You are on page 1of 53

Stuart McCready

University of Chemical Technology, Prague

Fluid Flow and Aggregation Modelling

July 2016

Dissertation presented to Faculdade de


Ciências e Tecnologia, Universidade Nova de
Lisboa for obtaining the master degree in
Membrane Engineering.
2)!Literature(Review#
!
Chapter(One#
–!Fluid"Mechanics!–#

Introduction*
This chapter charts the most direct navigable path from first principles to the mathematical models used to
provide the description of the velocity field in the mixing tank. The aim is to provide physical insight, and
to show how we arrive at the final model used for the flow field. Throughout, the goal will be to answer the
question:

"How can we provide a detailed heterogeneous description of the fluid velocity field in order to make
predictions about the behaviour of the particles therein?"

The following discussions assume the reader has an understanding of Cartesian tensors and Einstein
summation notation.

Stuart McCready (mccready.stuart@gmail.com) 7


2.1 – Equations of Fluid Motion
1.1.1 – Lagrangian & Eulerian Descriptions
To define a fluid flow, we use either a Lagrangian or Eulerian frame of reference – defining the rate of
change of fluid properties with time within these reference frames is called a Lagrangian or Eulerian
description respectively.

Lagrangian*Description*–*“riding*along*with*the*fluid*parcel…”*
Firstly, we define the continuum concept of a Lagrangian ‘fluid parcel’1. This is a very small volume
of fluid that moves according to the local velocity field, carrying its properties (such as momentum
and density) with it. Because it moves with the fluid, the relative velocity between the parcel and the
surrounding fluid is zero. Consequently, there is no entry or exit of fluid particles – rather, the fluid
parcel stretches and deforms according to the flow field.

• Describing the flow by recording the properties of


each fluid parcel over time is called a Lagrangian
description.
• It can be thought of as riding along with the fluid
parcel, recording its properties as we go. The man in
the fluid parcel could record a Lagrangian description
of his fluid parcel.
• An example is a weather balloon of neutral buoyancy
recording conditions as it moves through the air. Figure 1 – Lagrangian fluid parcel

The Lagrangian description is simple to deal with: the


fundamental principles of conservation of mass and Newton’s laws of motion apply directly to the
fluid parcel. As such, most laws of nature, including those of fluid mechanics are easily described in
Lagrangian quantities – that is, written in non-conservative form. However, it is not computationally
practical in dealing with a continuum - this would require keeping track of all the fluid parcels in the
flow!

Eulerian*Description*–*“staying*put,*watching*fluid*pass*through…”*
Instead of following the fluid parcel as it moves through the flow field, we can sit at a fixed point in
space and record how the fluid properties at that location change over time. To this end, we introduce
a new concept – that of the Eulerian 'control volume'. This is a very small volume at a fixed location
in space2, through which fluid flows.

• Describing the flow by recording the properties at a


fixed point over time is called an Eulerian description.
• It can be thought of as sitting down at some point and
recording the properties of the fluid as it passes
through. The man could now record an Eulerian
description of his control volume.
• An example of an Eulerian description is any fixed
laboratory probe recording conditions over time.
Figure 2 – Eulerian control volume
When dealing with a continuum, the Eulerian description
is more attractive – it is easier to take Eulerian measurements and monitor their change over time.

1
Often called the ‘material volume’.
2
Its position in space is fixed only for an inertial reference frame. For a moving reference frame, its position relative to the reference frame
will be fixed.

Stuart McCready (mccready.stuart@gmail.com) 8


Transforming*from*Lagrangian*to*Eulerian*–*the*“material*derivative…”**
Thus we have a dilemma – the laws of fluid mechanics are more easily written in a Lagrangian
reference frame, but monitoring the change in fluid properties with time is easier in an Eulerian
reference frame. Therefore, we must find a way of transforming from Lagrangian to Eulerian co-
ordinates.

Conveniently, at the moment a fluid parcel passes through a laboratory point, the Lagrangian
description (of that fluid parcel) and the Eulerian description (at that point) are instantaneously the
same. This will prove useful, as we shall now see.

Let us imagine we are riding along with the fluid parcel, and we are interested in some scalar quantity
– say, temperature – of the parcel. The change in temperature experienced by the fluid parcel due to
tiny little changes in time (!!) and space (!!, !!, !") is:

!" !" !" !"


!! + !! + !! + !!
!" !" !" !"

We’re interested in recording the changes in temperature of the fluid parcel as it moves through the
fluid (giving a Lagrangian description), therefore, !! = !! !!!, !! = !! !!! and !! = !! !!!.
Consequently:
!" !" !" !"
+ !! + !! + !
!" !" !" !" !

!" !"
+ !! ⋅
!" !!!

If we reduced these tiny little changes in space down to infinitesimally small changes, we would end
up describing how the temperature changes with time at a fixed point in space – in other words, we’d
end up with an Eulerian description! This concept can be extended to any other quantity besides
temperature, including vectors and tensors. In general:

! ! !
!!!! = + !! ⋅
!" !" !!!
Eq. 1

! !
= +!⋅!
!" !"

Material Derivative

So, the rate of change of a quantity with time of a fluid parcel is given by:
• Partial derivative of Lagrangian field
• Material derivative of Eulerian field

The equations that follow are mostly in non-conservative form (for the blue cube, the moving fluid
parcel) in terms of the material derivative. Some are written in conservative form (for the green cube,
the fixed control volume), in terms of surface and volume integrals. The bridge between the two forms
is the Reynolds Transport Theorem. In Chapter 4, we will see that in order to solve the system of
equations developed in this chapter, we must use the conservative form.

Stuart McCready (mccready.stuart@gmail.com) 9


2.1.1 – Continuity Equation
We have already stated that fluid particles may not leave or enter the fluid parcel – therefore, the
conservation of mass applied to the fluid parcel is:

!" !!!
+! =0
!" !!!
Eq. 2

!"
+ !!! ⋅ ! = 0
!"

Continuity Equation

Physically, this states that the change in density of the fluid parcel with time depends on the spatial
divergence of the velocity field it traverses3. In this work, we consider only incompressible fluid in
which case:

• There is no change in density of a fluid parcel:

!"
=0
!"

• This reduces the continuity equation to the solenoidal velocity condition:

! ⋅ ! = 0.

3
For those unfamiliar with Gauss’ Theorem, we can make intuitive sense of the equation: think that if fluid particles are diverging away
from a point, then the density must be decreasing (converging would increase the density). See Appendix A on velocity field.

Stuart McCready (mccready.stuart@gmail.com) 10


2.1.2 – Momentum Equation
Applying Newton’s second law of motion to a fluid parcel (of volume !" and surface area !") as it
moves through a fluid, we can define the change in momentum of the fluid parcel as:

! Eq. 3
!!!! ! = !!"#$!!"#$%& + !"#$%&'!!"#$!"
!" !

Momentum Equation

Figure 3 - Forces on a fluid parcel

Body*Forces*
!!
!"#$!!"#$% = !"#$%&#&%'(#)!!"#$% = − !!!!
!!!

Since the gravitational field (Ψ) is a scalar potential field, its gradient is a conservative force field
!!
(conservative body force). Note that close to the Earth, = !!~!9.81!!/! ! .
!!!

Surface*Forces**

!"#$%&'!!"#$%& = !"#$%&!!"!#$!%!!"!!"#$!! = !! !(!!" !!! + !!!! ) = !! ! !!! + !!!" !!!

where !! is the outward unit normal vector to the surface ‘j’. Using Gauss’ divergence theorem, we
can re-write this as:
! !
! !!! + !!!" ! !!" + !!!"
!! !(!!" + !!!" )!!! = !!" + !!!" !!! !!" = !!" = !!
!!! !!!
!! !!

!!! !!" !!"


!!" = !!" = !!" !!! !!"
!!" !!" !!!

! 0 0
!!!!!!!!!!!!!" = 0 ! 0
0 0 !

Figure 4
[Left] The stress due to strain and the stress due to pressure in matrix form.
[Middle] The surface stresses acting on a cubic fluid parcel.
[Right] Illustration that 'i' denotes the direction the force acts, while 'j' denotes the face on which it acts.

Stuart McCready (mccready.stuart@gmail.com) 11


There are two surface forces of interest: stress due to strain (!!" ) and stress due to pressure (!).

• The strain can be decomposed into an isotropic (!!" ) and anisotropic component (!!" )4:

1 !!!
!!" = ! = ! ⋅ ! = 0!
3 !!! !"
!
1 !!! !!!
!!" = + − !!" !
2 !!! !!!
!
Thus, due to incompressibility (! ⋅ ! = 0), we are left only with the anisotropic strain – and
its consequent anisotropic stress, which acts tangentially. This is called the shear stress (!!" ).5
We require a constitutive law6 to relate the strain rate experienced by the fluid to the shear
stress applied. For a Newtonian fluid, we can use Newton’s law of viscosity:

!!! !!!
!!" = !!! + !
!!! !!! Eq. 4

!!!" = !!!!(2!!!" )

Shear Stress

• In an incompressible fluid, pressure waves travel infinitely fast [11] – consequently, the
pressure at every point on the surface of fluid parcel is the same, and acts normal to the
surface. The pressure then, is a normal, isotropic stress.

Normal, Isotropic component Tangential, Anisotropic component

!!!!" !!!"
!"#$$%"#!!"#$% = (!!) ! !ℎ!"#!!"#$% = !!
!!! !!!

NavierDStokes*Equation*
Therefore, the momentum equation (Eq. 3) for an incompressible Newtonian fluid now becomes:

! !! !!!!" !!! !!!


!!!! !! = ! − !!!! − (!!) + (!")!!!! +
!" !!! !!! !!! !!!! !!! !!!! !
!"#$%&'())!*+(

! !! !! ! !!!
!!!! !! = ! − !!!! − (!!) + (!")!!!!!
!" !!! !!! !!! !!! !

Note that the pressure force becomes a (conservative) body force, which allows is to define the
modified pressure as:
! = !" + !

4
See Appendix A (Velocity Field).
5
For second order tensors such as this, isotropic ≣!spherical, anisotropic ≣!deviatoric
6
A constitutive law is simply a relation between two physical quantities that is specific for a given material. Typically this is the kinematic
response of the material to an external kinetic stimulus. [10]

Stuart McCready (mccready.stuart@gmail.com) 12


Therefore:

! !" ! !!!
! !! = ! − + !!!!!
!" !!! !!! !!!
Eq. 5

!
! ! = ! −!! + !!!!!! ! !
!"
Navier-Stokes Equation

The following boundary conditions are applicable:


• ! = 0, on any solid surface. This is a combination of the impermeability and no slip conditions
of stationary solid surfaces. At the fluid-solid interface, the force of attraction between the fluid
particles and solid particles (adhesive forces) is greater than that between adjacent fluid particles
(cohesive forces). This force imbalance brings down the fluid velocity to zero at a solid surface.

2.1.3 – Passive Scalars


We will return to the Navier-Stokes equations and our discussion of the velocity field shortly.
However, before doing so, let us consider a passive scalar quantity in a fluid parcel. We define a
‘passive’ scalar as one that has no effect on the fluid properties – for example, the number of particles
in a dilute sheared dispersion. Since the particles are dilute, we assume they have no effect on the
flow, so we are left only to consider the flow's effect on the particles (as shown in).

The number of particles in a fluid parcel can change due to particles being transported in or out, by the
addition from some ‘source’ or removal by some 'sink':

!" !"
= + !!"#$%!&'$% = !"##$%"&' + !"#$%&"'( + !"#$%& − !"#$
!" !"

Where it is assumed there is no electric field for migration. The diffusion, can be described using
Fick’s Law of Diffusion:

!" Eq. 6
+ !! ⋅ !! = !! ! ! + !"#$%& − !"#$
!"

Passive Scalar Equation


where ! is the (constant and uniform) diffusivity. It will be seen later that, for the particles in the
mixing tank, the source and sink terms are due to aggregation and breakage of the particles – which
are influenced by the flow field.

Figure 5 – The "number of particles" as a passive scalar in a fluid parcel. It is worth noting that the 'local change' is that experienced by a
control volume. The differential forms of the non-conservative and conservative equations are the same.

Stuart McCready (mccready.stuart@gmail.com) 13


2.1.4 – Viscous Dissipation
Let us now consider the dissipation of energy in a fluid flow – the conversion of mechanical energy to
internal energy (heat) by the action of friction.

Figure 6 – Viscous stress acting tangentially on the surface of a cubic fluid parcel of volume δV and surface area δS.

In a fluid flow, this conversion is done by the viscous stress (!!" ). Thus, let us consider the rate of
working of this viscous stress on the fluid parcel. We showed earlier that the viscous stress is:

!!" = !!!!(2!!!" )

The rate of working of this viscous stress is therefore:

Ẇ = (!! !!!" )!!! !!!

where !! is the unit normal vector to the surface ‘j’. Using Gauss’ Theorem, we can re-write this as:
! !
! !
(!! !!!" )!!! !!! = (!! !!!" )!!! !!" = (!! !!!" ) !" = !! (! ! )
!!! !!! ! !"
!! !!

Therefore, the rate of working by viscous forces per unit volume is:
! !!!" !!!
Ẇ= (!! !!" ) = !!! + !!"
!!! !!! !!!

The first term simply represents the rate of working of the viscous stress on the fluid parcel, acting to
change the kinetic energy of the fluid parcel:

!!!" Eq. 7
!!
!!! !
Total Working of Viscous Stress

The second term represents the rate that kinetic energy is converted to heat as a result of viscous
dissipation. This rate of viscous dissipation can be re-written as:

!!! 1 !!! 1 !!! !!!


!!" = !!" + !!" = !!" + !!" = !!! !!!"
!!! 2 !!! 2 !!! !!!

Therefore, we can finally define the rate of viscous dissipation (per unit mass) to be:

!!" !!!" Eq. 8


!= = !2!!!!!" !!!"
!
Viscous Dissipation Rate

Stuart McCready (mccready.stuart@gmail.com) 14


2.1.5 – Looking for Closure
Let us return to our discussion of fluid velocity (and momentum), and take stock of what we have:

• A single time-dependent equation for conservation of mass – Continuity Equation:


!"
+ !! ! ⋅ ! = 0
!"
• Three time-dependent equations for conservation of momentum (one for each component of
velocity7) – Navier-Stokes Equation:
! !" !!! !!! !!!
! !! = ! − + !!!! + !!! + !!!
!" !!! !!! !!!! !!! !!!! !!! !!!!
! !" !!! !!! !!!
! !! = ! − + !!!! + !!! + !!!
!" !!! !!! !!!! !!! !!!! !!! !!!!
! !" !!! !!! !!!
! !! = ! − + !!!! + !!! + !!!
!" !!! !!! !!!! !!! !!!! !!! !!!!

Thus we have 4 time-dependent equations, along with the boundary condition (U = 0 at stationary
solid surfaces) – but we have 5 unknowns (!, !, !! , !! , !! ). We almost have a closed system.

Relating*Pressure*to*Velocity*
Given that the velocity field is solenoidal (! ⋅ ! = 0), we can relate it to the pressure field by taking
the divergence of the Navier-Stokes Equation (Eq. 5). Doing so obtains:
!
! ⋅ ! ⋅ !!! = ! −! !
!

The Biot-Savart Law[12] may be used to invert this equation, giving us:
! ! ⋅ (! ⋅ !!) ′
! ! = ! !!!′
4! |! − !! |

Thus, by writing the pressure in terms of the velocity, we have reduced the unknowns to
(!, !! , !! , !! ).

“Out*of*the*woods…”*
We now have a closed system of coupled partial differential equations and associated boundary
conditions! In principle, this system could be solved using numerical methods, and we could give the
description of the velocity field that we require to predict the aggregation and breakage behaviour of
the particles! It appears as though we're out of the woods.

To understand why things may not be so straightforward, let’s consider an important characteristic of
the Navier-Stokes Equation:
!!! !!! !" !!!
! + !! ⋅ =!− + !!!!
!" !!! !!! !!! !!!!

The convective term is clearly non-linear (quadratic). This means that while our system may be
closed, it is non-linear dynamic – the hallmark of mathematical chaos.

As we shall see, this takes us very much out of the woods and into the jungle.

7
The Navier-Stokes Equation (Eq. 5) is often referred to as the ‘Navier-Stokes Equations’, since it can be separated into three velocity
components (in Cartesian co-ordinates, this would be the velocity in the x, y and z directions – !! , !! , !! ).

Stuart McCready (mccready.stuart@gmail.com) 15


2.2 – Turbulence
Recall from elementary fluid mechanics that the Reynolds number is the ratio of inertial to viscous
forces, and gives some indication of the presence of 'turbulence' in a flow. The Reynolds number takes
a particular form for a mixing tank (with impeller of diameter D revolving N times per second):

! ! ! !"#$%!&'!!"#$%& Eq. 9
!" = =
! !"#$%&#!!"#$%&
Reynolds Number Equation
The operating conditions in our mixing tank are such that 2000 < Re < 8000 – indicating that
turbulence is present.

What consequences will this 'turbulence' have for the quest of providing the description of the velocity
field in the mixing tank? To answer this question, we might first want to consider where this
'prediction' of turbulence comes from, and the meaning of the phrase 'inertial to viscous forces'.

2.2.1 – Origins
Consider the fate of the smoke from the burning
cigarette (Figure 7). As the smoke emerges, it
accelerates directly upwards in a controlled manner8. It
then begins to 'wobble', before suddenly busting into
seemingly chaotic and random behaviour. We call this
chaotic behaviour of fluids Turbulence.

Visually, it appears as though the upper plume is


experiencing significant disturbances. Where did these
disturbances come from, and why do they not affect
the lower plume?

Disturbances*
Even in the most rigorously controlled experiments,
there exists small disturbances and perturbations that
lie out with experimental control.

In the lower part of the flume, the smoke has not yet
gathered much velocity – consequently, the Reynolds
number is low:
!!! !"#$%!&'!!"#$%&
!" = ! ! = !!
! !"#$%&#!!"#$%&
Figure 7 – Transition to turbulence in a cigarette plume

In other words, the viscous forces dominate the inertial forces. These comparatively large viscous
forces act to dampen the small, unavoidable disturbances – through viscous dissipation, as seen in the
previous chapter. Consequently, they cannot grow and manifest into appreciable flow instabilities. We
say that the flow is 'laminar'.

However, as the smoke accelerates upwards, the inertial forces begin to dominate the viscous forces.
The small, unavoidable disturbances can no longer be sufficiently damped, and they reveal themselves
in the 'wobbling' of the plume. These un-damped disturbances interact with one another, producing
greater disturbances, which interact further still – this growth and interaction leads to the significant
disturbances which we see in the upper plume, and which we call turbulence.

8
The smoke is warmer and less dense than the surrounding air, and rises due to buoyancy effects.

Stuart McCready (mccready.stuart@gmail.com) 16


Thus, we continue with the following notion: if the inertial forces outweigh the viscous forces, small
disturbances in the flow can grow and interact, before causing flow instabilities and finally expressing
themselves in the chaotic behaviour we define as turbulence.

The mixing tank is subject to continual disturbances of varying amplitudes and frequencies, and the
Reynolds number indicates that the inertial forces do indeed outweigh the viscous forces – hence why
we predicted the presence of turbulence in the tank.

2.2.2 – Characteristics
To explore the characteristics of turbulence,
we will look at a classic experimental example
– flow across a cylinder. (Figure 8)

In this experiment, the cylinder is towed


through an initially still fluid at a fixed speed.
The velocity is then measured with time at a
fixed location (x0) downstream, in the
turbulent wake.[13] The experiment is Figure 8 - Flow around a cylinder towed through fluid, showing the
repeated a hundred times, rigorously ensuring turbulent wake
that the experimental conditions remain the
same.

Sensitivity*to*Disturbances*
The result is velocity profiles like those in
Figure 9. Clearly, the velocity profile is
markedly different for each experimental run.

This is because, between runs, there are small


differences in experimental conditions that lie
out with experimental control.

Figure 9 – Nominally identical experiments produce significantly


These small perturbations in initial conditions, different velocity traces*
boundary conditions and material properties are
amplified in turbulent flows – no matter how small these perturbations between experiments, the
resulting velocity profiles will look completely different. This is the first notable characteristic of
turbulence:

• Turbulent flows display an acute sensitivity to finite perturbations in initial conditions, boundary
conditions and material properties – meaning a precise deterministic prediction of their evolution
is impossible.9

Mean*&*Turbulent*Velocities*
To observe another important characteristic, let's
introduce the simplest of statistical quantities –
the mean (Figure 10). The blue line represents
the (time-averaged) mean velocity profile of
both experimental runs. It's clear that the mean
is steady with time and reproducible –
predictable, one could say.
Figure 10 – The mean velocity profile is reproducible in nominally
identical runs of the experiment.*

9
This acute sensitivity to initial conditions is popularly known as the 'butterfly effect', the mathematical study of which is known as 'chaos
theory'. It is due to the non-linear term identified at the end of the previous chapter – we will further discuss this non-linearity later.

Stuart McCready (mccready.stuart@gmail.com) 17


In spite of the velocity displaying wide variations on a range of timescales, significant variations from
the mean are not observed, and tend to be short lived. This points to a second, crucial characteristic of
turbulent flows:
• The velocity field in turbulent flows fluctuates randomly in time and is highly disordered in
space. This velocity field is composed of a steady, reproducible mean velocity component and an
unsteady, random turbulent velocity component.

––––––––––––––––––––––––––––––––––––––––––
Suppose we were able to visualise these mean
and turbulent velocity components – what would
they look like?

Sticking with the previous example of the flow


across the cylinder, we could take a time-lapse
photograph of the flow, which would give us a
visual representation of the mean flow. This
would look something like Figure 11.
Figure 11 – Mean flow as would be seen from a time-lapse
As we would expect, the spatial profile of the photograph of flow around a cylinder
mean velocity is smooth and ordered.

While a similar photographic representation of the


turbulent flow field isn't possible, an instantaneous
photograph would give us visual representation of
the total velocity field. This would look something
like Figure 12.

The spatial profile of the total velocity field, with


the influence of the turbulent component, is highly
Figure 12 – Total velocity profile as would be seen from an
instantaneous photograph disordered – as we would have expected.

Absolute*vs.*Statistical*approach*
So what approach should we take for describing the velocity profile in the mixing tank?
• Absolute:
We could attempt to provide a description of the total velocity field by numerically solving the
Navier-Stokes equation. This would, as we know, only give us one particular realisation of the
velocity field, which would look entirely different from one realisation to the next.
• Statistical:
Otherwise we could search for a solution based on the statistical properties (such as the mean
velocity), which we know are reproducible and predictable.

Consider an analogous dilemma – suppose we were interested in the behaviour of a gaseous system.
Would we be more concerned with the motion of every single gas molecule, or the statistical
properties of the system, such as pressure?

Before deciding, let us discuss three further fundamental characteristics of turbulence, which may
make the decision rather easy:

• Vorticity
• Scales
• The Energy Cascade.

Stuart McCready (mccready.stuart@gmail.com) 18


2.2.3 – Vorticity & Turbulence
We stated in the introduction to the Fluid Mechanics section that, given the scope of the thesis, a
proper discussion of vorticity dynamics would not be possible. However, the concept of turbulence
and that of vorticity are so inextricable that we owe at least a short discussion to their relationship. The
aim will be to develop a kind of mental image of turbulence. Given the brevity required, this image
will necessarily be crude – however it will prove useful for future discussions on important topics of
the energy cascade, and of the transfer of energy from the mean flow to the turbulence.

Figure 13 –– (a) Illustrative example of turbulence as a tussling tangle of vortex tubes and blobs; (b) Warping of a vortex tube by the fluctuating
velocity field; (c) Stretching of a vortex tube by the mean velocity strain

Turbulence can visualised as a tussling tangle of vortex tubes, sheets and blobs, continually being
teased and stretched out of shape by the local velocity field (Figure 13). This seething mess of
vorticity exhibits chaotic behaviour on a wide range of time and length scales. To endow some kind of
physical meaning to this rather hand-wavy definition, we have to introduce some dynamics.

Vorticity*Dynamics*
Let us consider a small blob of fluid, which we will define as being instantaneously spherical10 – in
effect, a spherical fluid parcel.

The angular momentum (!) of this blob is related to its moment of inertia (!) and vorticity (!) by:
1
!!!! !=
2
Now the angular momentum of this blob can only change due to surface forces (viscous forces) given
that the only body forces of interest – the 'modified pressure' !"/(!!! ) – is a conservative body force.
We have already stated that the pressure is the same at every point on the surface of the fluid parcel,
and acts normal to the surface. Therefore, given that the blob is spherical, pressure cannot change its
angular momentum – only the viscous forces can:
!" 1 ! !!!
!!!!!!!!!!!! !! = !! ! = (!"#$%&#!!"#$%&)
!" 2 !"
!! !"
! = −!! + 2!×! !"!"#$!!!"#$%&
!" !"

This tells us that the vorticity of a fluid parcel may increase:


• Through the action of viscous forces (viscous torque) Figure 14 –Forces acting on a spherical fluid
• By decreasing its moment of inertia (vortex stretching) parcel

Therefore, we can intensify the vorticity of fluid elements by stretching them (decreasing !), as in
Figure 13 (c) above. It will become clear, much later, that this vortex stretching transfers kinetic
energy from the mean flow and into the turbulence.

10
Note: we are free to define a spherical fluid parcel anywhere in the flow at any instant of time.

Stuart McCready (mccready.stuart@gmail.com) 19


2.2.4 – Scales & The Energy Cascade
"Big whirls have little whirls that feed on their velocity,
and little whirls have lesser whirls and so on to viscosity"
Lewis Fry Richardson
Weather Prediction by Numerical Process, 1922

The*Energy*Cascade*
The above poem actually represents one of the most useful and notable advances in the study of
turbulence – but what exactly did Richardson mean?

The idea is that the largest eddies (at the so-called integral scale) are created by instabilities in the
mean flow, which arise due to insufficient viscous damping as we discussed earlier. These eddies have
a size comparable to the characteristic length of the flow (like the diameter of the mixing tank).

Figure 15 – The length scale of successively smaller eddies effectively lowers the Reynolds number, until eventually, at the smallest
scales, viscous dissipation becomes significant and mechanical energy is converted to heat.

However, these large eddies are also subject to inertial instabilities and the distorting effects of other
eddies, and so are teased out into many smaller eddies. Crucially, this process is essentially inviscid –
this is because !" = !!/! is large, and so the viscous forces acting on the large eddies is negligible.
So the offspring essentially steal kinetic energy from the original eddy.

These smaller eddies suffer a similar fate; evolving into even smaller eddies. Hence, there is a
continual, inviscid cascade of energy from the large scales to the smaller scales.

Things change, however, when we get down to the very smallest scale (the so-called Kolmogorov
Microscale). Here, !" = !!/! is much smaller (being based on the smaller eddy size) – consequently,
the viscosity is no longer negligible, and finally the energy starts to be dissipated. Viscosity, therefore,
acts to mop up whatever energy happens to cascade down from above, and has little effect at the larger
scales.

Stuart McCready (mccready.stuart@gmail.com) 20


Scales*of*Turbulent*Motion*
So, how small do the eddies get before viscous stress starts to dissipate energy? To answer this
question, let's define the scales of the largest and smallest eddies, as in Table 1. Note that the lifespan,
as defined, comes from considerable experimental evidence.[14]

Table 1 – Length scales, time scales, turnover time and Reynolds number of the smallest and largest eddies in turbulent flows

Length Scale Typical Velocity Lifespan Re


(turnover time)
!"
Largest Eddies ! ! ~!!/! !!! ~
!

Smallest Eddies ! ! ~!!/! ~!1

Let's start with the largest eddies – those responsible for passing energy down the cascade. We know
that their kinetic energy (per unit mass) !! !~!! ! , and so the rate at which energy passes down the
cascade is:
!! !!
!!~ =
(!/!) !

–––––––––––––––––––––––––––––––––––––––––––
In a steady state situation (such as our mixing tank after some time), the rate at which energy passes
down the cascade must exactly match the rate at which energy is dissipated. We know from our earlier
discussion that the viscous dissipation rate (per unit mass) is:
! = 2!!!!!" !!!" = 2!!!!!" !!!" + 2!!!!′!" !!′!"
! = ! + !′

We also know that viscous dissipation is predominantly at the smallest eddies, and so the viscous
dissipation is simply:
!′ = 2!!!!′!" !!′!"

Where !′!" is the turbulent rate of strain associated with the smallest eddies, given of course by
!′!" !~!!/!. The viscosity is, as always, given by υ.
–––––––––––––––––––––––––––––––––––––––––––
Therefore, the energy passing down the cascade from the largest eddies balances with the energy
being dissipated by the smallest eddies:

!!!!! = !′
!! ! !
!!~!!!!
! !

Finally, using the Reynolds numbers for the


largest and smallest eddies, we can express the
Kolmogorov Microscales as:
!
!!~! !/!
!!!

!
!!~! !/!
!!!

So then, we find that the larger the Reynolds Figure 16 – The smallest scales of turbulent motion are smaller for
number, the smaller the smallest eddies will be! higher Reynolds number flows
*

Stuart McCready (mccready.stuart@gmail.com) 21


Smallest*Scales*in*the*Mixing*Tank**
So, what does all this have to do with the mixing tank? We know that the integral scale of the largest
eddies will be comparable to the diameter of the mixing tank, so 80mm.

It is well documented[15] that in mixing tanks, the turbulent component of the velocity is in the order
of 5-10% of the total velocity. Therefore, we can expect the Reynolds number of the largest scales,
based on this turbulent velocity, to be 200 < !!! < 700.

This means that the Kolmogorov Microscale of our mixing tank can be expected to be:

0.5!! < ! < 2!!

Therefore, the size of the eddies in the mixing tank range from as small as 0.5mm to as large as 80mm.
But what does this mean for providing the description of the velocity field, and how has it helped us
decide on which approach to take?

So*–*Absolute*or*Statistical?*
Let's make the case for numerically solving the total velocity field – that is, providing as accurate a
specification of the initial, boundary and material conditions as possible, and then integrating the
Navier-Stokes equation forward in time.

This would require us to numerically compute flow structures ranging from 80mm to 0.5mm, over a
range of time scales. If we did not resolve the flow structures of the Kolmogorov Microscales, then we
would essentially be spatially averaging, and in doing so would miss out on some of the physics. If we
did not resolve the largest scales, we would be doing the same – worse still, we would be spatially
averaging the scales containing the most energy! Resolving the full range of scales in this manner is
called direct numerical simulation (DNS). It is prohibitively computationally expensive for all but the
simplest of geometries, and for all but the lowest of Reynolds numbers. Therefore, such a calculation
for the mixing tank simply isn't practical on a desktop computer.[16]

We indicated earlier that a solution based on the total velocity field might not be so useful, given its
irreproducibility. Now we have shown that it isn't computationally practical either. Thus, we had
better begin searching for a statistical solution!

Stuart McCready (mccready.stuart@gmail.com) 22


2.3 – Mean Flow Equations
To this end, we will now attempt to re-define the entire closed system in terms of this most basic of
statistical quantities – the mean, or time-average.

We start with the unknown quantities from the closed system we presented at the end of the previous
chapter: (!, !! , !! , !! ). We can decompose these unknowns (along with the passive scalar, let's say
number of particles '!') into their mean and fluctuating components:

!!!! = !! + !! ′
Eq. 10
! = ! + !′ ! = ! + !′
! = ! + !′
Reynolds Decomposition

With the above definitions to hand, and the time-averaging rules given in Appendix B, we may now
set about re-defining the closed system of equations in time-averaged form – commonly referred to as
Reynolds-averaging.

ReynoldsDaveraged*Continuity*Equation*
We begin with Reynolds-averaging the continuity equation:

!"
+ !!! ⋅ (! + !) = 0
!"

For incompressible fluids:


!"
= 0!!!!!!!!!!!!"#!!":!!!!!!!!!!! ⋅ ! + ! = 0
!"

Therefore:
!⋅ ! +!⋅ ! =0

!⋅!=0

Thus, the turbulent velocity field is also solenoidal. So the mean and turbulent velocities satisfy the
same form of mass conservation as the total velocity field does – which we would have expected,
since the continuity equation is linear.

ReynoldsDaveraged*NavierDStokes*(RANS)*Equation*
We now repeat this procedure for the Navier-Stokes equation, using Reynolds decomposition to re-
express the equation, and then Reynolds-averaging:

! !! + !! ! !! + !! ! !+! ! !! + !!
! + !! + !! ⋅ =− + !!!!
!" !!! !!! !!! !!!!

After some mathematical manipulation, we end up with:

!!! !!! !! !! !! ! !!!


! + !! ⋅ + =− + !!! !
!" !!! !!! !!! !!! !!!

Stuart McCready (mccready.stuart@gmail.com) 23


As expected, this gives us three new equations (! = 1,2,3) for the mean quantities !! and ! –
however, the non-linear term has spawned a new quantity, involving turbulent velocities.
Rearranging, and noting that:

!!! !!!
!!" = !!!! + !
!!! !!! Eq. 11

!!!" = !!!! 2!!!"

Mean Shear Stress

And hence:
! ! !!! !!!
!!" = !!!! +
!!! !!! !!! !!!

Therefore, we finally obtain:

!!! !!! !! !
! + !! ⋅ =− + ! − !!!! !!
!" !!! !!! !!! !"
Eq. 12

! !
! (!! ) = −!! + ! !!" + !!"
!!

Reynolds-averaged Navier-Stokes Equation (RANS)

Where we have defined a new form of the material derivative – the mean material derivative – as:

! ! ! Eq. 13
= + !! ⋅
!! !" !!!

Mean Material Derivative

Thus, it seems as though these new unknowns may be interpreted as a surface force – a 'stress', in
other words. This new term is called the Reynolds Stress:

−!!!! !! −!!!! !! −!!!! !!


!
!!" = !!"! = −!!!! !! −!!!! !! −!!!! !!
−!!!! !! −!!!! !! −!!!! !!

It's immediate consequence for our goal of providing the velocity field description in the mixing tank
is obvious – it adds a further six unknowns. Given that we have only four new equations – one from
the conservation of mass (Reynolds-averaged continuity equation) and three from the conservation of
momentum (RANS) – the system is no longer closed.

This raises two immediate questions, which we will attempt to answer in the forthcoming sections:
(i) – What is the physical meaning of this so-called Reynolds "Stress"
(ii) – How can we provide additional equations for the Reynolds Stress to close the system?

Stuart McCready (mccready.stuart@gmail.com) 24


ReynoldsDaveraged*Passive*Scalar*Equation*
Before attempting to answer these questions, we will repeat the Reynolds-averaging procedure for the
passive scalar equation. For the number of particles '!' in the mixing tank:

!!
= !! ! ! + !"#$%& − !"#$
!"

Introducing the Reynolds averaged quantities and re-arranging, just like we did previously, we obtain:

! Eq. 14
! ! = !!! ! !! !! − !! ∇ ∙ !! !! + !"#$%& − !"#$
!!

Reynolds-averaged Passive Scalar Equation

Where the final term on the right (!! !! ) is the scalar flux. This arose in a similar way to they Reynolds
Stress, with one important difference – the turbulent velocity in the above equation is uncorrelated,
meaning it is not multiplied by another turbulent velocity in the way that the Reynolds Stress is. This,
of course, is because the passive scalar equation is linear. Nonetheless, it is an additional unknown.

Figure 17 – The Reynolds-averaged "number of particles" in a fluid parcel.

This equation will prove to be important in how we relate the velocity field to the number of particles
in the mixing tank – which is, after all, a scalar quantity. Since the volume fraction of the particles is
very small compared to the fluid itself, we assume that the solid particles in dispersion have no effect
on the flow field (meaning the ‘number of particles’ is a passive scalar). Therefore we are left to
consider only the effect of the flow on the particles, which of course is the primary consideration of
this work.

The unknown scalar flux term must be closed. We will attend to this matter later.

For now though, we return to consider the two proposed questions about the Reynolds Stress.

Stuart McCready (mccready.stuart@gmail.com) 25


2.3.1 – Reynolds Stress
To answer the first question, we will shift our focus from the fluid parcel to the control volume. The
equations will be conservative, and written in integral form. We will therefore be considering the
'budget' of quantities in the control volume.

(i)*What*is*the*physical*meaning*of*this*soDcalled*Reynolds*"Stress"?*
Let's consider the Navier-Stokes equation applied to a control volume, which we can obtain from the
non-conservative form by application of Reynolds Transport Theorem[17]:
! ! ! !
!
!!!! !!" = − !!!! ! !!!! !!" + (!!" !! )!!" − (! !! )!!"
!"
!! !" !" !"

where !! is the unit normal vector the surface 'j',


and !! is the unit normal vector in the 'i'
direction. The terms in this equation can be
understood as:

• The rate of change of momentum in a


control volume (of volume δV).
• Minus the rate at which momentum flows
out through the surface (of area δS).
• The net viscous force acting on the surface
(of area δS).
• The net (modified) pressure force acting
Figure 18 – Navier stokes applied to control volume involves
on the surface (of area δS). monitoring the budget of quantities in the control volume

Conservative*form*of*RANS*
If we now use the Reynolds decomposition and time-average (just like we did previously), then:
! ! ! ! !
!
!!!! !" = − !!!! ! !!!!! !!" + −!!! !! !! !!"! + !!" !! !!" − !!!! !"
!"
!! !" !" !" !"

This is the RANS equation in integral form. The terms in this equation can be understood as meaning:

• The rate of change of (mean) momentum in the cube (of volume δV).
• Minus the rate at which (mean) momentum flows out through the surface (of area δS), carried
by the mean velocity
• Minus the rate at which (mean) momentum flows out through the surface (of area δS), carried
the turbulent velocity
• The net (mean) viscous force (of area δS).
• The net (mean) modified pressure force acting on the surface (of area δS).

So the Reynolds Stress isn't actually a stress at all – it's a fraud! To better understand its true physical
meaning, let's consider the total momentum flux out of the control volume. We'll consider only the !-
direction for simplicity.

The momentum fluxes through the side of the cube which contribute to a rate of change of momentum
in the !-direction are:

!!!! !!" = !! !! + !! !! + !! !!"!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)


!!!!!!!!!!!!!!!!!!!!!+!!! !! + !! !! + !! !!"!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)
!!!!!!!!!!!!!!!!!!!!!+!!! !! + !! !! + !! !!"!!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)

Stuart McCready (mccready.stuart@gmail.com) 26


Thus, the time-averaged flux of momentum in the x-direction is therefore:

!!!! !!" = !! !! !! !!"!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)


!!!!!!!!!!!!!!!!!!!!!+!!! !! !! !!"!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)
!!!!!!!!!!!!!!!!!!!!!+!!! !! !! !!"!!!!!!!!!!!!(!ℎ!"#$ℎ!!"#ℎ!(!!!!)!!"#$%)

The Reynolds stress, then, is actually the flux of


momentum in the ! th direction through the !th plane (a
plane that is perpendicular to the !th direction) caused
by the turbulent velocities.

We will talk more about the physical meaning of the


Reynolds Stress later, and why we refer to it as a
'stress'. For now though, the more pressing matter is
answering the second question.
Figure 19 – Momentum flux on a control volume in the 'x'
direction..

(ii)*How*can*we*provide*the*further*Equations*we*require*to*close*the*system?*
To answer the second question, let's first consider how we provided the necessary equation for the
viscous stress.

The viscous stress can be thought of as mixing of momentum at the molecular scale. We then want to
determine the relationship this viscous stress has to the flow – in other words, a relation between a
kinetic property (viscous stress) and a kinematic property (strain) – in other words a 'constitutive'
relationship! Since the scale of the viscous stress is very small, and the scale of the flow is much
larger, we can easily find an intermediary scale with which to average the required constitutive
property (viscosity) over – in other words, the continuum hypothesis can be considered valid. There
are several reasons why this wouldn't work for the Reynolds stress.

Firstly, and most critically, it's not a real stress, it's a flux of momentum (a kinematic property, not a
kinetic property). Therefore, we're trying to relate one kinematic property of the flow to another
kinematic property of the flow – this is not the job of 'constitutive' relationships!

Secondly, the Reynolds Stress is the mixing of momentum at the scale of the flow. The mean strain
(by definition) is also on the scale of the flow. So even if we were to consider the Reynolds Stress as a
real kinetic property, we cannot find an intermediary scale with which to average some kind of
fabricated, ad hoc constitutive property over.

Both these points, of course, are a manifestation of the fact that the turbulence is a property of the
flow, not of the fluid!

So how can we provide the additional equations we require? One possibility would be to provide a
governing transport equation for the Reynolds Stress. Let's see if we can concoct such an equation.

2.3.2 – Reynolds Stress Transport Equation


To this end, we recall both the Navier-Stokes equation (written with Reynolds decomposition) and the
RANS equations for the fluid parcel:
! ! ! + !! !
! !! + !! = ! − + ! + ! ! !" !!!!!!!!!!!! !"#$%&– !"#$%&!!"#$%&'(
!" !!! !!! !"

! !! ! !
! (!! ) = − + ! + !!" !!!!!!!!!!!!!!!!!!!!!!!!!!(!"#$!!"#$%&'()
!! !!! !!! !"

Stuart McCready (mccready.stuart@gmail.com) 27


Step*1*–*Transport*Equation*for*the*turbulent*velocities,*!! *&*!! *
Obtaining a transport equation for the turbulent velocity !! is simple – just subtract the RANS
equation from the Navier-Stokes:
!
!"#$%&– !"#$%& ! − ! (!"#$)
!!! !!! !!! !!! !! ! ! ! !
!! + !! + !! + !! =− !+ ! !" ! − ! !!!!! !! !!!!!!!!!!! = 1,2,3
!" !!! !!! !!! !!! !!! !!!
Note that the '!' is a free index, and can be replaced by '!' by multiplying by !!" to obtain a second
equation:
!!! !!! !!! !!! !! ! ! ! !
!! + !! + !! + !! =− !+ ! !" ! − ! !!!!! !! !!!!!!!!!! = 1,2,3
!" !!! !!! !!! !!! !!! !!!

Step*2*–*Transport*Equation*for*the*turbulent*velocity* !! !! *&* !! !! *
Multiplying the first equation by !! and averaging, we obtain:
!!! !!! !!! !!! !! ! !! ! !"
!! !! + !! ! !! + !! !! + !! !! = −!!! ! + !! !!!!!!!!!!!!!, ! = 1,2,3
!" !!! !!! !!! !!! !!!

Similarly, multiplying the second equation by !! we get:


!!! !!! !!! !!! !! ! !! ! !"
!! !! + !! ! !! + !! !! + !! !! = −!!! ! + !! !!!!!!!!!!!, ! = 1,2,3
!" !!! !!! !!! !!! !!!

Step*3*–*Summation*of* !! !! *&* !! !! *Transport*Equations*


Now, finally, we can add the previous two equations together to obtain:
! !!! !!! ! !! ! !! ! !! ! !" !! ! !"
! ! (!! !! ) + !! !! + !! !! + !! ! ! = − !!! + ! !! + !! + !!
!! !!! !!! !!! ! ! ! !!! !!! !!! !!!

Step*4*–*ReDarrangement*
The above equation can be understood in a more physically intuitive form by simply re-arranging the
terms:
! !!! !!! ! ! ! ! !
! !! !! = !!
!" + !!
!" + !!! !! !! − ! !! − ! !! + 2!! ! !′!"
!! !!! !!! !!! !!! !!!
!"#
!"#$%&'(#) !"#$%&'"( !"#$%&'"(

!! !!! !!!
+! ! ! !!! !! − 2!"
!!! !!! !!!
!"#$%&'"( !"#"$%&%"'(11
!"##"$%&"'(

The physical meaning of these terms will become important shortly, in a discussion on kinetic energy.
For now we note that:
• Production – the production of Reynolds Stress by the mean velocity gradients.
• Dissipation – the mean viscous dissipation of Reynolds Stress due to fluctuating velocity.
• Transport – all these terms represent the mean spatial redistribution of the Reynolds Stress
by the turbulent velocity, the pressure fluctuations and the viscous stresses respectively.
• PSC – the 'pressure – strain-rate correlation' gives the correlation between
the fluctuating pressure (!′) and the fluctuating strain rate (!′!" ).

11
Note that this term only represents the viscous dissipation of homogeneous flow, which is not the case for sheared flows such as the mixing
tank. However, at the Kolmogorov microscales, the.[18]

Stuart McCready (mccready.stuart@gmail.com) 28


The most important physical take-away at this stage is:
• The mean velocity gradients are responsible for the production of the Reynolds Stresses.

We will see later that this 'production' is involves a transfer of kinetic energy from the mean flow to
the turbulence.

However, it should be immediately obvious what the problem is from our perspective – we have
generated a considerable number of new unknowns. Not all of these new unknowns are independent,
because some can be derived from others[19] – however, we've certainly opened a can of worms!

Let's focus on one of these new unknowns in particular – the triple turbulent velocity correlation,
!! !! !! – and how we could go about closing this term.

2.3.3 – The Closure Problem


It may occur to use same strategy as we did for the double turbulent velocity correlation (the Reynolds
Stress) – and indeed, this would provide the transport equation for !! !! !! . However, this would
contain a quadruple turbulent velocity correlation !! !! !! !! . A further transport equation for !! !! !! !!
would contain fifth-order turbulent velocity correlations !! !! !! !! !! . We don't have to make our way
through the alphabet to know we're chasing our tail!

This is the closure problem of turbulence.

Let's recall how we ended up in this slump. In the "Equations of fluid motion" chapter, we developed a
perfectly closed system capable, in principle, of providing the Eulerian description of the velocity field
we required. However, we detected the presence of turbulence in the mixing tank, and learned that the
velocity would behave chaotically, and over a wide range of scales. This meant that solving the system
numerically would be too computationally expensive.

Conversely, we saw that the statistical properties do not behave chaotically, and were in fact
predictable. To that end we developed a statistical system capable of making such predictions – only
to find that the non-linear term had come back to bite us, giving birth to the Reynolds Stress for which
we can provide no closed equation. This is what we meant when we said the non-linearity would take
us "out of the woods and into the jungle".

The closure problem has haunted all of the brilliant minds to tackle the subject, from its earliest
pioneer, Osborne Reynolds. Though many great explorers have mapped out important sections of this
mysterious jungle – Prandtl, Richardson, Taylor, Kolmogorov, Corrsin – we are still no closer to
understanding whether a grand theory of turbulence even exists.

When we first predicted the presence of turbulence, we asked: what consequences will this 'turbulence'
have for our quest of providing an Eulerian description of the velocity field in the mixing tank?

The answer, at this stage, is 'catastrophic'.

It is impossible to develop a predictive, statistical model of turbulence through manipulation of the


governing equations of fluid motion. To close the system we need to introduce some form of
additional information, and this information is necessarily ad hoc in nature – in other words, we must
propose some kind of turbulence closure model.

We should be warned, however, that this alternate path – that of proposing turbulence closure models
– is riddled with questionable hypotheses, which are often applicable to narrow classes of problems
and heavily reliant on empirical data.

Notwithstanding, we proceed to consider one such category of models – the eddy viscosity models.

Stuart McCready (mccready.stuart@gmail.com) 29


2.4 – Eddy Viscosity Models
Thus far, we have considered the mean flow and the turbulence somewhat independently, with
particular attention on the latter. However, in sheared flows like that in the mixing tank, there is a
continual and complex interaction between the mean flow and the turbulence – the mean flow
generates, maintains and redistributes the turbulence, which in turn shapes the velocity distribution of
the mean flow. This interaction occurs via the Reynolds Stress.
• The Reynolds Stress exists because of the presence of turbulence, and acts to shape the
evolution of the mean flow.
• The Reynolds Stress simultaneously acts as a conduit for kinetic energy to pass into the
turbulence from the mean flow.
Therefore, any turbulence closure model must describe both of these phenomena.

2.4.1 – Eddy Viscosity Hypothesis (effect of turbulence on mean flow)


When the Reynolds Stress first appeared during the
derivation of the RANS equation, and we sought an
equation for its closure, we gave a thorough
reasoning of why we could not use the same
constitutive approach as we had for the viscous
stress. However, having encountered the closure
problem by attempting to provide a transport
equation, we must now re-consider.

Boussinesq*Hypothesis*
To this end, we return to the notion that the Reynolds Stress is actually a kinetic property – that is, a
real stress. Recall that this is what the RANS equation actually suggested in the first place:

! !! ! !
! (!! ) = − + ! + !!"
!! !!! !!! !"
In this case, as far as the mean flow is concerned, the mixing of momentum caused by the Reynolds
Stress is analogous to the molecular mixing of momentum by the viscous stress. It might then occur to
provide a similar constitutive quantity – a viscosity – to quantitatively relate this stress to the
kinematic property of interest – this time the mean strain.

This ad hoc viscosity is called the eddy viscosity (!! ), and its proposed effect is to increase simply the
effective viscosity:
!!"" = ! + !!

Consequently, we can now re-define the Reynolds Stress as:

! !!! !!! 1
!!" = −!!!! !! = !!!!! + − !! !!! !! !!!"
!!! !!! 3

! 1
!!!!!!" ! + !!! !!! !! !!!" ! = !! !!!! ! 2!!!" Eq. 15
3
!"#$"%&'(& !"#$!!"#$

Boussinesq Hypothesis
!
Since the normal stresses of !!" must sum to −!!!! !! , the term on the right is required.

The question now is – how can we find a value for !! ?

Stuart McCready (mccready.stuart@gmail.com) 30


Prandtl*Mixing*Length*Theory*
The kinetic theory of gases studies the microscopic behaviour of gas molecules to explain
macroscopic properties of the gas as a whole. The theory posits that the pressure of a gas is due to the
collisions of the gas molecules with the walls of the container.
––––––––––––––––––––––––––––––––––––––
The theory can be adapted relatively successfully to make similar macroscopic predictions about
viscosity, using the mean free path ! of the molecules and their RMS speed ! :
1
!= !!!
3
When fluid flows in a laminar regime, layers of fluid slide over one another. The layers exchange
momentum due to thermally agitated fluid molecules hopping around between layers, carrying their
momentum with them (Figure 20). Therefore, the layers apply a mutual stress on one another, !!" .
When this microscopic jumping around of molecules is averaged out, we can express the macroscopic
result as:
1 !!!
!!" = !!! !!! !
3 !!!
!!!
!!" = !!!!!
!!!
–––––––––––––––––––––––––––––
Ludwig Prandtl extended this idea from laminar to turbulent regimes –
only this time, instead of imagining thermally agitated molecules
hopping between layers, we imagine portions of fluid being tossed and
Figure 20 – The hopping of molecules
pushed around by the fluctuating velocity field. Since these portions of between layers in laminar flow.
fluid carry their momentum with them, there will similarly be a mutual Alternatively, the throwing of fluid portions
stress between two fluid layers – the Reynolds Stress: between layers in turbulent flow.

!
!!" = −!!!!! !!

Therefore, the above diagram can be re-interpreted as portions of fluid being tossed between layers by
the turbulent velocity. This analogy between the mixing of momentum due to thermal agitation, and
mixing of momentum due to the turbulent velocity led Prandtl to propose an analogous macroscopic
kinetic theory for turbulent flows:

Eq. 16
!!!!!!!!!!!! = ! !! !!

Prandtl Hypothesis

Where !! is the so-called 'mixing length', while !! is an appropriate measure of the magnitude of the
turbulent velocity |!|. The physical intuition is rather clear:
• The portion of fluid being jostled around will conserve its properties for a characteristic length,
!! , before mixing with the surrounding fluid.
• The greater !! , the more energetic the turbulence, hence the greater the tossing around of fluid
portions, hence the greater mixing of momentum, hence the larger the 'viscosity' !! .

!
Thus, we started out with the question of "what is !!" ?"
We then used the eddy viscosity hypothesis to reformulate this question to "what is !! ?"
Now we have used Prandtl's mixing length theory to finally re-formulate it as "what are !! & !! ?"

We will seek to answer this question later, in Section 2.6.

Stuart McCready (mccready.stuart@gmail.com) 31


2.4.2 – Transfer of Kinetic Energy (from mean flow to turbulence)
Firstly though, let's consider the second phenomena –
how the mean flow transfers kinetic energy to the
turbulence. To do so, will consider the kinetic energy of
both the mean flow and the turbulence.

The hope is, of course, that a comparison of the mean and


turbulent kinetic energy will reveal the mechanism by
which this energy is transferred.

Kinetic*Energy*of*the*Mean*Flow*
To determine the kinetic energy of the mean flow, we must firstly recall the transport equation for the
mean flow itself – the RANS equation:

! !! !!!" !!!!! !!
! (!! ) = − + −
!! !!! !!! !!!

We can therefore find an expression for the kinetic energy (per unit volume) of the mean flow simply
by multiplying this equation by !! and halving as follows:

!
! 1 !! !!!" !!!!
! !!! !!! ! = −!! ! + !! ! + !! !
!! 2 !!! !!! !!!

Re-arranging, we obtain:

! 1 ! ! ! ! Eq. 17
! !! = −!!!!!" !!! ! + ! !!" !!! ! + ! !!" !!! − !!" !!!" − 2!"!!!" !!!"
!! 2 !!!
Kinetic Energy (per unit volume) of the Mean Flow

We can identify the first term on the right (the divergence term) as being the rate of working of the
mean pressure, the mean viscous stress and the Reynolds stress on the boundary. If we consider the
mixing tank as a whole, which is of course a closed domain, then this term will integrate to zero. The
last term on the right should be easily recognisable as the rate of mean kinetic energy dissipation rate
(!!). For the mean flow, this term will be very small – this is because at the scale of the mean flow,
the Reynolds number is high12. We will find out the meaning of the unknown term shortly.

!"#$%&"'!!ℎ!"#$ !"!"#$%&'!!" !
= − ! !!" !!!" ! − !! !!!!!"##"$%&"'(!!"!!"#$!!! !!!(!!) !!
!"!!"#$!!! !"!!"#$!!!
!"!!"#$"%"&'$(!!"##"$%&"'(!!"!!"#$%&'(!!"#$%

12
Note, however, that the viscous dissipation isn't negligible at the boundaries (i.e. near the walls), where the no slip condition (! = 0)
means that the Reynolds number (!" = !"/!) is small. Consequently, the viscous forces are important near the walls, and the mean viscous
dissipation also becomes important.

Stuart McCready (mccready.stuart@gmail.com) 32


Kinetic*Energy*of*the*Turbulence*
To consider the kinetic energy of the turbulence, we recall the Reynolds Stress transport equation:

! !!! !!! ! ! ! ! !
! !! !! = !!
!" + !!
!" + !!!! !! !! − ! !! − ! !! + 2!! ! !′!"
!! !!! !!! !!! !!! !!!
!"#
!"#$%&'(#) !"#$%&'"( !"#$%&'"(

!! !!! !!!
+! ! ! !!! !! − 2!"
!!! !!! !!!
!"#$%&'"( !"#"$%&%"'(
!"##"$%&"'(

To obtain an equation for the kinetic energy, we simply have to contract ! = !:

! !!! !!! ! ! ! ! !
! !! !! = !!
!" + !!
!" + !!!! !! !! − ! !! − ! !! + 2!! ! !′!!
!! !!! !!! !!! !!! !!!
!"#
!"#$%&'(#) !"#$%&'"( !"#$%&'"(

!! !!! !!!
+! ! ! !!! !! − 2!"
!!! !!! !!!
!"#$%&'"( !"#"$%&%"'(13
!"##"$%&"'(

Halving the above equation, re-arranging and simplifying, we obtain:

! 1 !
! 1 ! Eq. 18
! !! = −! ! !!" !!! ! + ! !! !!′!" + !!!! !! !! + !!" !!!" − 2!"!!′!" !!′!" 14

!! 2 !!! 2
Kinetic Energy (per unit volume) of the Turbulence

When we presented the Reynolds Stress transport equation, we gave some physical meaning to the
terms involved – and so these meanings can be understood in the new context of kinetic energy. The
first term on the right is the spatial redistribution of kinetic energy by pressure fluctuations, by mean
viscous stress and by the turbulent velocity itself. The last term on the right is the
mean turbulent kinetic energy dissipation rate (!!′). Unlike the mean rate of strain, this term is not
negligible, because viscous forces are important at these small scales (see Section 2.2.4). Once more
we have an unknown term in our equation.
!"#$%&"'!!ℎ!"#$ !"#$%&'"(!!"
= + ! !!!" !!!" ! − !! !"##"$%&"'(!!"!!"#$"%&'!!!! !!! !!′ ! !!
!"!!"#$"%&'!!!! !"!!"#$"%&'!!!!
!"##"$%&"'(!!"!!"#$"%"&"'!!"#$%!!!"!!"#$%&!"#

Transfer*of*Energy*from*the*Mean*Flow*to*the*Turbulence*
If we compare the equation for the kinetic energy of the mean flow with that of the turbulence:
!"#$%&"'!!ℎ!"#$ !!!"#$%&'"(!!"!! !
= − ! !!" !!!" ! − !! !!!!!"##"$%&"'(!!"!!"#$!!! !!!(!!) !!
!"!!"#$!!! !"!!"!"!!!
!"!!"#$"%"&'$(!!"##"$%&"'(!!"!!"#$%!!"!!"#$!!"#$

!"#$%&"'!!ℎ!"#$ !"#$%&'"(!!" !
= + ! !!" !!!" ! − !! !"##"$%&"'(!!"!!"#$"%&'!!!! !!! !!′ ! !!
!"!!"#$"%&'!!!! !"!!"#$"%&'!!!!
!"##"$%&"'(!!"!!"#$"%&'!!!"#$%!!!"!!"#$%&!"#

The meaning of the unknown term in both equations should now be obvious – it clearly represents the
transfer of kinetic energy out of the mean flow and into the turbulence!
13
As noted previously, when written in this form, the dissipation term is only valid for homogeneous flows.
14
However, written in this form, it represents the true viscous dissipation.

Stuart McCready (mccready.stuart@gmail.com) 33


• The Reynolds Stress (produced by mean velocity gradients as we seen earlier) is responsible
for transferring kinetic energy from the mean velocity gradients to the turbulence.
The true physical understanding of the mechanisms of this energy transfer is very poorly understood.
Two different perspectives on this mechanism are offered here, the latter more insightful than the first.

(1) If we take the physical interpretation of the Reynolds Stress as a real stress, then we can arrive at
an understanding by analogy with the viscous stress. The viscous stress acts to resist the deformation
of the fluid parcel caused by the instantaneous rate of strain. The energy expended in doing so ends up
as internal energy. By analogy, the Reynolds Stress resists the deformation of the fluid caused by the
mean rate of strain. The energy expended in doing so ends up in the turbulence (in the turbulent
velocity). Thus, the Reynolds Stress removes kinetic energy from the mean flow and transfers it to the
turbulence – but how can we physically interpret this transfer? Yes, it comes from the work the
Reynolds Stress does against the mean rate of strain – and so rain does come from the sky!
(2) To paint a simplified mental picture, let us return to the idea
of a tangled tussle of vorticity. Consider the fate of such a tangled
blob of vorticity in a simple sheared flow (Figure 21). The mean rate
of strain in this simple example acts to stretch the blob of vorticity –
thereby decreasing its moment of inertia. Through the conservation
of angular momentum (!), this leads to an increase in the angular
velocity – and so increases the kinetic energy of the turbulence: Figure 21 – The stretching of a portion of
vorticity by the mean strain
1
!= !!!(!!×!!)
2
This, of course, is the vortex stretching term we identified earlier, and we now realise its importance.

The*Energy*Cascade*Revisited*
With an insight into how kinetic energy is transferred, it would be useful to connect this to what we
already know about what happens to energy in turbulent flow – thus we recall the energy cascade.

We can now define the rate of generation of kinetic energy (per unit mass) in the energy cascade
(from the mean flow into eddies at the integral scale) as:
!
!!" !!!"
!= = ! −!!! !! !!!"
!
We have previously defined the rate of mean turbulent kinetic energy dissipation rate (per unit mass)
at the Kolmogorov Microscales to be:

Eq. 19
!′ = 2!!!′!" !!′!"

Mean Turbulent Kinetic Energy Dissipation Rate


We have also previously defined the rate at which energy passes down the cascade (per unit mass) as:
!! !!
!!~ =
(!/!) !
If we consider the turbulence to be steady and homogeneous, then we can state that the flux down the
cascade will be equal through all the scales (in other words, energy will not accumulate at an
intermediate scale), and so ! = ! = !. However, the mixing tank is an example of a shear flow,
which is very seldom homogeneous in nature. However, even for unsteady, homogeneous flows, the
following approximation holds well[19]:
!!
!!~!!!~! !~!!′
!
We will return to this approximation shortly, when we finally develop a turbulence closure model
based on the eddy-viscosity hypothesis. However, before doing so, we must critically analyse the
validity of the inherent assumptions of the hypothesis on which this closure model will be based.

Stuart McCready (mccready.stuart@gmail.com) 34


2.4.3 – Limitations of Eddy-Viscosity Hypothesis
Now that we have considered both phenomena required for any turbulence model, we will now
provide an initial critique of the eddy-viscosity hypothesis and Prandtl's theory of mixing length.
––––––––––––––––––––––––––––––––
The essence of the eddy-viscosity hypothesis is that, as far as the mean flow is concerned, the mixing
of momentum caused by the Reynolds Stress is analogous to the molecular mixing of momentum by
the viscous stress.

However, molecules have a fundamentally different nature of physical interaction. They are discrete,
and their collisions are intermittent. Turbulent eddies, on the other hand, are distributed and in a state
of constant interaction.
In addition, for molecular mixing of momentum, the molecular mean-free path is significantly smaller
than the characteristic length of the mean flow – in turbulent flows of course, the integral scale is
comparable to the characteristic scale of the mean flow.
––––––––––––––––––––––––––––––––
The eddy hypothesis extends the above notion by proposing that the relation between the Reynolds
!
Stress (!!" ) and the mean rate of strain (!!" ) can be defined in a similar constitutive manner as the
!
viscous stress. In effect, the two symmetric, anisotropic components !!" !&!!!" each have five
independent components, and the eddy viscosity hypothesis assumes these can be related using the
scalar coefficient, the eddy-viscosity (!! ).

!
By using this scalar quantity (!! ) to define a linear relationship between (!!" ) and (!!" ), we are
essentially assuming that the flow is isotropic, such that:

! ! !
!! = !! = !!

In mixing tank operations, particularly in the impeller discharge[20], the turbulence is known to have
an anisotropic nature – meaning that the assumption of such correlations is generally not valid.
––––––––––––––––––––––––––––––––
A further consequence of the eddy viscosity hypothesis is that is assumes the Reynolds Stress is
determined by the local mean strain rate (!!" ), and not by the history of the strain.

However, we know that the Reynolds Stress comes from the turbulence itself (it is a correlation of
!
turbulent velocites). Therefore, we can expect !!" to be increased in accordance with the turbulence.
We also learned that the turbulence can be increased by the stretching of vortex blobs and tubes.

With this in mind, consider again the simple sheared stretching of vorticity we presented earlier
(Figure 21). It is clear from this graphic that the vorticity will depend on the history of the mean rate of
strain – and so too, by extension, will the Reynolds Stress.

If the turbulent fluctuations occurred on a timescale which was much smaller than the timescale of the
meal flow, then we could make the case that the local conditions would have reached a relative
statistical equilibrium. However, the Reynolds Stress stems predominantly from the large eddies,
whose timescale is comparable to that of the mean vorticity.

Therefore, we cannot simply consider that at any instant, the flow will reach a statistical equilibrium
rapidly enough for us to consider only the instantaneous rate of strain.
––––––––––––––––––––––––––––––––
The take-forward message from the above limitations is clear: any turbulence closure model based on
the eddy-viscosity hypothesis must be viewed in light of these limitations, and regarded with caution.
With this in mind, we will consider one such model which has become the standard working
turbulence closure model used in engineering: the k-ε model.

Before doing so, we will briefly shift focus to consider a passive scalar quantity again.

Stuart McCready (mccready.stuart@gmail.com) 35


2.5 – Gradient-Diffusion Hypothesis
Recall that an unknown quantity – the scalar flux !! !! – appeared from the Reynolds averaging
procedure in a similar way to the Reynolds Stress. The key difference was that this term did not
involve turbulent velocity correlations.

The scalar flux is the transport (or flux) of the scalar quantity (!) resulting from the turbulent velocity
fluctuations – the gradient-diffusion hypothesis states that this flux is down the mean scalar gradient
(−∇!!). Therefore:
!! !! = −!!! !∇!!
Where !! is defined as the turbulent diffusivity. Thus, we can re-define our mean passive scalar
equation:
!
! ! = !!! ! !! !! − !! ∇ ∙ !! !! + !"#$%& − !"#$
!!
!
! ! = !!! ! !! !! − !! ∇ ∙ !! !∇!! + !"#$%& − !"#$
!!
!
! ! = ! ! + !! !! ! !! ! + !"#$%& − !"#!
!!
It worth noting that the form of the equation has returned to that of the original passive scalar
equation, with the exception that it now contains Reynolds averaged quantities. This is mathematically
analogous to Fick's law of diffusion.

Figure 22 – The Reynolds-averaged "number of particles" in a fluid parcel after the Gradient-Diffusion Hypothesis

The similarity to the Eddy-viscosity hypothesis is clear. We started with three unknown terms in this
case (one for each component of turbulent velocity) and ended up with a single unknown scalar
quantity. Thus we now have to close both the turbulent viscosity and the turbulent diffusivity.

However, it has been shown [16][14] that for most flows, the turbulent Schmidt number for a passive
scalar is approximately unity:

!!
!"! = !~!1
!!

Thus, we can use the gradient diffusion hypothesis to express:

! !! Eq. 20
!!=! ! + !! ! !! ! + !"#$%& − !"#$
!! !"!
Reynolds-averaged Passive Scalar Equation (Gradient-Diffusion Hypothesis)

Thus, we can proceed to search for closure of the turbulent viscosity alone. Firstly, however let's give
a brief critique of the gradient diffusion hypothesis.

Stuart McCready (mccready.stuart@gmail.com) 36


Limitations*of*the*Gradient*Diffusion*Hypothesis*

The gradient diffusion hypothesis assumes that the three components of the scalar flux (!! !! ) can be
related to the mean scalar gradient vector (! ! !!) through a scalar coefficient (!! ). This assumes these
vectors are aligned. However, it has been shown that for homogeneous turbulent shear flow, these
quantities are considerably misaligned.

2.6 – The k-ε Model


Recall that, using the Boussinesq Hypothesis, we had reformulated the closure problem from one of
!
determining the Reynolds Stress (!!" ), to one of determining the eddy-viscosity (!! ):

! 1
!!" = !!!!! !!!" − !! !!! !! !!!"
3

Then, using the Prandtl Hypothesis, we had re-formulated it once more to be a question of
determining the mixing length (!! ) and a suitable measure of the magnitude of the
turbulent velocity (!! ):
!!!!!!!!!!!! = ! !! !!
––––––––––––––––––––––––––––––––
Given the assumption that the Reynolds Stress is uniquely determined by the local conditions, it
would seem reasonable that the eddy-viscosity !! would be determined by the characteristics of the
large eddies at that the local position in question. We are also working on the assumption that there is
an absence of severe anisotropy. Therefore, it is natural to define:
!! !~! !
!! = !
!
1 !
!! !~! !! ! = ! !/!
2 ! !
Where ! is the kinetic energy (per unit mass) of the turbulence, and ! is the intregral scale associated
with the largest eddies in the turbulent flow. Hence we can re-express the turbulent viscosity as:
!! !~!! !/! !!
––––––––––––––––––––––––––––––––
We also noted at the end of our discussion on the energy transfer to turbulence that:
!! ! ! ! !/!
!!~!!!~! !~!!′!!!!!!!!!!!!!!!!!!!!!!"#!!"!!!!!!!!!!!!!!!!!!!!!!!!!!~! =
! !′ !′
Where !′ is the mean turbulent kinetic energy dissipation rate (per unit mass) at the Kolmogorov
Microscales. Therefore, re-expressing the eddy viscosity once more we obtain:
! !/!
!! !~
!′

!! Eq. 21
!! ! = !!
!′
k-ε Hypothesis
Where !! = 0.09 is a constant[21]. This constant has been determined by numerous iterations of data
interpolation and fitting for a wide range of turbulent flows.[14]
––––––––––––––––––––––––––––––––
Thus, our search for closure takes on yet another form, this time for the determination of the turbulent
kinetic energy (!) and the viscous dissipation rate (!′). As we will see, the !– ! model provides semi-
empirical transport equations for these terms.
Stuart McCready (mccready.stuart@gmail.com) 37
2.6.1 – The 'k' equation
Recall the equation for the kinetic energy of the turbulence:

! 1 ! 1 !
!!!! !! = −! ! !!" !!! ! + ! !! !!′!" + !!!! !! !! + !!" !!!" − 2!"!!′!" !!′!"
!! 2 !!! 2

We can express this equation as:


!
! ! = ! ⋅ ![!! ] + !!! − !!!′
!!

!
! = !!! ⋅ !!! + !!!!!!! − !!!!′!!!
!! !"#$%&'"( !"#"$%&'(# !"##"$%&"'(

––––––––––––––––––––––––––––––––
If we consider the newly defined transport term, we find that it contains three unknowns:
1
![!! ] = −! ! !!" !!! ! + !!!!!! !!′!" + !!!! !! !!
2
We know already that viscous dissipation is concentrated in the smallest scales of turbulence – hence
we can expect this term to be very small above. The key modelling step is that the k-ε model assumes
that the pressure fluctuations caused by the turbulent velocity fluctuations act to spatially redistribute
the turbulent kinetic energy in a diffusive manner, from regions of high intensity to low intensity. The
same applies to the triple turbulent velocity correlations. Specifically:

!"#$%&'"(! = ! = !! !!

where !! is an unknown turbulent diffusivity. This diffusivity is typically taken to be equal to the
eddy-viscosity, since we have seen that the turbulent Schmidt number (!"! ) for most flows is
approximately unity[14].
!!
!"! = !~!1!!!!!!!!!!!!"#!!ℎ!"!#$"!:!!!!!!!!!!! = !! !!
!!
––––––––––––––––––––––––––––––––
Thus, the turbulent kinetic energy can be expressed as a simple transport equation, with an associated
sink and source term:
!
! !!"
! = ! ⋅ !! !! + !! − 2!!!′!" !!′!"
!! ! !"

!
! = ! ⋅ !! !! + ! − !′
!!

This equation is then typically re-written in the following more general form, which allows for
adjustment of the diffusivity by specifying the turbulent Schmidt number (!"! ) [22]:

! !! Eq. 22
! =!⋅ !+ !!! + ! − !′
!! !"!

'k' Transport Equation

Therefore, the 'k' equation amounts to an exact equation, where the transport term is modelled as a
simple gradient diffusion. The other terms in the equation (the mean material derivative of 'k', the
generation and the viscous dissipation) are all in closed form. Thus the 'k' equation has been closed.

Stuart McCready (mccready.stuart@gmail.com) 38


2.6.2 – The 'ε' Equation
While the equation for the turbulent kinetic energy had an identifiable physical origin, the equation
used for 'ε' is totally fabricated, and will be discussed in more detail in the next page:

!
! !! !!!′ !′ Eq. 23
!′ = ! ⋅ (! + )!!!′ ! + ! !! − ! !!
!! !"! ! !
'ε' Transport Equation

Where !! , !! , !! are additional empirical constants, which normally take the following values:
!! = 1.3!;!!!!!!!!! = 1.44!;!!!!!!! = 1.92

Thus, with the inclusion of these empirical constants, the 'ε' equation has been closed.

2.6.3 – Summary & Interpretation


The k-ε turbulence closure model, therefore, can be summarised as:

Table 2 – Summary of k-ε model

! 1
Boussinesq Hypothesis !!" = !!!!! !!!" − !! !!! !! !!!"
3
!!
k-ε Hypothesis !! ! = !!
!′

! !!
'k' transport equation ! =!⋅ !+ !!! + ! − !′
!! !"!

!
! !! !!!′ !′
'ε' transport equation !′ = ! ⋅ (! + )!!!′ ! + ! !! − ! !!
!! !"! ! !

Empirical constants15 !! = 0.09 !"! = 1 !"! = 1.3 !! = 1.44 !! = 1.92

where:
!
!!"
!= !!!" !!!!!!!!!!!"#!!!!!!!!!!!′ = 2!!!′!" !!′!"
!

––––––––––––––––––––––––––––––––––––––––
It seems like the model relies on a totally fabricated equation along with some arbitrary coefficients,
designed to reproduce a few laboratory results. Indeed, the k-ε is essentially a highly sophisticated
interpolation between data sets! However, it has proven to be largely effective for a wide variety of
engineering applications, with the exception of flows containing large pressure gradients[14] (not
present in the mixing tank).

We stated already that while the 'k' equation had an identifiable physical origin – being based on the
derived equation for the kinetic energy of the turbulence – the 'ε' equation was simply plucked from
thin air. Is there any inherent rationale behind the equation?

15
Note that both the constants for the turbulent Schmidt numbers are approximately unity, as stated previously.

Stuart McCready (mccready.stuart@gmail.com) 39


Interpretation*of*'ε'*equation*
Consider that the largest eddies of the flow have a lifetime of:
! !
= ~
! !!′
Thus, we can state that, for the largest eddies, their characteristic vorticity is:

!!′
!!~
!
Suppose now that there is a homogeneous shear flow in one-dimension:
!!! !
! = !! ! ! ê! !!!!!!!!!!!!! = !!"
!"
Where ê! is the unit normal vector in the !-direction. We have seen already that these large scale
eddies can be distorted by the mean flow, and that this process increases the vorticity of these eddies.
Therefore, it might be concluded that after some time, the vorticity of the large scale eddies would
eventually settle down to a value in the order of the mean rate of strain:
!!"
!!~!!!" !!!!!!!!!!!!!!"!!!!!!!!!!!!!!! =
!
Where ! is a constant. Therefore, we might also conclude that if at any instant, the vorticity of the
large eddies is different from this value, then it would tend towards this value on a timescale of
!!
!!" such that we could define the overall behaviour in a heuristic equation:
!
!! !!"
= !! ! − ! ! !!!!!… … … … … … … … … … … !!!". (!)
!" !

Where ! is a constant. Now, from the !– ! model:

!
!" ! !!′ ! !!′ !′
= ! − ! !′!!!!!!!!!!!!!!!!"#!!!!!!!!!!!!!! = !! ! − ! !!
!" !" ! !

Where:
!
! !! ! ! !!"
! = !! ! !!" =
!′

In the case of the homogeneous, one-dimensional flow, these !– ! equations can be re-arranged:

!! !
!
= (!! − 1)!! !!" − (!! − 1)!! ! !… … … … … … … … !!!". (!)
!"

If we now compare !". (!) and !". (!), we find that:

!!
! ! = !! − 1 !!!!!!!!!!!!!!!!!!!!"#!!!!!!!!!!!!!!!!!!!!!!! = !! !! − 1
!!

Therefore, in conclusion, if !! ,!!! > 1 then the particular form of the '!' equation means that in one-
dimensional homogeneous shear flow, the vorticity has reasonable behaviour. Therefore, one
interpretation of the '!' equation is that it is a transport equation for the vorticity of the large eddies,
which causes their vorticity to tend towards that of the mean flow.

Stuart McCready (mccready.stuart@gmail.com) 40


2.6.4 – Conclusion
To conclude, we finally have a closed system capable of providing a description of the velocity field
in the mixing tank we require, and for passive scalar transport. It is important, however, to bear in
mind the limitations of the assumptions that made this closure possible.

2.7 – Reynolds Transport Theorem


It was stated at the beginning of the chapter that to solve the final set of equations, we require the
equations in conservative form. To obtain the conservative form of the equations, we apply Reynolds
transport theorem. For some quantity '!', this can be expressed as:
! ! !
! !"
! !!" = !!" + (! ⋅ !)!! !!"
!" !"
!" !! !!

Where ! is the unit normal vector to the surface of the fluid parcel. The terms of the equation can be
understood as:

• The time rate of change of ! within the fluid parcel of interest.


• The time rate of change of ! within the associated control volume (the control volume
occupying the same region of space as the fluid parcel at any instant of time).
• The net flux of ! through the surface of the associated control volume (known as the control
surface).

It is useful to note that there are a large number of variants of Reynolds transport theorem present in
the literature. This is because the formula is extremely general, and can be applied in a variety of
different contexts. As such, different literature will inevitably have equations that often look different
than the above equation in both appearance and complexity.

Stuart McCready (mccready.stuart@gmail.com) 41


4)!Materials)&)Methods#
#
–!Computational+Fluid+Dynamics+(CFD)!–#

Stuart McCready (mccready.stuart@gmail.com) 48


4.1 – Introduction
The closed system of differential equations presented thus far describes the continuous movement of
fluid in the mixing tank in space and time. To solve this system numerically, and provide the
description of the velocity field in the mixing tank, the whole system needs to be discretised – to be
changed from a continuous problem domain to a discrete problem domain. This will mean:

• Spatial and temporal changes are described by finite steps, rather than infinitesimal steps
• The closed system of equations can be written in algebraic form

There are several approaches to the spatial discretisation, including finite element (FEM) method,
finite difference method (FDM) and finite volume method (FVM). In these methods, each flow
variable is defined at every discrete point in the entire spatial domain. The general method of spatial
discretisation is different for each of the numerical methods – our focus will be on the finite volume
method, since the CFD software used to compute the solution (ANSYS Fluent) is based on this
method.

4.2 – Finite Volume Method (FVM)


The spatial discretisation approach used by the finite volume method is to split the entire spatial
domain into small control volumes. The conservative form of the system of equations is then
integrated over these control volumes, and Gauss' theorem is applied to convert volume integrals to
surface integrals. The terms of the equation are then evaluated as fluxes at the surfaces of the control
volume. The values used for the calculations are 'stored' in the centre of the control volume – in the
computational node. Therefore, in order to obtain the values at the faces of the cells – required for the
computation of the fluxes – interpolation is required.

The principle of the finite volume method is local conservation – and this is the key reason why it is
successful in computational fluid dynamics applications. As we have seen in the previous chapter,
fluid mechanics problems are governed by local conservation principles – the continuity equation is
based on the conservation of mass, while the Navier-Stokes is based on the conservation of
momentum. In this way, using Gauss' theorem, these laws of conservation can be recast for each cell
as linear algebraic equations – and these equations can be solved iteratively.

4.2.1 – Spatial Discretisation – Mesh Generation


In order to split the entire spatial domain into small control volumes, a
mesh is used. In three dimensions, this mesh defines the planar faces
between neighbouring control volumes. In this way, each control volume
is given a unique address in !, !, ! space. The shapes of the control
volumes (or cells) can be tetrahedral, hexahedral, prismatic or pyramidal,
as shown in Figure 24.

The mesh must be fine enough to capture the important details of the
flow and provide an appropriate resolution – however, a finer mesh also
means more cells, which means greater computational cost.

The mesh used for the mixing tank is shown in Appendix C. Figure 24 – Polyhedra possible from
3D meshing operation

Stuart McCready (mccready.stuart@gmail.com) 49


4.2.2 – Discretisation of Equations
Wish a suitably defined mesh, the closed system of equations can now be discretised, which requires
that the equations be expressed in conservative form.

To illustrate how these conservation equations can be discretised, we will consider the transport of
some passive scalar property in two dimensions. The conservative form of the mean passive scalar
equation, for some passive scalar '!' is given by:

!!" ! ! !!
+ !!!! !! ! = ! !! ! + !!"#$%&
!" !!! !!! !!!

Where Γ is the diffusion coefficient associated with the


passive scalar, and S is a source term. We have already
identified that the cell values are stored in a computational
node. We will consider the simplified two-dimensional
example. The cell of interest has a computational node
labelled 'P'. The computational nodes of adjacent cells are
labelled according to their cardinal direction relative to the
cell of interest (N, S, E, W). The faces of the cell of interest
are also labelled according to their relative cardinal direction,
with lower case letters (n, s, e, w). These faces are assigned a
Figure 25 – A simplified 2D mesh, with the faces
surface area of 'A'. labelled in lowercase and the computational nodes
in uppercase
To simplify the example further, we will consider only the
East and West cells. In addition, we will consider a steady state flow field, flowing west to east for
now. The time-discretisation will be considered later in this section.

Discretised*Equations*
The process discretisation begins with integrating the conservative form of the mean scalar transport
equation over the control volume (cell) of interest, and using Gauss' theorem to convert the volume
integrals to surface integrals. Doing so, we obtain the integral form of the conservative transport
equation:
!! !!
! !! ! !! ! !!! ! − ! !! ! !! ! !!! !! = !!! !! − ! !! ! !! + !! !"#$%&
!" ! !" !
Where the final term is the volume integral of the source term. The problem is, of course, that the
equation requires values at the faces, while the values themselves are stored in the computational
nodes in the cell centres. Therefore, interpolation – or a 'differencing scheme' – is required.

4.2.3 – Differencing Schemes


When considering a differencing scheme, there are four important requirements:

• Conservativeness – The flux out of a cell should be equal to the flux into the corresponding
neighbouring cell. We noted already that the finite volume method is based on the principle of
conservation, and so this criterion is automatically satisfied.
• Boundedness – The face value should be neither larger nor smaller than the cell values used for its
computation. In other words, the face value must lie within the cell values used to calculate it. This
ensures that no spurious minima/maxima are present within the domain.

Stuart McCready (mccready.stuart@gmail.com) 50


• Transportiveness – The interpolation should reflect how information is transported, which
depends on the ratio of diffusion to convection. If diffusion dominates, then information is
transported similarly in all directions. However, if convection dominates, then information is
naturally transported in the direction of the flow. An indication of the relative effects of diffusion
and convection is given by the Peclet number:
!!!!!
!" = !
!

• Accuracy – The differencing schemes assume shape for the variable of interest (Φ). This is
approximated by a Taylor series expansion:

!′ !! !′′ !! ! ! !!
! !! = ! !! + (!! − !! ) + (!! − !! )! !!…!! (!! − !! )!
1! 2! !!

If!the!scheme!takes!into!account!only!the!constant, and!ignores!the!derivatives, it!is!first order


accurate. If the first derivative is taken into account, then the differencing scheme is second order
accurate.

Upwind*Differencing*
For !" >> 1, convection dominates, as would be expected in the mixing tank – meaning an upwind
differencing scheme is used. This means that the values required for the faces can be assumed to be
dependent only on the values at the computational nodes further upstream (upwind).

In first order upwind differencing, the value at the face is equal to


the value in the computational node immediately upstream
(upwind). In our case, given the direction of flow, this would
mean:
!! = !!

!! = !!

This scheme is simple to implement, is convergence is very stable.


However, it is also very diffusive, leading to a 'smear' of the flow.
This is highlighted in the 'numerical diffusion' section that
immediately follows. Figure 26 – First order upwind
differencing scheme

Greater accuracy can be achieved by using second order


upwind differencing. In this scheme, quantities at cell faces
are computed using values from two upwind computational
nodes. For example, to determine the value at the eastern
face of the cell in question (!! ), the assumption is made
that the gradient between the cell in question and the
eastern face is the same as the gradient between the
western cell and the cell in question. Therefore:

!! − !! !! − !!
= !!!!!!!ℎ!"#!:
!! − !! !! − !!

Figure 27 – Second order upwind differencing !! − !! !! − !!


scheme !! = !!
!! − !!

The major disadvantage of the second-order upwind differencing scheme is that it is unbounded. This
can lead to convergence and stability issues.

Stuart McCready (mccready.stuart@gmail.com) 51


The first order upwind scheme will generally converge quicker and have greater stability, but less
accuracy. Therefore, in the simulations for the mixing tank, the first order upwind differencing is used
to achieve initial convergence, while the final solution uses second order upwind differencing for
greater accuracy. Higher order schemes are also necessary to minimise numerical diffusion.

Numerical*Diffusion*
Numerical diffusion is an inherent error in CFD calculations, and presents itself as an increase in the
diffusion coefficient. Therefore, in the mean passive scalar transport equation, the diffusion coefficient
would appear to be higher than it actually is. Similarly, in the RANS equation, the viscosity would
appear to be higher than it actually is. The relative effect of numerical diffusion is greater when
diffusion is small.

Numerical diffusion is more problematic in course meshes. Therefore, it is important that a suitably
fine mesh is used, particularly for the regions where accuracy is required, such as those where a high
velocity gradient would be expected. Accordingly, as the mesh in Appendix C shows, the regions near
the impeller and the wall have a finer mesh structure.

Figure 28 – Numerical diffusion

In addition, numerical diffusion is more pronounced when the flow is misaligned with the mesh.
Clearly, in the turbulent mixing tank, this misalignment will be significant, and the chaotic flow field
will rarely be aligned with the mesh. To minimise numerical diffusion, a higher order discretisation
scheme can be used. Hence, in the calculation of the final solution, the second-order upwind scheme is
used. This is highlighted in Figure 28.

Stuart McCready (mccready.stuart@gmail.com) 52


4.2.4 – Solution Methods
The above discretisation process produces a set of coupled algebraic equation that must be solved
simultaneously for every cell in the domain. However, as we learned previously, this set of coupled
equations contain non-linear terms. Therefore, the solution procedure must be an iterative process.

There are two approaches to this process.

• Segregated – This approach involves solving one variable at a time, for the entire spatial
domain. For example, !! is solved across the entire domain, followed by !! , then !! and so
on and so forth. The 'iteration' is complete only when all of the variables have been solved in
this way.
• Coupled – This approach solves all of the equations simultaneously, but for one cell at a time.
The 'iteration' is complete only when this process has been repeated for all of the cells in the
entire spatial domain.

Figure 29 – (a) Segregated solving approach; (b) Coupled solving approach

Since we are assuming incompressible flow, and the mixing tank operation is turbulent, the segregated
approach is preferred [29].

PressureDVelocity*Coupling*
Since there is no explicit equation for the pressure, other approaches have been devised to extract it –
these are so-called pressure-velocity coupling methods. The particular method used in these
simulations was the semi-implicit method for pressure-linked equations (SIMPLE).

In the SIMPLE approach, a guess is made for the value of pressure so that the RANS equation can be
solved. This 'guessed' value is of course the last updated value from the previous iteration. Using this
pressure, new velocities are calculated, which will not satisfy the Reynolds averaged continuity
equation. Therefore, the velocities are corrected accordingly, and used to determine a new value for
the pressure. Thereafter, the other variables are calculated, and if convergence is not met, a new
iteration begins.

Residuals*
The solution of the algebraic equations involves an iterative process – therefore, at any step in this
process, the calculation is based in the inexact values calculated from the previous iteration. These
inexact values originate from the initial 'guessed' value, and in a convergent solution, are refined
through repeated iterative calculations.

The total residual for a given equation is the sum of the residuals of all the cells in the entire domain.

Stuart McCready (mccready.stuart@gmail.com) 53


Convergence*Criteria*
As stated above, the residuals in converging solutions decrease with successive iterations. The
convergence criteria are pre-determined conditions that state the maximum value of residuals that
constitute a converged solution. In other words, when the residuals fall below the convergence criteria,
then the solution is considered converged.

In the final calculations (using the sliding mesh approach, for both flow field simulations and also for
the simulations involving the solution to the PBE), the following convergence criteria were used:

Table 3 – The absolute criteria specified for convergence

Equations. .Absolute.Criteria.
Velocities#(x,y,z)# 0.00001#
k# 0.00001#
ε# 0.00001#

It is important to note that this type of convergence criteria only indicates when overall balances are
achieved. It requires engineering discretion to verify that convergence has indeed been reached.

Since we are interested in providing a description of the steady state velocity field in the mixing tank,
it is wise to consider other convergence criteria in addition to the absolute residuals above. To this
end, the impeller moment, average velocity magnitude and average tangential magnitude were
monitored. These monitors, along with the residuals, are discussed in more detail in the results section.

UnderDrelaxation*
We have seen already that the iterative solution procedure involves using information calculated from
the previous iteration. Therefore, some small change in the value (∆!) is added to the value from the
previous iteration (!! ) to yield the new value (!!!! ):

!!!! = !! + ∆!

However, in iterative processes involving the solution of a set of couples equations, it is common to
introduce a 'relaxation factor' (!), such that:

!!!! = !! + !∆!

This is called 'under-relaxation', and helps to reduce unstable oscillations in the flow solution, but
naturally increase the time required for convergence. The values of the relaxation factors typically
range from 0 to 1. In the case of the mixing tank, the standard Fluent relaxation factors were used:

Table 4 – Standard under-relaxation Factors in Fluent

.. UnderMrelaxation.Factors.
Pressure# 0.3#
Density# 1#
Body#Forces# 1#
Momentum# 0.7#
Turbulent#Kinetic#Energy# 0.8#
Turbulent#Dissipation#Rate# 0.8#
Turbulent#Viscosity# 1#

Stuart McCready (mccready.stuart@gmail.com) 54


Temporal*Solutions*
We now consider the solution of temporal problems, such as the mixing tank operation we are
interested in.

In order to solve a temporal solution, the time derivative must be discretised. For the example of the
mean passive scalar transport equation:

!!" ! ! !!
!!! !!! + !!!! !! ! = ! !! ! + !!"#$%&
!" !!! !!! !!!
!"#$
!"#$%&'$%"

The first order discretisation of this time derivative for incompressible flow is:

! !!! − ! !
= !(!)
∆!

Where ! ! is the value of the variable at time '!', while ! !!! is the value at time '! + ∆!'. The spatially
discretised part of the equation is denoted by !(!), and has been described previously. This spatially
discretised part can be evaluated at time '!' or at time '! + ∆!' – this is referred to as an explicit and
implicit approach respectively. For the mixing tank operation, the preference is to use an implicit
approach, and as such:

! !!! − ! !
= ! !!!!
∆!

The underlying assumption is that the newly calculated value for the variable (!) is constant
throughout the entire time step – that is, from time '!' to time '! + ∆!'.

Stuart McCready (mccready.stuart@gmail.com) 55


4.3 – Approaches to Stirred Tank Modelling
The mixing tank operation involves the rotation of the impellers in a stationary tank, with stationary
baffles. Therefore, a mesh must be created which fills the space in between these physical elements.
However, the mesh must also be built to allow the solution to account for the motion of the impeller
relative to the remaining stationary elements.

To do so, an implicit or explicit approach may be used. In the implicit approach requires experimental
data of the time-averaged velocity in the fluid region immediately adjacent to the impellers. The
explicit method involves modelling the modelling the impeller geometry directly, and specifying either
a time-averaged or steady state approach to meshing and solving.

We consider only explicit approaches, using the physical geometry of the impellers and shaft.

Since mixing tank operation involves the rotation of the impellers in a stationary tank in an otherwise
stationary geometry, a stationary mesh specified at time '!' of the operation will no longer conform to
the new geometric configuration at time '! + ∆!' – at this point the impeller will be at a different
position relative to the remaining geometry of the tank.

This requires us to split our mesh into multiple fluid zones – one for the region surrounding the
moving impeller and shaft (hereafter called the "impeller zone"), and another for the remaining
stationary tank and impellers (hereafter called the "stationary zone"). This naturally requires
information to be shared between zones at the zone interface. There are two approaches to this
problem – the multiple reference frame approach and the sliding mesh approach.19 Both of these
approaches will be discussed, before the final solution strategy is described.

4.3.1 – Multiple Reference Frame Approach


This is the simplest of these is the multiple reference frame approach. This is a steady state
approximation in which the reference frame for the impeller zones moves at a rotational rate equal to
that of the impeller, while the stationary zone reference frame has zero rotational speed. Effectively, if
the rotation of the reference frame for the impeller zone exactly matches the rotation of the impeller,
then the impeller is at rest within that reference frame. This allows the problem to be solved as a
steady state problem. A reference frame transformation is applied at the interface between the
stationary zone and the impeller zone. This allows the flow variables from the impeller zone to be
used to determine fluxes at the interface with the stationary zone, and vice versa.

It is important to realise that this approach cannot account for the relative motion of the impeller zone
with respect to the stationary zone – this is because the mesh remains fixed. Therefore, the solution
calculated using this approach is valid only for one particular impeller-baffle orientation. In other
words, we are freezing the motion of the impeller at some point, and calculating the instantaneous
flow field at that orientation.

Consequently, if the interaction between the impeller and the baffles is strong, then this solution
method would only be valid for the particular impeller orientation used. Considering the geometry of
the mixing tank in question, the impeller-baffle interaction can be expected to be significant. The flow
at the interface of the impeller zone and stationary zone is therefore unlikely to be uniform.

As such, this approach is not used for the final solution required to determine the aggregation and
break-up behaviour of the particles. We will see later, however, that this approach can still prove very
useful in the overall solution strategy.

19
There is s third approach, known as 'mixing plane model', which was not used and it therefore not discussed.

Stuart McCready (mccready.stuart@gmail.com) 56


4.3.2 – Sliding Mesh Approach
The sliding mesh approach is a time-dependent solution approach, whereby the mesh surrounding the
impeller zone physically moves at a rotational rate equal to that of the impeller. This means that the
impeller and shaft are stationary relative to the rotating mesh. The mesh in the stationary zone does
not rotate, and hence the tank and baffles are also stationary relative to the surrounding mesh.
However, this naturally means that the two mesh zones move relative to one another – in other words
they slide past one another, hence the nomenclature. This requires that the mesh interfaces remain in
contact, which will be ensured if the interface is a surface of revolution about the axis of symmetry.
This is visible in the picture of the mesh presented in Appendix C.

Unlike the multiple reference frame approach, in which the solution is valid only for the particular
orientation of the impeller relative to the baffles, the sliding mesh model allows a time-accurate
solution in which the equations are solved for successive impeller-baffle orientations as the baffle
rotates.

The time-iteration for a sliding mesh model requires further comment. The 'sliding' of the mesh
interfaces past one another is not continuous. Rather, the set of conservation equations is solved
iteratively for a particular impeller-baffle orientation until convergence is achieved. The impeller and
surrounding mesh then rotate – the rotational distance dictated by the designated 'time step' size (e.g
0.01s) and the rotational rate of the impeller (e.g. 100rpm) – and the equations are once again solved
iteratively until convergence. During each of the quasi-steady calculations, information is exchanged
at the interface between the impeller zone and the stationary zone.

Unlike the multiple reference frame approach, the


mesh of the impeller zone and that of the
stationary zone are inherently non-conformal, as
Figure 30 clearly shows. Consequently,
interpolation is required to ensure appropriate
exchange of information from cells on one side of
the interface with its neighbours on the other side.

The sliding mesh is the most rigorous and


Figure 30 – Sliding mesh approach, showing non-conformality of
informative approach for mixing tank operations, mesh interface
giving a truly time-dependent solution reflecting
the interaction of the impellers with the baffles. However, it is more computationally expensive than
the multiple-reference frame approach.

4.3.3 – Solution Strategy


The multiple reference frame approach, as we seen, is not valid for the mixing tank on account of the
impeller-baffle interaction. However, it does offer several advantages applicable to the overall solution
strategy, which must also take into account computational expense, and hence computational time.

Recall that the goal is to determine the aggregation and break-up behaviour of particles in the mixing
tank after at steady state conditions. Therefore, we are not particularly interested in the start-up
behaviour of the mixing tank (from when the impeller is initially at rest in the fluid). As such,
performing a time-accurate solution (using the sliding mesh approach) would require many
computational 'time-steps' before the mixing tank arrived at steady state.

Therefore, the multiple reference frame approach was used to determine the instantaneous solution at a
particular impeller-baffle configuration. The converged solution was then simply used as initial
conditions for the sliding mesh approach, thereby reducing the overall computational load (and time).

Operation at 50rpm, 100rpm and 150 rpm impeller speeds were modelled.

Stuart McCready (mccready.stuart@gmail.com) 57


References
[1] X. Meng, H. Wu, and M. Morbidelli, “Shear-driven aggregation of binary colloids for randomly
distributing nanoparticles in a matrix,” Soft Matter, vol. 12, no. 16, pp. 3696–3702, Apr. 2016.
[2] X. Sheng, D. Xie, X. Zhang, L. Zhong, H. Wu, and M. Morbidelli, “Uniform distribution of
graphene oxide sheets into a poly-vinylidene fluoride nanoparticle matrix through shear-driven
aggregation,” Soft Matter, vol. 12, no. 27, pp. 5876–5882, Jul. 2016.
[3] C. M. Sorensen, “Light Scattering by Fractal Aggregates: A Review,” Aerosol Sci. Technol., vol.
35, no. 2, pp. 648–687, Jan. 2001.
[4] J. C. Flesch, P. T. Spicer, and S. E. Pratsinis, “Laminar and turbulent shear-induced flocculation
of fractal aggregates,” AIChE J., vol. 45, no. 5, pp. 1114–1124, May 1999.
[5] T. Serra and X. Casamitjana, “Modelling the Aggregation and Break-up of Fractal Aggregates in a
Shear Flow,” Appl. Sci. Res., vol. 59, no. 2–3, pp. 255–268, Jun. 1997.
[6] F. Xiao, H. Xu, X. Li, and D. Wang, “Modeling particle-size distribution dynamics in a shear-
induced breakage process with an improved breakage kernel: Importance of the internal bonds,”
Colloids Surf. Physicochem. Eng. Asp., vol. 468, pp. 87–94, Mar. 2015.
[7] V. Runkana, P. Somasundaran, and P. C. Kapur, “Mathematical modeling of polymer-induced
flocculation by charge neutralization,” J. Colloid Interface Sci., vol. 270, no. 2, pp. 347–358, Feb.
2004.
[8] E. D. Hollander, J. J. Derksen, O. S. L. Bruinsma, H. E. A. van den Akker, and G. M. van
Rosmalen, “A numerical study on the coupling of hydrodynamics and orthokinetic
agglomeration,” Chem. Eng. Sci., vol. 56, no. 7, pp. 2531–2541, Apr. 2001.
[9] J. J. Ducoste and M. M. Clark, “The Influence of Tank Size and Impeller Geometry on Turbulent
Flocculation: I. Experimental,” Environ. Eng. Sci., vol. 15, no. 3, pp. 215–224, Jan. 1998.
[10] “Continuum Mechanics: Constitutive Laws.” [Online]. Available:
http://www.brown.edu/Departments/Engineering/Courses/En221/Notes/Constitutive_Laws/Consti
tutive_Laws.htm. [Accessed: 28-Jun-2016].
[11] P. A. Davidson, An Introduction to Magnetohydrodynamics. Cambridge University Press, 2001.
[12] W. H. Westphal, A Short Textbook of Physics: Not Involving the Use of Higher Mathematics.
Springer Science & Business Media, 2012.
[13] P. A. Davidson, Turbulence: An Introduction for Scientists and Engineers. OUP Oxford, 2004.
[14] D. C. Wilcox, Turbulence Modeling for CFD. DCW Industries, Incorporated, 1994.
[15] R. C. Dorf, The Engineering Handbook, Second Edition. CRC Press, 2004.
[16] J. O. Hinze, Turbulence. McGraw-Hill, 1975.
[17] R. B. Bird, W. E. Stewart, and E. N. Lightfoot, Transport Phenomena. John Wiley & Sons, 2007.
[18] J. F. Wendt, Ed., Computational Fluid Dynamics. Berlin, Heidelberg: Springer Berlin Heidelberg,
2009.
[19] S. B. Pope, Turbulent Flows. Cambridge University Press, 2000.
[20] R. Escudié and A. Liné, “Analysis of turbulence anisotropy in a mixing tank,” Chem. Eng. Sci.,
vol. 61, no. 9, pp. 2771–2779, May 2006.
[21] B. Andersson, R. Andersson, L. Håkansson, M. Mortensen, R. Sudiyo, and B. van Wachem,
Computational Fluid Dynamics for Engineers. Cambridge University Press, 2011.
[22] Y. Tominaga and T. Stathopoulos, “Turbulent Schmidt numbers for CFD analysis with various
types of flowfield,” Atmos. Environ., vol. 41, no. 37, pp. 8091–8099, Dec. 2007.
[23] D. L. Marchisio, M. Soos, J. Sefcik, M. Morbidelli, A. A. Barresi, and G. Baldi, “Effect of Fluid
Dynamics on Particle Size Distribution in Particulate Processes,” Chem. Eng. Technol., vol. 29,
no. 2, pp. 191–199, Feb. 2006.
[24] D. Ramkrishna, Population Balances: Theory and Applications to Particulate Systems in
Engineering. Academic Press, 2000.
[25] P. Sandkühler, J. Sefcik, M. Lattuada, H. Wu, and M. Morbidelli, “Modeling structure effects on
aggregation kinetics in colloidal dispersions,” AIChE J., vol. 49, no. 6, pp. 1542–1555, Jun. 2003.
[26] D. L. Marchisio, R. D. Vigil, and R. O. Fox, “Quadrature method of moments for aggregation–
breakage processes,” J. Colloid Interface Sci., vol. 258, no. 2, pp. 322–334, Feb. 2003.
[27] H. M. Hulburt and S. Katz, “Some problems in particle technology,” Chem. Eng. Sci., vol. 19, no.
8, pp. 555–574, Aug. 1964.

Stuart McCready (mccready.stuart@gmail.com) 68


[28] R. B. Diemer and J. H. Olson, “A moment methodology for coagulation and breakage problems:
Part 2—moment models and distribution reconstruction,” Chem. Eng. Sci., vol. 57, no. 12, pp.
2211–2228, Jun. 2002.
[29] F. Moukalled, L. Mangani, and M. Darwish, The Finite Volume Method in Computational Fluid
Dynamics, vol. 113. Cham: Springer International Publishing, 2016.
[30] D. L. Marchisio, M. Soos, J. Sefcik, and M. Morbidelli, “Role of turbulent shear rate distribution
in aggregation and breakage processes,” AIChE J., vol. 52, no. 1, pp. 158–173, Jan. 2006.

Stuart McCready (mccready.stuart@gmail.com) 69


7)!Appendices#

Stuart McCready (mccready.stuart@gmail.com) 70


Appendix A – Velocity Field
Throughout the discussion of fluid mechanics, and subsequently of turbulence and its modelling, we
had a clear goal in mind – to give an Eulerian description of the velocity field in the mixing tank.

This velocity field (in other words, the spatial gradient of velocity in all three dimensions) is described
by the second-order tensor:

!!! !!! !!!


!" !" !"
!!! !!! !"! !!!
!!! = = !" !" =
!!! !" !" !"
!!! !!! !"!
!" !" !"

As with any second-order tensor, this can be decomposed into symmetric and skew components. To
highlight the physical meaning of these components, we will consider an example flow in two
dimensions only (the third component of velocity, which would point out of the screen, is assumed to
be zero everywhere). Note that the same ideas apply to the mean velocity (!! ).

Stuart McCready (mccready.stuart@gmail.com) 71


Total Velocity Field

Symmetric Component Skew Component


– describes the deformation – – describes the rotation –

1 !!! !!! 1 !!! !!! 1 1


!"#!" = + !!" = − = (!!×!!) = !!
2 !!! !!! 2 !!! !!! 2 2

Rate-of-Strain Tensor Rate-of-Rotation Tensor


This symmetric component is the pure strain of This represents a rigid-body rotation of the fluid
the fluid parcel, which can be further decomposed parcel about the point, which causes no
into an isotropic and anisotropic component: deformation. The vorticity of the fluid is !! , which
is twice the angular momentum Ω.

Anisotropic Strain Isotropic Strain

1 !!! !!! 1 !!!


!!" = + − !!" !!" = ! =!⋅!=0
2 !!! !!! 3 !!! !"

This gives the rate of shear of the fluid parcel, This gives the rate of expansion (divergence) of the
without any associated change in volume. fluid parcel, which produces a change in volume.

For incompressible fluids, ! ⋅ ! = 0, and hence the isotropic component is zero: !!" = 0. For this
reason, we often refer to the symmetric, anisotropic component simply as the rate-of-strain.

Stuart McCready (mccready.stuart@gmail.com) 72


Appendix B – Averaging Laws
After our important discovery of the statistical nature of the turbulent velocity field, we set about
defining the time-averaged version of the closed system presented at the end of "The Equations of
Fluid Motion" chapter. The following time-averaging proofs are helpful to those wishing to follow the
derivation themselves.

Firstly, for the mean velocity:

!!! !!!
!! 1 !! 1 !!
= !" = ! !" =
!" ∆! !" ∆! !"
! !

Secondly, for the fluctuating velocity:

!!! !!!
!" 1 !" 1
= !" = ! !" = 0
!" ∆! !" ∆!
! !

And finally, some general rules:

!=!

!+!=!+!

!⋅! =!⋅!

!" !!
=
!" !"

! !" = ! !"

!∙! ≠!⋅!

Stuart McCready (mccready.stuart@gmail.com) 73


Appendix C – The Mesh

Here, the interface between the impeller zone and the stationary zone is clearly visible.

Stuart McCready (mccready.stuart@gmail.com) 74

You might also like