇k = ∇ · (ν (t) ∇k) + G − ε (4.85)
∂t
Comparisons of the model predictions with experimental data shows a modest improvement with
respect to the mixing-length model (Wilcox, 2006).

51
4.6 Current models of turbulence
4.6.1 The k − ε model
This is a type of two-equation models, because it requires to solve equations for the turbulent kinetic
energy, k and the viscous dissipation, ε. This type of model is one of the most popular nowadays
and it is incorporated in software such as Fluent and Comsol. It departs from the same approach
to the turbulent viscosity given in Eq. (4.83) and the governing equation for k is still Eq. (4.85).
However, the equation for ε is entirely empirical and it is proposed by analogy from Eq. (4.85):
!
Dε ν (t) Gε ε2
=∇· ∇ε + Cε1 − Cε2 (4.86)
Dt σε k k

where the closure constants are

Cε1 = 1.44; Cε2 = 1.92; σε = 1.3 (4.87)

this is one of the simplest models of turbulence and it is acceptable to use it for simple flows.
However, for complicated situations such as flow around obstacles, it can be quite inaccurate
because of the assumptions involved in the turbulent viscosity and in the equation for ε. It is thus
necessary to make modifications to this model in order to apply it to the viscous near-wall region.

Homework
In homogeneous turbulence there are no spatial velocity gradients and the k − ε model
reduces to
dk
= −ε (4.88a)
dt
dε ε2
= −Cε2 (4.88b)
dt k
Solve this system of equations subject to k(0) = k0 and ε(0) = ε0 . The result is

k = k0 (1 + t/t0 )−n (4.89)

where t0 = nk0 /ε0 and n = (c2 − 1)−1 ≈ 1.09.

4.6.2 The k − ω and SST models


The second most widely used two-equation model is the k − ω model developed by Wilcox (2006).
This model uses the same equations for ν (t) and k. The difference is that ad hoc an equation of the
vorticity (interpreted as the specific rate of dissipation of the turbulent kinetic energy into internal
energy: ω = ε/k), is introduced:
!
Dω ν (t) Gω
=∇· ∇ω + Cω1 − Cω2 ω 2 (4.90)
Dt σω k

With this model it is a bit more complicated to achieve convergence than in the k − ε model.
For boundary-layer flows, the k − ω model is superior in the near-wall region and in accounting for

52
pressure gradients than the k−ε model. However, this model has problems to reproduce free-stream
flow, which is actually not an issue for the k − ε model.
Menter (1994) proposed a Shear-Stress Transport (SST) turbulence model that combines the
k−ω and the k−ε models. In this model, the k−ω model is applied in the inner zone of the boundary
layers and then it switches to the k − ε model in the free flow by means of a position-dependent
blending function.

4.6.3 Spalart-Allmaras model


Spalart and Allmaras (1992) proposed a one-equation model for aerodynamic applications involving
wall-bounded flows that consists of a single equation for ν (t) :
!
Dν (t) ν (t) (t)
=∇· ∇ν + Sν (4.91)
Dt σν

where σν = 2/3 and the source term Sν is a function of the laminar and turbulent viscosities, the
mean vorticity and the distance to the nearest wall. It has been applied with success to aerodynamic
flows and in turbomachinery but it has been proved to be incapable of accounting for the decay of
ν (t) is isotropic turbulence.

4.6.4 Reynolds stress model


This is probably the most sophisticated turbulence model in the sense that it attends many of the
deficiencies of the k − ε model and its variants. It involves the RANS equation:

Dv
ρ = −∇p + ∇ · (τ + τ (R) ) (4.92)
Dt
In addition, the equation for the Reynolds stress, previously obtained (see Eq. 4.26) is considered:

Dτ (R)
= −2p0 S0 − τ (R) · ∇v − (∇v)T · τ (R) + ρε + ∇ · H (4.93)
Dt
where, the third-order tensor H is defined as:

H = ρu0 u0 u0 + ν∇τ + Ip0 u0 + p0 u0 I (4.94)

and the rate of dissipation tensor is

ε = 2ν(∇v0 )T · ∇v0 (4.95)

Clearly, at this point the model is unclosed because the relation of p0 S0 , H and ε with the stress
tensor is unavailable and few progress seems to have been gained by writting the stress equation in
this form. To address this issue, the following relations have been proposed:
2
ε = εI (4.96a)
3
k
H = 0.22α · ∇τ (R) ; α = uu (4.96b)
ε

53
hε  
i 2
2p0 S0 = cR τ (R) + ρε − ρĉR P − GI (4.96c)
k 3
with cR = 1.8, ĉR = 0.6 and
1
ρP = τ (R) · ∇v + (∇v)T · τ (R) , G = Pii (4.97)
2
Finally, a dissipation of energy equation is needed, which is

Dε Gε ε2
= ∇ · (0.15α · ∇ε) + Cε1 − Cε2 (4.98)
Dt k k
A salient feature of this model is that anisotropies can be taken into account by means of the tensor
α. The weakest point of this approach is the empirical relation for the pressure given in Eq. (4.96c).
Nevertheless, it has been found that this approach works well for homogeneous turbulence and free
shear layers.

4.6.5 DNS and Large-eddy simulation


DNS
Direct numerical simulations (DNS) consist on directly solving the NS equations in order to compute
the complete cascade of eddies. This is basically equivalent to perform numerical experiments.
Current codes are being improved year after year as well as the computational capabilities that
may, in the future, make this a viable modeling approach that would not involve any type of
averaging at all. However, nowadays there is an obstacle for its application. To better understand
this issue, let us recall that the characteristic size of the smallest eddies, η, is estimated as:

η = O(Re−3/4 `) (4.99)

where ` is the size of the largest eddies. Evidently, as the Reynolds number is increased, the size η
becomes even smaller and the mesh size should also decrease if we accept the estimate ∆x = O(η).
Furthermore, the computational time can be estimated according to the following expression:
   3
tmax L
Computer time = O Re3 (4.100)
`/u `

where tmax and L represent the maximum computer time and the maximum length of the system.
For tmax = 10`/u and Re = 104 and L = 100`, 1 TB computer would take 2 centuries to solve the
problem of flow without boundaries. Therefore, for the time being DNS is not a feasible alternative.

LES
The Large Eddy Simulations (LES) is an approach that falls in between DNS and turbulence closure
schemes. The main idea in LES is to compute the mean flow and the large energy-containing eddies
in an exact manner without computing the energy dissipation from the smaller eddies. This seems
reasonable because usually energy (and information) tends to travel from the larger eddies towards
the smaller eddies. Obviously, this approach is recommendable when the interest is in the largest
eddies as it is the case, for example, in wind farms and in chemical pollutants transport in cities.

54
(b)

LES, (b) the

perhaps
Figure 4.6:the large
Sketch of scales will of
the filtering nota random
notice variable
the absence of the
to produce small scales
a spatially smooth field
(Figure 7.2(a)).s
This approach has been gaining popularity in engineering and meteorological modeling. The key
The attraction of LES lies in the fact that often (though not always)
idea in LES is filtering. The aim is to distinguish between small and large scale variations, in this
way,itthe
is spatial
the large scalesoperator
averaging which isare the most
introduced as important.
the convolutionForproduct:
example, they
dominate the transfer of momentum, Z∞ heat, and chemical pollutants.
Yet in DNS virtually all of uLthe
(x, t)effort
= goes
u(x −into computing the small-to- (4.101)
r, t)G(r)dr
intermediate scales. This can be −∞ seen from the following simple

The example. Suppose


filtered velocity uL that femax is
represents thethe maximum
motion waveeddies.
of the large numberInused in a G(r) is a
addition,
spectral (i.e.,
homogeneous simulation. Todepend
it does not obtainofresolution at theand
x) filter function, Kolmogorov scale we
it is defined as:
needfemaxí?~ n. However, the G(r)wave numbers
= 1/L, |r < L/2|which characterize the
large, energy-containing scales, (4.102)
G(r) k=B,0, satisfy
|r| > L/2kBl^n. It follows that
4
kEr^kmax(ri/l)^kmaxR.e . Suppose, for example, that Re = 103
Note that it satisfies the normalization condition:
(i.e. R¿~120). Then k3E ~2 X 10"7fe^ax and we conclude that only
Z∞
0.01% of the modes used in the1 =simulation G(r)dr
are required to calculate (4.103)
the behaviour of the all-important −∞ large eddies (defined, say, by
kE < k < 10feB), the other 99.99% of the modes going to simulate the
Evidently, the average is performed over a distance L that should be large enough so that the
small-to-intermediate
fluctuations scales.
taking place at scales Clearly,
smaller than for
L are those interested
spatially smoothed in as
theillustrated
effects in Figure
4.6.
The residual field is defined as:
Actually they probably do! See Lesieur (1990), or Hunt et al. (2001). Lesieur, for
u0 (x, t) =inu(x,
example, notes that random fluctuations − uL (x,
thet)small t) will, eventually, lead to
scales (4.104)
random fluctuations in the large scales, over and above those generated by the large-
This decomposition seems similar to the one used for the RANS. However, the velocity is now a
scale dynamics themselves. These are not captured by an LES as the information con-
random field, therefore
tained in the small scales is suppressed. In this sense LES is ill-posed. Nevertheless, we
(u0 )L 6= 0 (4.105)
might still expect an LES to capture the statistical trends and typical coherent structures
of the large-scale motion.
55
Consequently,
(uL )L 6= uL (4.106)
Clearly, this type of spatial averaging is not the same as the typical volume averaging explained
by Whitaker (1999). Another important difference with volume averaging is that, since integration
is performed in the complete domain, the operations of spatial and temporal differentiation and
integration are commutable, i.e.,
 L
∂vL ∂v
= ; ∇vL = (∇v)L (4.107)
∂t ∂t
The result of performing the averaging of the NS equations is simple:
∂vL 1
+ ∇ · (vv)L = − ∇pL + ν∇2 vL (4.108)
∂t ρ
obviously, vL is solenoidal. Without decomposing the velocity in the inertial term, the above
equation can be reformulated as follows:
∂vL 1 1
+ ∇ · (vL vL ) = − ∇pL + ν∇2 vL + ∇ · τ R (4.109)
∂t ρ ρ
where,
τ R = ρ vL vL − (vv)L
 
(4.110)
where τR is the residual stress, which is expressed in an analogous manner as the eddy viscosity
model:
1
τ R = 2ρνR SL + IτiiR (4.111)
3
where νR is the eddy viscosity of the residual motion and τiiR denotes the diagonal components of
the τ R tensor. In this way, the average equation takes the form
∂vL 1
+ ∇ · (vL vL ) = − ∇p∗L + 2∇ · [(ν + νR )SL ] (4.112)
∂t ρ
where p∗L is a modified pressure, defined as
τiiR
p∗L = pL − (4.113)
3
The final element is the residual viscosity, which is related to the viscosity of the larger eddies and
Smagorinsky (1963) proposed to use
νR = CS2 L2 (2ST : S)1/2 (4.114)
where the Smagorinsky coefficient CS = 0.1. This model seems to work well for isotropic turbulence
and free shear flows and it has been modified to be able to study near-wall problems. From a
computational viewpoint, a shortcoming of this method is that it requires much more computational
resources than the k − ε model and its variants. Evidently, a shortcoming of LES is the study of
boundary layers. To circumvent this issue, one may use a very fine mesh, so that the model results
resembles those from DNS (although at a high computational cost) or study the boundary layer
separately with a suitable eddy viscosity model and then couple it with LES. This last hybrid
approach is known as the detached eddy simulation (DES), and has gained popularity in the
recent years. When the computational resources are limited, the Reynolds stress model is a better
alternative than LES.

56
5 | The phenomenology of Taylor-
Richardson and Kolmogorov

Contents
5.1 Richardson’s theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.2 Kolmogorov’s theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3 Turbulent mass transport . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4 Random walk and Taylor diffusion . . . . . . . . . . . . . . . . . . . 67
5.5 Richardson’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

5.1 Richardson’s theory


In previous discussions we have established that mechanical energy is transferred from the largest
scale and it is extracted at the smallest scale. The largest vortices arise because of an instability
in the mean flow vortex lines. Their size then is comparable to the characteristic length associated
to the mean flow. Richardson addressed the question about how does energy is transported from
the largest to the smallest vortices by introducing the idea of an energy cascade. He suggested that
the largest eddies pass their energy to the smaller ones and these in turn pass their energy to even
smaller eddies and so on. The essential feature of this idea is that this a multistage process involving
a hierarchy of eddies. What we mean by eddies is a blob of vorticity which can be spherical, tubular,
sheet-like or even more complex as sketched in Figure 5.1. This energy cascade is a form of energy
Richardson revisited

e 'blobs' of vorticity.
al, tubular or sheet-like
re complex.

Figure 5.1: Sketches of the different shapes of eddies containing vortices.


distortion or instability of the mean flow vortex lines. Their size then
57
corresponds to a length scale characteristic of the mean flow, for
example, the length associated with gradients in the mean velocity field.
Therefore, we have mechanical energy transferred to the turbu-
lence at a large scale, and extracted at a much smaller one. The
transfer from one scale to another through a distortion of the eddy shape.
Richardson suggested that at sufficiently large values of Re, the energy cascade is driven by
inertial forces only so that viscosity plays no part in the cascade, except at the smallest scales. In
this way, the smallest eddies are dissipated by viscous forces because the Re value decreases to the
order of unity.
Let us put these ideas in mathematical terms. The rate of inlet of turbulent energy from the
mean flow to the largest eddies is
ρG = (τ (R) )T : S (5.1)
The rate at which passes down the energy cascade is denoted by Π. In this way, there is a finite
number of transitions of energy that we denote by N . For the first element of the cascade we have
the following order of magnitude estimate for the rate of energy:
 2  3
u u
ΠA = O ∗ = O (5.2)
t `

here we took `/u to be the characteristic time of the first eddy. As defined in other chapters, u and
` denote the velocity and characteristic length of the largest eddy.
Finally, the smallest eddies dissipate energy at a rate that can be estimated to be
 2
νv
ε=O (5.3)
η2

where v and η denote the velocity and characteristic length of the smallest eddies, respectively.
Consequently, we have

u3 νv 2
G = ρ−1 (τ (R) )T : S = = ΠA = ΠB = · · · = ΠN = ε = 2 (5.4)
` η

while this equality is likely to hold for homogeneous and statistically steady turbulence. However,
in terms of orders of magnitude estimates, the above expression may be expected to hold.
Of course some exceptions to Richardson’s theory exist. For example, there are cases in which
the flow of energy goes from the smaller to the larger eddies or when the energy goes directly from
the largest to the smallest eddies without going through the whole cascade. These typical and
atypical energy cascades are illustrated in Figure 5.2.

5.2 Kolmogorov’s theory


Let us commence by recalling the structure function:

h[∆v(r)]2 i = h[u0x (x + rex ) − u0x (x)]2 i (5.5)

We have discussed about this function before and we noted that is of the order of all the energy of
eddies of size r or less. In homogeneous and freely-decaying turbulence, after sufficient time, the
memory effects of the initial condition should no longer affect the flow. The functionality of the
structure function may be proposed to be

h[∆v(r)]2 i = F̂ (u, l, r, t, ν) (5.6)

58
Richardson revisited

Energy cascade
Log (energy)

Dissipation of
energy at rate Ê

menology
Schematic of Taylor, Richardson,
representation of the and Kolmogorov
ade. Eddies depend on / and « Eddies depend on v

Log(£)
Evidently, in the case of Batchelor's data we have H = |Aw3/!,
where A is approximately constant during the decay and of the order
of unity, A ^1.1 ±0.2. Note, however, that the value of Re in these
early experiments is modest. The more recent data of Pearson is at
much higher Re and suggests that the asymptotic value of A is A ~ 0.3.
This is consistent with Kaneda et al.'s (2003) findings.
rocesses which violate NowLog consider the smallest scales. Suppose
(wave number) they have
Log (wave a characteristic
number)
cascade. Why do these not (i) Can energy move to larger (ii) Can energy transfer
velocity v and length scale t]. Since the rate of dissipation directly to of
ly, sometimes they do.) Figure 5.2:through
scales Typical andmerger?
vortex atypical
2 (but still
samll possible)
scales, energy
bypassing cascades.
the cascade?
mechanical energy is i/(w ) we have

e ~ i/v2 / v f . 5.3
We also know that
In homogeneous, statistically steady turbulence the rate of extrac-
v/7/z/~ 1 (5.6)
tion of energy from the mean flow, T^Sy/p, must equal the rate at
and
which so energy
from (5.5) and down
is passed (5.6) we can derive
the energy estimates
cascade from theforlarge
t] and v:
scales,
3
HA ~ u /l. This 3must also equal the rate of transfer of energy at all
d/!/)- /4- (S/ef
r\ ~ l(ul¡v) 4
(5.7)
points in the cascade since we cannot lose or gain energy at any
i/4
v ~ u(ui/^r r
particular scale in a steady-on-average flow. In particular, if YIA,5.8
YiB,..., YIN represents the energy flux at various stages of the cascade
These
then weare, haveofYiAcourse,
= YIB = the• • • =Kolmogorov microscales
59N ~ u3¡I (Figure
YI 5.3). So hitroduced
the energy hi
Chapter
transfer even 1. When
in the the flow
small eddiesis neither homogeneous
is controlled by the ratenor statistically
of break-up
steady
of the largethen eddies.
G and Finally
e needwe notnotebalance.
that theHowever, it is
energy flux at usually
the end found
of
that H and e are of the same order of magnitude
the cascade, IIN, must equal the viscous dissipation rate, e. In sum- and so (5.5)-(5.8)
still
mary, hold.
then, for homogeneous, statistically steady turbulence,
For r  η, the viscous forces can be neglected and the above functionality can be rewritten in a
dimensionless form as follows:
h[∆v(r)]2 i = u2 F (r/l, ut/l) (5.7)
where F is now a dimensionless function. Furthermore, in decaying free turbulence, the time
dependence can be dropped and thus:

h[∆v(r)]2 i = u2 F (r/l) (5.8)

where
F = 2(1 − f ) (5.9)
with f being the longitudinal correlation function.
Actually, Kolmogorov was interested on eddies of size r  `. He claimed that these small
vortices are statistically isotropic, in statistical equilibrium and of universal form. So that, for these
eddies, an equation similar to the one given in Eq. (5.8) can be postulated:

h[∆v]2 i = v 2 F (r/η) (5.10)

where, as shown in Chapter 1, v = (νε)1/4 and η = (ν 3 /ε)1/4 and F is expected to be a universal


function valid for all forms of turbulence. This is the basis of Kolmogorov’s universal equilibrium
theory, from which we have the first similarity hypothesis:
When Re is large enough and r  `, the statistical properties of ∆v have a universal
form which depends only on ε, r and ν.
To have a more clear view of these features, let us consider the following points, which are
applicable to eddies of size r  `:
1. They do not seem to retain information from two or three generations before them. Moreover,
it seems unlikely that they are influenced by the velocity field of large-scale eddies, which
seems to be almost uniform at the scale of small eddies. This is because large-scale eddies
evolve very slowly in comparison with the small eddies.

2. Large-scale eddies are anisotropic and statistically unsteady. The origin of this behavior is the
mechanism that gives rise or maintains the turbulence. However, since smaller scale eddies
do not feel directly the changes of large-scale eddies, it seems plausible that they do not feel
the changes with position and time that the larger eddies experience. This means that, at
any instant, the small eddies are in approximate statistical equilibrium and they are more
less isotropic. Hence, the regime in which r  ` is known as the universal equilibrium range.
The universal equilibrium theory works remarkably well since it can reproduce experimental
data arising from many sources. Moreover, there is experimental evidence that local isotropy exists
at very small values of r. However, recent evidence shows that the universal equilibrium theory is
applicable before local isotropy is achieved.
Let us now consider a sub-domain of the universal equilibrium range called the inertial subrange
in which η  r  `. In this range Kolmogorov proposed his Second Similarity Hypothesis, which
states that

When Re is large and in the range η  r  `, the statistical properties of ∆v have a


universal form which is usually determined from r and ν alone.

60
Kolmogorov revisited

Log (E)

Inertial subrange

Log(fc)
Figure 5.18 Different ranges in the energy
spectrum. Universal equilibrium range

Figurecan derive
5.3: (5.22) either
Sketch of thebydifferent
dimensional arguments,
ranges as we
in the did to obtain
energy spectrum.
(5.21), or else by noting that

Let us return to Eq. (5.10) and 2write itE(k)


<[Av(r)] in therms
dk of ν, r and ε:
n/r
!
(see Section 2.5 of Chapter
2
3). 1/2 rε1/4
h[∆v]wei may
In summary, then, = (νε) F spectrum
divide the (5.11)
ν 3/4 of eddies up into
three ranges (Figure 5.18), as shown below:
To cancel ν, the only possibility is that F (x) = x2/3 , in this way the famous two − thirds law of
Name Range Form of ([Av"I2}
Kolmogorov is recovered:
2
Energy containing eddiesh[∆v] i = βε2/3
r~ I r 2/3 {[Av]2)=«2FC (5.12)
(freely decaying turbulence only)
Where the universal constant β has been found to r¡have
Inertial sub-range
a value{[A
<C r <C I
of V2. When
]2}=£ E
2/ this law is formulated
in terms of the energy (all
spectrum, it is known as Kolmogorov’s five-thirds law :
types of turbulence)
Universal equilibrium range r <C I {[AV]2)=v2F(r/tf
(all types of turbulence)
E(k) = αε2/3 k −5/3 (5.13)

where α ≈ 0.76β. We shall return to this table in Chapters 6 and 8 where we see that
In summary, we may divide
it must the spectrum
be modified. of eddies
In the meantime intothatthree
we note ranges
very high (see
values of also Figure 5.3):
Re are necessary in order to obtain an inertial sub-range. Recall that
t] ~ (ul/i/)~3 4!. If we are to obtain a range of r in which r\ <C r^C !,
then
Namewe need Re3 s 3> 1. This is difficult to achieve in a h[∆v]
Range wind-tunnel,
2i
though it can and has been done.
Energy-containing
There is an alternative eddies
means of deriving r =theO(`) u2law
two-thirds F (r/`)
which
isInertial subrange
often attributed to Obukhov. Obukhov's η theory βε2/3
r is `usually r2/3in
framed
2
 `stay invreal2 F (r/η)
Universal
terms of E(K), equilibrium range
rather than ([Av(r)] ), but wer shall space in
our description. Suppose that eddies of size rhave a typical velocity vr.
We have seen that the flux of energy down the cascade is constant
As a final note, it is(provided
worth we have statistical
recalling that equilibrium) and so the cascade of energy
at each point in the universal equilibrium range, H(r), must be equal
p p
to £. Also, we h[∆v]
havei seen
= h[u that + rex3)/!−because
x (xI1(!)^M ux (x)]thei large eddies (5.14)
evolve (break up?) on a timescale of their turn-over time. Now
From the developments presented thus far, it may appear that the following equation is applicable:

h[∆v]p i = βp (εr)p/3 229 (5.15)

Clearly, for p = 2, we recover the two-thirds law and β2 = 2. For p = 3 we have

h[∆v]3 i = β3 εr (5.16)

61
where β3 = −4/5 for globally isotropic turbulence. The above result is the famous four-fifth law of
Kolmogorov. However, for powers larger than 3, Eq. (5.15) does not perform that well.
Landau and Lifshitz (1987) argued that it is not the globally averaged dissipation, hεi, which
is important, but the local dissipation averaged over a volume larger than r but smaller than l.
This local average is a random function of time and space and can change from one type of flow to
another. This problem arises because ε is extremely intermittent in space, i.e., there are regions
of large dissipation and regions of small dissipation. It thus follows that, for eddies of size r  `,
(which are the eddies responsible of the energy flux cascade) should be controlled by ε = 2νST : S
averaged over a volume of size r instead of the global average used above.
In this way, let us define the finite volume average as:
Z
1
εAv (r, x, t) = 2νST : SdV (5.17)
Vr
Vr

with Vr being an averaging domain of size r centered on x. Note that hεAv i = ε. In terms of this
average, we may rewrite Eq. (5.15) as

p/3
h[∆v]p i = βp hεAv irp/3 (5.18)

Certainly, for p = 3, the four-fifth law is recovered. However, for other values of p, we need to know
something about the relation between εAv and ε. Kolmogorv (1962) proposed a simple statistical
model for εAv called the log-normal model that leads to the following correction

h[∆v]p i = Cp (εr)p/3 (`/r)µp(p−3)/18 (5.19)

where µ is the intermittency exponent 0.2µ < 0.3 and Cp is non-universal. The latter is consistent
with Landau’s point of view that intermittency changes from one flow to another.

5.3 Turbulent mass transport


Let us now direct the attention to turbulent mixing of a scalar quantity such as the concentration C
that is diluted in a stream of fluid undergoing turbulent transport. In the absence of homogeneous
chemical reactions, the governing equation for the contaminant concentration is

∂CA
+ v · ∇CA = DA ∇2 CA (5.20)
∂t
Clearly, this equation may also be used to model turbulent heat transport for cases in which there
is no homogeneous source or sink of heat. Since the concentration is a scalar that does not influence
the flow, we to it as a passive scalar. Fur sufficiently large values of the Péclet number (u`/DA ),
the chemical species act like a tracer of the fluid particles. There are three interesting problems
that can be studied with passive scalars (see Figure 5.4):

1. Taylor diffusion: This problem corresponds to the continuous release of a contaminant in


a turbulent fluid stream from a single point source. The question that needs to be answered
is about the time it takes for the contaminant to spread.

62
Kolmogorov revisited

Problem 1: Taylor's problem

,, /

Problem 2: Richardson's problem

5.20 In Taylor's problem we consider


ntinuous release of a contaminant from
point and estimate the average rate of
h of the contaminant cloud. In
dson's problem a small puff of
minant is released at t = 0 and we aim
ermine the rate of growth of the puff
ounces around in the turbulent flow.
blem 3 the contaminant is unevenly
ed throughout the medium. Problem 3: Local variations in scalar intensity

Figure 5.4: Sketches of Taylor diffusion, Richardson’s law and local turbulent fluctuations.
scale of die large eddies. The scalar C then acts like a marker which
tags the fluid particles. Diffusion is important only at scales which are
characterized by Pe^ 1. We shall take both M!/«^> 1 and M!/Í/^> 1
throughout this section so that diffusive effects are restricted to the
microscales of the turbulence. We shall also assume that there is no
mean flow. Although this second restriction is rarely achieved in
practice, it has the merit of greatly simplifying the analysis.
There are three interrelated problems which often arise in the study
of passive scalars (Figure 5.20):
63
Problem 1 Taylor diffusion
Imagine that dye is released continuously into a turbulent flow from a
single point source. As time progresses the cloud of dye will spread by
turbulent mixing. A natural question to ask is; how large, on average,
will the cloud be after a time t? Since ul/ot.^> 1 this is equivalent to
2. Richardson’s law: Reconsider the above problem but instead of having a continuous release
of contaminant supose that we only release the contaminant once at t = 0. The initial puff
should have a size that is larger than the Kolmogorov scale but smaller than the integral
scale. The centroid of the puff will move with the fluid and its size should increase with
the movement as the result of small-scale turbulence. In this problem, we are interested
on determining the average rate of spreading of the puff. To be more specific, we want to
determine the average rate of separation of two adjacent fluid particles located at opposite
sides of the puff.

3. Local turbulent fluctuations in scalar intensity: Now supose that the passive scalar is
distributed unevenly throughout the field of turbulence. An example of this situation is the
problem of mixing in a CSTR. In this case we are interested on knowing the time it takes to
achieve perfect mixing.

We shall begin with the last of these problems. To this end, it is convenient to form the
time-average of Eq. (5.20), the result is:

∂C A
+ ∇ · vC A = DA ∇2 C A − ∇ · v0 CA
0 (5.21)
∂t
where, by analogy with turbulent flow and Fick’s law, let us define the turbulent molar flux as:
(t) 0 = −D ∇C(t)
J A = v0 CA A A (5.22)

In this way, the time-averaged mass conservation equation for species A is:

∂C A (t)
+ ∇ · (vC A ) = ∇ · [(DA + DA )∇C A ] (5.23)
∂t
Let us now direct the attention to the equivalent of the mechanical energy equation for mass
transport. To this end, let us first multiply Eq. (5.20) by CA to obtain
   
∂ 1 2 1 2
C + ∇ · v CA = DA ∇ · (CA ∇CA ) − DA ∇CA · ∇CA (5.24)
∂t 2 A 2
For the developments that follow, we are interested on the ensemble average of the variance of C,
defined as hC 2 i. For this reason, we make the ensemble average of the above equation to obtain
   
∂ 1 2 1 2
hC i + ∇ · v CA = DA ∇ · hCA ∇CA i − DA h∇CA · ∇CA i (5.25)
∂t 2 A 2
For spatially-homogeneous turbulence and mass transport, this result is considerably reduced to
 
d 1 2
hCA i = −DA h∇CA · ∇CA i (5.26)
dt 2
Or, after splitting the concentration in the right-hand side term into its ensemble average and
deviations and taking into account the spatial homogeneity assumption, we obtain:
 
d 1 2 0 0
hCA i = −DA h∇CA · ∇CA i (5.27)
dt 2

64
This last term denotes the rate at which contaminant fluctuations fade away due to diffusion and
it is an analogue of ε, that we denote as:
0 0
εc = DA h∇CA · ∇CA i (5.28)

Let us now denote the length-scale over which CA0 exhibits significant variations to be η , in this
c
way, we can estimate the order of magnitude of εc to be

CA02 
εc = O DA 2 (5.29)
ηc
By analogy of the energy cascade described before, we can now propose that there is a corresponding
cascade for hCA02 i. The idea is the following: when a large eddy breaks up into smaller eddies, the

contaminant becomes more finely mixed over and over until the moment in which the characteristic
length of the fluctuations is ηc and diffusion sets in. In this way, for r  ηc , the details of the
cascade are independent of diffusion.
We are interested on determining the analogue of Kolmogorov’s two-thirds law for mass
transport. In this way, let us define the structure function

h[∆CA ]2 i = h[C(x + r) − C(x)]2 i (5.30)

Now consider the intermediate range of length scales, which is known as the inertial-convective
subrange:
ηmax = max[η, ηc ]  r  min[`, `c ] = `min (5.31)
with `c being the characteristic length of the large-scale variations of CA . In this range of length-
scales it is to be expected that neither viscous stress nor mass diffusion would play a role over the
structure function h[∆CA ]i. Thus, in this range, the structure function should be influenced only
by r and the dissipations ε and εc :

h[∆CA ]2 i = f (ε, εc , r) (5.32)

Now we use dimensional analysis and note that only εc has units of concentration square, hence:

h[∆CA ]2 i = O(εc f (ε, r)) (5.33)

since εc [=]mol2 /s, ε[=]m2 /s3 and r[=]m, the only possible combination is

h[∆CA ]2 i = O(εc ε−1/3 r2/3 ), ηmax  r  `min (5.34)

which is the analogue of Kolmogorov’s 2/3 law and was first suggested by Obukhov (1949) and
Corrsin (1951) and it has been confirmed with experimental data.
At this point, it is worth recalling that the ratio between the kinematic viscosity and the
diffusivity defines the Schmidt number:
ν
Sc = (5.35)
DA
When Sc > 1, diffusion of CA is less important than the diffusion of vortices and we should expect
that the fine scale structure of CA will develop in ηc < η. To verify this, let us estimate the order
of magnitude of ηc in this condition, which has been found to be
 r 
1/2 η
ηc = O α , Sc > 1 (5.36)
v

65
Since, Re Reynolds number is of order 1 in the Kolmogorov microscale, we have
 
ν
v=O (5.37)
η
hence:  r 
α ν
ηc = O η , Sc = >1 (5.38)
ν α
This proves that indeed, ηc < η in the regime.
Let us now consider the opposite situation, i.e., Sc < 1. In this case, we should expect that
ηc > η and this range is known as the inertial-diffusive subrange. In this range, we require diffusion
to compete with convection, this requires that the Péclet number should be of order one:
DA
ηc = , Sc < 1 (5.39)
vc
with vc being the characteristic velocity at the length-scale ηc and it is estimated to be

vc = O (εηc )1/3 (5.40)

Consequently, we have
DA
 
ηc4/3 =O (5.41)
ε1/3
Or, since ε = O(νv 2 /η 2 ) and v = O(ν/eta), we have
3/4 !
DA

ηc = O η , Sc < 1 (5.42)
ν

The above gives us an idea of the characteristic length scale for mass transport in which CA 0

exhibits significant variations for conditions in which Sc > 1 or Sc < 1, which is similar to our
developments about Kolmogorov’s theories. As a final point, let us estimate about the speed of
mixture of the contaminant, which is our main objective.
From the 2/3 law for mass transport, we can obtain the estimate:
2
hCA i = O(εc ε−1/3 r2/3 ) (5.43)

from which we can estimate εc to be


2 1/3 −2/3
εc = O(hCA iε r ) (5.44)

Or, for r = ` ≥ `c
2 1/3 −2/3
εc = O(hCA iε `c ) (5.45)
and taking into account that in this regime ε = u3 /`:
2
εc = O(hCA iu`−1/3 `−2/3
c ) (5.46)

Using this estimate in Eq. (5.27) yields


 
d 1 2 2
hCA i = O(hCA iu`−1/3 `c−2/3 ) (5.47)
dt 2

66
Certainly, for the case in which `c = O(`), we have:
   2 
d 1 2 hCA i
hCA i = O (5.48)
dt 2 `
which is similar to the estimate of the rate of kinetic energy in free-decaying homogeneous
turbulence:  
d 1 2 u 
v = −ε = O hv2 i (5.49)
dt 2 `
2 i and hv2 i decay at the same characteristic time, which corresponds
Clearly, in this case, both hCA
to the large-eddy turnover time.
For the more general case in which `c 6= `, we have:
2i  2/3
d lnhCA `
=m (5.50)
d lnhv2 i `c

where m is presumably a number of O(1).

5.4 Random walk and Taylor diffusion


Before moving on to study Taylor’s diffusion and Richardson’s law, it is convenient to study random
walks. To understand this concept consider a drunken man who exits from a bar and takes a
sequence of steps of length `. However, each step is taken at a random direction. The question is,
after taking N steps, how far has the man gone from the bar?
To answer this question, let RN be the position of the man after the N th step and LN the
displacement vector of the N th step, then the magnitude of this vector is:
2
RN = |RN | = RN · RN (5.51)

Since RN = RN −1 + LN , we have:
2 2 2
RN = (RN −1 + LN ) · (RN −1 + LN ) = RN −1 + 2RN −1 · LN + LN (5.52)

Since LN is a random variable, it follows that hLN i = 0. Furthermore, the direction LN is


uncorrelated (i.e., independent) of the position where the man was RN −1 , therefore hLN · RN −1 i =
0. After a large number of steps, we can take the ensemble average of the above equation to obtain:
2 2 2
hRN i = hRN −1 i + hLN i (5.53)

Since the length of the steps is `, it is easy to deduce that

hL2N i = `2 (5.54)

Therefore, we have the recurrence formula:


2 2 2
hRN i = hRN −1 i + ` (5.55)

from which we can obtain that:


2 2
hRN −1 i = (N − 1)` (5.56)

67
Therefore,
2
hRN i = N `2 (5.57)
So the distance advanced by the man is on the order of N 1/2 , rather than N for a sober man
walking in a straight line. Now let us apply this concept to turbulence. Let x(t) be the position of
a fluid particle at a given time. The Lagrangian velocity is defined as: v = dx/dt. In the following,
we assume that the flow into which the particle is released has zero mean velocity, is homogeneous
and statistically steady.
In this way, the time-rate of change of the magnitude of the square of the displacement is:

dx2
= 2x · v (5.58)
dt
However, since
t0 =t
Z
x(t) = v(t0 )dt0 (5.59)
t0 =0

we have
t0 =t
dx2
Z
= 2v(t) · v(t0 )dt0 (5.60)
dt
t0 =0

At this point, it is convenient to introduce the change of variables

τ = t − t0 (5.61)

so that
Zτ =t
dx2
=2 v(t) · v(t − τ )dτ (5.62)
dt
τ =0

Now imagine releasing many particles and then taking the ensemble average to obtain:

Zτ =t
dhx2 i
=2 hv(t) · v(t − τ )idτ (5.63)
dt
τ =0

At this point, it is convenient to introduce the Lagrangian velocity correlation tensor :

QL (τ ) = hv(t)v(t − τ )i (5.64)

Note that QL is only a function of τ because we have assumed turbulence to be statistically steady.
With this definition at hand, we have now:

Zτ =t
dhx2 i
=2 QL
ii (τ )dτ (5.65)
dt
τ =0

Note that at t = t0 , we have τ → 0 and thus QL 2


ii (0) = hv (t)i. Conversely, if τ  t, the two
velocities are uncorrelated and thus QL
ii ≈ 0.

68
Let us now define the Lagrangian correlation time to be
τZ=∞
1
tL = 2 QL
ii (τ )dτ (5.66)
hv i
τ =0

This quantity is a measure of of the time in which a fluid particle retains some memory of its
release. In terms of this time scale we have the following extremes:

• for t  tL (early times, τ → 0):

Zτ =t
dhx2 i
=2 hv 2 (t)idτ = 2hv 2 it (5.67)
dt
τ =0

Notice that we have omitted the time dependence of hv 2 i on the basis of the time-steady
assumption. Taking as initial condition x(0) = 0, the above equation yields

hx2 i = hv 2 it2 (5.68)

Therefore, at early time the distance varies linearly with time and velocity. In other words,
at early time, the contaminant cloud will grow at a rate R = O(t).

• for late times, t  tL , we have


τZ=∞
dhx2 i
=2 QL 2
ii (τ )dτ = 2hv itL (5.69)
dt
τ =0

Or, after time-integration,


hx2 i = [2hv 2 itL ]t (5.70)
In this case, the contaminant cloud growth is proportional to the square root of time as in the
random walk example. Note that tL = O(`/v 0 ), therefore, for t  tL , we have hx2 i  `2 . In
other words, the mass spreading (or the contaminant cloud) is much larger than the integral
scale.

5.5 Richardson’s law


Let us now consider the discrete, rather than the random release of contaminant. This is called
relative diffusion and relates to the diffusion of a single puff released at t = 0. In other words,
the average rate at which two particles separate from each other gives a measure of the rate of
spreading of the cloud.
Consider two particles that are adjacent at t = 0 and then they are swept around by turbulent
eddies. At first they move as a pair, but as they bounce around they will tend to separate. We are
interested on the relative separation rather than the movement of the pair.
To this end, let R(t) be the mean radius of the contaminant cloud at time t. We are interested
in clouds of size η  R  ` and of eddies of size R as illustrated in Figure 5.5, because they make
the most significant contribution to the growth of the cloud. As a consequence, we may expect

69
menology of Taylor, Richardson, and Kolmogorov

Eddies smaller than R

Eddies of size R

The effect of different eddy sizes


Eddies much lar
of contaminant. ger than R

Figure 5.5: Effect of eddies size over the spreading of a contaminant cloud.
It is important to understand that this problem is very different to
the diffusion of a single particle. To understand why, consider two
that the time rate of change of R is of the same order of magnitude of the velocity vR , i.e., the
particles which are adjacent at t = 0. To some extent they tend to
velocity associated to eddies of size R:
move as a pair as they are swept around by the turbulent eddies.
However, as they bounce around dR they will also tend to separate and it
= O(vR ) (5.71)
is that relative separation, rather dt than the movement of the pair,
Since ε = O(vR 3 which
/R), weis haveof interest here. If the particles are initially very close we
might anticipate that this relativedR separation is much slower that the
= O(εR)1/3 (5.72)
absolute movement of the pair, dt and indeed this is the case. So a small
This expressioncloud of contaminant
is usually rewritten as will spread by turbulent diffusion, but at a rate
somewhat slower than the average rate of movement of its centre of
dR2
mass. The question, though, is = how
O(ε1/3 much
R4/3 slower?
) (5.73)
dt
Let R(t) be the mean radius of the cloud at time t. We are interested in
this is known ascloudsRichardson’s
for which four-thirds law. Although
t] <^R<^1, where he was
! is the integral interested
scale. Evidentlyineddies
the dispersion of a
small cloud of contaminants,
of size R are most it is often more for
important convenient
the growth to study
of thethe relative
puff, since separation
smaller of a pair of
fluid particles. Let δx(t) be the instantaneous distance between two fluid particles that we released
eddies merely ripple the surface of the cloud while larger ones tend to
at t = 0 from two fixed adjacent points. According to Richardson’s four-thirds law, we have:
advect it without a change in shape. This is illustrated in Figure 5.34.
Now we have already d suggested
h|δx| 2
i = O(εthat1/3 eddies
h|δx|2 i2/3of
) size r tend to break (5.74)
dt
up on a timescale of their turn-over time, and so e ~ v^/r where e is
In statistically the
steadyrate flows,
of dissipation
we haveofthat
turbulent energy per unit mass.
ε is time-independent andItthefollows
abovethat
equation can be
1 3
time-integratedvusing
r ^(er) the .initial
Consider now ah|δx|
condition cloud
2 i(0)which
= 0 tohas grown sufficiently to
obtain
have forgotten the conditions of2its initial3 release, yet whose radius is
h|δx| i = gεt (5.75)
still significantly smaller than the integral scale, !. Since only eddies of
size R contribute
where g is Richardson’s constant,towhich
the rate
hasofbeen
change
foundof to
R, vary
we might
betweenexpect
0.06that
andthe
O(1). This wide
average rate of change of R should be independent of v, u and
range of values is still under debate nowadays. This result should be expected to fail for cases in ! and
depend only on VR and t: dR/dt=J(vn,t). The only dimensionally
consistent possibility is 70
which the cloud diameter exceeds the large eddy size, i.e., for R > ` because there would no longer
be eddies of size R. In this case, the dispersion of the cloud is essentially the same as in Taylor’s
diffusion, i.e., R = O(t1/2 ).
As a final comment, it is worth stressing that turbulence is such an unusual field of science,
in which we have plenty of experimental data, some simple empirical relations that seem to work
well but are not easy to justify in a rigorous manner. The closure problem for turbulence is still
an open matter as it remains unsolved. Perhaps the best hope is to wait for computational power
to increase and make any simulation of turbulent flow feasible by direct numerical simulations.
Despite these comments, the material covered here should serve as a motivation to study in more
detail this interesting field and to make significant advances. The fact that turbulence is an open
field of research should not be regarded as a drawback but as an opportunity to make innovative,
intelligent, imaginative and most probably simple solutions to this complex subject.

71
72
Bibliography

Batchelor, G., 1953. The theory of homogeneous turbulence. Cambridge University Press.

Boussinesq, J., 1877. Essai sur la théorie des eaux courantes. Imprimerie Nationale.

Corrsin, S., 1951. On the spectrum of isotropic temperature field in isotropic turbulence. Journal
of Applied Physics 22, 469–473.

Davidson, P., 2015. Turbulence: An Introduction for Scientists and Engineers, 2nd Edition. Oxford
University Press.

Kolmogorv, A., 1942. The equations of turbulent motion in an incompressible fluid. Izvestia Acad.
Sci., USSR; Phys. 6, 56–58.

Kolmogorv, A., 1962. A refinement of previous hypotheses concerning the local structure of
turbulence in a viscous incompressible fluid at high Reynolds number. Journal of Fluid Mechanics
13, 82–85.

Landau, L., Lifshitz, E., 1987. Fluid Mechanics, 2nd Edition. Butterworth-Heinemann.

Lorenz, E., 1963. Deterministic non-periodic flow. Journal of Atmospheric Science 20, 130–141.

Menter, F., 1994. Two-equation eddy-viscosity turbulence models for engineering applications.
AIAA Journal 32, 1598–1605.

Obukhov, A., 1949. The structure of the temperature field in a turbulent flow. Izv. Akad. Nauk.
SSSR, Ser. Geogr. and Geophys. 13, 58–69.

Pope, S., 2000. Turbulent flows. Cambridge University Press.

Prandtl, L., 1925. Ueber die ausgebbildete turbulenz. ZAMM 5, 136–139.

Prandtl, L., 1945. Ueber ein neues formelsystem fur die ausgebildete turbulenz. Nachr. Akad. Wiss.
Gottingen Math-Phys K1, 6–19.

Smagorinsky, J., 1963. General circulation experiments with the primitive equations: I, the basic
equations. Mon. Weather Rev. 91, 99–164.

Spalart, P., Allmaras, S., 1992. A one-equation turbulence model for aerodynamic flows. AIAA
Paper 92, 0439.

Whitaker, S., 1999. The Method of Volume Averaging. Kluwer.

73
Wilcox, D., 2006. Turbulence Modeling for CFD, 3rd Edition. D C W Industries.

Wood, B., Cherblanc, F., Quintard, M., Whitaker, S., 2003. Volume averaging for determining
the effective dispersion tensor: Closure using periodic unit cells and comparison with ensemble
averaging. Water Resources Research 39, 1210.

74

View publication stats

You might also like