Sunteți pe pagina 1din 78

A THESIS SUMMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF MASTER OF SCIENCE


IN CHEMICAL AND PROCESS ENGINEERING
SIRINDHORN INTERNATIONAL THAI-GERMAN GRADUATE SCHOOL OF ENGINEERING
(TGGS)
GRADUATE COLLEGE
KING MONGKUTS UNIVERSITY OF TECHNOLOGY NORTH BANGKOK
ACADEMIC YEAR 2007
COPYRIGHT OF KING MONGKUTS UNIVERSITY OF TECHNOLOGY NORTH BANGKOK
THE SIMULATION OF HYDROCYCLONE NETWORK FOR SEPARATING YEAST AND
CALCIUM IN ETHANOL PRODUCTION
MR.SUCHAIN TONGSIRI


ii


Name : Mr.Suchain Tongsiri
Thesis Title : The Simulation of Hydrocyclone Network for Separating Yeast
and Calcium in Ethanol Production
Major Field : Chemical and Process Engineering
King Mongkuts University of Technology North Bangkok
Thesis Advisor : Dr.Karn Pana-Suppamassadu
Co-Advisor : Assistant Professor Dr.Phavanee Narataruksa
Academic Year : 2007

Abstract
This thesis investigated the operating parameters governing the flow structures
of fermented broth inside a 10-mm hydrocyclone which in turn had a strong influence
on the separation of yeast and calcium in ethanol production. In doing so, a
Computational Fluid Dynamics approach (COMSOL MULTIPHYSICS 3.3) adopting
the k- turbulence model demonstrated the capability in the analysis of the
hydrodynamic of hydrocyclone, fluid-particle interaction, and effect of arrangements
of hydrocyclones. From the simulation results, the increasing mass flow rate of
fermented broth increased the axial velocities in both upward and downward
directions, and also increased the static pressure differential along the radial positions
inside the hydrocyclone. Together with the obtained flow structures, the particle
trajectory generated by Khan-Richardson force model gave further insights about the
separation mechanism and performance, which were essential for the design and
selection of hydrocyclone or hydrocyclone network. Finally, the hydrocyclone in
parallel configuration increased the production rate, and was efficient and suitable for
the separation process at hand.
(Total 66 pages)

Keywords : Hydrocyclone, Computational Fluid Dynamics, Hydrodynamic Stability,
Fluid-Particle Interaction
Advisor

iii

- a.- .+ .
-.aur .+..+.+..aaa.+++.:
..:..aaa.a+.a++.+a.!.....
++..r+
.- .+.....+.a+...:..
..a+a.++au....+u..+..
.av..aur+ .a a.. u+..a
.av..aur... +-.a+...a a... ..a.
v.+ :a

unaaa
. a ur+ a .u.. .... .+ .+.:+ ...+. .a!
aa.+++a a . ++ .... !- .:.aa .+a .a+.a++.+ a.!..:..
++ .. r+ aa+ a.+ .a. - .. .. +.u++.. .+ covso|
vu|.i|u\sics u. .:a:: +. . ...-...... aa.! . a .+a.u!.
..+..aa.a.+.aa.+++ ...a.....+a+. .+a+.++.
..+..aaa.+++ ++..+.a.-....+u:. ...+ .u ..
.++!+.....!a.a ..u .! +..+.a+..+.+. a+.a. .++
!. +..a....+..aa...a..+..a!aa.+++.u.a.a +....
..+ aa+.....++ . .+a .a+.a++.+ a. + .aa:: +.
.a... x........+.. a!... a.:+a+.v.. r .u..+aaa.v
.++!.a::a+..+!-aa.+++. ..::aa.+++ !.aa.
u:. .. aa.+++a::..a.u . ...++ .a+. ! .+ aaav..:
++..a a.
.aur... --

++ aa.+++ .+.a.-....+.u++...+ .aa..u.
u++...+ ... a.....+a+. .+
.av..aur+

iv

ACKNOWLEDGEMENTS

First of all I wish to thank my parents because without their support I would not
have studied at the university at all. So this thesis would never have come into
existence without their help.
Special thanks go to Dr.Karn Pana-Suppamassadu and Assistant Professor Dr.
Phavanee Narataruksa who are the advisors of this work. On one hand they supported
me very well during the various stages of this work, especially with a thoroughly
planned schedule, on the other hand they gave me enough freedom to make decisions
on my own and follow my own way.
I am very thankful to everyone of TGGS-CPE advisors, King Mongkuts
University of Technology North Bangkok and RWTH Aachen University.
Thanks also go to my friends for making my life good fun and for reading and
correcting this work.


Suchain Tongsiri

v
TABLE OF CONTENTS
Page
Abstract (in English) ii
Abstract (in Thai) iii
Acknowledgements iv
List of Tables vii
List of Figures viii
List of Abbreviations and Symbols xi
Chapter 1 Introduction 1
1.1 Background and general statement of the problem 1
1.2 Purpose of the study 2
1.3 Scope of the study 2
1.4 Benefits of the study 2
1.5 Methodology of the study 2
1.6 Methodology 3
Chapter 2 Hydrocyclone Principles 4
2.1 Hydrocyclone 4
2.2 Separation principle of hydrocyclone 5
2.3 Velocity distributions 7
2.4 Separation theories of hydrocyclones 9
2.5 Literature reviews 14
Chapter 3 Methodology 20
3.1 Governing Equations 20
3.2 Procedures in flow simulation 23
Chapter 4 Results and Discussions 31
4.1 Hydrodynamics of hydrocyclone 31
4.2 Fluid-particle interaction 50
4.3 Hydrocyclone network in a series connection 58
4.4 Hydrocyclone network in a parallel connection 59
Chapter 5 Conclusions 62
5.1 Conclusions 62
5.2 Suggestions 63

vi
TABLE OF CONTENTS (CONTINUED)
Page
References 64
Biography 66

vii
LIST OF TABLES

Table Page
2-1 Geometric proportions of two well-known families of
hydrocyclones 5
3-1 Physical properties of fermented broth 25
3-2 Mesh Statistics 28
4-1 Comparison of predicted water splits with experimental value in
a 101-mm hydrocyclone 43
4-2 Percentage of particle entering underflow and overflow when they
were injected into the hydrocyclone at different positions 57
4-3 Residence time of particle inside the hydrocyclone that injected
into the hydrocyclone at different positions and finally discharging
from underflow 57
4-4 Residence time of particle inside the hydrocyclone that injected
into the hydrocyclone at different positions and finally discharging
from overflow 57
4-5 Pressure at inlet, overflow and underflow of three hydrocyclone in
series 59
4-6 Pressure at inlet, overflow and underflow of hydrocyclone in
parallel 61


viii
LIST OF FIGURES

Figure Page
2-1 Schematic diagram of a hydrocyclone 4
2-2 Essential feature of the hydrocyclone 6
2-3 Tangential velocity distribution in a hydrocyclone 7
2-4 Vertical (axial) velocity distribution in a hydrocyclone 8
2-5 Radial velocity distribution in a hydrocyclone 9
3-1 Procedures in flow simulation by COMSOL Multiphysics 23
3-2 Geometry of a hydrocyclone 24
3-3 Geometry of a hydrocyclone in COMSOL Multiphysics 24
3-4 Boundary condition at the feed inlet 26
3-5 Boundary condition at walls 26
3-6 Boundary condition at the overflow and the underflow outlets 27
3-7 Boundary condition between sub-domains 27
3-8 Mesh grids in COMSOL Multiphysics 28
4-1 Pressure distribution within the hydrocyclone at different
transverse planes and at central vertical plane 32
4-2 Cell Reynolds pattern at the central vertical plane varies with the
feed velocity (a) 10 m/s, (b) 15 m/s, (c) 20 m/s and (d) 25 m/s 32
4-3 Static pressure (a) at z = 44 mm below the vortex finder bottom
end and (b) at z =0 underflow exit plane 33
4-4 Axial velocity (a) contour at central vertical plane and
(b) distributions along the radial positions and axial heights 35
4-5 Axial velocity contours in the vertical plane (a) Positive (upward
vertical flow) and (b) negative (downward vertical flow) 35
4-6 Magnified view of the flow near the hydrocyclone wall: (a) x-y
plane at z = 30 mm (b) center vertical plane in the connection
region between the cylindrical and conical parts 36

ix
LIST OF FIGURES (CONTINUED)

Figure Page
4-7 Velocity vectors (a) in a plane close to the bottom end of vortex
finder (b) in a plane close to the underflow opening. Feed velocity
of 20 m/s 37
4-8
Axial vorticity contour
Z
(a) in a plane close to the bottom end
of vortex finder (b) in a plane close to the underflow opening.
Feed velocity of 20 m/s 37
4-9
Axial vorticity contour
Z
(a) in a plane close to the bottom end
of vortex finder (b) in a plane close to the underflow opening.
Feed velocity of 25 m/s 38
4-10 Tangential velocity distribution on the center-plane of
hydrocyclone (a) contour plot (b) value plot 39
4-11 Tangential velocity distribution on the central plane of
hydrocyclone (a) 10 m/s (b) 15 m/s (c) 25 m/s (d) 30 m/s 40
4-12 Velocity vector projected on the planes x-y at different heights of
hydrocyclone 40
4-13 vector projected on the central plane (z-x) of hydrocyclone 41
4-14 Cell Renolds Number at feed inlet velocity of 15 m/s (a) z = 6 mm
(b) z = 12 mm (c) z = 18 mm (d) z = 24 mm (e) z = 30 mm
(f) z = 36.2 mm (g) z = 42.4 mm (h) z = 48.6 mm 42
4-15 Typical air-core occur in the conventional and the commercial
Hydrocyclones 44
4-16 Centrifugal instability of flow in a boundary layer on a concave
wall 45
4-17 The critical values of the Gortler number for (1) the Blasius,
(2) the asymptotic suction and (3) the straight-line profiles,
Respectively 47
4-18
Circumferential vorticity contour

at feed velocity of 20 m/s


(a) in the center-plane at = 90 and (b) in the offset plane
48

x
LIST OF FIGURES (CONTINUED)

Figure Page
4-19 Turbulence kinetic energy at a 0-plane at feed velocity of 15 m/s 48
4-20
Turbulent kinetic energy in a plane = 0 at feed velocity of
15 m/s: (a) z = 48.6mm, (b) z = 42.4 mm, (c) z = 36.2 mm,
(d) z = 30 mm, (e) z = 24 mm, (f) z = 18 mm, (g) z = 12 mm and
(h) z = 6 mm 49
4-21 Turbulent kinetic energy in a plane z = 30 mm at feed velocity of
15 m/s (a) = 0, (b) = 45, (c) = 90 and (d) = 135
50
4-22 Pathline trajectories of particles with different feed velocities
that discharging from underflow: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s 54
4-23 Pathline trajectories of particles with different feed velocities
that discharging from overflow: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s 54
4-24 Pathline trajectories of particles with different feed velocities
that discharging unspecification: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s 55
4-25 Feed inlet position of particles at feed velocity of: (a) 10 m/s
(b) 15 m/s, (c) 20 m/s, (d) 25 m/s and (e) 30 m/s 56
4-26 Three hydrocyclones in series 58
4-27 Pressure distribution of three hydrocyclones in series 59
4-28 Six hydrocyclones in parallel 60
4-29 Pressure distribution in main pipe of hydrocyclone network in
parallel 60

xi
LIST OF ABBREVIATATIONS AND SYMBOLS

C Feed volumetric concentration


d Particle diameter, m
50
d Cut size, m
50
d Reduced cut size, m
c
D Hydrocyclone diameter, m
i
D Feed inlet diameter, m
o
D Overflow diameter, m
u
D Underflow diameter, m
Eu Euler number
T
E Total efficiency
T
E Reduced total efficiency
F
v
Volumetric force vector, kg
2
m s
-2

G Reduced grade efficiency
k Parameter of the model
K Turbulent kinetic energy, J kg
-1
L Hydrocyclone length, m
inl
L Length of the inlet channel, m
l
L Height of hydrocyclone in cylindrical part, m
l Vortex finder length, m
m Particle mass, kg
n Constant
p Pressure, kg m
-1
s
-2

inl
p Pressure at the end of the inlet channel, kg m
-1
s
-2

P Pressure drop, kg m
-1
s
-2

Q Feed volumetric flow rate, m
3
s
-1

r Radius of hydrocyclone, m
p
r Particle radius, m

xii
LIST OF ABBREVIATATIONS AND SYMBOLS (CONTINUED)

0
R Radius of curvature of the wall
W
R Water flow ratio
Re Reynolds number
p
Re Particle Reynolds number
50
Stk Stokes number
t Time, s
r
u Particle terminal velocity in radial direction, m s
-1

u
v
Velocity vector, m s
-1

U Velocity field or the modulus of the velocity vector, m s
-1

r
v Radial velocity, m s
-1

t
v Tangential velocity, m s
-1
z
v Velocity in z component, m s
-1

y Cumulative particle size distribution (undersize) of feed suspension
Dissipation rate
Dynamic viscosity, kg m
-1
s
-1

Angle of the hydrocyclone cone, degree
Fluid density, kg m
-3

S
Solid density, kg m
-3

Residence time, s
Momentum thickness of the boundary layer
Boundary-layer thickness
Rayleigh discriminant

CHAPTER 1
INTRODUCTION

1.1 Background and General Statement of the Problem
In ethanol production yeast obtained is added to a fermentor tank in converting
starch and sugar into alcohol. The product is fermented broth containing alcohol and
yeast, which is processed through filtration and distillation processes and is
concentrated. In the fermentor tank yeast still breeds continuously so that when the
fermentation is complete, it contains an amount of juvenile yeast that still works to be
used in the next fermentation. If it is possible to find a suitable technique to separate
juvenile yeast and reuse it in the next fermentation and to bring back old yeast to use
without leaving yeast, it will be a benefit for alcohol industries. In addition a
separation of yeast is an essential process because if there is no process to separate
yeast from fermented broth before feeding it to the distillation tower, fouling in the
distillation tower and decreased efficiency of the distillation tower can occur and can
result in shutting down the process and necessitate clean up. So before fermented
broth is fed into the distillation tower, it has to be put through a separation process
such as sedimentation, filtration, and hydrocyclone etc. Sedimentation and filtration
are the only two processes used so far in separating yeast. In sedimentation process it
takes at least a week to make yeast sediment settle to the bottom of the separator.
However, a recently previous research (Cilliers and Harrison, 1997) found that mini-
hydrocyclones can separate some amounts of yeast from fermented broth. From that
result, this research adopt the idea of using mini hydrocyclones in separating yeast
and believes that this approach can be applied in the alcohol industries which usually
have a high production rate. Therefore, this research will study the flow
configurations of fermented broth within hydrocyclone.


2
1.2 Purpose of the Study
This research aims to find out the flow configurations of the fermented broth
within a 10-mm hydrocyclone in separating yeast in ethanol production using the
computational fluid dynamics as COMSOL MULTIPHYSICS 3.3.

1.3 Scope of the Study
1.3.1 Find the flow configurations of the fermented broth within the
hydrocyclone by getting good efficiency based on a proper utilization.
1.3.2 Find parameters which influence on the flow configuration of fermented
broth within the hydrocyclone.
1.3.3 Integrate the results of the computational study in order to propose a
general hydrocyclone design guideline.

1.4 Benefits of the Study
1.4.1 The optimum configuration of the hydrocyclone could be properly applied
in the ethanol industries.
1.4.2 The knowledge of a design and operating conditions of hydrocyclone
using computational fluid dynamics, which is applied in separating particles for
several industries.
1.4.3 A creation of an alternative technique in the particle separation, which
increases an efficiency of continuous production process such as an increased
efficiency of the distillation column and a decreased waste product occurring from the
old separating processes.

1.5 Methodology of the study
1.5.1 Investigate and evaluate the flow configurations of fermented broth within
the hydrocyclone using a commercial software as COMSOL MULTIPHYSIC 3.3.
1.5.2 Conduct the results of the hydrocyclone to evaluate its performance.
1.5.3 Analyze and make a conclusion about the findings of the present study.
3
1.6 Methodology
This research studied mainly in flow configuration of the fermented broth
within a hydrocyclone using the computational fluid dynamics. This research will
describe the following topics:
Chapter 2 will describe the hydrocyclone principle including the characteristics
of yeast.
Chapter 3 will describe methodology of flow simulation by using computational
fluid dynamics software.
Chapter 4 will describe the results obtained from computational fluid dynamics.
Chapter 5 will describe the conclusions of this research.
CHAPTER 2
HYDROCYCLONE PRINCIPLES

2.1 Hydrocyclone
Traditionally, cyclones have found wide application in various fields of
technology such as gas cleaning, burning, spraying, atomizing, powder classification
etc. They are also used for solid-liquid separation. The cyclones specially designed for
liquids are so-called hydrocyclones, hydraulic cyclones or hydroclones. The basic
separation principle applied in cyclones is centrifugal sedimentation, i.e. the
suspended particles are subjected to centrifugal acceleration which makes them
separate from the fluid. Unlike centrifuges (which use the very same principle),
cyclones have no moving parts and the necessary vortex motion is performed by the
fluid itself.

FIGURE 2-1 Schematic diagram of a hydrocyclone
D
o

D
i

D
u

D
c

L
1

l
L

unit: mm
5
Figure 2-1 shows a cross-section of a hydrocyclone of conventional design. It
consists of a cylindrical section joined to a conical section. The suspension of
particles in a liquid is injected tangentially through the inlet opening in the upper part
of the cylindrical section and, as a result of the tangential entry, a strong swirling
motion is developed within the cyclone. A portion of the liquid containing the fine
fraction of particles is discharged through a cylindrical tube fixed in the center of the
top and projecting some distance into the cyclone; the outlet tube is called the
overflow pipe or vortex finder. The remaining liquid and the coarse fraction of
material leaves through a circular opening at the apex of the cone, called the
underflow orifice or spigot (Svarovsky, 1984).

TABLE 2-1 Geometric proportions of two well-known families of hydrocyclones
(Castilho and Medronho, 2000)
Hydrocyclone D
i
/D
c
D
o
/D
c
L/D
c
L
1
/Dc l/D
c

Bradley 1/7 1/5 - 1/2 1/3 9
Rietema 0.28 0.34 5 - 0.40 20

As with all separation principles involving particle dynamics, a knowledge of
the flow pattern in the hydrocyclone is essential to understand its function and
subsequently for the optimum design and evaluation of the particle trajectories, which
in turn allows prediction of the separation efficiency. A short account of the flow
pattern within typical hydrocyclones and the known or probable behaviour of solid
particles in the flow are given in the following section.

2.2 Separation Principle of Hydrocyclone
When the fluid is tangentially fed through the inlet opening, a primary vortex is
formed by the fluid itself in the cylindrical section of the hydrocyclone. This vortex
creates a centrifugal force and results in the dispersion of the particles across the
whole cross-sectional area. Two equally important stages can be distinguished in the
subsequent separation process; separation of solids from the main flow and their
migration to the boundary layer on the wall, and the removal of the separated solids
from the wall into the apex and out of the cyclone. A small fraction of the flow will
6
form a cross-circuit flow below the top cover and bypass the vortex; consequently,
any particles in this flow will then directly pass through into the overflow. On the
other hand, most particles will go through the primary vortex under the influence of
the centrifugal forces. The centrifugal forces must be greater than drag forces to drive
particles radially towards the wall; otherwise they tend to move radially inward. Since
centrifugal and drag forces are proportional to particle volume and size, respectively,
therefore the separation performance of hydrocyclone is supposed to be strongly
particle-size dependent. Thus, large particles are expected to be readily separated than
fine ones, which may not reach the boundary layer by the time the primary vortex
reaches the bottom of the cone. Upon reaching the bottom of the cone, the primary
vortex flow and its fine particles feed across into the secondary vortex. Some of the
fine particles may also be able to leave the secondary vortex resulting in a certain
amount of solid circulation, while others will leave with the overflow. The boundary
layer and its particle content move downwards into the apex of the cone where they
leave from the spigot. The separation in hydrocyclone is not affected by gravity. It is
the flow itself that drives particles into the boundary layer (Cilliers and Harrison,
1997).


FIGURE 2-2 Essential Feature of the Hydrocyclone
(Cilliers and Harrison, 1997)
7
2.3 Velocity Distributions
Velocities can be considered under the three component dimensions into which
they all can be resolved. The most useful and significant of these components and the
one which has been subject to the most study is the tangential velocity. Data on the
other components, the vertical and radial velocities will, however, be reviewed.
2.3.1 Tangential Velocity (Svarovsky, 1984)
At level below the rim of the vortex finder, the tangential velocity,
t
v , increases
considerably with decreasing radius down to a give radius, which is smaller than the
exit radius of the vortex finder as shown in Figure 2-3. This can be described by the
relationship

const r v
n
t
= Eq.2-1

where n is normally 0.6 n 0.9
As the radius is further increased, the tangential velocity decreases and is
proportional to r , this relationship holds until the cylindrical air core which normally
forms in the hydrocyclone discharging at atmospheric pressure is reached. At levels
above the rim of the vortex finder, the break in the rise of
t
v occurs at a larger radius
as can be seen in Figure 2-3. Apart from this phenomenon and the wall effects,
t
v is
independent of the vertical position so that envelopes of constant tangential velocity
are cylinders coaxial with the cyclone.

FIGURE 2-3 Tangential velocity distribution in a hydrocyclone
(Svarovsky, 1984)
8
2.3.2 Axial Velocity (Svarovsky, 1984)
As can be seen from Figure 2-4, there is a strong downward flow along the
outer walls of both the cylindrical and conical portions. This flow is essential for
cyclone operation since it removes the particles that have been separated into the
underflow orifice. It is for this reason that it is not essential to build cyclones with the
apex pointing downwards. The cyclone efficiency is only very slightly influenced by
its position relative to the gravity field.
The downward current is partially counterbalanced by an upward flow in the
core region, depending on the underflow-to-throughput ratio. There is a well-defined
locus of zero vertical velocities (LZVV) which follows the profile of the cyclone.
Above the rim of the vortex finder, the largest downward velocities occur again
near the cyclone wall. At radii between the cyclone wall and the vortex finder, the
axial velocity becomes upward. Around the vortex finder strong downward flow may
be observed. This is due to wall-induced flow which runs inward along the top of the
cyclone.

FIGURE 2-4 Vertical (axial) velocity distribution in a hydrocyclone
(Svarovsky, 1984)

2.3.3 Radial Velocity (Svarovsky, 1984)
The radial velocity components are normally much smaller than other two
components and, as such, they are difficult to measure accurately. As can be seen
from Figure 2-5, the radial velocity is inward and its magnitude decreases with
decreasing radius. The radial position of zero radial velocity is not known.

9
const r v
m
r
= Eq.2-2

At level above the rim of the vortex finder, there may be outward recirculatory
flows, and near the flat top of the cyclone there are strong inward radial velocities
directed towards the root of the vortex finder, thus causing the above mentioned short
circuit flow down the outside wall of the vortex finder.
It should be pointed out here that this short account of velocity profiles in a
hydrocyclone is only qualitative. The flow patterns are highly complex even for water
with a low specific gravity and viscosity, and it may be incorrect to assume that
precisely similar profiles occur in cyclones with a considerably different geometry or
with liquids of high viscosity.


FIGURE 2-5 Radial Velocity Distribution in a Hydrocyclone
(Svarovsky, 1984)

2.4 Separation Theories of Hydrocyclones
There have been many theories developed to explain the separation process
within hydrocyclones. Some of these theories are based on actual physical models
derived from first principles while others are the so-called model derived as empirical
equations from linear regression analysis. The major two theories applied for the
separation of particles in hydrocyclones are the equilibrium orbit and the residence
time theories.

10
2.4.1 Equilibrium Orbit Theory
This theory is based on the concept of equilibrium radius of the particle, since
the motion inside the hydrocyclone is developed by the swirling liquid flow itself. The
tangential velocity and the radial fluid velocity toward the center are both assumed to
vary exponentially with radius:

const r v
n
t
= Eq.2-1

const r v
m
r
= Eq.2-2

The balance between these forces corresponds to an equilibrium orbit position
of the particle within the hydrocyclone. If Stokes law is assumed, the governing
equation is:

r
v
v
t
r
2
= Eq.2-3

where can be obtained from

18
2
d
S
= Eq.2-4

The equations can be combined to give
r r
v u = at equilibrium

m n
r d
+

2 1 2
Eq.2-5

and d is a diameter of particle, and is the dynamic viscosity of carrier fluid. The
equations can be combined to give, at equilibrium. Since the exponent m n + 2 1 is
always greater than 0, therefore larger particles will be found close to the wall and
entrained to the outer vortex swirling downwards toward the underflow outlet.
Smaller particles, on the other hand, will be found toward the center and entered the
inner fluid core swirling upward (Cilliers and Harrison, 1997).
11
However, the equilibrium orbit theory has been criticized on the bases that it
does not consider the residence time of particles inside the hydrocyclone. Since the
residence time of most particles inside the hydrocyclones is typically short normally
in millisecond, it is unlikely will reach its equilibrium position. The theory also takes
no account of the turbulences, which evidently affect the separation process
(Svarovsky, 1984).
2.4.2 Residence Time Theory
As proposed by Rietema (1965), it takes into account non-equilibrium
conditions considering whether a specific size particle, d , will reach the hydrocyclone
wall within its residence time, T , and report to the underflow, i.e.,

dr
D v
L
dt
c z
2
= Eq.2-6

For the inlet diameter,
i
D , the necessary velocity,
r
v , is expressed as:

i
T
r
D dt v 5 . 0
0
=

Eq.2-7

Based on Stokes analysis, the radial velocity in a centrifugal field of particles of the
cut size is:

( )
r
v d
u
r S
r
2 2
50
18

= Eq.2-8

( )


2
50
9
d
v D D
P L
R
z c i
S
S

= Eq.2-9

Based upon these ideas, Cilliers and Harrison (1997) established the relationship
between solid recovery, cut size, and viscosity for a single particle size as in Eq.2-9;
they also applied to mono-disperse particle in 10-mm hydrocyclones, and determined
12
the feed pressure changes on solids recovery. For poly-disperse feed size distribution,
d is some representative size (Cilliers and Harrison, 1997). Nevertheless, the
residence time theory does not take into account the so-called hindered settling, and
neglect inertial effects and radial fluid flow.
Both theories were developed under the assumption that the feed suspension is
diluted (excluding hindered settling). This led to the conclusion that the product of the
Stokes number and the Euler number, Eu Stk
50
, is constant for geometrically similar
hydrocyclones. However, recently, Coelho and Medronho (2001) reported that the
product Eu Stk
50
can be expected as a function of water flow ratio
W
R and volumetric
feed concentration

C since the feed concentration reduces the terminal settling


velocities of the particles. Also, it is expected to vary with
u
D or of an operational
variable greatly affected by parameters
u
D , such as the water flow ratio
W
R . Based on
experimental works of Coelho and Medronho (2001), a semi-empirical model, based
on dimensionless groups for Reitemas optimum design given by Eq.2-10 Eq.2-12
was adopted in this study for analysis and verification purposes.
For analysis and qualitative verification purposes, the estimated cut size from
the present CFD simulation was compared with that obtained from the semi-empirical
correlations of Coelho and Medronho (2001). Based on experimental works and
literature reviews of Coelho and Medronho (2001), a semi-empirical model based on
dimensionless groups for Reitemas optimal design given by Eq.2-10 Eq.2-12:

( )
(

|
|

\
|
|

\
|

|
|

\
|
= C
R l L
D
D
D
Eu Stk
W
c
o
c
0 . 12 exp
1
ln 12 . 0
79 . 0
33 . 1
95 . 0
50
Eq.2-10

( )

\
|

|
|

\
|
+
|
|

\
|
= C
l L
D
D D
D
D
D
D Eu
c
u o
c
i
c
c
51 . 0 exp Re 5 . 43
12 . 0
98 . 0
42 . 0
2 2
61 . 2
57 . 0
Eq.2-11

54 . 0
10 . 3 97 . 5
18 . 1

|
|

\
|
|
|

\
|
= Eu
D
D
D
D
R
c
u
o
c
W
Eq.2-12

13
where the product Eu Stk
50
, the Euler number Eu , the Reynolds number Re , and the
water flow ratio
W
R can be found from Eq.2-13 Eq.2-16, respectively.

( )
Q
d PD
Eu Stk
c S


36
2
50
50

= Eq.2-13
2
4 2
8 Q
PD
Eu
c


= Eq.2-14

c
D
Q

4
Re = Eq.2-15

( )
( )

=
C Q
C Q
R
u u
W
1
1
Eq.2-16

The reduced grade efficiency curve G as expressed in Eq.2-17 was given by a
modification of the Rosin-Rammler distribution function as proposed by Plitt (1976).
The reduced grade efficiency and the particle size distribution of the feed as
cumulative undersize fraction y as expressed in Eq.2-18 will describe the
performance of hydrocyclone (Castilho and Medronho, 2000).

(
(

|
|

\
|

=
45 . 2
50
693 . 0 exp 1
d
d
G Eq.2-17

|
|

\
|
|

\
|
=
m
k
d
y exp 1 Eq.2-18

Furthermore, the reduced total efficiency
T
E and the total efficiency
T
E can be
determined from equations Eq.2-19 Eq.2-20.


14

=
1
0
dy G E
T
Eq.2-19

W
W T
T
R
R E
E

=
1
Eq.2-20

By considering Eq.2-10 and Eq.2-13, the reduced cut size as expressed in Eq.2-21
will be calculated.

( )
( )
( )
(

|
|

\
|
(

= C
R P
Q
l L D
D
d
W S
o
c
0 . 6 exp
1
ln
173 . 1
395 . 0
5 . 0
665 . 0 475 . 0
64 . 0
50


Eq.2-21

2.5 Literature Reviews
Cilliers and Harrison (1997): Small diameter hydrocyclones have had an
increasing use in performing difficult separations between phases, due to the large
centrifugal forces generated in them. The potential use of hydrocyclones in the
concentration of microbial suspensions is attractive as they are continuous, high
capacity devices requiring low maintenance while having the additional benefit in that
they can be readily sterilised.
Results are reported on the de-watering of Bakers yeast in a 10 mm diameter
hydrocyclone to quantify the separation process. The form of the model equation for
recovery has been derived based on the non-equilibrium residence time theory. This is
shown to represent experimental data in that increasing pressure and temperature
exhibit a positive effect on both the recovery and the concentrating effect while an
increase in the feed concentration exhibits a negative effect on these. In addition, the
influence of cyclone geometry on the recovery and concentration ratio has been
illustrated. Increasing the vortex diameter results in an increasing concentration ratio
and a decreasing recovery. Increasing the diameter of the spigot shows the opposing
trends.
Typical results from a single stage separation combine a recovery of 60% with a
concentration ratio of 1.25 and a recovery of 30% with a concentration ratio of 2.0.
15
Concomitant improvement of the recovery and concentration ratio will be attainable
through the use of multi-stage hydrocyclone circuits.
Yang, Shin, Kim and Kim (2004): A three-dimensional simulation was
performed to predict the flow field and the separation efficiency for particles in a
hydrocyclone to be used for the sludge separation in water purifying plants. The
Reynolds averaged NavierStokes and Reynolds averaged continuity equations
employing renormalization group k model were solved to calculate the turbulent
flow field in a hydrocyclone. The particle trajectories were computed by integrating
the force balance equations on particles based on the predicted flow field. The
separation efficiency defined as the fraction of particles recovered to underflow was
obtained by using the calculated particle trajectories. The separation efficiency of a
hydrocyclone for the sludge separation in water purifying plants was measured
experimentally as a function of particle size. Simulation results agreed favorably with
experimental data.
Slack, Del Porte and Engelman (2004): Resolution of the complex multiphase
swirling flow field in a hydrocyclone using computational fluid modelling techniques
is not a trivial task. CFD Modelling technology is not perfect and it is certainly still
possible to improve our understanding of the fundamentals and the models needed to
describe them. Nevertheless, computational modelling techniques are being used to
compare and understand the workings of different hydrocyclone designs and modes of
operation.
This paper describes how automated CFD tools have been developed based on
documented best practices. The bespoke tool development described will enable a
non-CFD analyst to carryout hydrocyclone simulations. The paper will also consider
the advantages and limitations of different computational fluid dynamic techniques
applicable to hydrocyclone modelling.
Delgadillo and Rajamani (2005): In the computational fluid dynamics study of
hydrocyclones, the air-core dimension is key to predicting the mass split between the
underflow and overflow. In turn, the mass split influences the prediction of the size
classification curve. Three models, the renormalization group je model, the Reynolds
stress model, and the large-eddy simulation model, are compared for the predictions
of air-core dimension, mass split, and axial and tangential velocities. The large-eddy
16
simulation model, since it produces some detailed features of the turbulence, is clearly
closer in predicting the experimental data than the other two. It is shown that particle
tracking done with the velocity field obtained from the large-eddy simulation model
accurately predicts the experimental sizeclassification curve.
Delgadillo and Rajamani (2006): The hydrocyclone has been widely used in the
mineral industry for over hundred years, yet the standard geometry of hydrocyclones
has remained almost unchanged. The exploration of new designs is time-consuming
and costly to do by experimentation. In this paper, the computational fluid dynamics
tool is used to explore alternative geometries in a way to manipulate the
hydrodynamics to achieve the desired classification. Fluent 6.0 was used to solve
the governing equations. The large-eddy simulation model was used for the
turbulence closure and the Lagrangian particle-tracking method was used to predict
the particle classification. Six new geometries are explored and compared with the
standard design. The mass balance and the classification curve are the variables used
to evaluate the performance of each of the novel designs.
Narasimha, Brennan and Holtham (2006): Hydrocyclones are widely used in the
mining and chemical industries. An attempt has been made in this study, to develop a
CFD (computational fluid dynamics) model, which is capable of predicting the flow
patterns inside the hydrocyclone, including accurate prediction of flow split as well as
the size of the air-core. The flow velocities and air-core diameters are predicted by
DRSM (differential Reynolds stress model) and LES (large eddy simulations) models
were compared to experimental results. The predicted water splits and air-core
diameter with LES and RSM turbulence models along with VOF (volume of fluid)
model for the air phase, through the outlets for various inlet pressures were also
analyzed. The LES turbulence model led to an improved turbulence field prediction
and thereby to more accurate prediction of pressure and velocity fields. This
improvement was distinctive for the axial profile of pressure, indicating that air-core
development is principally a transport effect rather than a pressure effect.
Wang and Yu (2006): This paper presents a numerical study of the gasliquid
solid multiphase flow in hydrocyclones with different dimensions of body
construction, which include the lengths of cylindrical and conical parts and cyclone
body size. The turbulent flow of gas and liquid is modelled using the Reynolds stress
17
model, the interface between the liquid and air core is modelled using the volume of
fluid multiphase model, and the results are then used in the simulation of particle flow
described by the stochastic Lagrangian model. The flow features are examined in
terms of flow field, pressure drop, split ratio reported to the underflow, particle
trajectories and separation efficiency. The proposed model is first validated by the
good agreement between the measured and predicted results, and then used to study
the effects of cyclone size and length. The results show that the flow fields in the
hydrocyclones with different size and length are different, which results in different
performance. A smaller cyclone is helpful to higher efficiency. The cylindrical section
plays an inessential role in collecting particles. A long conical section can improve the
performance of hydrocyclone considerably.
Mainza, Narasimha, Powell, Holtham and Brennan (2006): Simplicity in design
and minimal floor space requirements render the hydrocyclone the preferred classifier
in mineral processing plants. Empirical models have been developed for design and
process optimisation but due to the complexity of the flow behaviour in the
hydrocyclone these do not provide information on the internal separation
mechanisms. To study the interaction of design variables, the flow behaviour needs to
be considered, especially when modelling the new three-product cyclone.
Brennan, Narasimha and Holtham (2007): A comprehensive multiphase model
of cyclone separators using Computational Fluid Dynamics is under development.
The model is capable of predicting velocity profiles, flow splits, air core position and
efficiency curves in classifying hydrocyclones. The model approach uses the Mixture
model with the granular options and large eddy simulation (LES) to resolve the
turbulent mixing of the particles. Multiphase simulations of Hsiehs [Hsieh, K.T.,
1988. A phenomenological model of the hydrocyclone, Ph.D. thesis, University of
Utah] data show a very good prediction of the cyclone efficiency curve. Whilst further
model development is needed, the approach is showing promise as a cyclone design
tool.
Bhaskar, Murthy, Raju, Tiwari, Srivastava and Ramakrishnan (2007):
Hydrocyclone is a key unit operation in mineral process industry and simulation of
which using CFD techniques is gaining popularity in process design and optimization.
The success of the simulation methodology depends primarily on how best the results
18
are matching with the experimental values and the computational time it requires for
obtaining such results. In the present investigation, attempts are made to develop a
methodology for simulating the performance of hydrocyclone. Initial work included
comparison of experimental and simulated results generated using different
turbulence models i.e., standard k, k RNG and RSM in terms of water throughput
and split with the help of suitably designed experiments. Among the three modeling
methods, predictions using RSM model were found better in agreement with
experimental results with a marginal error between 4% and 8%. Parametric studies
have indicated that a decrease in the spigot opening increased the upward vertical
velocity of water more compared to a decrease in the downward vertical velocity. An
increase in the inlet pressure has increased the axial velocities of water in both the
upward and downward directions and increased the mass flow rates through the
cyclone. An increase in the inlet pressure has also increased the static pressure
differential along the radius within the cyclone body and hence more water split into
overflow. Further, an increase in the inlet pressure has also increased the tangential
velocities and reduced the cyclone cut size. The simulated particle distribution values
generated using the particle injection technique are found matching with the
experimental results while achieving cut sizes between 4.9 and 14.0 m.
Bhaskar, Murthy, Ramakrishnan, Srivastava, Sarkar and Kumar (2007): The
investigation pertains to establishing a simulation methodology for understanding the
flyash classification characteristics of a 76 and 50 mm diameter hydrocyclone where
the work was carried out using commercially available CFD software. Comparative
results on the simulated and experimental water throughput, split values are presented.
Results indicted that there is a good match in water split between the experimental
and simulated values with error values below 10% at different hydrocyclone designs.
Further a discussion is made on the flow features at comparable ratio of cyclone
diameter to spigot opening in the 76 and 50 mm designs. The vertical core region
around the cyclone axis having static pressure equal to or below the atmospheric
pressure is examined to be increasing in diameter from bottom of the spigot opening
till the interface where vortex finder joins the main cylindrical cyclone body and
remains more or less similar at the vortex finder outlet. The diameter of this zone at
the spigot outlet is 0.6 and 5.4 mm at 3.2 and 9.4 mm spigot openings in case of 50
19
mm diameter hydrocyclone. The diameter of the core at spigot outlet is found to be
around 9.2 and 11 mm at 15 and 20 mm for spigot openings in case of 76 mm
diameter hydrocyclone.
Doby, Nowakowski, Nowak and Dyakowski (2007): Swirl flow is the primary
mechanism for separation of particles in a hydrocyclone and hence influences the
performance. The study of examining the basic swirl flow mechanism is
accomplished by using a cylindrical apparatus with rotating lids. The upper and lower
lids of the cylinder are able to rotate in the counter and clockwise direction at
different angular velocities. In this study, the basic mechanisms of swirling in the
laminar regime are investigated using an optically clear cylinder with rotating lids.
The cylinder is filled with a liquid and the upper and lower lids of the cylinder are
allowed to rotate in the clockwise and counter-clockwise directions. The induced flow
is measured with a laser doppler anemometer and is compared to the velocity
predictions from an in house developed software which solves the equations of
motion. The results presented here constitute the first step in a series of experiments
with the ultimate goal to improve the separation mechanism of hydrocyclones by
means of controlling swirling flow characteristics.
Murphy, Delfos, Pourquie, Olujic, Jansensa and Nieuwstadt (2007): Two
commercial CFD codes were used to simulate the strongly swirling single-phase flow
with core recirculation within an axial hydrocyclone. Both packages used a
Differential Reynolds Stress Model with default constants for the turbulence closure.
The effect of omitting wall reflection terms was also investigated and it was generally
found that this lead to better agreement with experiment. Interestingly, the predicted
velocity profiles from the two CFD codes did not agree with each other. Possible
reasons for this are different turbulence modeling approaches with different terms for
the turbulent diffusion and rapid pressurestrain terms.


CHAPTER 3
METHODOLOGY

The flow simulation conducted in this investigation used the Computational
Fluid Dynamics Software (CFD) as COMSOL Multiphysics 3.3, which utilized the
Finite Element Scheme to carry out coupled governing non-linear partial differential
equations in fluid mechanics. The CFD simulation was performed with an Intel (R)
Pentium (R) 4 CPU 2.93 GHz HP workstation xw6400 with 512 cache-memory, 6.00
GB RAM, and 110 GB hard disc memory. The major steps involved are explained in
Figure 3-1, and the geometry of hydrocyclone is shown in Figure 3-2. The underlying
flow physics was also explored by the finite element method.

3.1 Governing Equations
The Navier-Stokes equations describe the basic phenomena of mass and
momentum transport. These equations can also be used for turbulent flow simulations.
The Navier-Stokes equations are the basic equations describing movement in
incompressible fluids, which are an important class of fluid type. They are derived
simply by assuming that the momentum of the fluid must be conserved unless forces
are acting upon the fluid:

( ) ( ) F p u u u u
t
u
T
v
v v v v
v
= + + +

) ( Eq.3-1

0 = u
v
Eq.3-2

where denotes the dynamic viscosity, u
v
the velocity vector, the density of the fluid,
p the pressure, and F
v
is the volumetric force vector. With the governing equations, the
boundary conditions can be any of the ones described below. For an imposed velocity,
the velocity vector normal can be specified to the boundary by:
21
0
u n u =
v v
Eq.3-3

which is the Inflow/Outflow boundary condition. A certain pressure can instead be
imposed in the Outflow/Pressure boundary condition:


0
p p = Eq.3-4

There is special boundary condition that combines both these descriptions. This is
the Normal flow/Pressure or straight-out boundary condition, which sets the velocity
components in the tangential direction to zero and sets the pressure to a specific value.
It is available to simulate channels that are long in length and where flow is assumed to
have stabilized so that all velocity occurring in the tangential direction is negligible.
Imposing this condition removes the necessity of simulating this length.

0 = t u
v
v
Eq.3-5

0
p p = Eq.3-6

The Slip/Symmetry condition states that there are no velocity components
perpendiculars to a boundary:

0 = n u
v v
Eq.3-7

The Neutral boundary condition states that transport by shear stresses is zero
across a boundary. This boundary condition is denoted neutral, since it does not put any
constraints on the velocity and states that there are no interactions across the modeled
boundary.

[ ] 0 ) ( ( = + n u u
T
v v v
Eq.3-8


22
The No-slip boundary condition eliminates all components of the velocity vector:

0 = u
v
Eq.3-9

The laminar inflow/outflow boundary condition is appropriate for low Reynolds
number flow regimes, where a fully developed laminar profile is expected

( ) [ ]
inl
T
t t t inl
p n u u I p L
r v v
r
= + ) ( Eq.3-10

0 = u
t
v
Eq.3-11

In the equation above,
inl
L gives the length of the inlet channel and
inl
p the pressure at
the end of the inlet channel (outside of the model).
The equation for incompressible flow can be used for small changes in density
through thermal expansion. This is usually accomplished through the Boussinesq
approximation, which implies that gravitational forces can be accounted for explicitly in
the force term in the equations.
Fundamental to the analysis of fluid flow is the Reynolds number:

UL
= Re Eq.3-12

where U denotes the velocity field or the modulus of the velocity vector (m s
-1
), and L
denotes a representative length (m). The Reynolds number represents the ratio
between inertial and viscous forces. At low Reynolds numbers, viscous forces
dominate and tend to damp out all disturbances, which leads to laminar flow. At high
Reynolds number, the damping in the system is very low giving small disturbances
the possibility to increase. The nonlinearity of the equations forces different
disturbances to interact with each other. If the Reynolds number is high enough, the
fluid flow field eventually ends up in a state of statistical equilibrium called
turbulence.
23

FIGURE 3-1 Procedures in flow simulation by COMSOL Multiphysics

3.2 Procedures in Flow Simulation
The details of procedures in flow simulation as geometry of hydrocyclone,
material settings, boundary settings, mesh parameters, solve problem and post-
processing by COMSOL Multiphysics will be described in the following sections.

Material Settings
(Subdomain Settings)

Boundary Conditions
(Boundary Settings)



Geometry of Hydrocyclone
Mesh Parameters
Solve Problem
Post-processing
24
3.2.1 Geometry of the Hydrocyclone

FIGURE 3-2 Geometry of a hydrocyclone

Figure 3-3 shows geometry of the hydrocyclone that is shaped in COMSOL
Multiphysics by the details from Figure 3-2.

FIGURE 3-3 Geometry of a hydrocyclone in COMSOL Multiphysics
2.8
2.8
2
10
20
4
50
15
unit: mm
25
3.2.2 Material Settings (Subdomain Settings)
The material used in flow simulation is fermented broth that has physical
properties as shown in Table 3-1.

TABLE 3-1 Physical properties of fermented broth (Prongsantia, 2004)
Fermented Broth Value
Density (kg m
-3
) 1,086
Viscosity (Pa s) 0.028
Molecular Weight (kg kgmol
-1
) 18.0152
Thermal Conductivity (W m
-1
K
-1
) 0.6
Specific Heat (kJ kg
-1
K
-1
) 4.182
Temperature (K) 298

In the flow simulation, the two important properties used are density and viscosity.

3.2.3 Boundary Conditions (Boundary Settings)
Boundary conditions in flow simulation of the hydrocyclone which consists of
the feed inlet, the all walls, the overflow and the underflow outlets, and sub-domain
can be set into 4 boundary conditions as shown in Figure 3-4, 3-5, 3-6 and 3-7,
respectively.
The boundary condition at the feed inlet of the hydrocyclone is set to inflow
velocity as shown in Figure 3-4.
26

FIGURE 3-4 Boundary condition at the feed inlet

The boundary condition at the walls of the hydrocyclone is set to no slip as
shown in Figure 3-5.

FIGURE 3-5 Boundary condition at the walls

The boundary condition at the overflow and at the underflow outlets is set to
normal flow/pressure of 0 Pa as shown in Figure 3-6.
27

FIGURE 3-6 Boundary conditions at the overflow and the underflow outlets

The boundary condition between sub-domains is set to neutral as shown in
Figure 3-7.

FIGURE 3-7 Boundary condition between sub-domains

3.2.4 Mesh Parameters
A mesh is a partition of the geometry model into small units of simple shapes. A
Detail of mesh parameters is listed in Table 3-2 and a shape of mesh girds is shown in
Figure 3-8.

28
TABLE 3-2 Mesh Statistics
Parameters Value
Number of degrees of freedom 48922
Number of mesh points 1412
Number of elements 5899
Tetrahedral 5899
Prism 0
Hexahedral 0
Number of boundary elements 1771
Triangular 1771
Quadrilateral 0
Number of edge elements 333
Number of vertex elements 43
Minimum element quality 0.193
Element volume ratio 0


FIGURE 3-8 Mesh grids in COMSOL Multiphysics

29
3.2.5 Solve the Problem
After the first four steps of procedures in flow simulation are complete, the
problem can now be solved to get the answer.
3.2.6 Post-processing
In addition COMSOL Multiphysics has many tools for post-processing and
visualizing model quantities, which creates a wide variety of plots, i.e. surface plots,
slice plots, isosurface plots, contour plots, streamline plots, pathline plots that
describe particle traces, principal stress/strain plots and combinations of these plots.
Pathline plots mode is a significant tool that describes particle tracing in the
fluid. The equation describing particle tracing is called Khan and Richardson Force
that is available for particle tracing plots in the Incompressible Navier-Stokes
application mode and other application modes for fluid dynamics in the COMSOL
Multiphysics products.
The force expression that the software uses is derived partly using experimental
results and it is valid for a wide range of Reynolds numbers, stretching from creeping
flow toward the turbulent regime. The following equation describes the total force
that a fluid exerts on an immersed spherical particle:

( ) ( ) ( ) ( )
45 . 3
06 . 0 31 . 0 2 2
Re 293 . 0 Re 84 . 1
p p p p
u u r F + =

Eq.3-13

where the definition of the particle Reynolds number
p
Re is

( ) / 2 Re
p p p
r u u = Eq.3-14

As a result, a very small particle reported to the overflow, while a rather large
particle reported to the underflow, but the effective range of size separation was
smaller compared to the experiments. The orientation of a given force component (for
example, positive or negative x-component) is determined by the sign of the
corresponding component in the vector difference
p
u u , because this determines
whether the fluid is accelerating or slowing down the particle in that direction.
30
There is one parameter in this force expression: the particle radius, r
p
, which has
to be defined. Particle tracing plots can be separated into 2 types. First are particle
tracing plots for particles with mass and second are particle tracing plots for massless
particles.
For particles with mass, the software generates the pathlines by solving the
fundamental equation of motion

( ) x x t F x m & & , , = Eq.3-15

For x(t) being the pathline. Here, m is the particle mass, F equals the force
acting upon the particle, and t is time. This is a system of ODEs for x, which
COMSOL Multiphysics solves using a pair of Runge-Kutta methods of orders four
and five. The solver advances the algorithm with the solution of order five and uses
the difference between the order-five and order-four solutions to obtain the local error
estimate.
For massless particles, the equation of motion is

) , ( u t v x = & Eq.3-16

CHAPTER 4
RESULTS AND DISCUSSIONS

4.1 Hydrodynamic of Hydrocyclone
The flow field of the continuous phase fluid within a hydrocyclone is three
dimensional, and turbulent in nature. The flow analysis is further made more
complicated by the condition of multiphase components, in which, the different
treatments must be applied for different phases. In general, the turbulence is often
modeled by the Reynolds Stress Model (RSM), among other models, and the interface
between the liquid and gas (air core) can be modeled by the Volume of Fluid (VOF)
scheme. The results are then used in the simulation of particle trajectory described by
the stochastic Lagrangian model. For the present research, only single phase i.e., the
liquid phase was simulated because of the limitation of the COMSOL Multiphysics
3.3, in which the k model was available for turbulence modeling. However, this
software allowed the tracking of massless particles motion based on the Khan-
Richardson force model. Multiphase modeling were studied and discussed elsewhere
e.g., Brennan et al. (2007), and Wang and Yu. (2006).
In this section, the flow structures within a single Rietema hydrocyclone will be
verified prior to the simulation of the hydrocyclone network should be made.
4.1.1 Static Pressure
Although the present version of COMSOL was not equipped with the feature of
multiphase simulation, the evidence of air core characteristic was indirectly captured
by considering the static pressure distribution. Apparently from Figure 4-1, the
minimum or even negative values of static pressure occupy the central domain of the
hydrocyclone implying the existence of air core. At each axial height, the static
pressure is maximum at the wall and decreases toward the center core. The maximum
values of the static pressure tend to decrease from the inlet (i.e. near vortex finder) to
the exit (i.e. near spigot). Therefore, the particles wandering close to the wall of
hydrocyclone will passage helically downward, and report to the underflow port.

32

FIGURE 4-1 Pressure distribution within the hydrocyclone at different transverse
planes and at the central vertical plane

The pressure profile around the center of the hydrocyclone also shows the
relatively lower pressure with the decreasing height; as a result, suspended particles
residing near the central region will swirl vertically upward, and exit through the
overflow opening. Furthermore, the pressure difference between the hydrocyclone
wall and its axis decreases from the top cylindrical body to the bottom conical body.
The feed fluid accelerates and rather moves in the axial direction than in the radial
direction near the underflow, and so the lower pressure field occurs.


(a) (b) (c) (d)
FIGURE 4-2 Cell Reynolds pattern at the central vertical plane varies with the feed
velocity (a) 10 m/s, (b) 15 m/s, (c) 20 m/s and (d) 25 m/s
33
The presumed shape of the air core based on the static pressure will be more
noticeable from the Cell Reynolds pattern as shown in Figure 4-2. The air core is
more or less axisymmetric about the hydrocyclone axis, and occupies a large volume
along the center core. The degree of axisymmetry depends on the turbulence level of
the flow field, which in turn depends on the feed condition i.e., the size of the air core
reduces with increasing feed velocity, and the air core becomes more stable. The
origin of air core relates to the hydrodynamic instability of the boundary layer flow
over the concave wall of hydrocyclone. The proposed scenario of such phenomena
will be discussed in details in section 4.1.6.1.


(a)

(b)
FIGURE 4-3 Static pressure (a) below the vortex finder bottom end (z = 44 mm.)
and (b) underflow exit plane (z = 0 mm.)
34
From pressure difference illustrated in Figure 4-3, the slope of the contour
illustrates that the swirl and radial motions of the fluid persist even at the lower
portion of hydrocyclone. The steep slope can be noticed and the discontinuity at the
plateau of zero static pressure below which is expected to be the air core (Bhaskar,
2007). A large volume of the continuous phase fluid should be mainly classified
within the cylindrical and the upper conical portion of the hydrocyclone.
4.1.2 Axial Velocity and Vorticity
Another feature of counterflow is observed from the contour of vertical axial
velocity in the middle plane of the hydrocyclone as demonstrated in Figure 4-4. In
Figure 4-4, the negative downward axial velocity occurs in the vicinity of wall and
covers the large part of the hydrocyclone. The counterflow in the positive upward
direction consumes the core region. The interface between the two regions suggests
the locus of zero vertical velocity (LZVV). From the cross sectional view in Figure 4-
4(a), the loci of changing of flow direction can be estimated; from the center vertical
plane, the locus of zero vertical velocity (LZVV) is also apparent. A size of particle
whose path is in the vicinity of LZVV can approximately be considered as a particle
cutsize. The description of LVZZ dictates the separation performance of the
hydrocyclone via the motion of particles since the magnitude of axial velocity plays
an important role in determining the residence time i.e., the crease of downward
vertical velocity eventually reduces the residence time of the coarser particles and so
reduces the hydrocyclone cutsize (Bhaskar, 2007). Obviously, the value of the LZVV
when the air core present and absent should be different; thus the prediction of cutsize
could not be determined from the LZVV conclusively. Figure 4-4(b) exhibits the plot
of vertical velocity as a function of radial locations in various axial planes.
A better flow visualization can be obtained from by separately plot the positive
and negative axial velocities as shown in Figure 4-5(a) and 4-5(b), respectively. From
these plots, the thickness of the wall layer depends strongly upon the operating
condition e.g., the feed velocity. At high feed velocity, the boundary layer flow
becomes turbulent, and the oscillating air core is due to the fluctuation of the turbulent
boundary layer.


35

(a) (b)
FIGURE 4-4 Axial velocity: (a) contour at the central vertical plane and
(b) distributions along the radial positions and axial heights


(a) (b)
FIGURE 4-5 Axial velocity contour in the vertical central plane (a) positive (upward
vertical flow) and (b) negative (downward vertical flow)



36

(a) (b)
FIGURE 4-6 Magnified view of the flow near the hydrocyclone wall: (a) x-y plane
at z = 30 mm (b) center vertical plane in the connection region
between the cylindrical and conical parts

To investigate the flow field within the hydrocyclone more further, for inlet
velocity of 20 m/s, the velocity distribution at two cross sections i.e., very close to the
vortex finder and the underflow outlet are given in Figure 4-7(a) and 4-7(b),
respectively. As indicated, the cross sectional views of the magnitude of flow velocity
at different axial locations along the hydrocyclone, the counterrotating or parallel
swirling flow structures are recognized i.e., one is close to the wall and another
moving towards the overflow port. According to the high swirling but relatively low
velocity zone around the core region, a region of negative gauge pressure along the
hydrocyclone axis normally developed, and in fact, the central zone should represent
the actual air core formation (Narasimha, 2007). Indeed, the exact interface of air-
water within the hydrocyclone might need a multiphase modeling, which has gained
many attentions from researchers, and this should be the scope of a future research.
The axial vorticity
Z
as depicted in Figure 4-8(a) and 4-8(b) reveals the
clockwise and counterclockwise rotational fields at the same axial planes as in Figure
4-9 (feed velocity of 25 m/s). Evidently, by comparison Figures 4-8(a) and 4-8(b)
with Figures 4-9(a) and 4-9(b), the vortex strength increases with feed velocity on
both planes, and this indicates that the centrifugal force field and so the classification
of the continuous fluid phase from the suspended discrete phase should be promoted.
37
The asymmetry of the vortices in the plane near the vortex finder is apparent, and it is
due to the asymmetry of the inlet pattern of the hydrocyclone. The asymmetry
subsides at the lower axial height (toward the spigot opening), when the velocity
vector orients axially at the underflow port.


(a) (b)
FIGURE 4-7 Velocity vectors at feed velocity of 20 m/s: (a) in the plane close to
the bottom end of vortex finder (b) in the plane close to the
underflow opening


(a) (b)
FIGURE 4-8 Axial vorticity contour
Z
at feed velocity of 20 m/s (a) in a plane
close to the bottom end of vortex finder, (b) in a plane close to the
underflow opening

38

(a) (b)
FIGURE 4-9 Axial vorticity contour
Z
at feed velocity of 25 m/s (a) in a plane
close to the bottom end of vortex finder, (b) in a plane close to the
underflow opening.

4.1.3 Tangential Velocity
The tangential velocity is mainly responsible for inducing the centrifugal force
field within the hydrocyclone. In Figure 4-10, the tangential velocity profiles are
plotted against the axial locations. The simulation results indicate that the tangential
velocity increases in the radial direction starting from the hydrocyclone axis. The
maximum value of the tangential velocity is achieved at a certain location between the
core and the wall of hydrocyclone. A similar pattern persists also at other axial planes.
The present results are consistent with previous experiments and simulations reported
in the literatures (Slack and Wraith, 1997).
The absolute maximum value of the tangential velocity occurs in the cylindrical
part of hydrocyclone, whereas the minimum value appears in the conical part close to
the underflow outlet, and it is expected that the induced centrifugal force is weaker at
that location. Due to the relatively lower centrifugal force near the underflow, this
should allow the re-orientation of the entrapped fine particles towards the secondary
vertical upward flow in the core region, and progress to the overflow opening
(Cullivan, 2004).
39

(a) (b)
FIGURE 4-10 Tangential velocity distribution on the center-plane of hydrocyclone
(a) contour plot (b) value plot

Figure 4-11 compares the tangential velocity at various feed velocity. The sign
of the tangential velocity is positive everywhere i.e., fluid rotates in the same sense.
As the feed velocity is increased, the higher the feed velocity the lower the minimum
value of the tangential velocity. This could mean that part of input energy promotes
and strengthens the secondary vortex and the air-core.
Once again, the flow field and the induced force field depend strongly on the
feed condition. At low feed pressure, the centrifugal force field is rather weak,
therefore, the small and the large particle rotate almost in the identical orbit. When the
pressure is increased and so the centrifugal force, which is high enough to move the
large particle outward and the small particle inward (with liquid) in the radial
direction. Here in this case, the separation does occur. At higher feed pressure, the
force is extremely strong so that it can bring both the small and large particles towards
the wall. Thus, the separation performance declines. The separation phenomenon is
apparently reflected through the various information of flow structures presented so
far.

40

(a) (b) (c) (d)
FIGURE 4-11 Tangential velocity distribution on the central plane of hydrocyclone
at: (a) 10 m/s (b) 15 m/s (c) 25 m/s (d) 30 m/s


FIGURE 4-12 Velocity vector projected on the planes x-y at different height
of hydrocyclone

From the velocity vector plot projected on the plane x-y at different heights and
on the central plane (z-x) of the hydrocyclone, as shown in Figure 4-12 and 4-13,
respectively, both primary and secondary flow structures can be observed. The
velocity vector clearly shows upward vertical flow in the central portion and the
41
downward flow along the wall. The reverse flow between the outer and inner zones
can be identified from the vector presentation. In Figure 4-12 and 4-13, a recirculation
region exists beneath the inlet, and the reverse swirling flows are relatively more
intense near the vortex finder and spigot ends.

FIGURE 4-13 Velocity vector projected on the central plane (z-x) of hydrocyclone

4.1.4 Cell Reynolds
A strong evidence of the existence of the air core region can be depicted from
the Cell Reynolds plots. In Figure 4-14, the perspective views of the Cell Reynolds
indicate the low pressure zone in the core region in consistent with the fact that at
high feed velocity (and so the Cell Reynolds) the air core is induced. The cross
sectional views of the Cell Reynolds presented in Figure 4-14 also illustrate the
variation of the size of air core at feed velocity of 15 m/s. The Cell Reynolds varies
with the feed velocity and so expectedly the size of the air core.
In Figure 4-14(e) 4-14(h), the asymmetry becomes more appealing at the cross
section close to the vortex finder due to the location of the inlet. On the contrary, the
symmetric Cell Reynolds patterns can be observed in the conical portion of the
hydrocyclone as shown in Figure 4-14(a) 4-14(d). The asymmetry takes place near
the connection between the cylindrical and the conical portion. This implies that the
shape of air core depends on the proportion of both parts.
42

(a) (e)


(b) (f)


(c) (g)


(d) (h)
FIGURE 4-14 Cell Renolds Number at feed inlet velocity of 15 m/s (a) z = 6 mm
(b) z = 12 mm (c) z = 18 mm (d) z = 24 mm (e) z = 30 mm
(f) z = 36.2 mm (g) z = 42.4 mm (h) z = 48.6 mm

43
4.1.5 Water Split
In Table 4-1, the mass flux at and the water split into vortex finder are
determined from the simulated umbrella type velocity distribution at the spigot for
each feed velocity. Both the mass flux and the water split increases with the
increasing feed velocity. This observation is similar to previous experimental and
computational studies (Narasimha, 2005). The higher feed velocity strengthens the
centrifugal force and the secondary vortex, and results in a higher flow rate of the
continuous phase through the vortex finder. The difference between the simulated and
experimental values is within 2.26%.

TABLE 4-1 Comparison of predicted water splits with experimental value in
a 101-mm hydrocyclone (Narasimha, 2005)
Predicted values Experimental values Inlet
velocity,
m/s
Inlet flow
rate, kg/s
VF flow
rate, kg/s
Water split
into VF, %
Inlet flow
rate, kg/s
VF flow
rate, kg/s
Water split
into VF, %
06.10 1.903 1.876 98.58 1.896 1.865 98.36
09.50 2.963 2.934 99.02 2.966 2.940 99.12
11.35 3.540 3.512 99.21 3.868 3.850 99.53
12.35 3.850 3.820 99.22 3.868 3.850 99.53
10.70 3.337 3.309 99.16 3.378 3.325 98.43
08.25 2.574 2.544 98.83 2.604 2.563 98.42
05.91 1.844 1.816 98.48 1.937 1.894 97.78

4.1.6 Air Core
Besides the operating velocity, the appearance of the air core is affected by the
length and geometry of the vortex finder as investigated by Mainza et al. (2004). They
used the conventional hydrocyclone and three-product cyclones with short and long
inner vortex finders. From the velocity field vectors, they reported about the flow
reversals that occurred over the entire volume of both the conventional and the three-
product hydrocyclone, and a minimum velocity of zero was noticed in the air-phase
region for all the cases simulated. Their results are consistent with the observations of
44
Dabir and Petty (1984) that the geometry of the vortex finder affects the flow patterns
in the core region of the cyclone.


FIGURE 4-15 Typical air-core occur in the conventional and the commercial
hydrocyclones (Mainza, 2004)

From Mainza et al. (2004), Figure 4-15 shows the results indicated that the air-
core is present only at the central core of the inner vortex finder. Simulation results
indicated that the air-core in the vortex finder region was generally found to be
cylindrical in shape and along the central axis for both types of hydrocyclones. The
diameter of the air-core was found to be fairly even in the vortex finder region
From the air-core profiles shown in Figure 4-15, it can also be inferred that for a
fixed spigot diameter the air-core diameter and shape is depended upon the diameter
and length of the inner vortex finder for the three-product hydrocyclone, and on the
vortex finder diameter and length for the conventional hydrocyclone. They further
45
suggested that the presence of the air-core imposed a limit on the minimum inner
vortex finder diameter that can be used to maintain a reasonable area active for flow
in the inner vortex finder.
4.1.6.1 Onset of Air Core: Centrifugal Instability
The onset of the air core might due to the centrifugal instability originate close
to the underflow opening and propagate (convective instability) upward to the vortex
finder and overflow opening. The investigation of the onset of the air core based on
the hydrodynamics stability approach should be fully and carefully investigated.
The flow occurs along a curved surface of the inner wall of the hydrocyclone. In
this case of a velocity distribution which increases monotonously from the wall a
concave wall can be shown to give rise to instability in the form of streamwise
oriented counter rotating vortices. This happens if the so-called Gortler number is
high enough and such vortices are usually called Gortler vortices. The Gortler number
is defined as

2 / 1
0

=

R
U
G

Eq.4-1

where is the momentum thickness of the boundary layer.


FIGURE 4-16 Centrifugal instability of flow in a boundary layer on a concave wall
(Drazin and Reid, 1981)

46
From Drazin and Reid (1981), three assumptions were applied in deriving the
theory. First, it was assumed that the boundary-layer thickness is much smaller
than the radius of curvature of the wall
0
R , this being equivalent to the narrow gap
approximation. Second, it was assumed that the basic flow is nearly parallel to the
wall i.e., centrifugal effects are neglected in the description of basic flow, although
they are partially retained in the disturbance equations. Third, it was a local stability
analysis for which the basic flow is assumed to be independent of x and the y -
component of the basic flow is neglected.
The basic velocity then has the form ] 0 , 0 ), ( [ U where

y
= and 1 ) ( U
as . It was proved that the Rayleigh discriminant given by

2 2
3
) (
1
) ( = r
dr
d
r
r Eq.4-2

is proportional to
d
dU
U and instability is to be expected therefore locally, near the
wall, in a thin layer, the thickness of which is of order . Consider now the perturbed
flow with velocity ] , , [ w v u U + and suppose that it can be resolved into normal
modes of the form

ikz st
e p w v u p w v u
+
= ) , , , ( ] , , , [ Eq.4-3

where i s + = , and k is the wave number. On eliminating w and p from these
four equations in the standard form

Uu a v a D a D
2 2 2 2 2
) )( ( = Eq.4-4

and

v U u a D = ) (
2 2
Eq.4-5
47
where
d
d
D , k a = ,

2
s
= ,
0
2
2
R
U

=

and DU U = . The boundary
conditions are 0 = = = Dv v u at 0 = and as .
Hammerlin (1955) reported some results for the three representative velocity
profiles as shown in Figure 4-17. From Figure 4-17, it is evident that when the results
are expressed in terms of G they are not sensitive to the detailed form of the basic
velocity profile.

FIGURE 4-17 The critical values of the Gortler number for (1) the Blasius,
(2) the asymptotic suction and (3) the straight-line profiles,
respectively (Drazin and Reid, 1981)
4.1.5 Vorticity
The swirling flow is investigated from the axial vorticity contour
Z
as
previously shown in Figure 4-8. However, the flow field within hydrocyclone is fully
three-dimensional, so the rotational flow in the plane that perpendicular to the vertical
plane is also studied. Figure 4-18 illustrates the circumferential vorticity indicating the
twisting of the vortex filaments. From the vorticity contours shown, the helical motion
of the vortex filaments is intrigue. The elongated shape of different sign vortices
occupies the large portion of the cylindrical body of the hydrocyclone. The vortex
pairs have their top ends at the bottom end of the vortex finder. A pair of vortices can
also be seen near the spigot outlet, however, they are more rounded in shape but lower
in strength.
48

(a) (b)
FIGURE 4-18 Circumferential vorticity contour

at feed velocity of 20 m/s


(a) in the center-plane at = 90 and (b) in the offset plane


FIGURE 4-19 Turbulent kinetic energy at a 45-plane at feed velocity of 15 m/s

49

(a) (b)

(c) (d)

(e) (f)

(g) (h)
FIGURE 4-20 Turbulent kinetic energy in a plane = 0 at feed velocity of
15 m/s: (a) z = 48.6mm, (b) z = 42.4 mm, (c) z = 36.2 mm,
(d) z = 30 mm, (e) z = 24 mm, (f) z = 18 mm, (g) z = 12 mm and
(h) z = 0 mm
50

(a) (b)

(c) (d)
FIGURE 4-21 Turbulent kinetic energy in a plane z = 30 mm at feed velocity of
15 m/s (a) = 0, (b) = 45, (c) = 90 and (d) = 135

4.1.6 Turbulent Kinetic Energy
The asymmetry also reflects through the turbulence kinetic energy as reported in
Figure 4-19, 4-20 and 4-21. As mentioned earlier, the asymmetry rises from the
asymmetry at the hydrocyclone inlet. The turbulence kinetic energy is approximately
5 times higher relatively the turbulence kinetic energy of other locations. Besides the
value of turbulence kinetic energy close to the inlet, a general distribution is exhibit
rather symmetry about axis of hydrocyclone.

4.2 Fluid-Particle Interaction
The turbulent flow within the hydrocyclone comprises various scale flow
structures, which in turn, influence the performance of hydrocyclone. Many recent
studies paid attention to developing models capable of simulating the flow field, water
splits, commencement of the air core, and separation efficiency curves, among other
characteristics, of various models of hydrocyclone. The centrifugal force filed
51
developed from the intrinsic swirling flow play a key role in separation of particles,
and density or size classification of particles in slurries and particle laden flows.
Herein, the available turbulent models and the induced forces on particles are
considered prior to the particle tracking (trajectory) can be investigated.
4.2.1 Choice of the Turbulent Models
Generally, to compute the turbulent stresses and other unknown second
moments occurring in the governing equations for the mean flow of a turbulent flow,
the linear eddy viscosity model, and simultaneously solving transport equations and
equation both provide the time- and length-scale of turbulence. In isotropic
turbulence, the dynamical equation of the energy spectrum ) (k E is (Batchelor, 1953)

) ( 2 ) ( ) (
2
k E k k T k E
t
=

Eq.4-6

where ) (k T is the spectral transfer term, which has a role of redistributing energy in
wave number space i.e.,

0 ) (
0

= dk k T Eq.4-7

In term of ) (k E , the turbulent kinetic energy and the dissipation rate can be described
as

=
0
) ( dk k E K Eq.4-8

and

dk k E k

=
0
2
) ( 2 Eq.4-9

52
In practice, the regions near solid walls imposes challenge on the modeling
since the structure of the turbulence is highly anisotropic and strongly influenced by
the presence of the wall (Hallback et. al., 1996). Due to the enhanced mixing of the
turbulence, away from solid walls, the mean velocity gradients are moderate whereas
in the near-wall region strong mean velocity gradient are formed. As a consequence,
the associated strong turbulent production and high turbulence levels in the near wall
region can occur. Furthermore, the no-slip condition at the wall causes virtually all
turbulence parameters to exhibit very steep gradients in the vicinity of the wall.
In global scenario, there is a disparity of turbulence scales in the near-wall
region where the viscosity strongly influence, and away from the wall region where
the viscosity is secondary importance.
Since the turbulent flow inside a hydrocyclone is anisotropic, and the Reynolds
stress model (RSM) is claimed to predicted turbulence swirling flow within the
hydrocylone with better accuracy (Slack et al., 2003; Slack and Wraith, 1997;
Suasnabar, 2000; Cullivan et al., 2003; Delgadillo and Rajamani, 2005). Nevertheless,
for the fermented broth system, it might be considered as sufficiently dilute flow with
mono-dispersed suspended particles (generally, concentration below 10% by weight)
orbiting considerably not so close to the hydrocyclone wall, and the effects of the
discrete particle phase are negligible. Therefore, for the present study the k model
is available for turbulence modeling, and Khan-Richardson force model is provided
for particle tracing i.e.,

( ) ( )
45 3
06 0 31 0 2 2
293 0 84 1
.
.
p
.
p p p
Re . Re . u u r F + =

Eq.4-10

where
p
Re is equal to / 2
p p
r u u .
4.2.2 Flow-Induced Forces on Particle
The Khan-Richardson force model is provided as a function of
p
Re for particle
tracing complying with the stochastic Lagrangian multiphase flow model. The
momentum is transferred between the continuous and discrete phases through force
interaction only i.e., drag and lift with no mass transfer involved in this case. In real
flow, many factors influence the drag and lift on the particle especially under the
53
unsteady, three-dimensional flow. Nonetheless, in a Lagrangian reference frame, the
buoyancy force and liquid drag force on particles are calculated as

P
P
p D
P
g
u u F
dt
u d

) (
) (

+ =
r
r r
r
Eq.4-11

where ) (
p D
u u F
r r
is the drag force per unit particle mass and
24
Re 18
2
P
D
P P
D
C
d
F

= Eq.4-12

Here,
p
u
r
is the particle velocityu
r
is the fluid phase velocity,
P
is the density of
the particle, and
P
d is the diameter of the particle.
P
Re is the relative Reynolds
number and
D
C is the drag coefficient.
4.2.3 Particle Trajectory
In this research, the massless particle was used to investigate the trajectory, and
the resulting particle trajectory depends strongly upon the injected location consistent
with the observations of Wang et. al. (2007). They used a high-speed motion analyzer
(HSMA) system to track the trajectory of solid particles fed into the hydrocyclone
with Rietemas optimum geometry, and studied the stochastic and statistical behaviors
of the motion of solid particles by adopting a Lagrange method.
From Figure 4-22 to Figure 4-24, it is evident that the initial feed location
dictate the particle path. A particle may orbit only in the top portion of the cylindrical
part of hydrocyclone prior to exiting from the overflow opening or wander from the
inlet following the path and exit through the underflow opening. The particle
trajectory relates to the residence time of the particle inside the hydrocyclone which in
turn affects the overall performance of the hydrocyclone. Figure 4-25 illustrates the
particle injection locations in which the red, blue and black spots indicate that the
particle will discharge from the overflow, underflow or unspecified (software
limitation), respectively.
54

(a) (b) (c) (d) (e)
FIGURE 4-22 Pathline trajectories of particles with different feed velocities
that discharging from underflow: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s


(a) (b) (c) (d) (e)
FIGURE 4-23 Pathline trajectories of particles with different feed velocities
that discharging from overflow: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s
55

(a) (b) (c) (d) (e)
FIGURE 4-24 Pathline trajectories of particles with different feed velocities
that discharging unspecification: (a) 10 m/s, (b) 15 m/s,
(c) 20 m/s, (d) 25 m/s and (e) 30 m/s

From Figure 4-25, the percentage of particle entering underflow, overflow and
neither underflow nor overflow can be listed in Table 4-2. Table 4-3 listed the
residence time of particles discharged from underflow with different feed inlet
velocities. With increasing the feed inlet velocity, both the longest residence time and
shortest residence time of particle within the hydrocyclone decreased. With increasing
feed inlet velocity, the separation process finished faster. However, a particle with
feed inlet velocity of 30 m/s held the shortest residence time. Table 4-4 listed the
residence time of particle discharged from overflow with different feed inlet
velocities. With increasing the feed inlet velocity, both the longest residence time and
the shortest residence time of particle within the hydrocyclone decreased.
56

(a) (b)


(c) (d)


(e) (f)

FIGURE 4-25 Feed inlet position of particles at feed velocity of: (a) 10 m/s
(b) 15 m/s, (c) 20 m/s, (d) 25 m/s and (e) 30 m/s

* All particles are massless
Particle entering overflow
Particle entering underflow
Particle unspecified outlet
57
TABLE 4-2 Percentage of particle entering underflow and overflow when they were
injected into the hydrocyclone at different positions
Particle percentage, % Injected velocity,
m/s Underflow Overflow Unspecification
10 41.21 15.88 42.91
15 37.62 18.71 43.67
20 49.53 13.23 37.24
25 47.64 14.18 38.18
30 47.45 13.04 39.51

TABLE 4-3 Residence time of particle inside the hydrocyclone that injected into the
hydrocyclone at different positions and finally discharging from
underflow
Residence time of particles inside the hydrocyclone, s Injected velocity,
m/s Longest time Shortest time Average time
10
0.138512 0.080033 0.109421
15
0.102405 0.055881 0.079143
20
0.052153 0.038447 0.045300
25
0.049754 0.033180 0.041467
30
0.040013 0.027485 0.033749

TABLE 4-4 Residence time of particle inside the hydrocyclone that injected into the
hydrocyclone at different positions and finally discharging from
overflow
Residence time of particles inside the hydrocyclone, s Injected velocity,
m/s Longest time Shortest time Average time
10
0.114684 0.042352 0.078518
15
0.071420 0.023167 0.0472935
20
0.069743 0.019716 0.0447295
25
0.056485 0.012852 0.0346685
30
0.043622 0.012518 0.0280700
58
4.3 Hydrocyclone Network in Series Connection
The series connections of hydrocyclones are a common way of improving the
performance of single units. In the case of hydrocyclones, due to their low capital and
running costs, multiple series arrangements are quite frequently used (Svarovsky,
1994). In this study, the hydrocyclone network in series in the direction of the
overflow is studied, such as the arrangement shown in Figure 4-26. In Figure 4-26,
the overflow from the first or subsequent hydrocyclone is fully fed to the feed channel
of the next hydrocyclone.


FIGURE 4-26 Three hydrocyclones in series

In Figure 4-27, the feed pressure of 1
st
hydrocyclone was 30 bar. Pressure
distribution inside the 1
st
-, 2
nd
- and 3
rd
hydrocyclones is similar to Figure 4-1 except
that pressure at the overflow of the first two hydrocyclone is not equal to atmospheric
pressure because their overflow are connected with the feed inlet of the next
hydrocyclone in network. Pressure drop in a hydrocyclone reflects the energy
necessary for a separation process. An each hydrocyclone in series has its own
pressure drop, as a result, inlet pressure of 2
nd
hydrocyclone has decreased. The inlet
pressure of 2
nd
hydrocyclone came from pressure of overflow of 1
st
hydrocyclone but
both pressure values are not identical because of pressure drop in piping and fitting
between the 1
st
- and 2
nd
hydrocyclones. This effect also occurred in the 3
rd

hydrocyclone. Pressure distribution of three hydrocyclones obtained from simulation
is listed in Table 4-5.
59

FIGURE 4-27 Pressure distribution of three hydrocyclones in series

TABLE 4-5 Pressure at inlet, overflow and underflow of three hydrocyclone in series
Pressure (Bar) Hydrocyclone No.
Feed inlet Overflow Underflow
1 49.464 17.078 27.078
2 5.781 0.773 1.325
3 0.226 4.5410
-5
0.023

4.4 Hydrocyclone Network in Parallel Connection
Another arrangement of hydrocyclone network is the parallel connection as,
shown in Figure 4-28, the six hydrocyclones are connected in two parallel
connections in a direction of the feed inlet. An each parallel connection contains three
hydrocyclone. In Figure 4-28. This kind of arrangement gains an increase in capacity
of hydrocyclones.

60

FIGURE 4-28 Six hydrocyclones in parallel

The fermented broth at feed pressure of 30 bar is fed through a feed pipe of
hydrocyclone network. The pressure distribution inside the feed pipe of hydrocyclone
network is simulated in COMSOL Multiphysics without six hydrocyclones connected
with feed pipe. The feed pressure of six hydrocyclones is obtained from the
simulation of feed pipe of hydrocyclone network as shown in Figure 4-29. In Figure
4-29, it shows that pressure distribution of two parallel connections in main pipe of
hydrocyclone network has decreased along the length of feed pipe except that
pressure at the end of feed pipe that is not the smallest pressure in feed pipe, resulting
from backward pressure in the end of feed pipe.


FIGURE 4-29 Pressure distribution in main pipe of hydrocyclone network
in parallel
61
At feed pipe, it contains six reduction feed pipes that are six hydrocyclones
connected with. From simulated result of pressure distribution in feed pipe, pressure
at reduction feed pipe is the feed pressure of hydrocyclone. An each feed pressure is
brought to simulate one by one in COMSOL Multiphysics. The results of pressure of
six hydrocyclones in two parallel connections are listed in Table 4-6. In Table 4-6, it
shows that pressure at feed inlet of six hydrocyclone is proximity as well as pressure
at overflow and underflow.

TABLE 4-6 Pressure at inlet, overflow and underflow of hydrocyclone in parallel
Pressure (Bar) Hydrocyclone No.
Feed inlet Overflow Underflow
1 10.13 0.010 6.309
2 10.82 0.011 6.735
3 09.33 0.009 5.815
4 09.72 0.010 6.052
5 09.90 0.010 6.167
6 10.52 0.011 6.552

CHAPTER 5
CONCLUSIONS

5.1 Conclusions
The hydrocyclone for separating yeast and calcium in ethanol production has
been modelled using computational fluid dynamics as COMSOL MULTIPHYSICS
3.3. The turbulence terms of the Navier-Stokes equations are calculated on the basis
of k- model. This research is presented the applicability in the analysis of the
hydrodynamic of hydrocyclone, fluid-particle interaction and hydrocyclone network
in a series connection and in a parallel connection
In the hydrodynamic of hydrocyclone result, it shows that an increase in feed
velocity of fermented broth has increased the axial velocities in both upward and
downward directions and also increased the static pressure differential along the radial
positions inside the hydrocyclone (Bhaskar, 2007). The measured turbulent kinetic
energy was high in a plane (x-y) of the feed channel. The radial profiles of the
turbulent kinetic energy had local maxima near the core, the wall and in the annulus
(Murphy, 2007).
In the fluid-particle interaction result, it shows that the initial position of
particles at the feed channel of hydrocyclone strongly affects the motion trajectory of
particles inside hydrocyclone and, as a result, the separation efficiency of particles in
the hydrocyclone was heavily influenced by the initial position of the particle at the
feed channel of hydrocyclone.
In the hydrocyclone network, it is very difficult to simulate hydrocyclone
network in parallel connection. It has to be separated simulation in two parts, feed
pipe and individual hydrocyclone. Hydrocyclone network in series is advantageous to
increase recovery and hydrocyclone network in parallel connection is useful to
increase capacity.
Finally, the simulated results in this study are of great worth for understanding
the behaviors of flow field and particle motion in the separation process inside the
hydrocyclones
63
5.2 Suggestions
A CFD simulation needs a hi-specification computer to solve a problem. In
present work, an error of a solution occurred might result from mesh grid that cannot
be created to fine size. An amount of computers used in this work is inadequate to
work.
REFERENCES

Bhaskar, U.K., et al. CFD simulation and experimental validation studies on
hydrocyclone. Minerals Engineering. 20 (2007) : 60-71.
_______. CFD validation for flyash particle classification in hydrocyclones.
Minerals Engineering. 20 (2007) : 290-302.
Brennan, M.S., et al. Multiphase modeling of hydrocyclones-prediction of cut-size.
Minerals Engineering. 20 (2007) : 395-406.
Castilho, L.R., and Medronho, R.A. A sample procedure for design and
performance prediction of Bradley and Rietema hydrocyclone. Minerals
Engineering. 13 (2000) : 183-191.
Cilliers, J.J., and Diaz-Anadon, L. Temperature, classification and dewatering in 10
mm hydrocyclones. Minerals Engineering. 17 (2004) : 591-597.
Cilliers, J.J., and Harrison, S.T.L. The application of mini-hydrocyclone in the
concentration of yeast suspensions. Chemical Engineering Journal. 65
(1997) : 21-26.
Coelho, M.A.Z., and Medronho, R.A. A model for performance prediction of
hydrocyclones. Chemical Engineering Journal. 84 (2001) : 7-14.
Cullivan, J.C., et al. New understanding of a hydrocyclone flow field and separation
mechanism from computational fluid dynamics. Mineral Engineering. 17
(2004) : 651-660.
Delgadillo, J.A., and Rajamani, R. K. Exploration of hydrocyclone design using
computational fluid dynamics. Int J. Miner. Process. (2006).
_______. A comparative study of three turbulence-closure models for the
hydrocyclone problem. Int J. Miner. Process. 77 (2005) : 217-230.
Doby, M.J., et al. Numerical and experimental examination of swirl flow in a
cylindrical container with rotating lids. Mineral Engineering. (2007).
Mainza, M., et al. Study of flow behaviour in a three-product cyclone using
computational fluid dynamics. Minerals Engineering. 19 (2006) : 1048-
1058.
65
Murphy, S., et al. Prediction of strongly swirling flow within an axial hydrocyclone
using two commercial CFD codes. Chemical Engineering Science. 62
(2007) : 1619-1635.
Narasimha M., et al. CFD modeling of hydrocyclone-prediction of cut size. Int. J.
Miner. Process. 75 (2005) : 53-68.
_______. Prediction of magnetite segregation in dense medium cyclone using
computational fluid dynamics technique. Int. J. Miner. Process. 82 (2007) :
41-56.
_______. Large eddy simulation of hydrocyclone-prediction of air-core diameter
and shape. Int. J. Miner. Process. 80 (2006) : 1-14.
Neesse, Th., and Dueck, J. Dynamic modeling of the hydrocyclone. Mineral
Engineering. 20 (2007) : 380-386.
Prongsantia, S., et al.. Prediction of the optimum flow patterns in a mini-
hydrocyclone for yeast recovery application by using computational fluid
dynamics technique. In The Joint International Conference on Sustainable
Energy and Environment (SEE). Thailand, 2004 : 576-580.
Slack, M.D., and Wraith, A.E. Modelling the velocity distribution in a
hydrocyclone. In 4th International Colloquium on Process Simulation.,
Finland, 1997 : 6583.
Slack, M.D., et al. Designing automated computational fluid dynamics modeling
tools for hydrocyclone design. Minerals Engineering. 17 (2004) : 705-711.
Svarovsky L., Hydrocyclones. New York : Holt, Reinehart and Winston, 1984.
Wang, B., and Yu, A.B. Numerical study of particle-fluid flow in hydrocyclones
with different body dimensions. Minerals Engineering. 19 (2006) : 1022-
1033.
Yang, I.H., et al. A three-dimensional simulation of a hydrocyclone for the sludge
separation in water purifying plants and comparison with experimental data.
Minerals Engineering. 17 (2004) : 637-641.



66

BIOGRAPHY

Name : Mr.Suchain Tongsiri
Thesis Title : The Simulation of Hydrocyclone Network for Separating Yeast and
Calcium in Ethanol Production
Major Field : Chemical and Process Engineering

Biography
I was born on November 15, 1982. I graduated Bachelor degree in Chemical
Engineering, Faculty of Engineering, King Mongkuts University of Technology
North Bangkok (KMUTNB) in 2004 and graduated Master of Science in Chemical
and Process Engineering, The Sirindhorn International Thai-German Graduate School
of Engineering, KMUTNB.
A contact address is Division of Chemical and Process Engineering, The
Sirindhorn International Thai-German Graduate School of Engineering, King
Mongkuts University of Technology North Bangkok, 1518 Pibulsongkram Road,
Bangsue, Bangkok 10800, THAILAND.

S-ar putea să vă placă și