You are on page 1of 262

University of Wollongong

Research Online
University of Wollongong Thesis Collection University of Wollongong Thesis Collections

1997

Velocity distribution of single phase fluid flow in


packed beds
Subagyo
University of Wollongong

Recommended Citation
Subagyo, Velocity distribution of single phase fluid flow in packed beds, Doctor of Philosophy thesis, Department of Materials
Engineering, University of Wollongong, 1997. http://ro.uow.edu.au/theses/1537

Research Online is the open access institutional repository for the


University of Wollongong. For further information contact Manager
Repository Services: morgan@uow.edu.au.
VELOCITY DISTRIBUTION
OF SINGLE PHASE FLUID FLOW IN
PACKED BEDS
UNIVERSITY <
4GOMC
WOILONGOMC
U8RAF
_X I
A thesis submitted in fulfilment of the requirements
for the award of the degree of

Doctor of Philosophy

from
University of Wollongong

by
SUBAGYO
(Ir., U G M Yogyakarta)

Department of Materials Engineering


July 1997
Gusti Allah Mboten Sare

(God never sleeps)

n
DECLARATION

The work presented in this thesis is, to the best of m y knowledge and

belief, original except as acknowledged. I hereby declare that I have not

submitted this material, in whole or in part, for a degree at this or any

other university.

in
ACKNOWLEDGMENTS

I would like to express my sincere appreciation and gratitude to all who


contributed to this thesis with their experience, expertise and support. In
particular I a m deeply grateful to m y supervisor Professor Nick Standish
for his invaluable guidance, deep enthusiasm and encouragement
throughout the course of this project. Sincere appreciation is extended to
m y co-supervisors Dr. G.A. Brooks and Professor R J . Dippenaar for their
invaluable guidance, useful suggestions and encouragement throughout
this research work. I also thank Dr. S. Nightingale for providing a
computer facility.

I am greatly indebted to the board of management of PT Krakatau Steel,


Cilegon, Indonesia for giving opportunity and financial support, which
allowed this research to be carried out. Grateful acknowledgment is also
extended to ICMINET-ICMI, Jakarta, Indonesia for providing formative
financial support.

My sincere thanks are extended to the workshop and laboratory staff of


the Department of Materials Engineering, especially G. Hamilton and R.
Kinnell, for their assistance with various aspects of the research. Thanks
are also extended to m y student colleagues: D. Muljono, S. Street, W .
Setiadharmaji, D. Phelan, D. Rosawinarti, J. Jones, Tridjaka, Supramu, E.
Rukman, S. S o e d o m o and N. Ross for their constant encouragement and
support during the academic years. Grateful acknowledgment is also
extended to J. Giarini and A. Moerwanto for their invaluable help.

Last but not least, I would like to express my deepest gratitude to my


parents, Hardjo Suwito and Suharti, for their patience, constant support
and encouragement, which have helped m e to keep going throughout.

IV
ABSTRACT

A detailed knowledge of the flow distribution is required for further stu

of rate processes and their mechanism taking place in the packed beds. A

better understanding of the rate processes is essential for proper proces

design and process optimisation of packed bed systems to maximise the

comparative advantage and safety factor in industrial operations.

A new mathematical model of velocity distribution of single phase fluid

flow in packed beds was developed by assuming the flow characteristic is

a combination of a continuous and a discontinuous systems of fluids

between voids in the bed. In order to allow a comparison with data

measured at the downstream of the bed, the model was completed by a

new mathematical model of a developing flow profile in an empty pipe.

The model can be applied for both compressible and incompressible fluid.

The validity of the model has been checked using previous experimental

data and new measurement results. The agreement between measured

data and results predicted by the mathematical model is good. The new

model favourably compares with previous models in terms of accuracy

and its simplicity does not require new empirical constants.


It is clearly demonstrated that the fluid flow distribution in a packed bed is

influenced by the Reynolds number and the bed characteristics. However,

when the Reynolds number is higher than 500, the flow profile is mostly

determined by the bed characteristics. Moreover, it is also demonstrated

that the disagreement of previous investigators in the effect of Reynolds

number and particle diameter on fluid flow distribution in packed beds is

mostly due to limitation experimental.

Similar to the macroscopic view of fluid flow in packed beds, it can be

shown at the microscopic level that models based on discontinuous

systems are only successful for local porosity less than 0.5. In conditions

when the local porosity is higher than 0.5, especially in the vicinity of the

wall, a model based on the continuous systems approach, as used in the

present case, is more accurate.

The restriction for the flat flow profile assumption for packed bed systems

was investigated by using the present model. It is clearly shown that the

deviation of flat profile condition not only depends on the D/DP ratio but

also depends on the Reynolds number. The deviation of flat flow profile

condition was also used to investigate the possibility of generating a

distorted physical model in terms of D/DP ratio and L/D ratio.

VI
CONTENTS

DECLARATION iii

ACKNOWLEDGMENTS iv

ABSTRACT v

CONTENTS vii

LIST OF SYMBOLS ix

1 INTRODUCTION 1

2 LITERATURE SURVEY 7
2.1 THE EXPERIMENTAL W O R K S 7
2.2 MATHEMATICAL MODELLING 33
2.2.1 The Phenomenological Approach 34
2.2.2 The Theoretical Approach 46

3 CHARACTERISATION OF A PACKED BED 76


3.1 PARTICLE SIZE 77
3.2 THE BED POROSITY 79
3.2.1 Mean Bed Porosity 80
3.2.2 Radial Distribution of the Bed Porosity 94
3.3 PERMEABILITY 119

4 DEVELOPMENT OF A MATHEMATICAL MODEL FOR 125


VELOCITY PROFILE OF FLUID FLOWING IN PACKED
BEDS
4.1 THE EQUATION OF FLOW THROUGH A SINGLE PIPE 126
4.2 THE EQUATION OF FLOW THROUGH A PACKED BED 130
4.2.1 The Equation of Continuity 135
4.2.2 The Incompressible Fluid 135

VII
4.2.3 The Compressible Fluid 140
4.2.4 Pressure Drop Correlation for Packed Beds 141
4.3 FLUID F L O W AT THE OUTLET O F THE BEDS 143

5 EXPERIMENTAL TECHNIQUES 157


5.1 EXPERIMENTAL APPARATUS A N D MATERIALS 157
5.2 EXPERIMENTAL P R O C E D U R E 161

6 EXPERIMENTAL RESULTS AND MATHEMATICAL 166


M O D E L VERIFICATION
6.1 EXPERIMENTAL RESULTS 167
6.1.1 Reproducibility of Data 170
6.1.2 Measurement Results of Velocity Profile 170
6.2 MATHEMATICAL M O D E L VERIFICATION 175
6.2.1 Validation of the Mathematical Model 175
6.2.1.1 Uni-Sized Particle Packed Beds 178
6.2.1.2 Multi-Sized Particle Packed Beds 190
6.2.2 Comparison of the Mathematical Models of the Velocity 197
Profile in Packed Beds

7 DISCUSSION 205
7.1 C O M P U T E D VELOCITY PROFILES 205
7.2 SIMILARITY CRITERIA OF PACKED BED SYSTEMS 220

8 CONCLUSIONS 226

REFERENCES 228

APPENDIX 241
ALGORITHMS OF VELOCITY PROFILE CALCULATION 241

VI
LIST OF SYMBOLS

A = area {L~
A = constant defined by equation (4-66)

Am = constant defined by equation (2-41)


At = transition region area {L2
Aw = wall region area {L2
Az = constant defined by equation (2-93)
Ao = constant defined by equation (2-23)
Ai = constant defined by equation (2-24)
A2 = constant defined by equation (2-25)
AE = constant defined by equation (2-91)
a = constant defined by equation (2-28)
a = constant defined by equation (3-34) or (3-35)
a = aspect ratio
at - constant in equation (4-44) {ML"Y
a, = constant in equations (6-1) and (6-3)
av = surface are per unit volume of particle {L~
B = constant defined by equation (4-67)

Bp = function defined by equation (2-71) {ML"Y


Bm = constant defined by equation (2-42)
Bm = constant defined by equation (2-42)
Be = constant defined by equation (2-89)
Bz = constant in equation (2-94)
B(n) = constant in equation (2-57)
Bb(n) = constant in equation (2-61)
Bt(n) = constant in equation (2-63)
B w (n) = constant in equation (2-66)

IX
b = constant defined by equation (2-6) {-}
b = constant defined by equation (3-37) {-}
b{ = constant in equation (4-44) {ML" 4 }
bi = constant in equations (6-1) and (6-3) {-}
h = constant defined by equation (3-49) {-}
C = constant defined by equation (4-68) {-}

*—m = constant defined by equation (2-43) {-}


C = constant defined by equation (2-8) {-}
Cj = constant in equations (6-1) and (6-3) {-}
C\ = calculation result of ith sample H
c' = constant defined by equation (3-24) {-}
D = column diameter (L)
De = effective channel diameter (L)
DP = particle diameter {L}
Dpe = equivalent packing diameter {L}
D Pi = particle diameter of ith component {L}
Dpm = m e a n size diameter of multi-sized particles {L}
Dpv = equivalent volume diameter {L}
L-'vs = volume-surface m e a n particles diameter (L)
di = constant in equations (6-1) and (6-3) {-}
di = data of ith sample {-}

d = sample m e a n {-)

E = constant defined by equation (2-22) {-}


F = force {MLt"2}
Fh = friction head {L}
Fk = friction force {MLf 2 }
Fi = first randomising factor defined by equation (3- {-}
41)
F2 = second randomising factor defined by equation {-)
(3-42)
f = constant defined by equation (2-82)
fk = constant defined by equation (4-21)
fi = constant defined by equation (2-47)
f2 = constant defined by equation (2-48)
Q = constant defined by equation (3-18)

g = gravitational constant

gi = constant defined by equation (3-46)

h = constant defined by equation (3-50)


I = inertia parameter of bed

Io =- modified Bessel function of first kind,


Ic = constant defined by equation (3-14)
Id = constant defined by equation (3-13)
Jo = Bessel function of first kind, order 0

K = constant defined by equation (2-18)


Ko = constant in equation (2-55)
k = coordination number
k' = constant defined by equation (2-4)
ki = constant in equation (2-1)
k2 = constant in equation (2-1)
k0 = constant defined by equation (4-39)
ki = constant defined by equation (4-40)
k2 = constant defined by equation (4-41)
4 = constant defined by equation (5-1)
fi2 = constant defined by equation (5-1)
L = bed length

Le = equivalent length
LM = Prandtl's mixing length
Liu = inlet length
1 = radial position in the bed that has porosity {L}
equal to 0.5
k = effective path length {L}
lw = loss work {ML2."2}

M = constant defined by equation (4-16) {-}


m = mass {M}

N = transformation function as given by equation {MLY 3 }

(2-53)
o¥c = digital computer number [-:

Nj = number of spheres with centres lying in the fh {-}


cylindrical concentric layer
NRe = Reynolds number {-]
n = number of segments {-
nc = constant in equation (2-27) {-]
nd = number of data {-
n, = particles number of ith component {-
P = dimensionless pressure defined by equation [-
(4-61)
Ps = pressure defect per unit length {ML"Y2

P = pressure {ML"Y2

P = particular layer of particles {-


Q = flow rate {LV

Q. = local flow rate {LY1


= heat {MLY 2
q
R = column radius {L

X
SH = residual defined by equation (4-38) {Lf1

/? = reproducibility {-

#" = average error {-

9. = residual defined by equation (4-72) {-

Re equivalent radius {L
radial position (L
distance from wall in particle diameters {-
r radius of outer edge of central core {L
r constant defined by equation (3-38) {-
TH hydraulic radius {L
TM constant defined by equation (2-26) {-
rc starting point radial position of the core region {L
re equivalent radial position (L
rm radial position of m a x i m u m superficial velocity {L
S entropy {MLY2T'
S dimensionless axial position defined by {-
equation (4-60)
s constant in equation (2-21) {-
T temperature {T
U internal energy {ML 2 f 2
7i dimensionless radial velocity defined by {-
equation (4-54)
uM cross section average of superficial velocity {Lf'
u velocity {Lf1
uz local velocity {Lf1
Ub superficial velocity at the bulk region {Lf1
u, local velocity at t {Lf1

Urn local superficial velocity {Lf1

XIII
u = local superficial velocity at the edge of central {Lf }
core
ut = superficial velocity at the transition region {Lf'}
uw = superficial velocity at the wall region {Lf'}
umb = superficial velocity at bypass section {Lf'}
umc = superficial velocity at core section {Lt"1}
Umr = radial direction of superficial velocity {Lf1}
umz = axial direction of superficial velocity {Lf1}
um0 = superficial velocity at the centre of the bed {Lf'}
uzo(r) = velocity profile at the top (outlet) of the bed {Lf1}
UZLW = fully developed flow profile in the empty pipe {Lf1}
V = volume {L }
V = dimensionless axial velocity defined by {-}
equation (4-53)
Vj = initial specific volume of ith component {-}
-p = dimensionless velocity defined by equation (4- {-}
64)
-t i

V = specific volume {L M }
V. = total volume of solid in the ith cylindrical {L3}
concentric layer
Vy = volume of the solid in the ith cylindrical {V}
concentric layer due to a sphere with centre in
the jth cylindrical concentric layer
W = constant defined by equation (2-19) {-}
Xb = starting point of bulk region {L}
Xi = volume fraction of ith component {-}
Xt = starting point of transition region {L}
x = distance from wall in particle diameters {-}

z = axial position {L}


T = wall distance defined by equation (3-25) {L"1}

Xl\
a = constant defined by equation (2-15)
ag = constant defined by equation (2-73)
ak = constant in equation (3-27)
d = constant in equation (4-71)
psj = quadratic coefficient of binary synergism
Ac!} = m a s s fraction of ilh component
e = local porosity
eb = average porosity at the bypass section
ec = average porosity at the core section
ep = constant defined by equation (2-10)
et = porosity at the transition region
e'0 = over cross section average porosity
e L0 = average porosity at the core region

ecb = porosity at the bulk region


siw = average porosity at the i-region
e0w = average porosity at the wall region
o) = function defined by equation (2-11) {L2
y = constant defined by equation (2-5)
y = surface area of material {L
Yij = cubic coefficient of binary synergism
(pm = bypass cross-sectional fraction
K = bed permeability {L
K' = constant in equation (3-55)
KS = constant in equation (2-35)
Xi = inertia factor defined by equation (4-33) { M L-4
\2 = permeability factor defined by equation (4-34) {ML"Y'
3
t
.
X3 = pressure factor defined by equation (4-35) { M L "iY
i.
p. = viscosity {MLU't

XV
= effective viscosity {ML'Y'

V = kinematic viscosity {LY1


ve = effective kinematic viscosity defined by {LY'
equation (2-88)
v< = turbulent kinematic viscosity {LY1

vr = function defined by equation (2-90) {LY1

= density {MU 3
p
o = surface tension {Mf2

_ = shear stress tensor {ML"'f2


= shear stress tensor {ML"'f2

T1 = turbulence shear stress tensor {ML'Y2

= laminar shear stress tensor {ML"Y2

= laminar shear stress tensor ;ML"Y

r_o = constant defined by equation (4-79) or (4-81)


= correction factor defined by equation (4-18)
= sphericity {-
= dimensionless radial position defined by
equation (4-55)
= the "del" or "nabla" mathematical operator {-
= increase in internal energy due to chemical {ML-f
j^dnij
effects or changes in component or substance
i, between states 1 and 2
= the s u m of square errors {-}
X*

x
CHAPTER ONE

INTRODUCTION

Many engineering fluid flow problems fall into one of three broad

categories, namely flow in channels, flow around submerged objects, and

the transition between flow in channels and flow around submerged

objects. Examples of fluid flow in channels are pumping oil in pipes, flow

of water in open channels and flow of fluids through a filter. Examples of

fluid flow around submerged objects are the motion of air around an

aeroplane wing, motion of fluid around particles undergoing sedimentation

and flow across tube banks in a heat exchanger. The fluid flow in packed

beds is an example of the transition fluid flow between flow in channels

and flow around submerged objects.

In many chemical and metallurgical process operations, a fluid phase

flows through a particulate-solid phase. Examples include gas-solid phase

reactors, filtration, heat transfer in regenerators and pebble heaters, mass

transfer in packed columns, chemical reactions using solid catalysts, and

gas absorption and chemical reactions in packed columns. In many

cases, the solid phase is stationary, as it is in a packed absorber column.

In some cases, the bed moves counter-current to the gas stream, as it

does in a pebble heater or in some gas-solid phase reactors. In some

1
Introduction

cases, the fluid velocity is great enough that the m o m e n t u m transferred

from the fluid to the solid particles balances the opposing gravitational

force on the particles and the bed expands into a fluid-like phase, as it

does in a fluidized bed reactor. In still other applications, the fluid phase

carries the solid phase with it, as it does in pneumatic conveying.

The fluid flow in a packed bed system has important applications as in

heat and mass transfer equipment and in chemical reactors. The constant

effort to realise the profit improvement in industrial operations has

prompted extensive studies of mathematical and physical models for this

system. In order to improve the accuracy of the evaluation of packed bed

system performance, new study efforts should be undertaken in a

microscopic view, eg. heat and mass transfer coefficient distribution over

a packing of particles, rather than a macroscopic view, eg. pressure drop

or overall mass and heat transfer coefficient.

One of the crucial factors in the study of packed bed systems is the

velocity distribution of fluid flow across the bed. A knowledge of velocity

profile in the packed bed systems is required for further studies of rate

processes and their mechanisms taking place in a packed bed. This

thorough study is required as a sound basis for process evaluation of

more complex situations where flow distribution is accompanied by heat

1
Introduction

and mass transfer with chemical reactions, as for example, in hot spot

formation in a packed bed chemical reactor.

It has been extensively studied and reported in several papers that the

velocity profile in packed beds has a significant influence on the

performance of mathematical models of the packed bed systems [Schertz

and Bischoff, 1969; Choudhary et al., 1976ab; Lerou and Froment, 1977;

Kalthoff and Vortmeyer, 1980; Vortmeyer and Winter, 1984; Vortmeyer

and Michael, 1985; Cheng and Vortmeyer, 1988; Vortmeyer and

Haidegger, 1991; McGreavy et al., 1986; Delmas and Froment, 1988;

Ziolkowski and Szustek, 1989; Kufner and Hofmann, 1990]. Figure 1-1

illustrates the influence of velocity distribution across the bed on the

performance of reactor evaluation as reflected in the temperature profile

calculation [Kufner and Hoffmann, 1990]. The results in Figure 1-1 only

account for the convective contribution arising from the variation in the

residence time over the radial cross section. Strictly, account should also

be taken of the consequential changes in the local film coefficients due t

the velocity distribution, but this only tends to exacerbate the situation.

The work undertaken in the present study investigated the mathematical

model of velocity distribution of single-phase fluid flow in packed beds.

This problem has been widely studied; however, the results obtained have

T
1
s
•2

2 o
cn

o c
m c
(0

Ut
E
Tt
•c
d £
ra
c o ._
o
• H
0)
c
c
o
4—>
H—

<J1 a
M-J -C 3
T3 M-J
>. • i-H _2
.o-a T3
T3
O
"o>.3 oo
• —

__ f
m F E
cd
• -H <U (w)

T3 > d __ u
Cr 13 •M
Ui

e_ c
Kl M—1

M—1
C. 3
O —
i/i s-u ro
"O -C wG '• CN -
T3
Mr—I
•mM
•w—'
• 1-H
,
O u E
1)
rH
£ £ —o
3 (U u
c. X)
t/i u
T3 «
CJ u
<u o OH 1
o
CN 06 cs
rt T^
. o
(U
Sr.

0 c
o
c
o
'-l-l

3
JO
'_Z
4-1
._.
•5
>
4--
o
o
>
o o re
CN o u
T3
.2 51 '8JIHBJ8dui3X I B ! X V OJ
ro
LU

(1)
i_

3
cn
il
Introduction

not yet satisfactorily considered the models performance and the methods

of solving the mathematical equations.

According to Reid and Sherwood [1966], the value of a mathematical

model of the physical phenomena depends on at least three parameters:

its accuracy, its simplicity, and the type of information necessary for its

use. Actually, it is a difficult problem to generate perfect models but the

continuous improvement of the existing models will increase their value.

Figure 1-2 shows schematically the flow chart for mathematical model

development of velocity distribution of single-phase fluid flow in packed

beds.

5
Introduction

Physical
Phenomena

1
Development of
Mathematical Model

I
Results of
Calculations

I
Compare to
Experimental Data

No

Yes

Compare to
Existing Models

No

Yes

New/Improved Model

Figure 1-2: Flow diagram of model development.

6
CHAPTER TWO

LITERATURE SURVEY

In the 1950s, mass, heat and momentum transfer in packed bed systems

has received much attention. This condition has prompted extensive

studies of mathematical and physical models for these systems. One of

the crucial factors in the study of packed bed systems is the velocity

distribution of fluid flow across the bed.

The investigations of velocity distribution of fluid flow in packed beds can

be done by two main methods: experimental investigation and

mathematical modelling. The results of these investigations are discussed

in detail in the following section.

2.1 THE EXPERIMENTAL WORK


Since 1950, numerous experimental works have been done to investigate

the velocity distribution of fluid flowing through packed bed systems.

Arthur et al. [1950] investigated the radial velocity distribution of fluid f

in packed beds by measuring the flow distribution of air through a bed of

charcoal granules (0.0009 to 0.0028 m) in a glass tube of 0.0483 m

diameter. In this investigation, the flow was separated into several parts

7
Literature Sun'ex

by the inserting of thin rings, concentric with the tube wall, at the top of

the bed.

In their experiment, the flow rate in each section was measured

simultaneously using the soap bubble technique. In this technique, the air

flow rate is measured by allowing the flow to drive a soap bubble along a

tube and observing the time taken for the bubble to sweep out a

calibrated volume.

The essential result of Arthur ef al. [1950] is that the fluid flow is not

uniform across a packed bed. Although the results obtained from this

experiment did not give point velocities but integrated flow rates over

small cross-sectional areas of the bed, they indicated that the fluid

velocity reached a maximum at a short distance from the tube wall and a

minimum at the centre of the tube. It is considered that a good qualitative

representation of non-uniform fluid flow over the cross section of a packed

bed has been shown in these results. Actual flow distribution above a

rectangular packed bed depicted by Vortmeyer and Schuster [1983] is

shown in Figure 2-1.

The authors [Arthur etal., 1950] also obtained similar results by using the

following other methods:

8
Literature Survey

Figure 2-1: Actual flow distribution above a rectangular packed

bed, NRe = 8 and Dp = 1.25 mm [Vortmeyer and

Schuster, 1983].
Literature Survey

1. Chemical estimation of the products of reaction or adsorption on

various parts of charcoal when a stream of air containing a gas which is

absorbed by, or reacts with the charcoal, is passed through the bed.

2. Qualitatively, by observing the passage of a gas-laden air stream

through a bed of granules stained with an indicator.

3. By observing the concentration of emergent gas from the charcoal

column at various parts of the cross-section by means of test papers, a

gas-laden air being used.

4. By measuring the temperature attained at the side and the middle of a

wide column.

The measured data are listed in Table 2-1. Referring to their experimental

results, Arthur et al. [1950] remarked that the main factor affecting flow

rate distribution is bed porosity. Higher flow rates of fluid at the region

near the tube wall are strong indications that this occurs because the

smooth wall increases the bed voidage near it.

Morales et al. [1951] employed a series of circular hot wire anemometers

of various diameters to measure the fluid velocity at a series of the radial

positions of the bed. Their measurements were made over a range of air

velocities (0.123 - 0.533 m/s) in a 52.5 mm diameter tube packed with

three sizes (3.175, 6.35, and 9.525 mm) of the equilateral particle

diameter and height of the cylindrical pellets. The results indicated that

io
Literature Survey

velocity distribution in packed beds is a function of air velocity and bed

height as shown in Figure 2-2.

The authors believed that two important factors bring about a velocity

distribution of the kind obtained. These are skin friction at the tube wall

and variation in the void space in the bed with radial position. These

conclusions seem reasonable since it is known that with flow in an empty

tube, a wall friction causes the velocity to decrease sharply near the wall.

In fact, for streamline flow in an empty tube, a wall friction causes the

velocity to parabolically decrease from the centre of the tube. It may

therefore be expected that when packing is introduced, the effect of wall

friction would be dampened and become negligible near the centre of the

tube. Close to the wall, however, wall friction would again become

important and is the probable cause of decreasing velocities near the

wall.

The characteristic of a maximum in the velocity profile was found to occur

for all heights of the packing. This, together with the subsequent decrease

as the centre of the bed was approached, was thought to be due to

variation in bed porosity with radial position and the resultant losses of

pressure energy. Referring to their experiment results, Morales et al.

[1951] also remarked that the velocity distribution is not independent of

the distance of the anemometer above the top of the bed.

3 0009 03201226 7
ii
>X
to

CO
->
-.
3
--.
s o
rr
in
-3 c.
q CO o. oo h-
CO Tt
CM
q
T—
q q
CD
]_
3
-C
<
CO CM
TJ
OJ
°
d. Tt CM CM CM
CM OJ T—

«
O
o
_.
o CM CO CO CO
ro CO d CO CO h-
-C CM
d d d
o
"ro
o
'_Z CO q LO cn a.
s
T3 -o CO CO h- I-
C T—
d d d
u
ro
CO
(A
O o CO in CO o
CD CD
i_ LO q
u d d d
ro
O)ro
CO CO
cu
4-< CO
CO
CO CO
c E E
oo E E E E E
c E E E o o
o tn to o o
o o
o
'•3 3 3 h- CO
TJ ^ CO CO
ffl CO CO o II II
_.
ro JL.
CD _.
CD
O
o O O
> MJ-r LO
CM C II
3
0 c O g q
-Q o --_.
CO o o o
>x

r _

2 m
cn
T~
•>
M M ^

CO
*->
CD
(0
) / . CD
CO / / CO
L_ CO C_ . _.
o O c3 // // o o
CO
co" : 1
1
' / F E CO
1? : 1 ;/ / CM
E
E "
_.
<u 0)
TJ
13 /// \< ".CO O Q.
c
•'// LO CM CO Tf Q
>
^ ^/As n II
^II Ln __
r -i—i rr -o
>-
Q. o
TJ
v/
/ /'
CD
T3
"O uj
O
C\J CD
__
CD CL CO
*-.-,,. CQ Q Q CD
- >
' ' 1—
o o
Ll. LO d OCO
o
d d
E
m/-n E
CO
; / - LO
O.
: /
CD • /
CO 0 o .
depth = 152.4 mm

Lf) - / CD
52.5 mm

d d c3 ' ' /
CD
CO i E >
E l , / / jD
l E ro
Z3
I ' /
yS CO O £ CO
'jS co U
Tf £
II s__
o
/V DC 4-1
3
O
-Q
"D o CM CO
CD a. TJ
CQ Q Q >
4-1
' , H — , —' ' 1 ' ' '"—' 1 '—'
O
"o
oCD
LO LO
o
o >
d d CM
Ti/zn •
CM
OJ
_.
3
U)
Literature Survey

A more comprehensive investigation of velocity profiles in packed beds

was undertaken by Schwartz and Smith [1953] with the objectives of

determining under what conditions the uniform velocity assumption would

be valid, and, of correlating and explaining the magnitude of the observed

velocity profiles. Data were obtained in 50.8, 76.2 and 101.6 mm pipe,

using 3.175, 6.350, 9.525 and 12.7 mm spherical and cylindrical pellets,

corresponding to the range of D/DP from 5 to 32. The length of the

cylindrical pellets was equal to the diameter.

Their measurement velocity distribution is tabulated in Table 2-2. The

reproducibility of the data was tested without repacking of the bed; and

the maximum deviation of 2% was reported. If, however, a point velocity

was rechecked after the pipe was emptied and repacked, the variability in

packing increased the deviations. The maximum deviation of velocity

distribution data with repacking was 25%, although average deviation for

the three replications was 9%.

One of the main limitations of Schwartz and Smith [1953] study is related

to the velocity measurement points at the downstream of the bed by

assuming that the velocity distribution does not change. This distance

should be sufficient to smooth out the severe variations of velocity found

near the exit face of the bed and to eliminate non-axial components of

velocity without permitting any gross changes in velocity profile.

14
Literature Smre\

Table 2-2: Velocity distribution above the bed [Schwartz and Smith, 1953].

(Anemometer position 50.8 mm above the bed, 0.584 m bed depth).

Column Diameter = 50.8 m m .


UZ/UM
1
r/R 6.35 m m * 9.525 m m
0.30**' 0.46 0.67 0.82 0.30 0.46 0.67 0.82
0.32 1.05 0.96 0.89 0.86 1.06 1.02 0.94 0.92

0.55 1.25 1.12 1.07 1.05 1.26 1.19 1.15 1.14

0.71 1.23 1.16 1.17 1.22 1.26 1.13 1.16 1.20

0.84 1.09 1.12 1.11 1.11 1.06 1.03 0.99 0.98

0.95 0.45 0.68 0.74 0.76 0.46 0.59 0.63 0.64

Column Diameter = 76.2 m m .


UZ/UM
r/R 6.35Imm*' 9.525 m m 12.7 mm
0.49**' 0.80 0.49 0.80 0.49 0.80
0.32 0.62 0.61 0.81 0.79 0.77 0.82

0.55 0.80 0.70 0.88 0.82 0.97 0.90

0.71 1.08 1.12 1.04 1.04 1.09 1.11

0.84 1.20 1.20 1.06 1.17 1.01 1.04

0.95 0.99 0.93 0.80 0.82 0.89 0.83

: Diameter of spherical packing.


: Average velocity, m/s.
Literature Sur\>e\

Table 2-3: Velocity distribution above the bed [Schwartz and Smith, 1953].

(Anemometer position 50.8 mm above the bed, 0.584 m bed depth).

(Continued)

Column Diameter = 101.6 m m .


Uz/UM
r/R 6.35imm*' 9.525 m m 12.7 m m
0.31**' 0.49 0.31 0.49 0.31 0.49
0.32 0.69 0.71 0.69 0.67 0.70 0.62

0.55 0.83 0.71 0.98 0.83 0.89 0.81

0.71 0.99 1.00 1.09 1.05 1.06 1.07

0.84 1.38 1.31 1.25 1.20 1.27 1.24

0.95 1.19 1.10 1.14 0.97 1.08 1.03

Column Diameter = 101.6 m m .


UZ/UM
r/R 6.35imm*' 9.525 m m 12.7 mm
0.65**' 0.80 0.65 0.80 0.65 0.80
0.32 0.70 0.69 0.64 0.64 0.63 0.65

0.55 0.75 0.75 0.85 0.85 0.83 0.83

0.71 1.03 1.02 1.05 1.05 1.09 1.09

0.84 1.34 1.18 1.16 1.15 1.21 1.20

0.95 1.10 1.07 0.97 0.98 1.02 1.00

: Diameter of spherical packing.


: Average velocity, m/s.

16
Literature Suirey

Schwartz and Smith [1953] overcame the problem of measurement

distance from the packing by assuming the ratio of point velocity to

average velocity equal to 1.0 at r/R equal to 0.55. Based on this

assumption and regarding to preliminary experimental investigations, they

concluded that in their system, a distance of 50.8 mm between the bed

and the anemometer would minimise the errors and so all data were taken

at this position.

Based on their experimental data, these authors concluded the results as

follows: The maximum or peak velocity ranges up to 100% higher than the

centre velocity as the ratio of pipe diameter to pellet diameter decreases.

The divergence of the profile from the assumption of a uniform velocity is

less than 20% for ratios of pipe diameter to pellet diameter of more than

30. The maximum in the velocity profile occurred at a distance of

approximately one particle diameter from the wall, regardless of pipe and

packing size. The deviation from flat profile became more pronounced as

the ratio of D/DP decreased.

Dorweiler and Fahien [1959] used a series of circular hot wire

anemometers to determine velocity profile at the exit of packed beds. The

test column was a vertical 101.6 mm diameter pipe, packed with 6.35 mm

spherical, ceramic catalyst-support pellets. The anemometer was

17
Literature Survey

operated at 25.4 m m above the bed, after this distance w a s determined

experimentally to be an optimum height.

Their measurement result of the velocity distribution across the test bed is

shown in Figure 2-3, and this profile was reported to be independent of

total flow rate above an average superficial velocity of 0.122 m/s. Briefly,

the radial fluid velocity profiles obtained by these investigators exhibited

the similar basic trend of having a maximum near the wall as those of

previous workers. The values of the ratio of (uz)max/uM were greater than

those reported before.

In order to overcome the problem of flow changes in the open tube where

velocity profile measurement was carried out at the exit of the bed, Cairns

and Prausnitz [1959] measured velocity profile at inside of the bed. They

used electrode techniques to measure mean axial velocities over a length

of bed, using both fixed and fluidized beds with water as the fluid.

A salt tracer solution was injected into the water main stream over the

entire cross-section of the bed and the time interval necessary for

detecting a sudden change in the rate of injection at zero position was

measured at some distance downstream. Electrical conductivity cells at

various radial and axial positions in the bed detected the change in the

injection rate.

18
Literature Sune\

Column diameter = 101.6 m m


Spheres = 6.35 m m
Superficial velocity =0.12 m/s

Figure 2-3: Velocity distribution 25.4 m m above a bed of spheres

[Dorweiler and Fahien, 1959].

19
Literature Swvey

Typical results obtained by Cairns and Prausnitz [1959] are shown in

Figure 2-4. ln each case a slight maximum was found at the centre of the

tube, but apart from this the profiles were found to be flat. It was not

possible to make measurements closer than two particle diameters from

the wall. The dotted extensions to the profiles shown are based on a mass

balance and indicate the type of behaviour that the authors expected at

the wall. The limitation of these experimental results in generating the

velocity profile near the wall by employing a material balance due to the

value of local bed porosity is assumed constant over the cross section of

the bed.

Price [1968] used a 3.8 mm diameter pitot-static tube to measure the air

velocity at the exit of packed beds. In order to overcome the problem of

flow changes in the open tube at the outlet of the bed, Price [1968]

divided the exit flow area with a honeycomb of concentric splitters with

intersecting radial vanes placed between the exit face of the bed and the

plane of measurement. The nose position of the Pitot-static tube is

approximately 1.6 mm downstream of the exit face of the honeycomb. This

is to minimise the flow blockage caused by the Pitot-static tube.

Measurements were made for all compartments of the honeycomb over

the whole cross section of the bed, as many as 1,000 readings were taken

in a single run to confirm the reproducibility of measurement.

20
Literature Su/rev

10000

Pipe diameter =50.8 m m


Spheres diameter =3.2 m m
e'o = 0.38

uM/e'0 = 1440 mm/s


— o o-'
1000 --
630
—o o-

320
230
160
100 -- 80
SI
3

25
-o o*'

10-- 8
-t =8= =0-"
7.4

6.0 4.0 2.0 0.0


(R-r)/DP

Figure 2-4: Velocity profiles for a randomly packed bed [Cairns and

Prausnitz, 1959].

21
Literature Sun-ey

The author investigated the parameters that were suspected to have

influence on velocity profile, such as Reynolds number, bed length,

packing method, sphere material and D/DP ratio. Tests also were made to

assess the reproducibility of the measurements with repacking of the bed

between tests.

A typical velocity distribution above the bed measured by Price [1968] is

shown in Figure 2-5. The results from these investigations may be

summarised as follows: The velocity profiles, normalised with respect to

uM, were independent of Reynolds number (1,470 < NRE < 4,350), beds

length (9 < L/D < 36), and spheres material over the range tested. Slight

systematic effects were observed near the walls due to packing method,

sphere properties, and vessel to sphere diameter ratio (12 < D/Dp < 48).

The maximum velocity was found to exist within one-half sphere diameter

from the walls of the containing vessel.

Newell and Standish [1973] used thermistor anemometers to measure the

velocity distribution of gas streams flowing in packed beds. Velocity

distributions for a number of air velocities (0.04, 0.08, and 0.09 m/s) wer

determined in two columns, having square and rectangular cross-section,

respectively. The square column (101.6 mm by 101.6 mm) was used to

measure velocity distributions for packing consisting of 6.35 and 16.93

IT
Literature Surrey

m m spheres, 6.35 m m Raschig Rings and 6.35 m m coke resting on a wire

gauze support. The rectangular column (533.4 mm by 152.4 mm) was

used to determine the profile for 6.35 mm spheres.

In addition, velocity distributions were determined in a slice model of a

copper blast furnace. This model was packed with 6.35 mm and 12.7 mm

coal and velocity distributions were measured at various heights above

the tuyere level [Newell, 1971].

The essential result from Newell and Standish [1973] was that the fluid

flow profile in square and rectangular packed beds is similar to that in

circular beds. Their measurements also indicated that the fluid velocity

reached a maximum at less than one particle diameter from the wall as

shown in Figure 2-6.

Szekely and Poveromo [1975] also used hot wire anemometers to

measure the velocity distribution at the exit of gas streams flowing in

packed beds. This investigation was undertaken to elucidate the

mechanism of flow maldistribution or non-uniform flow through packed

beds system. Their measurements were made, over a range of Reynolds

number (100 - 400) in 101 and 152.5 mm diameter tube packed with 1 - 6

mm diameter glass spherical particles.

23
>x
ri
-_.

o
d

CO
CO
CO
o
oi o

0)
_Q
T3
p F o CD
l/t x.
£ E £3 u
Ik
VO
CN 00 vo c a
O 00
vo o CN <N CO
OT CN
o cn
II II II II II
Ul
1)
•mm* to
M—»
o
£ vd •g
E D_
T j rC T3
Cr 3
1
r-
C-M
O
ui
-1 <u
? o
T3 -C
O UJ £--,
c
3 co U cq on g
q
od 3
-Q
'iZ
4->

w
TJ
o O
d o
CD
>
LO
CN
o OJ
CN _.
ON U. 3
d
z
wn/ n
Literature Survey

1.5

101.6 x 101.6 m m square column


16.9 m m Spheres

1.0-

s
3
-_
3

0.5 -

Superficial velocity, m/s


- - o - - 0.04
-^^0.06
- -o- - 0.08
- • o - 0.09

0.0 *-» I l I I L_J I i i i


-H-1-
0.0 0.5 1.0 1.5 2.0 2.5 3.0
(Re-re)/Dp

Figure 2-6: Velocity distribution of fluid flow 25.4 m m above square

bed of spheres [Newell and Standish, 1973].

25
Literature Suivex

The results of parallel flow measured by Szekely and Poveromo [1975]

were also similar to previous investigations on measurements at the exit

of the beds, the smoothed profile again showed the characteristic rise in

velocity near the wall as shown in Figure 2-7. Experimental

measurements were also made of the pressure distribution at the inlet and

at the exit of the column. It was also found that for the experimental

conditions used the pressure was uniform.

In order to provide a more comprehensive data on fluid flow inside a

packed bed, Stephenson and Stewart [1986] employed the optical

measurement technique to measure the velocity and porosity distribution

inside the beds. The experiments were done in a vertical 75.5 mm

diameter fused quartz tube, randomly packed for a length of 145 mm with

cylinders cut from fused quartz rod. They used tetra-ethylene glycol, tetra-

hydropyran-2-methanol, and a mixture of cyclo-octane and cyclo-octene

as a fluid in order to fulfil the requirement of the range of Reynolds

number and Newtonian fluid characteristics.

The composite of superficial velocity across the test bed is shown in

Figure 2-8. The velocity profile has a peak near the wall, where the

porosity is largest and its fluctuations correspond to those of local

porosity. These results are therefore similar to the measurement at the

exit of the beds by employing flow separator of Price [1968].

26
Literature Suirey

3.00

2.50 --

2.00 -•

s
3
""Si
3
1.50-

1.00 --:

0.50 --

0.00
0.0 5.0 10.0 15.0 20.0 25.0
(R-r)/DP

Figure 2-7: Velocity profiles of fluid flow 10 m m a b o v e a circular bed

of spheres [Szekely and Poveromo, 1975].

27
Literature Surx'ev

DPe = 7.035 m m

D/Dpe = 10.7

L/Dpe = 20.6

0 1 2 3 4 5

(R-r)/Dp

2-8: Composite superficial velocity profile inside the bed

[Stephenson and Stewart, 1986].

T
Literature Surcey

The conclusion of Stephenson and Stewart [1986] that the bed porosity

distribution determines the velocity profile is also supported by

experimental work carried out by McGreavy ef al. [1986]. These authors

measured the velocity distribution inside and at the exit of the bed by

using laser Doppler anemometry. It has the advantage that it is capable of

giving good spatial resolution so that the flow distribution can be related

to the structure of the bed. However, this method is only good to take

measurements for small values of D/DP ratio because the need to provide

a suitable optical path poses problems for fixed beds.

The characteristic of a non-single maximum or an oscillation in the

velocity profile was found to occur for all measurement points, namely in

the inside and at the exit of the beds as shown in Figure 2-9. The most

striking feature of Figure 2-9 is that the observed flow profiles at the exit

are different from those inside the bed. This is of some significance as it

can reject the assumption that velocity profile inside the bed is similar to

that at the exit, as being appropriate.

In a more recent study, Ziolkowska and Ziolkowski [1993] used thermo-

anemometric techniques to measure the velocity distribution at the exit of

gas flow in packed beds. The experiments were done in a vertical 94 mm

inside diameter tube, packed randomly with uniform porcelain spheres to

a height of 1050 mm and diameter of spherical particles (4.11 - 8.70 mm).

29
Literature Survey

Distance from wall, particle diameter


0.5 1 1.5 2 2.5
2.8 J L. _• I i i_ -J 1 I 1 L

Inside the bed.

6-
Exit of the bed. it A

4-
N
3
%x
2 - 1/ \
Atr ___-_^_mm^_____-_ \ *
3^r*"" ^^*-4r-_.

:__r IV _-W

0- ^ T — ! — I I I I I — i — r — T — r — i — i — -1 1 1 1 r—i 1 1 1 r-i 1 1 1 1 1 1 P

0 0.5 1 1.5 2 2.5


Distance from wall, particle diameter

Superficial velocity, u M (mm/s)

34 28 20 11

Figure 2-9: Comparison between corresponding velocity profiles at

the exit and inside of the packed bed (DP=16 mm, D=50

mm, Packing height = 220 mm, Measurement height =

140 m m and 222 m m ) [McGreavy etal., 1986].

30
Literature Survey

Ziolkowska and Ziolkowski [1993] measured the radial distribution of air

flowing through over a range of superficial velocity (0.4 - 1.0 m/s). A

typical velocity profile above the bed of their measurement result is shown

in Figure 2-10. Similar to previous investigations, these authors found

that, with an accuracy of ±3.2%, the shape of the local gas velocity radial

profile does not depend on flow rate while the average reproducibility of

these profiles after repacking the bed was 4.2%. The pellet diameter had

a more pronounced effect than the flow rate on the shape of velocity

profile that the smaller the pellet diameter, the flatter the profile.

From the foregoing brief summary of the results of experimental

investigations of velocity distribution in packed beds, it can be concluded

that the fluid flow is not uniform across a packed bed. Although the

number of observed flow maxima points is dependent on the

measurement technique that was employed; however, generally the

maximum value in velocity occurs near the wall of the container. This is

due to the opposing effects of wall friction and the variation in bed

porosity with radial position relative to the wall.

Generally, the dimensionless wall distance in terms of particle diameter,

(R-r)/DP, is more representative to explain the velocity distribution than

dimensionless radius, r/R. This condition agrees with the behaviour of bed

porosity in packed beds [Goodling etal., 1983; Roblee etal., 1958].

31
Literature Surve
rev

uM
0.4 m/s
0.8 m/s

1.5

2
3.

0.5
Pipe diameter = 94 m m
Bed depth = 1050 m m
Spheres diameter = 8.7 m m

0 l i i i i i i i i i i i i i i i i i i—1_

0 2
(R-r)/DP

Figure 2-10: Velocity distribution at 15 m m above the beds of

spheres [Ziolkowska and Ziolkowski, 1993].

32
Literature Survey

According to this condition, it is evident that the bed porosity distribution

has significant influence on the velocity distribution in a packed bed

system. Any account of experimental investigations of velocity profile in

packed beds must therefore also include investigations of the porosity

distribution in packed beds.

Regarding to the phenomena of the developing fluid flow in a channel

[Poirier and Geiger, 1994], it is reasonable to conclude that the velocity

profile at the exit of the bed without flow separator is a developing flow

condition. This is a transition profile between a velocity profile inside the

bed and a fully developed flow profile in an empty tube, and possesses a

developing flow distribution. Therefore, the velocity distribution data that

were measured at the outlet of the bed can not directly be used to

represent the velocity profile at inside of the bed. Based on this condition,

the study of velocity distribution inside the beds by using the velocity

profile that was measured at the outlet of the bed without flow separator

needs to be corrected with developing flow phenomena.

2.2 MATHEMATICAL MODELLING

As mentioned earlier, the fluid flow problem in packed beds is a transition

between flow in channels and flow around submerged objects. According

to discontinuity of this system, an exact representation of the fluid flow

distribution in packed beds is impossible [Ziolkowska and Ziolkowski,

33
Literature Survey

1993]. The fluid mechanics equations cannot be easily applied to describe

the flow field within the whole space bounded by the tube wall. They might

be applied for the space between granules [Mickley er al., 1965], but they

could not be integrated within that space because the boundary

conditions are indefinable.

For this reason, a number of mathematical models of fluid velocity profile

in packed beds is based on experimental data, or on theoretical

considerations adjusted by relatively simplified assumptions. Although the

accuracy of phenomenological approach is good for a particular set of

data, it is difficult to apply this type of model with any confidence to oth

systems/conditions. On the other hand, although numerous theoretical

approaches have been done in this area, a consensus has not been

reached and there is a need for further investigations.

2.2.1 The Phenomenological Approach

On the basis of velocity distribution that was measured at the exit of the

beds, Schwartz and Smith [1953] tried to develop a mathematical model of

velocity profile in packed beds. The model was developed by applying two

main assumptions: uniform pressure drop over all radial positions of the

bed and the variation of bed porosity due to the wall effects over cross

section of the bed.

34
Literature Survey

In order to develop the mathematical model, Schwartz and Smith [1953]

employed the Prandtl's expression of shearing force [Bird et al., 1960] and

Leva's correlation of pressure drop [Leva, 1992]. Schwartz and Smith

[1953] assumed the driving force for velocity distribution or momentum

transfer is the difference in pressure drop between the center of the pipe

with local porosity minimum, where no momentum transfer occurs, and

that at any radial position. This pressure defect per unit length of bed, P5,

can be derived from Leva's correlation as follows:

P — 2
' um2(1-e)
-^--u
3 "m0
2(1-0
3
(2-1)
8
"DP e £
LO

The total pressure drop correlation proposed by Leva is due to skin

friction between the fluid and solid particles and pipe wall and orifice

losses, as well as that due to turbulent shear in the fluid alone. Then, if

is assumed that the fraction of the total pressure drop due to turbulent

shear in the fluid is constant, the Leva's correlation can still be used to

determine the pressure defect by introducing a constant factor, k2, that

was obtained by adjusting with their experimental data to give the value of

k2 equal to 0.0096 [Schwartz and Smith, 1953].

In order to reduce the errors involved in using Leva's pressure drop

correlation, Schwartz and Smith [1953] measured the pressure drops

each time the velocity profile was determined. Considering the force

35
Literature SUITCX

balance over the bed to develop the correlation between the pressure

defect and the shear stress gives:

k 2 Ap fl-^^ 1-8 LO du. du.


u. -u mO = PL, (2-2)
l-e r tL dr dr
uM . ° yJ

Considering equation (2-2), Schwartz and Smith [1953] failed to

distinguish between true local fluid velocity, uz for Prandti's shearing fo

on the right hand side of equation (2-2) [Bird et al., 1960] and local

superficial fluid velocity for Leva's pressure drop correlation. This

condition is acceptable only if the value of bed porosity is uniform over t

whole cross section of the bed.

Equation (2-2) defines the velocity gradient in terms of bed porosity e,

radius r, and mixing length LM (the radial distance that a small mass of

fluid travels before losing its identity). Equation (2-2) is basically a

theoretical equation other than the semi-empirical expression for Leva's

pressure drop correlation, and Prandti's shearing force [Bird et al., 1960].

The problem in using this equation is that the integration of the expressio

to obtain the velocity profile varies in complexity with radial position.

However, simplifications have been developed to overcome the above

problems. Flow through packed beds is analogised to flow through a

36
Literature Suirey

bundle of tubes, which consists of a central core, containing tubes of

constant diameter, surrounded by tubes of progressively larger diameter.

The bed porosity at central core is assumed constant and equal to e0. Also

in this region the mixing length, LM, will be assumed equal to DP/2. With

these simplifications equation (2-2) may be integrated and then

substituting dimensionless parameter um/uM, to give:

2 V21
( xx \
'u^
u. • + < u mO
i/
uM VUM ; VUM J
In
U mO
{-
= k' (2-3)

UM

Where:

(
D \K k 2 D p ApY (2-4)
3 KD*J 8u M

ri- £ ' 0 Y £ ^ (2-5)


Y=

Although the value of the constant k', m a y be determined theoretically

from equation (2-4), however, the errors involved in using this model is

significant. The deviation may arise from the inappropriate assumptions in

developing the model, namely, LM = DP/2 [Schwartz and Smith, 1953], fluid

phase shear stress being a constant fraction of the pressure loss [Newell,

1971], the use of the poor analogy of mixing length by Prandtl [Bird ef al.

1960; Mickley etal., 1965], the failure to distinguish between uz and uM,for

37
Literature Suirey

Prandlt mixing length and Leva's pressure drop correlation, and the

constant properties of central core which are dependent on D/DP ratio

[Roblee etal., 1958; Benenati and Brosilow, 1962].

In order to reduce the errors involved in using this model, Schwartz and

Smith [1953] offered the experimental value of k'. The problem in using

this model is due to the lack of constant k' data of other

system/conditions. For this reason, it is appropriate to bring this model

into phenomenological model category.

With improvements in shear stress correlation and the concept of

pressure defect of the Schwartz and Smith [1953] model, Price [Price,

1968; Newell, 1971] tried to develop a mathematical model of velocity

profile in circular packed beds. By extending the Price result, Newell and

Standish [1973] employed the model for rectangular packed bed and non-

ferrous blast furnaces.

Price [Price, 1968; Newell, 1971] employed Boussinesq's shear force

correlation [Bird et al., 1960] to replace the Prandtl correlation that was

used in Schwartz and Smith's [1953] model. With reference to work by

Dorweiler and Fahien [1959], the correlation of eddy viscosity to fluid

velocity is performed by assuming constant value of Peclet number, then

introducing this result into Boussinesq's shear force correlation, gives:

38
Literature Surx-ey

x'=pbDpuM-^ (2-6)

Considering a cylindrical packed bed of unit length and radius, r, Price

[Price, 1968; Newell, 1971] developed a force balance on the fluid flow as

follows :

27.r.t + ..r2—= F (2-7)


dz '

In equation (2-7), F is the force resisting motion per unit length of bed and

arises from the interactions between the solid packing and the fluid. Since

velocity varies with radius across the cylindrical section considered, the

total resistive force acting on the fluid within the cylinder is:
R
F = J27crcuM2dr (2-8)
0

In developing the force balance, Price [Price, 1968; Newell, 1971] failured

to distinguish between shear stress momentum in continuous systems, for

example, fluid flow in an empty tube, and discontinuous systems, for

example, fluid flow in packed beds. The shear stress momentum part in

equation (2-7) is an expression of continuous systems rather than

discontinuous systems, which must be corrected by local bed porosity

factor.

39
Literature SuiTev

Substituting equation (2-8) into equation (2-7), and assuming that the

value of the pressure drop is independent of the radius and then

rearranging gives:

2c d2uv If du-^j 2c ( , ldp^


2 U + 2 U
MTT : MM ^ v
dr' r. dr J pbDF v

For the central core, the true velocity, uz, is related to superficial velocity

in the empty tube at the exit from the bed, um, by the expression

L
U eU
m=77 Z
(2-10)
p
= £ uz

Rewriting equation (2-9) in terms of u m , putting 2c/(pepbDP) equal to B 2

and making the substitution gives:

£p2d
AX 2 P ,0.,-n
6 = um (2-11)
y m V ;
c dz
By assuming (dum) «umd2um, which is reasonable when the value of

du m is between 0 and 1, results in

- i + --f-B 2 ()) = 0 (2-12)


dr r dr

Equation (2-12) is a Bessel equation [Mickley et al., 1957]. By assuming

the symmetric condition of velocity profile, which is reasonable, the

solution of equation (2-12) is:

40
Literature Survey

ep dp
((> =
um„ -• I„(Br) (2-13)
m0
c dz

Before equation (2-13) can be solved, a further boundary condition has to

be specified. This boundary condition relates to the outer edge of the

central core where the increased velocity provides the potential for

momentum transfer into the central core of the bed. These high velocities

arise primarily from the relatively high bed porosity fraction near the wal

[Newell, 1971].

Equation (2-13) predicts a superficial velocity profile whose gradient

increases with radius, as a central core is assumed to extend to a radius

r, where the gradient of the velocity profile is no longer increasing. The

velocity at this radius is denoted by um. Substituting these boundary

ep2 dp
conditions into equation (2-13), and solving for — — - in terms of u m , f

and um0, and then normalising with respect to um, gives:

r 2
o [l„(B.)-l]= '-- _^. * [(a2-l)l0(Br) + I 0 (Br)-a 2 ]
.U M J
(2-14)
m

Where

u. (2-15)
a =
u mO

41
Literature Survey

It should be noted that the use of equation (2-14) to predict the velocity

profile requires a knowledge of B, f, and a, as empirical constants. These

values must be determined by experiment with measuring of velocity

distribution. An evaluation of experimental data obtained by previous

workers [Schwartz and Smith, 1953; Dorweiler and Fahien, 1959], Price

[Newell, 1971; Newell and Standish, 1973] showed that B may be

reasonably assumed to have a value of (DP)"1. However, the values of

rand a remain to be determined by measuring the superficial velocity

profiles for the particular packing structure under consideration and over

the Reynolds number range involved. Since a number of empirical

constants in this model are required to be determined by experimental

work of velocity distribution, it is evident that the model is in the

phenomenological category.

Although the Price model was originally derived from fundamental fluid

dynamics theory, the result has a number of empirical constants. The

reason for this condition is because the urge for an analytical solution led

to oversimplification of the mathematical equations.

Newell and Standish [1973] tried to extend the Price model of fluid flow

distribution in circular packed beds to rectangular packed beds by

employing equivalent diameter concept. They assumed an equivalent

diameter term based on the bed cross section to represent a packed bed

42
Literature Survey

of non-circular geometry. Velocity profile of fluid flow in a non-circular

packed bed was developed by substituting the equivalent radius re into

equation (2-14) to give:

( V ( V
[l0(B?J-l]= ^=2. [(a2-l>0(Bre)+I0(Bfe)-a2]
I 11
(2-16)
.UiM, V U
M;

The validity of equation (2-16) w a s investigated by using square and

rectangular packed beds [Newell and Standish, 1973]. The agreement

between measured data and results predicted by this model was good for

physical model of square and rectangular packed beds. However, the

model failed validation for non-ferrous blast furnaces by using a V3-slice

model of the copper blast furnace. In the central region the measured

velocities differed widely from predicted values. This is considered to be

due to a combination of causes, which include the failure emanating from

original Price model development, as mentioned earlier, and also the

changing of the boundary conditions, for example, movement of the

burden, method of charging, and segregation in the burden.

Based on the data at the exit of the beds that was measured by Schwartz

and Smith [1953], Hennecke and Schlunder in 1973 [Tsotsas and

Schlunder, 1988; 1990] proposed an empirical correlation to predict the

velocity distribution in packed beds. The empirical correlation for circular

packed beds of spheres is as follows:

43
Literature Survey

u K + [(W + 2)/2](r/R)'
(2-17)
uM K +l

K = 1.5+0.0006 (2-18)

_D_ (2-19)
W = 1.14 -2
DT

The local superficial velocity at the centre of the bed can be predicted by

substituting r=0 into equation (2-17) to give:

K _____ (2-20)
uM K+l

Similar to Hennecke and Schlunder's model, Fahien and Stankovic [1979]

proposed an empirical correlation of velocity distribution based on the

data at the exit of the beds that were measured by Schwartz and Smith

[1953]. The empirical correlation for circular packed beds of spheres is as

follows:

A0+A,rs+I+A2rs+2 (2-21)
um = 2E

Where

An A, (2-22)
E= —
2 + s+3 s + 2

1 l
M (2-23)
0
s+2 s+1
Literature Sur vev

r
M
A (2 24
'=i7T " »
r
M
A 2 = ^ - (2-25)
s+ 2

rM = l - 2 ^ (2-26)

All constants in equations (2-17) to (2-26) are empirical. These values

must be determined by experiment if the condition, such as particle size

distribution, column diameter, etc., is not similar to Schwartz and Smith's

[1953] data. Thus, using this model is not practical because of the

requirement of the experimental data of the velocity distribution.

Tien [Vortmeyer and Schuster, 1983] has derived a general analytical

expression of flow profiles in semi-infinite packed beds which are

bounded on one side by a rigid wall. He found the general analytical

expression due to the porosity function near the wall. The equation for a

tubular packed bed with R^>°° was given as:

. \ I r-R
f
=1 r-R'e v Dp
(2-27)
U C
M 1- n DP
p
. J
where

4n.
a= (2-28)
4-n,

45
Literature Survey

Unfortunately, the coefficient n cannot be determined theoretically. This nc

value must be determined by experimental work, and as an approximation

Vortmeyer and Schuster [1983] have developed a formula for n as a

function of NRe.The value of nc varies from 0.1 to 27.

Because of the infinite packed bed assumption in developing this model,

error is able to occur for finite packed bed calculation. Consequently,

D
calculation of fluid flow in packed beds with small value of — by using
-Up

this model must be examined carefully.

From the foregoing brief summary of the modelling of velocity profile of

fluid flow in packed beds by using phenomenological approach it can be

concluded that this approach works better if it is calibrated with measured

data for a given system; however, the applicability may still vary

significantly for another system.

2.2.2 The Theoretical Approach

Numerous studies have been done to develop a mathematical model on

the basis of physical phenomena rules, or by extending the well-known

macroscopic model, have led to a number of mathematical models.

Because of the discontinuity of the systems a mathematical manipulation

46
Literature Survey

was required to apply the equations, so that a number of mathematical

models depended on the simplification of the system.

Szekely and co-workers [Stanek and Szekely, 1974; Szekely and

Poveromo, 1975; Poveromo et. al, 1975; Choudhary etal., 1976a; 1976b;

Szekely and Propster, 1977; Szekely and Kajiwara, 1979] tried to explain

the channelling and maldistribution of fluid flow in a packed bed reactor b

extending the Ergun [1952] pressure drop correlation into vector terms. A

good result in application of this model to the gas-solid reactor with larg

value of D/DP ratio was reported [Poveromo et ai, 1975; Szekely and

Propster, 1977; Morkel and Dippenaar, 1992],

In vector form, the Ergun [1952] equation may be written in microscopic

term [Stanek and Szekely, 1974] as:

-Vp = um(f,+f2um) (2-29)

By assuming incompressible fluid and then employing operator Vx on

equation (2-29) to eliminate the pressure term gives:

Vxum-umxV(ln(f1 + f2uj) = 0 (2-30)

The components of the velocity vector also have to satisfy the equation of

continuity [Bird etal., 1960; Poirier and Geiger, 1994] as follows:

V.um=0 (2-31)

47
Literature Survey

Upon finding the velocity field through the solution of equations (2-30) and

(2-31), the pressure distribution may be evaluated [Stanek and Szekely,

1974] from:

V2p = -um.V(f1+f2uJ (2-32)

For incompressible fluid flow through a cylindrical bed with axial symmetry

(that is, d/d§ = 0 and um$ = 0), equations (2-30) and (2-31) may be

rewritten [Stanek and Szekely, 1974; Bird et al., 1960; Jenson and

Jeffreys, 1963] as follows:

du du 3ln(f,+f7um) 3ln(f,+f,um)
" UJT IJQ2 , \ I 2 m/ \ 1 2 my- - rx
L / 0 nri\
_ u
i r - ^ r + u
- — ^ — u
-« ~ 3r
•- —=° (2 33
- »
l3(rumr) ( 5u m z (2-34)
=0
r dr 3z

In general, equations (2-33) and (2-34) have to be solved numerically

[Stanek and Szekely, 1974]. For special cases, the analytical solutions do

exist when dfjdz = df2/dz = 0, when the resistance does not vary in the

direction of flow. If the bed is sufficiently long, all radial components of

velocity vanish and the flow becomes parallel. On putting umr = 0 and

Um=Umz, equation (2-33) is readily solved to obtain:

1
U =
» -_T; + 'O
Uf; 2J
+^ (2-35)

48
Literature 5i.rv_v

The integration constant K S is determined from an overall balance on the

fluid.
R

uMR2=j2rumdr
o

R y
= }2r +
r ^ r -
dr (2-36)
0 2f2 + 2f 2 ;

The basis of this model is the Ergun [1952] macroscopic equation for

pressure drop of fluid flow in packed beds. Basically, the Ergun equation

was developed based on the approach of the packed bed being regarded

as a bundle of tangled tubes [Bird et al., 1960], and this approach yield

good results for bed porosity less than 0.5 [Bird et al., 1960; Cohen and

Metzner, 1981]. Another important limitation of the Ergun equation, as

cited by Gauvin and Katta [1973], is the fact that it is not appropriate for

systems containing particles of low sphericity. It follows that the uses of

this equation, or the use of relationships derived from it, become

automatically influenced by these limitations.

Comparison of Stanek and Szekely's [1974] model with experimental data

[Poveromo, et al., 1975; Szekely and Poveromo, 1975] shows a good

qualitative representation of velocity profiles in packed beds. However,

because of the limitation of the Ergun equation, the numerical values of

calculations are questionable, especially in the vicinity of wall area where

49
Literature Smrey

usually the bed porosity is greater than 0.5 [Goodling et al., 1983]. The

approach of dividing the bed into a number of incremental beds, as a

means of generating a theoretical velocity profile based on the Ergun

macroscopic approach, is also problematic because it means that the wall

effect is accounted for many times over and leads to a significant error.

In a more simple approach, Martin [1978] also developed the velocity

distribution of fluid flow in packed beds model based on the Ergun [1952]

equation. Martin [1978] divided the packed bed into a core section of

porosity, ec, and by-pass section of porosity, eb, where ec < eb. Assuming

radially constant velocities and the validity of the Ergun equation in each

of the sections the ratio umb/umc can be obtained.

Based on Benenati and Brosilow's [1962] bed porosity data, Martin [1978]

assumed that the by-pass section area is one particle diameter from the

wall of container. Since the two parts of the model packed bed offer

different resistances to the flow, a non-uniform flow distribution will resul

This can be calculated by applying the Ergun equation [Ergun, 1952] to

both parts of the packed bed:

i_P __ 150^1B^ + 1.75^^=1 (2-37)


L ec" Dp ec- Dp

and

50
Literature Survey

2
^= l50(±^}^+lJ5lz3^^ (2-38)
Dr D.

The superficial velocities umc and u m b are defined as the volume flow rates

divided by the empty cross-sections (1-<pm)A and cpmA. Substituting this

definition into equation (2-38) and rearranging gives:

u N
1 + 0.0117u mc R :
' c 1-e. Y Y _
l-e V
u mb M
(2-39)
u. u N 1-eb )
1 + 0.0117-u mb
M
Re
l-e b

For further use of the model, equation (2-39), can be solved explicitly for

the positive values of velocity ratio:

umb (pma + B m )-l +A/((pm(l + B ro )-l) 2 -4((p m +C m A m Xl-( Pm +A m )B r


(2-40)
u
»c " 2((pm+CmAm)

where

N,
A m =0.0117—Ss_ (2-41)
m -.
1-e.
f' Y
B. (2-42)
vfcc;
(2-43)
1-e.

Actually, the purpose of the simplifications by Martin [1978] are to

overcome the problem of calculation when employing the model to heat

51
Literature Survey

and mass transfer calculations in packed beds and the limitation of

applicability of Ergun equation which is restricted to bed porosity less than

0.5. Because of the increased availability of computing power and

software utilising numerical methods to solve the mathematical equations,

the first reason for the simplifications is not essentially required any

longer.

By using two zone area of fluid flow in packed beds as proposed by Martin

[1978], it may overcome the original limitation of Ergun equation for bed

porosity higher than 0.5 at the vicinity of the wall. However, this

simplification reduces the microscopic view advantage, because of the

results of investigation by using this approach are more pronounced as

macroscopic view, therefore it should not be considered for systems which

are sensitive to local velocity variation.

Vortmeyer and co-workers [Kalthoff and Vortmeyer, 1980; Vortmeyer and

Schuster, 1983; Vortmeyer and Winter, 1984; Vortmeyer and Michael,

1985; Cheng and Hsu, 1986; Cheng and Vortmeyer, 1988; Vortmeyer and

Haidegger, 1991] tried to develop the mathematical model of velocity

profile in packed bed by extending the Brinkman's equation [Brinkman,

1947] into microscopic view. The model was applied in a study of mass

and heat transfer [Vortmeyer and Michael, 1985; Cheng and Vortmeyer,

52
Literature Survey

1988] and a runaway chemical reaction [Kalthoff and Vortmeyer, 1980],

which have shown good results, qualitatively.

Brinkman's [1947] equation is developed from macroscopic view of fluid

flow in a porous medium by interpolating the Stokes equation and Darcy's

law [Durlofsky and Brady, 1987]. By neglecting the potential energy

change, in microscopic view the Brinkman equation may be written

[Wilkinson, 1985] as:

Vp = li'V2um-^um (2-44)

where K is the permeability, u\ is the viscosity of the fluid, and u' is an

effective viscosity. The permeability K and viscosity ratio ji'/p are

properties of the porous material [Wilkinson, 1985].

On the macroscopic level the flow of a single fluid through permeable

materials may be describe by Darcy law [Larson and Higdon, 1986]:

^ = -^u (2-45)
M v
dz K

Darcy flow is an expression of the dominance of viscous force applied by

solid porous matrix on the interstitial fluid and is of limited applicability.

Post-Darcy flow is effected by inertia forces and turbulence [Kececioglu

and Jiang, 1994], for example the Ergun [1952] equation:

53
Literature Surcey

^ = -f.UM-f2UM2 (2-46)

Where

(1_8o) H
f, = 1 5 0 - — r - — — (2-47)
p D
t
o ^P
(1_eo) P
f, = 1.75- ^ — (2-48)
e • D

Substituting equation (2-45) into equation (2-46) gives:

I1 . __. 2
(2-49)
- U M = flUM+f2UM
K

Since the validity of the Brinkman equation is restricted to low flow rates,

Vortmeyer and Schuster [1983] extended it to higher flow rates by

incorporating the Ergun equation. By assuming the value of \i'/\i = 1, and

the macroscopic view result of Darcy law and Ergun equation treatment

(equation (2-49)) can be applied into microscopic view, then substituting

equation (2-49) into equation (2-44) to give:

Vp = uV2um-f.um-f2um2 (2-50)

For incompressible fluid flow through a cylindrical bed with axial symmetry

(that is, 8 /3<|> = 0 and um$=0), no radial flow direction (that is umr=0), an

no pressure gradient over the cross-section of the bed, and then equation

(2-50) may be rewritten [Vortmeyer and Schuster, 1983] as follows:

54
Literature Survey

3p fav i9u
= -fiUm-f2um2+u. (2-51)
dz dr' + r dr

In order to fulfil the assumption that velocity profile has one m a x i m u m at

near the wall, Vortmeyer and Schuster [1983] proposed an exponential

relation of porosity distribution, as follows:

R-r
e = e' 1 + c'Exp 1-2 (2-52)
157
where c' has to be adjusted according to e'0.

Equation (2-51) is an elliptic partial differential equation. In order to

ensure the stability of the calculation of equation (2-51), Vortmeyer and

Schuster [1983] proposed a variational method. The solution of

mathematical equation by using variational method is based on the

optimisation of an integral [Courant and Hilbert, 1953].

Derived from equation (2-51) by employing the variational method, the

equation to calculate the velocity profiles in circular packed beds is

[Vortmeyer and Schuster, 1983]:


2
2 Au. A
N = LJ.]£ [(ri+ArJ-rr'_C 1 mi
i 2 r 3
+ 2rsM
i= l Ar

= minimum (2-53)

with regard to the m a s s balance equation as follows:

55
Literature Su/rev

Qicai = 2 ^ X u m i A r i r i
i

= constant (2-54)

Comparison of calculation results by using equations (2-53), (2-54), and

(2-52) with experimental data has shown that the predicted maximum

value is far higher than that measured [Vortmeyer and Schuster, 1983].

Similar result was also reported by Johnson and Kapner [1990], who

developed a model velocity profile based on Brinkman equation without

modification for the Darcy term with Ergun equation. A poor agreement

between predicted and measured results also has shown on more

advanced investigation by using this model for more complex situations

where flow distribution is accompanied by heat and mass transfer, and

chemical reaction, for example, runaway reaction [Kalthoff and Vortmeyer,

1980; Vormeyer and Winter, 1984]. This is due to a combination of

causes, which include the limitation of the Brinkman equation, the using of

macroscopic level of Ergun equation in microscopic view, and the

assumption that the value of viscosity ratio equals to 1.0.

The Brinkman equation is a superposition of Darcy's law and Stokes

equation [Saleh et al., 1993b; Durlofsky and Brady, 1987; Wilkinson,

1985]. The diffusion of momentum in the bed, via the effective viscosity u',

is given predominance to Darcy term (that was replaced by Ergun

56
Literature Survey

equation for this model). The effective viscosity is a function of bed

properties [Lundgren, 1972]. Consequently, the assumption of |i'=u would

reduce the validity of Brinkman equation. Another limitation is the validity

of Brinkman equation which is restricted to highly porous media [Saleh ef

al., 1993b] and it is the fact that the bed porosity of major packed beds is

generally lower than 0.5 except in the wall region [Vortmeyer and

Schuster, 1983].

The important effect of bed porosity on the velocity profile of fluid flow in

packed beds has been recognised by numerous investigators [Schwartz

and Smith, 1953; Cairns and Prausnitz, 1959; Newell and Standish, 1973;

Stanek and Szekely, 1974; Vortmeyer and Schuster, 1983; Standish,

1984]. Cohen and Metzner [1981] followed the question by attempts to

analytically quantify the effect of bed porosity variation on radial

distribution of fluid velocity in packed bed.

These authors developed a parallel channel model of the packed bed

which they regarded as a porous medium divided into wall, transition and

bulk regions, as shown in Figure 2-11. The region extending from the wall

to xt is considered to be the wall region. The second region, which

extends from xt to xb, is defined as the transition region. The remaining

region, which extends from xb to the centre of the bed, is considered to be

the bulk region.

57
Literature Survey

Considering a Newtonian fluid flowing in a packed bed and assuming that

inertial effects can be neglected, the relationship between interstitial

velocity and pressure drop becomes [Christopher and Middleman, 1965;

Cohen and Metzner, 1981]:

Ap De2
= 2 55
^ W.t < - >
Where le is the length of the channel or alternatively, the effective path

length followed by the fluid, and De is the effective channel diameter. The

factor K0 essentially accounts for the inadequacy in the choice of a proper

effective diameter, De. The usual choice of an effective diameter is based

on the concept of hydraulic radius [Bird et al, 1960], where the effective

diameter is replaced by four times the hydraulic radius.

Cohen and Metzner [1981] defined the interstitial velocity as follows:

"-it (2-56)

Substituting equation (2-56) into equation (2-55) and replacing the

effective diameter by means of the hydraulic radius, equation (2-55) may

be rewritten to give :

u =^^- (2-57)
m v
L uB(n)
where

58
Literature Suney

l^2
B(n) = K 0 l £ l (2-58)

Although the constant B(n) in equation (2-57) w a s claimed as a universal

constant, Cohen and Metzner [1981] suggested that constant B(n) be

determined by experimental in order to assure the accuracy of fluid flow

calculations.

In order to account for the wall effect, Cohen and Metzner [1981]

employed a different kind of hydraulic radius for each region of the bed.

For wall region, they employed the hydraulic radius definition that was

developed by Mehta and Hawley [1969], as follows:

Volume of voids
Volume of bed
H
~ Wetted surface area of spheres Wetted surface of wall
+
Volume of bed Volume of bed

and for transition and bulk regions, they employed the hydraulic radius

definition that was developed by Bird etal. [1960], as follows:

Volume of voids
Volume of bed (2-60)
H
" Wetted surface area of spheres
Volume of bed

59
Literature Survey

bulk
region

transition
region
bulk wall
region region

R/2
X=0 X, xh

Figure 2-11: A schematic representation of the tri-regional model

[Cohen and Metzner, 1981].

60
Literature Survey

For the bulk region, the equation to predict the velocity of fluid w a s

derived by substituting equation (2-60) into equation (2-57) to give [Cohen

and Metzner, 1981]:

Ap DPV (2-61)
ub =
36uL(l-e cb ) 2 B b (n)

The average superficial velocity, ut, in the transition region from xt to xb,

can be written as [Cohen and Metzner, 1981]:

u t =^-£u m dA (2-62)

W h e r e A t is the cross sectional area of the transition region. Substitution

of the capillary model for um (equation (2-57)), together with the definition

of the hydraulic radius (equation (2-60)) yields:

_Ap_ 1 V £ t dA
u, •D, (2-63)
36uL A, Jv 1-e.

In the wall region, the usual definition of the hydraulic radius (equation (2-

60)) cannot be used since the presence of the wall is not considered

[Cohen and Metzner, 1981]. As the porosity tends to the value of unity

[Roblee etal., 1958 ; Benenati and Brosilow, 1962], the hydraulic radius

tends to infinity. The applicability of the capillary model (discontinuous

system) is restricted to bed porosities less than 0.5 [Bird et al., 1960]. In

61
Literature Survey

order to minimise the error from limitation of the capillary model, Cohen
and Metzner [1981] employed the average porosity to wall region rather

than the local porosity. The average porosity in the wall region, £0w, is

defined as:

= — iedA (2-64) eow A JAxx,

The hydraulic radius for the wall region can be derived from equation (2-

59) to give:
(
' D X X
DP t t^0w
VD P (2-65)
r
H - D - r nD A
+ 6x (l-e 0w )
-x,
DP 'ID,

The average superficial velocity in the wall region can then be determined
by using equation (2-65) for hydraulic radius in equation (2-57), hence

[Cohen and Metzner, 1981]:

Ap %2 r (2-66)
^* Wl _ -~-wx I \
uLB w (n)

The limitations of Cohen and Metzner's model [Cohen and Metzner, 1981]
are the presence of empirical constants (Bw(n), B,(n), and Bb(n)) and the
prediction results that are average superficial velocities at each region
rather than the local velocities. Because of the requirement of velocity

62
Literature Suivev

distribution measurement to determine the empirical constants, using this

model is not practical. In cases where heat transfer or residence time

considerations are critical, as in chemical reactors, this velocity profile

prediction also may not adequately fulfil the requirement, as for example,

analysing of hot spot formation and runaway reactions.

The three region model of Cohen and Metzner [1981] was then adapted

by Nield [1983], who assumed that near the wall only fluid exists, whose

flow may be represented by Stokes equation with the unrealistic slip

boundary condition proposed by Beaver and Joseph [1967]. For the core

region, Nield [1983] assumed that the fluid flow obeys Darcy's law.

Although the mathematical equations that were developed by Nield [1983]

may be solved analytically, however, because of the limitation of the

assumption, the results are questionable. The assumption that near the

wall only the fluid exists is unrealistic for packed beds, because the

porosity never reaches unity for any random packed bed, except for r = R,

where the velocity is zero for no-slip condition. Additionally, the Nield

equation still needs an empirical constant to be determined

experimentally.

Similar to Cohen and Metzner [1981] and Nield [1983], McGreavy et al.

[1986] also divided the flow phenomena in packed beds into regions.

Their model was based on an assumption that the flow of fluid can be

63
Literature Suivey

divided into two zones, shown as the core and annulus in Figure 2-12. In

the core the velocity is assumed to be constant, umc, and governed by the

usual equations for flow through packed beds. The annular region

extends to approximately one to two particle diameters from the wall and

because of the enhanced porosity the velocity is higher. It is also

influenced by the boundary, which will cause the fluid velocity to approach

zero at the wall.

McGreavy ef al. [1986] assumed a continuum model can be applied for

the annular region, and they derived from a momentum balance that:

Ap 1 d . . ,„ ,.
—f- + -—M =0 (2-67)
L r dr

The pressure drop is assumed constant over the cross section of the bed,

and the shear stress is assumed can be expressed by two components,

as follows [McGreavy etal., 1986]:

du
x = x_-\x- (2-68)
dr

Where xd is the drag due to particles that is constant and analogous to

that arising in the core, but is based on different bed porosity.

Substituting equation (2-68) into equation (2-67) and then integrating

gives:

64
Literature Survey

Bed Centre

maximum

u m = f(r)

umm = u"mc

Figure 2-12: A two zoned bed [McGreavy etal, 1986].

65
Literature Survey

du.
2L i_-V dr r = C (2-69)

If the general profile in the annulus is to show a maximum at some radius

rm, then dum/dr = 0 at this point and this defines the constant C.

Integrating equation (2-69) gives the equation of u m , as follows:

um= r2 R2 rm (2_70)
pHf( " )~ 4^
where

p
(2-71)
2L

Which, w h e n r=R gives u m =0, as required. T h e other condition if that at

r=rc the velocity is equal to umc, and this then enables a relationship for rm

to be obtained as follows:

V
MUU, 2T, R
rm =• +- + 1+ (2-72)
m 2B. 2B n B p (R + rc) ^
B p ln B p ln
y^JJ

In order to reduce the errors involved in using the constant value of xd,

McGreavy etal. [1986] proposed that the value of xd proportional to um, as

follows:

66
Literature Survey

x
d = agum (2-73)

Where Og is a constant, that is determined by experiment.

The main limitations are that this mathematical model has empirical

constants and uses an inconsistent interpretation of shear stress.

Basically, the shear stress is function of fluid properties and true velocity

(interstitial velocity) difference [Bird et ai, 1960; Poirier and Geiger,

1994], so the definition of shear stress by McGreavy et al. [1986] in

equation (2-68) is unrealistic and in using it, it is possible to introduce

significant errors. This condition was proved by inconsistency in

interpretation of the value of xd [McGreavy et ai, 1986].

In order to accommodate the oscillation profile of velocity distribution

measured in packed beds, Ziolkowska and Ziolkowski [1993] developed a

model based on fluid kinetic energy dissipation that is not uniform over

the cross section of the bed. The kinetic energy dissipation is a result of

fluid friction against the interface and between fluid molecules. All of these

contributions are represented in a local effective viscosity, which is an

empirical parameter.

In order to overcome the problem of discontinuity of the packed beds

system, Ziolkowska and Ziolkowski [1993] proposed a dimensionless

distance parameter as follows:

67
Literature Survey

R-r
T* — (2-74)
DP

Where the parameter r* was chosen so that the reduced distances from

the tube wall, r*are multiples of DP/4 and the reduced thickness, AT*, of

each individual ring fulfils the conditions:

Ar;=r;-r;_p i = 1,2,3 n (2-75)

The model consists of fluid dynamics equations describing fluid flow

through an arbitrary annular segment of packed tube with a volume :

V.=2.iLr._.Ar. (2-76)

The shear force per unit area was assumed appropriate to modified

Newton's law by Boussinesq [Ziolkowska and Ziolkowski, 1993] as

follows:

x = x1 + x'
v
rz rz rz
p, t /du^ (2-77)
v
Dp•(v + v\dr*
^ k jJ

The turbulent kinematic viscosity coefficient v1, valid for an interval of

radial position Ar* within the system, may be approximated by a constant

function. The density and the laminar viscosity v1 are assumed to be

constant because of the isothermal conditions and negligible effects of the

68
Literature Survey

pressure variations due to the flow. The local bed porosity of an arbitrary

annular bed segment of a region near the wall and extending to r* < 5 of

the tube-to-pellet diameters, was assumed as a linear function

[Ziolkowska and Ziolkowski, 1993].

The equation given by Ziolkowska and Ziolkowski [1993] for the radial

component of interstitial velocity vector of a bed of spherical particles is:

ur=±^(l-e)u (2-78)

The coordination number is positive (+1) when ur is directed toward the

tube wall and negative (-1) when ur is directed towards the symmetry axis

of the tube [Ziolkowska and Ziolkowski, 1993].

The continuity equation given by these authors to develop the

mathematical model of velocity profile across the bed may be written as

follows:

^il = 0 (2-79)
Sr*
and the equation of motion in the axial direction:

13/ r , _, lN dp
,{purruzz+e[x\
L rz z+T[ z])-£+PF^0
rzJ/ (2-80)
Dpdr ^ dz

69
Literature Survey

The axial pressure gradient is determined by using the Ergun equation

[Ergun, 1952], Ziolkowska and Ziolkowski [1993] defined the Ergun

equation as the external force, and the friction (internal) force Fk was

calculated by using Ranz [1952] correlation, given as:

Fk=f^uz2 (2-81)
k
DP

where

f==1 . 75+1 50n-ftlai (2-82)


pDPuM

The boundary conditions of equations (2-79) and (2-80) are

- at the tube inlet, when z = 0:

0<r*<|-, u = uM, p = Po (283)


p

r* = 0, u = 0, p = p0

along the tube wall, when 0 < z < L:

u
1 < K n, r = r- , uz - £ (2-84)

i = n. r*=0, uz =0

at the tube outlet, when z = L:


. R
0<r < — , P = PL
Dp (2-85)
i = n, r*=0, u = 0, P = Pi
2 :
1 < i < n, r. . <r* < r.\ u = -Juz +u,.
Literature Survey

The model solution, determining the radial distribution of the gas

interstitial velocity within a circular packed bed, has been performed by

Ziolkowska and Ziolkowski [1993] for the region near the wall and the

central region, separately. This has been done because the bed porosity

fluctuates along the tube radius up to a distance of r* < 5 from the wall and

approaches a constant value within the bed core when r* > 5 [Roblee et al,

1958; Benenati and Brosilow, 1962],

When r* > 5, the model solution for the bed core is based on the

assumption that the bed porosity within this region is uniform. The

solution of the mathematical model for this region is [Ziolkowska and

Ziolkowski, 1993]:

ApE)
p (2-86)
pLf(l-e0i)

The model solution for the region near the wall of the bed, up to r* < 5 is

[Ziolkowska and Ziolkowski, 1993]:

1
1 ve ve ] ApDpA£
Uz = •+ . . + 4 _ (2-87)
2 D P e oi
D £
v p 0i J pLfe0l

where

v e = -_iL( v > + v . + v n ) (2-88)

71
Literature Survey

f AeV
(2-89)

ri _ 7 . U M D p k £n;e
Oi'
V (2-90)
~4 f_^L
I Ar*
e0i(e2+e Oi'
e= (2-91)
2(1-e]

Within the limits of assumptions made and of developing methods of

equations, equations (2-86) to (2-91) define the velocity profile in terms of

bed porosity, particle diameter, radial position, and local effective

kinematic viscosity, ve. The effective viscosity is a function of the radial

gradient of the bed porosity, the local bed porosity, the laminar

(molecular) and turbulent viscosities, the radial dispersivity, and the

Reynolds number.

The value of the effective viscosity, unlike the molecular and turbulent

viscosities, may be positive or negative depending upon the sign of the

bed porosity gradient, and thus the effective viscosity coefficient may also

be negative or non-negative [Ziolkowska and Ziolkowski, 1993]. The

difficulty arises from the fact that the effective viscosity is an empirical

parameter that is determined from measuring of superficial velocity

distribution at the outlet cross section of the bed.

72
Literature Survey

Based on the measured data at the downstream of the tube for the tube-

to-particle diameter ratio in the range 10.8 < D/DP< 22.9, Ziolkowska and

Ziolkowski [1993] proposed an empirical correlation to predict the

effective viscosity, as follows:

ve
— = Az(Exp(r*Bz))(Cos(2.0557ir*) + 0.45) (2-92)

where

D 'D V
A z =3.419-0.148 + 0.011 ,D ; (2-93)
P
DD

D 6n DV
(2-94)
B z = -0.668 + 0.048 + 0.004VDP ;
D
^p

Comparison of the model proposed by Ziolkowska and Ziolkowski [1993]

with their measured data shows the reported deviations in the range of

20- 30%. The errors probably are due to both the failures in developing

the mathematical model from the fundamentals of the physical correlation

and to turbulent fluid flow behaviour assumption. Another limitation of this

model is the empirical constant, effective viscosity. Thus, using this model

is not practical because of the requirement of the experimental data of the

velocity for the system being considered.

According to Bird etal. [1960], the frictional energy losses are included in

the Ergun pressure drop correlation [Ergun, 1952] rather than separately

73
Literature Suney

accounted by using Ranz correlation [1952], as shown in equation (2-80).

Basically, the Ergun and Ranz correlations are complementary equations

to account for energy losses of fluid flow because of friction which use

different theoretical approach in developing the correlations [Bird et ai,

1960; Poirier and Geiger, 1994]. However, the value of the parameter Fk

in equation (2-80) must be replaced by the acceleration of gravity, g, to

make this equation consistent with the basic equation of motion [Bird et

ai, 1960]. In addition, Ziolkowska and Ziolkowski [1993] also neglected

the term of -(e/(R-r*DP))xrz from the left-hand side of equation (2-80)

without any reasonable reasons. The above factors may also introduce a

significant error into the model.

The use of the turbulent fluid flow behaviour assumption is unjustified in

packed beds, as the work of Mickley et al. [1965]. Their has shown that

momentum transfer inside a packed bed is a function of the gross

properties of the bed, depending on the sidestepping of the fluid stream

as it passes between packing particles, rather than the actual turbulent

structure of the flow. The effect of turbulence is more pronounced on

flow of fluid in the voids between particles [Merwe and Gauvin, 1971a;

Matsuoka and Takatsu, 1996] rather than the flow profile over cross

section of the bed. This explanation is supported by the very much higher

74
Literature Surx-ev

deviation of effective viscosity (22%) measured with repacking, as

reported by Ziolkowska and Ziolkowski [1993],

From the above brief discussion of the available models of the velocity

distribution of flowing fluid through a packed bed reported in the literature,

it is clear that the basic requirement of a mathematical model according to

Reid and Sherwood [1966] has not been reached. This is due to the

nature of packed beds system in which the phenomena of fluid flow

through packed beds include a transition between flow in channels and

flow around submerged objects. There is, therefore, still a continuing

need for research in this area.

75
CHAPTER THREE

CHARACTERISATION OF A PACKED BED

The study of the packing behaviour of granular materials has been carried

out since many years ago over a very broad range of topics [German,

1989], whether the specific objective be the achievement of a dense

packing or the establishing and maintenance of a freely flowing condition.

In practice, there are a variety of factors, which influence the packing

behaviour as summarised by Macrae and Gray [1961] eg. particle,

container, deposition and treatment after deposition. The focus of the

following discussion is restricted to the behaviour of particle packing

related to the velocity distribution of the fluid flowing through a packed

bed.

The importance of the character of the packed beds for the velocity

distribution of fluid flowing in the bed, has been increasingly recognised

over the years by numerous investigators [Arthur et ai, 1950; Morales et

ai, 1951; Schwartz and Smith, 1953; Dorweiler and Fahien, 1959; Cairns

and Prausnitz, 1959; Newell and Standish, 1973; Szekely and Poveromo,

1975; Stephenson and Stewart, 1986; McGreavy et ai, 1986; Ziolkowska

and Ziolkowski, 1993]. According to Brown et al. [1950], the packing

76
Characterisation of a Packed Bed

variables that have significant influence on fluid flow in a packed bed are

as follows: porosity of the bed; size and shape of particles; packing

arrangement of the particles; and roughness of the particles.

3.1 PARTICLE SIZE

It is well known that the particle size, including the size distribution, is th

basic parameter of the packed beds system. For a smooth dense sphere,

the particle size can be accurately defined by a measurement of its

diameter. In most applications, however, as cited by Davies [Fayed and

Often, 1984] accurate particle size measurement is difficult, because most

practical particles are irregular in shape; and therefore, the diameter is a

function of the measurement method.

Generally, the method of particle size measurement is dependent on the

size of particle [Fayed and Otten, 1984; Levenspiel, 1984]. For a non-

spherical particle, the equivalent diameter is used to represent the size of

particle. The equivalent diameter is defined by the diameter of a sphere

having the same volume as the particle [Brown et ai, 1950; Levenspiel,

1984]. Usually, for the packed bed which consists of a number of different

size particles, the size of particles is characterised by mean size or

diameter [Brown et ai, 1950]. The mean size (diameter) of multiple sized

particles, DPm, can be based on diameter, on area, or on volume

[Standish, 1979; Goodling etal, 1983];

77
Characterisation of a Packed Bed

_£n,Dpi
D p (diameter) = (3-1)
Zni

DPm(area) = (3-2)

D _(volume) = (3-3)

and the harmonic m e a n diameter [Standish, 1979; Yu and Zulli, 1994] is:

n-i
n;
D p (harmonic) (3-4)
^ D

Although the consensus has not been reached on the best definition for

the mean size diameter, considering the study of fluid flow in packed

beds, the widely used m e a n size diameter is the volume-surface m e a n

diameter D vs [Standish, 1979; Bird et ai, 1960; Poirier and Geiger, 1994].

This is reasonable because mixtures which have the s a m e value of the

volume-surface m e a n diameter have the s a m e surface area (the total

particle surface/the volume of the particles) as cited by Standish [1979].

The volume-surface m e a n diameter is defined as [Standish, 1979]:

D _!___&: (3-5)

In practice, the mixture data are often performed in mass fraction terms,

and therefore the volume-surface m e a n diameter may be defined as

78
Characterisation of a Packed Bed

[Fayed and Often, 1984; Poirier and Geiger, 1984; Morkel and Dippenaar,

1992]:

3.2 T H E B E D P O R O S I T Y

German [1989] defines the bed porosity or voidage as a volume fraction of

void space in a powder mass. The porosity equals one minus the

fractional density of the bed. The mathematical expression of the bed

porosity as cited by Dullien [Fayed and Otten, 1984] is:

volume of voids in packing


e= (3-7)
bulk volume of packing

The bed porosity or voidage could be broadly categorised by two terms,

that is, mean voidage, e'0, and local voidage, e. The mean voidage is the

fractional free volume in a packed bed. The local voidage is the fractional

free volume in a point at the bed; however, because the point voidage is

not readily measurable, usually the local voidage is defined as a fractional

free volume in an element of bed volume, as in a thin strip or shell.

The nature of packed beds are random systems and cannot be exactly

duplicated [Schwartz and Smith, 1953; German, 1989], and hence, the

experiment is a key factor to describe the characteristics of random-

79
Characterisation of a Packed Bed

packed beds [Blum and Wilhelm, 1965]. But obviously, a mathematical

correlation of packed bed characteristics is often required to solve the

mathematical description of systems, which involve packed beds. A

mathematical correlation of the bed characteristics, with proper error level,

is perhaps expedient and satisfactory for several process-engineering

calculations. In general, a successful mathematical correlation of packed

bed characterisation is based on experimental data which have been

correlated by using statistical approach [Blum and Wilhelm, 1965;

German, 1989; Yu and Standish, 1993ac; Zou and Yu, 1996].

3.2.1 Mean Bed Porosity

In principle, the value of the mean bed porosity depends on size consist

(particle size distribution), handling method and container. The container

has significant influence on the mean bed porosity and is called the wall

effect [Haughey and Beveridge, 1966; Fayed and Often, 1984;

Cumberland and Crawford, 1987; German, 1989]. For example, the value

of the mean porosity of uniform sized spheres in a tube that has a

maximum value at the tube to diameter ratio of about 1.62 and gradually

tends to a constant value for the tube to diameter ratio higher than 10

[McGreavy etal., 1986] as shown in Figure 3-1.

80
- oc

-Mr
CD
_
OO
CD

0 TO
o.
CM CO

>
5 cc
a 03
1.
3 LO O
CM O
•a
03
A
CO
03
q O
c\i CO
a
ro
o
o 'sz
0)
Q. JZ
Q a.
co
o
>
o "co
CO
o
o
Q.
c
q ro
03
0.
E
c
o
*-•

o o
LO 0)
M—
03
i_
03
03
E
ro
T31
CO
0)
03
13i_
33
+J
O)
0 3
LL
£
Characterisation of a Packed Bed

Considering the random packed beds in which particles are randomly

arranged [Blum and Wilhelm, 1965], or all particles of the same size and

shape have the same probability to occupy each unit volume of the

mixture [Debbas and Rumpf, 1966], there are two reproducible states of

packing. These are random dense arrangement and random loose

arrangement [Brown et ai, 1950]. These terms characterise the

configurations, which result when a bed of particles is packed in an

apparently random manner to its densest and loosest conditions,

respectively. The packing arrangement is called dense random packing,

when the particles are poured into container then shaking for about 2

minutes to reduce the total volume. The packing arrangement is called

loose random packing, when it is tipped horizontally then slowly rotated

about its axis and returned gradually to the vertical position [Cumberland

and Crawford, 1987].

In a more detailed categorisation, Dullien [Fayed and Otten, 1984] divided

the random arrangement of packing into four categories; close random

packing; poured random packing; loose random packing; and very loose

random packing. When the bed was vibrated or vigorously shaken down,

the resulting arrangement was called close random packing. Pouring

particles into a container, corresponding to a common industrial practice

of discharging powders and bulk goods, was termed poured random

packing. Loose random packing resulted from dropping a loose mass of

82
Characterisation of a Packed Bed

particles into a container, or packing particles individually and randomly

by hand, or permitting them to roll individually into place over similarly

packed particles. The packing arrangement of the fluidised bed particles

at the minimum fluidisation is called very loose random packing.

The spherical particles is a simplest geometry of packing particles, that is

not a surprising condition if it is taken into account as a basis to develop a

method to predict the packing character [Brown et ai, 1950; Lamb and

Wilhelm, 1965; Levenspiel, 1984; Yu and Standish, 1993b; Zou and Yu,

1996]. Considering the infinite packing for which voidage (the bulk mean

voidage, e'0) is not affected by the presence of external surfaces

[Haughey and Beveridge, 1966], that has been experimentally found for a

tube to particle diameter ratio of from 10 to 15 [Lamb and Wilhelm, 1965;

McGreavy et ai, 1986; German, 1989], the mean voidage of mono-sized

spheres is only dependent on packing arrangement and independent from

particle size [Standish, 1990]. The measured data of the mean voidage of

the mono-sized spheres as a function of packing arrangement are listed in

Table 3-1.

Usually, for non-spherical particles, they are characterised in terms of an

equivalent spherical diameter [Brown et ai, 1950; Levenspiel, 1984; Yu

and Standish, 1993b; Zou and Yu, 1996], called as sphericity.

83
Characterisation of a Packed Bed

Table 3-1: Mean porosity of the packed beds of spheres.

No Arrangement Porosity, e 0 Reference

1 Cubic 0.476 Brown et ai, 1950


2 Face-centered cubic 0.2595 - 2880 German, 1989
3 Body-centered cubic 0.3198 German, 1989
4 Orthorhombic 0.3954 Brown etal., 1950
5 Tetragonal sphenoidal 0.3019 Brown etal., 1950
6 Rhombohedral 0.2595 Brown et ai, 1950
7 Diamond 0.6599 German, 1989
8 Close random packing 0.359 - 0.375 Fayed and Otten, 1984
9 Poured random packing 0.375 - 0.391 Fayed and Often, 1984
10 Loose random packing 0.40 - 0.41 Fayed and Otten, 1984
11 Very loose random 0.44 - 047 Fayed and Otten, 1984
packing Blum and Wilhelm, 1965

84
Characterisation of a Packed Bed

The sphericity concept is more applicable than the concept of packing

size which was proposed by Meloy [Fayed and Otten, 1984], as cited by

Yu and Standish [1993b].

The sphericity, \|/, is defined as the surface area of a sphere having a

volume equal to that of the particle, divided by the surface area of the

particle [Brown et ai, 1950; Levenspiel, 1984], and the maximum value of

the sphericity is equal to 1.0 [Levenspiel, 1984]. On the basis of the

similarity between the packing systems of spherical and non-spherical

particles, the characteristics of the non-spherical particle can be defined

and determined [Yu and Standish, 1993b].

A graphical correlation between the particle sphericity, x\r, and the bulk

mean voidage, e'0, of mono-sized particles have been proposed by Brown

etal. [Brown et ai, 1950; Levenspiel, 1984] as shown in Figure 3-2. Zou

and Yu [1996] have proposed a mathematical formulation for estimating

the bulk mean voidage of cylinder particles and disk particles as follows:

- For the loose random packing:

ln e
( 'o)cyl!nder = ¥558ExP[5.89(l - ¥)]ln(0.40) (3-8)

ln(e'0)dlsk = Y 60Exp[o.23(l - \|/)°45]ln(0.40) (3-9)

85
Characterisation of a Packed Bed

1.0

\\\
0.8-
V\*M Loose packing

/
V \ *
\ \
x *
0.6 x *
x *
-- x *
>_ N

Dense packing

/ s
0 . 4--
-

Nomal packing

0.2 --

1 . . 1 1 . 1 1 L • • 1 • i i
0.0 —1—'—'—'—
0.0 0.2 0.4 0.6 0.8 1.0
Particle sphericity

Figure 3-2: Effect of particle shape on voidage for random-packed

beds of uniform-sized particles [Levenspiel, 1984].

86
Characterisation of a Packed Bed

For the dense random packing:

ln e
( 'o)cylind.. = V674Exp[8.00(l - ¥)]ln(0.36) (3-10)

ln e
( 'oL = V°63Exp[0.64(l-¥)°45|ln(0.36) (3-11)

In order to develop a mathematical correlation to predict the m e a n bulk

porosity of non-spherical particles, Zou and Yu [1996] used the cylindrical

and disk particles as an extreme condition. The mean porosity of non-

spherical particles then may be predicted by using the equation as

follows:

e ' = — — (e'n) +——(e'J (3-12)


V
I + I ^cylinder J _|_ J \ °/disk '

Where

1,.= vi/-Vdisk (3-13)

I = W V cylinder (3-14)

For packed beds that consist of a mixture of particle sizes, the m e a n bulk

porosity principally depends on particle size distribution and handling

method [Standish, 1990]. A general quantitative representation of porosity

of a multi-sized packed bed system is impossible because of the nature

of this system having almost unlimited probability of particles

arrangements for a particular size distribution. It is not surprising that

numerous investigators [German, 1989] in this area tended to be

87
Characterisation of a Packed Bed

concerned either with theoretical, unreal (simplified) conditions or to be

entirely empirical to fulfill the requirement for a quantitative prediction

[Macrae and Gray, 1961]. Fortunately, for a uniform mixture of multi-sized

particles in which the number of size components is more than two, the

feature of the changing of the mean bulk porosity as a function of the

volume fraction of the components is similar to the binary system

[Standish, 1990].

Considering uniform mixtures of multi-sized particles, generally the

investigation of the mean bulk voidage is developed from the results of

investigation of two-sized mixture (binary system) of spherical particles

[Furnas, 1931; Ridgway and Tarbuck, 1968b; Standish and Borger, 1979;

Standish and Yu, 1987a'b; Yu and Standish, 1988; 1991; 1993abc]. By

introducing any arbitrary factor or function into the results of binary

spheres system then the correlation may be extended to characterise both

of multi-sizes mixture systems, namely spheres and non-spheres particles

[Yu and Standish, 1988; 1993b].

For uniform mixtures of the binary spherical particles, the mean bulk

porosity is lower than initial porosity of the former uni-sized particles

[Ridgway and Tarbuck, 1968b; Standish and Borger, 1979; Fayed and

Otten, 1984; German, 1989] as shown in Figure 3-3. The explanation of

Figure 3-3, as given by Standish [1990], is that on addition of fines to

88
Characterisation of a Packed Bed

coarse particles the voids among the coarse particles gradually fill up until

they are all filled and no more fine particles can fit in, and the voida

decreases. If more fines are still forced into the already filled space,

do that by forcing the coarse particles apart, and this increases the to

volume, therefore voidage increases.

Yu and Standish [1988] have applied a general thermodynamics concept

of solutions [Smith and Van Ness, 1975] to determine the mean bulk

porosity of the multi-sized particles bed. They introduced a definition

the initial specific volume of particles, Vi, that may be expressed

mathematically as follows:

1
V
i = 1-e'
(3-16)
Oi

Considering binary system of the bed with the fractional volume of

particles, Xi, equation (3-12) is satisfied [Yu and Standish, 1988]:

'V-V-X-V V V-X.-V2X2
+ 2$ v-v.x,
V, V, v,-i
(3-17)
^V-X,-V2X2^
+ =2
v,-i

89
Characterisation of a Packed Bed

0.1 H — ' — ' — ' — I — ' — ' — ' — | — ' — ' — ' — | — ' — ' — ' — I — ' — ' — ' —
0.0 0.2 0.4 0.6 0.8 1.0
% Volume of large particles

Figure 3-3: Voidage mixtures of binary system for spheres [Yu and

Standish, 1988].

90
Characterisation of a Packed Bed

Where the coefficient Q in equation (3-17) is an unknown parameter of

the Westman equation [Yu and Standish, 1988; Yu et ai, 1993]. For

spherical particles, Q has been reported to be dependent on the size ratio

of large particle diameter (DP2) to small particle diameter (DPi) and this

dependence can be determined empirically [Yu et ai, 1993]. Yu et ai

[1993] proposed the following general correlation to predict the value

of#

f rxx V-566
D^ Lrfrx,

1.355 < 0.824


V Dp2 J ID P2 J
(3-18)
-^Pl
> 0.824
VDP2 )

For given initial mean bed porosities of the binary systems, it is evident

that the mean voidage can be determined by employing equations (3-16)

to (3-18), simultaneously. In applying this method to predict the mean

voidage of the non-spherical systems, Yu and Standish [1993b] proposed

a concept of equivalent packing diameter. The equivalent packing

diameter, Dpe, of a non-spherical particle may be expressed as a function

of its equivalent volume diameter, DPv, and sphericity, as given by [Yu and

Standish, 1993b]:

D Pe 3.6821 1.5040
(3-19)
D Pv 3.17811- V + l|T=—

91
Characterisation of a Packed Bed

Considering multi-sized system of the bed with the fractional volume of

particles, Xi, Yu and Standish [1988] introduced some arbitrary terms into

the thermodynamics equation of solutions, that is called the binary

synergism of the mixture. For binary mixtures, Yu and Standish [1988]

approximated the specific volume equation by the equation:

V = V1X1 + V2(l-X1) + (3I2X1(l-X1) + y12X1(l-X1)(2X, -1) (3-20)

Where the coefficients p 12 and y12 are called the quadratic coefficient and

the cubic coefficient of the binary synergism, respectively [Yu and

Standish, 1988]. These coefficients are only dependent on the initial

specific volume and size ratio of binary systems, being constant for given

initial specific volumes and size ratio [Yu and Standish, 1988].

By extending equation (3-18), the specific volume of n-component

mixtures may then be represented by the following equation [Yu and

Standish, 1988]:

v = Evlxi + £ Spax.xj+ X Xy^x^-xJ (3-21)


1 l<] Kj

Where

v... + v... - v. - v.
p..=-^ * ! L
(3-22)
H,J V
0.4032 '
V
iii-Viii-0.44Vi+0.44 V.
! L
v, = — - (3-23)
r,J
0.177408

92
Characterisation of a Packed Bed

Vy and Vyj are the specific volumes corresponding to the two points:

X; = 0.72, Xj = 0.28, and X| = 0.28, Xj = 0.72 (i < j), respectively, which can

be calculated from equations (3-17) and (3-18).

As mentioned earlier, random packing is a random (stochastic) system,

which can almost certainly not be possible to be explained in an exact

correlation. In other words, all of the representations of packing behaviour

are an approximation. This has prompted numerous types of models to

predict the mean bulk porosity for which the applicability is strictly

dependent upon the basic assumptions of the model development. Beside

the successful model that was developed on the basis of the solution

thermodynamics theory, the correlation which was developed from the

coordination number (the number of neighbouring particles forming

contacts with a given particles) also gave satisfactory results [Ouchiyama

and Tanaka, 1980; 1981; 1989; Fayed and Otten, 1984].

Actually, the components used to construct a particle mixture in

engineering practice are themselves particle mixtures and not mono-sized

particles [Yu and Standish, 1993a]. That is, the particle mixture is usually

a mixture of a number of sub-mixtures of particles and its particle size

distribution is thus a mixture of distributions. Based on intensive studies

in this area, it has been shown that the application of models which were

93
Characterisation of a Packed Bed

developed on the basis of the solution thermodynamics theory [Yu and

Standish, 1988] and the coordination number approach [Ouchiyama and

Tanaka, 1981] are satisfactory to predict the bulk mean porosity of the

multi-sized particles mixture consisting of a number of sub-mixtures of

particles [Yu and Standish, 1991; 1993ac; Ouchiyama and Tanaka, 1989],

From the above brief discussion of the bulk mean porosity of multi-sized

packed bed, it has been shown that the general assumption for model

development of the uniform particle mixture has been reached. For

conditions that the particles mixtures are not homogeneous, i.e., there is

particle size segregation, the applicability of the models still remain a

question mark. As stated by Standish [1990], there are two requirements

that must be met simultaneously for a size segregation to occur, namely,

difference in particle sizes and relative motion between particles. For

conditions for which a size segregation may be suspected to occur, then

the statement of Blum and Wilhelm [1965] that the experiment is the key

factor to describe a random packed bed, is still relevant.

3.2.2 Radial Distribution of the Bed Porosity

The wall of container used to hold a random packing material will induce a

local area of order at the region near the wall [Blum and Wilhelm, 1965;

German, 1989]. The effect is more pronounced for flat, smooth containers

[German, 1989], giving local regions of oscillating porosity in first few

94
Characterisation of a Packed Bed

particle layers near the wall and, it has been shown experimentally,

almost independent of the bulk region [Roblee et ai, 1958; Benenati and

Brosilow, 1965; Thadani and Peebles, 1966; Kondelik et ai, 1968; Scott

and Kovacs, 1973; Goodling et ai, 1983]. A knowledge of this local

variation is important since the microscopic evaluation of fluid flowing

through a packed bed can only be obtained from a knowledge of the local

bed structure and not from the use of bulk properties.

The earliest reported comprehensive experimental investigation of radial

bed porosity distribution was carried out by Roblee et al. [1958]. They

designed experiments to study the influence of the confining wall on bed

voidage in a cylindrical column with randomly packed uniform particle

beds of spheres, cylinders, Raschig rings, and Berl saddles. They

investigated the radial voidage distributions by using the following

method. A packing material was poured into a cardboard cylinder, which

was then filled slowly with hot wax, and then allowed to solidify. After the

wax had solidified, the bed was sawed into circular slabs, which were in

turn sawed into concentric rings. Analysis for bed porosity was made by

first removing the wax from the packing material by dissolving the wax in

boiling benzene, then distilling the benzene to recover the wax. The void

fractions was then determined by calculating the mass of wax recovered

and its density.

95
Characterisation of a Packed Bed

By employing the similar method and filling materials to Roblee et ai

[1958], the porosity distribution of random close packing of uni-sized

spheres was investigated by Scott [1962]. He used about 4,000 steel balls

with 3.175 mm diameter that were poured into a 45 mm diameter and 150

mm long cylinder column. The study was continued [Scott and Kovacs,

1973] to investigate the porosity distribution of an equal number of two

sizes (3.172 mm and 3.567 mm diameters) of steel balls in 45 mm and

125 mm of cylinder columns.

Benenati and Brosilow [1962] investigated the radial bed voidage

variation of spherical particle beds in cylindrical, concave, and convex

columns. They used epoxy resin as filling material, which was introduced

into the bed from the bottom and allowed to flow upwards through the bed.

The function of this method is to avoid the air from being trapped inside

the filling agent, so the more accurate result of bed porosity may be

achieved. After curing the resin, the bed was machined and the layer of

approximately one sixth particle diameter removed each time. Porosity

was determined for each layer by means of the simple material balance

based on the mean density. The packing used consisted of uniform sized

lead spheres and measurements were made for D/DP ratios varying from

2.6 to infinity.

96
Characterisation of a Packed Bed

A non-destructive method on the basis of different absorption of the X-

rays or y-rays in the sphere material and the matrix material was

employed by Thadani and Peebles [1966] to determine the variation of the

local bed porosity over a cross section of spherical particles in cylinder

column. The cylinder vessel was charged with 9.525 mm diameter red

Plexiglas spheres which was then filled slowly with epoxy resin mixed with

araldite catalyst, and then allowed to cure. After curing the resin, the bed

was sawed into slices two particles diameter thick of circular slabs.

Analysis for bed porosity was made by scanning on the micro-photometer

scanning unit.

The similar non destructive technique was applied by Mueller [1992] to

study the radial porosity distribution of randomly packed beds of uniform

sized spherical particles in a cylindrical container. The local bed porosity

of Lucite Plexiglas spheres was analysed by using X-ray radiography. He

investigated the radial porosity distribution of 12.751 mm diameter

Plexiglas spheres that were packed into four sizes of cylindrical

containers. The diameters of cylindrical columns were 25.75 mm, 50.50

mm, 76.00 mm and 101.88 mm (corresponding to D/Dp ratio of 2.02, 3.96,

5.96 and 7.99) and with the height of each of the different cylindrical

columns is approximately 100.00 mm (corresponding to H/Dp ratio of

7.84). The study was continued [Mueller, 1993] to investigate the angular

porosity variation in randomly packed beds of uniformly sized spheres in

97
Characterisation of a Packed Bed

cylindrical containers. The materials, equipment and analysing methods

for the investigation of the angular distribution bed porosity were similar to

his investigation for radial distribution.

The minimum local bed porosity in the near wall region of a cylindrical

column packed with equilateral 7x7 mm cylinders was investigated by

Kondelik et ai [1968]. They used a technique which consisted of pouring

cylindrical particles into a container and then filling all the interstices with

a solution of poly (methyl methacrylate) in methyl methacrylate (Dentacryl,

Dental, Prague). Upon curing the resin, the bed was cut into cylindrical

layers 2-3 mm thick and 10 DP long. The removal of poly (methyl

methacrylate) was carried out by using acetone, the quantity of bed

particles in each fraction was determined by directly weighing.

A non-destructive method on the basis of the fluorescence of a slightly

impure organic liquid and on the refractive index matching of the packed

bed components was employed by Buchlin et ai [1977] to determine the

local voidage of uniformly sized spherical particles in a rectangular

vessel. The vessel was charged with glass spheres which was then filled

with a liquid having same refractory index as the glass spheres to allow

a light beam to cross the bed without scattering. With ethyl salicylate

liquid the light excites a fluorescent re-emission. The interstitial volumes

are therefore selected by taking advantage of this property and then the

98
Characterisation of a Packed Bed

local voidage distribution can be m a d e by manually marking of the photo

that was observed using a camera.

Goodling et ai [1983] investigated the radial bed voidage variation of

uniform and non-uniform spherical particle beds in a cylindrical column.

They used polystyrene spheres as particles and an epoxy resin was used

to fill the void matrix. The packing material was poured into a plastic pipe

of 50.8 mm inside nominal diameter, fixing a small-mesh wire screen over

the top to prevent flotation of the particles in the denser liquid and then

filled with liquid epoxy together with hardener, from the top. After curing

the resin, the bed was cut from the outer periphery over the entire length

of the sample. The local bed porosity was determined for each layer by

means of the simple materials balance based on the mean density. The

measurements were made for D/DP ratios varying from 7 to 17 for uniform

size of spheres and from 7 to 13.5 for multi-sized of spheres.

In a more recent study Stephenson and Stewart [1986] studied the bed

porosity distribution over the cross section of a cylindrical column packed

with cylindrical particles, by employing the optical measurement

technique. Analysis for bed porosity was made by manually marking of the

photo that was observed using a television camera. The markings were

transferred to punched cards via digitising, and then reduced to standard

coordinates by data reduction program.

99
Characterisation of a Packed Bed

Based on the above cited reports of experimental measurements of bed

porosity distribution in packed beds it has been shown that the local

voidage has non-constant or oscillation pattern over the bed cross

sections, especially in the region close to the walls [Roblee et al, 1958;

Benenati and Brosilow, 1962; Scott, 1962; Thadani and Peebles, 1966;

Kondelik et ai, 1968; Scott and Kovacs, 1973; Goodling et al, 1983;

Stephenson and Stewart, 1986]. The oscillation pattern is dependent on

the shape and size distribution of particles [Roblee et al, 1958; Scott,

1962; Scott and Kovacs, 1973; Goodling et ai, 1983] but almost

independent of the shape of the container wall [Benenati and Brosilow,

1962]. For spherical particles, the local bed porosity with ratio of the

particle diameter to container diameter greater than 6.0 is independent of

angular positions [Mueller, 1993].

In the case of uniform size spherical particles, it would be expected the

measured porosity would have a limiting value of unity at the wall,

reaching a minimum at one particle radius from the wall, and a maximum

at one particle diameter. The porosity continues cycling until four to five

particle diameters from the wall before the constant value is reached

[Roblee etal., 1958; Benenati and Brosilow, 1962; Scott, 1962; Thadani

and Peebles, 1966; Goodling et ai, 1983] as shown in Figure 3-4.

100
a
o
c d
•2
a

tj
2
a

q CO
oo a.
a
-«->
_.
re
<N
_.
a
_^
ON re
vo
•xO o
os
,—1 _.
£ a
-C
o s/s cn a. Q.
_3
c_
<U 00
00 OS CO
u. n
I-I
J-t r-H q Q
SO __T, i_
CQ a)
"c. CN o
H—
.3 a, 75
a T3
M—<

a.
Os •t c
M-J
OJ a E o
r-l
C o 3
a. c c.
T3 -Q
<u <u
d)
C- O 0) 'v.
o .r, a 3
3
tr. CQ E- a o (fl
c T3
CQ >
q to
*5 CO
o
--
o
a
ro
TJ
ro
OC
•sf

o CO
cs QJ
5-
3
a>

in
ir-
es
Characterisation of a Packed Bed

The radial variations of the voidage are due to the confining effect of the

wall of the bed. In a randomly packed bed, the layer of spheres nearest to

the wall tends to be highly ordered, in which most of the spheres make a

point contact with the wall of the container with the result of the unity

value of voidage [Goodling et ai, 1983]. The next layer builds up on the

surface of the first, in a less ordered fashion. The subsequent layers are

less and less ordered, until a fully randomised arrangement is attained in

regions far removed from the wall. In condition the particles are

surrounded by a container wall of small ratio of D/Dp, the opposite wall

also affects the particle arrangement. It explains why the measured data

of oscillation pattern [Benenati and Brosilow, 1962; Goodling et ai, 1983]

for the wall distance greater than one particle diameter is dependent on

the ratio of D/Dp.

The similar results also were reported for uniform size of cylindrical

particles [Roblee et al, 1958; Kondelik et ai, 1966; Stephenson and

Stewart, 1986]. The bed porosity has a value of unity at the wall, and then

reaching a minimum at 0.5 - 0.7 particle diameter from the wall [Kondelik

et ai, 1966], and a maximum at about one particle diameter. The porosity

continues cycling until four to five particle diameters from the wall before

the constant value is reached [Roblee et ai, 1958] as shown in Figure 3-5.

102
Characterisation of a Packed Bed

1.25

Stephenson and Stewart, 1986


1.00 - Kondelik et al., 1966
Roblee etal., 1958

0.25 -

0.00 + J I I I I I I I L.
' ' I
0.0 1.0 2.0 3.0 4.0 5.0 6.0
Distance from wall in particle diameters

Figure 3-5: Radial variation of bed porosity for cylindrical particles.

103
Characterisation of a Packed Bed

For highly irregular shapes such as Berl saddles and Raschig rings,

results indicate that the bed porosity decreases regularly from unity at the

wall to the constant porosity at about one particle radius from the

container wall [Roblee et ai, 1966], as shown in Figure 3-6. The constant

bed porosity almost over all of the cross section of the bed could be

expected because of the irregularity in the shape of materials, which does

not allow any appreciable orientation which might result in a definite

pattern.

For multi-sized spherical particles in a cylindrical column, the

measurements indicate that the bed porosity oscillation over the cross

section of the bed is function of the number of sphere sizes that were

mixed as shown in Figure 3-7. Based on Scott and Kovacs [1973] and

Goodling et al. [1983] data, obtained on equal number and equal volume

of multi-sized spherical particles, it may be concluded that for mixtures of

two sizes, regular oscillations are detected only up to 2 or 3 diameters

from the wall and for three sizes the effect of the wall is observed only

within a distance of one particle diameter. The behaviour of multi-sized

spherical particles bed approached the behaviour of the highly irregular

shape particle together with the increasing of the number of particle sizes.

104
Characterisation of a Packed Bed

2.0

Raschig rings
1.5
Berl saddles

1.0

CO

0.5-•

0.0 -I I I I I I I I I L. _i • • i i i i i i i i_
-H- +
0.0 2.0 4.0 6.0 8.0
Distance from wall in particle diameters

Figure 3-6: Radial variation of bed porosity for Raschig rings and

Berl saddles in cylindrical columns [Roblee etal., 1958].

105
Characterisation of a Packed Bed

1.0
Binary-mixture
0.8

0.6
Oo
0.4
o ^ A / *_
0.2 4

0.0
0.0 1.0 2.0 3.0 4.0 5.0 0.0 1.0 2.0 3.0 4.0 5.0
(R-r)/Dp (R-r)/Dp
Dpi = 6.35 mm Dpi - 4.76 mm
Dp2 = 7.94 mm Dp2 = 7.94 mm
Dvs = 7.06 mm Dvs = 5.95 mm
D = 52.6 mm D = 52.6 mm

1.0 1.0
Ternary-mixture Quaternary-mixture
0.8 0.8

0.6 0.6-
Qo
0.4 0 0.4 °>Do r°r ° Ci-p <

Oo 0 9> o

0.2 0.2

0.0 4 0.0
0.0 1.0 2.0 3.0 4.0 5.0 0.0 1.0 2.0 3.0 4.0 5.0
(R-r)/Dp (R-r)/Dp
Dpi = 4.76 mm Dpi = 3.18 mm
Dp2 = 6.35 mm Dp2 = 4.36 mm
Dp3 = 7.94 mm Dp3 = 6.35 mm
Dvs = 6.08 mm Dp4 = 7.94 mm
D = 52.6 mm Dvs = 4.86 mm
D = 52.6 mm

Figure 3-7: Radial variation of bed porosity for multi-sized spherical

particles in a cylindrical column [Goodling etal., 1983].

106
Characterisation of a Packed Bed

From the foregoing brief summary of the experimental work on radial bed

porosity distribution it can be concluded that the behaviour of the bed

porosity distribution near the wall is deterministic. This behaviour is due to

the wall effect, in that the presence of the wall allows to order the

arrangement of packing particles into a certain condition [Blum and

Wilhelm, 1965; German, 1989; Mueller, 1992]. The first layer of particles

in contact with the wall tends to be well ordered with most of the particles

touching it. Subsequent layers are less and less ordered as one moves

away from the container wall. Particles in layers far removed from the wall

display a randomised configuration or called as stochastic system. Thus,

any mathematical description of bed porosity distribution over the cross

section of the bed must be based on a combination of deterministic

approach for the vicinity of wall area, and stochastic approach, for the

centre of the bed area, to achieve a satisfactory explanation of this

system.

In order to develop a method of correlating the local porosity behaviour,

especially in the vicinity of the wall area, an extensive study has been

carried out by many investigators [Haughey and Beveridge, 1966;

Kondelik etal., 1968; Ridgway and Tarbuck, 1968a; Pillai, 1977; Gotoh, et

ai, 1978; Martin, 1978; Cohen and Metzner, 1981; Vortmeyer and

Schuster, 1983; Govindarao et ai, 1986; 1988; 1990; Kubie, 1988;

Johnson and Kapner, 1990; Kufner and Hofmann, 1990; Mueller, 1991;

107
Characterisation of a Packed Bed

1992; 1993]. There are two main categories in mathematical correlations,

namely, phenomenological approach and theoretical approach.

Numerous empirical correlations have been proposed to predict the local

bed porosity of the packed bed systems. Some representative models are

summarised below. Each of these equations contains empirical constants,

which were fitted based upon a particular set of experimental data, and

therefore, the validity of these kinds of correlation is restricted to the

original source of experimental data.

The simplest empirical correlation of radial bed porosity distribution is an

exponential function that was proposed by Vortmeyer and Schuster

[1983]. They employed the exponential function to fit their measured data

of a packed bed consisting of glass spheres with small deviation from the

spherical structure, which in this case only shows one oscillation in order

to approach the average porosity [Vortmeyer and Schuster, 1983]. The

porosity distribution was expressed by the following exponential function:

— = 1 + c'Exp' i - _ * - ' x (2-52)


. DP

where the constant c' can be determined as

l £
c'= '° (3-24)
2.71828e',o

108
Characterisation of a Packed Bed

In order to accommodate the oscillation porosity profile at near the wall

region of a packed bed, Martin [1978] employed the combination of

exponential function and cosinus function to fit the Benenati and Brosilow

[1962] measured data of spherical particles bed. Martin [1978] divided the

cross section of the bed into two zones and introduced a new wall

distance term, Z', that was defined by the following equation:

R-r
Z'=2 -1 (3-25)
Dt

and for - 1 < Z < 0 :

e = e . +(l-e . )Z'2 (3-26)


min V min /

for 0 < Z :

f \
e = e'o+(emin -e'„)Exp ( Z'^Cos 7. (3-27)
4
r J . k
J

where the value of ock is equal to J— for the ratio of D/D p equal to °o

[Martin, 1978].

A similar treatment for fitting the mathematical correlation of porosity

distribution has been proposed by Cohen and Metzner [1981]. They

developed a correlation of experimental data for spherical particles that

109
Characterisation of a Packed Bed

were measured by Roblee et ai [1958] by dividing the cross section of the

bed into three regions, and the correlation for each region is as follows:

R-r
For — — < 0.25:
DD
2^
1-e R-r 7 R-r
= 4.5 (3-28)
1-e' D, v DP ,

R-r
For 0.25 < — — <8.0:
DD

6-6, f R-r A ( R-r A


= 0.3463Exp -0.4273- Cos 2.4509- •-2.2011 n (3-29)
Dp / D, )

R-r
For 8.0 <
DD

6 = 6' (3-30)

A single equation, by combining the exponential function and cosinus for

fitting the experimental data of porosity distribution was proposed by

Johnson and Kapner [1990] for spherical particles and Kufner and

Hofmann [1990] for cylindrical particles. The measurement data of

Benenati and Brosilow [1962] were fitted by the following equation by

Johnson and Kapner [1990]:

/ 0.434 N -\U}\
R-r R-r
(3-31)
e = 0.38 + 0.62Exp -1.70 Cos 6.67
DD Dp _
V

no
Characterisation of a Packed Bed

For cylindrical particles, Kufner and Hofmann [1990] extended the

exponential correlation of Vortmeyer and Schuster [1983] by introducing a

cosinus function term to fit their measured data, as follows:

A f R - r ^ ( R-01
1+ ' 1--V, Exp 1-
v2.71828e'oy Cos 27,
V DPe ,
I DPe )

As mentioned earlier, the D/Dp ratio has significant influence on the

oscillation patterns of local voidage, especially for the wall distance

greater than one particle diameter [Benenati and Brosilow, 1962;

Goodling et ai, 1983]. Therefore, the applicability of empirical correlations

become strictly restricted to systems that have the same value of the

aspect ratio as the original data for fitting the correlation [Mueller, 1991;

1992; Govindarao et ai, 1986; 1988; 1990]. Mueller [1991; 1992]

overcame this problem by incorporating his correlation with an aspect

ratio parameter. Based upon the experimental data from Roblee et al.

[1958], Benenati and Brosilow [1962], Goodling et ai [1983] and Mueller

[1992], he has described the following correlation for predicting the radial

variation of voidage:

e = eb+(l-eb)J0(flr)Exp(-^r), for D/Dp> 2.02 (3-33)

Where

3.15
a = 7.45 - — — , for 2.02 < D/D P < 13.0 (3-34)

111
Characterisation of a Packed Bed

a = 7.45 - — — , for D/D P > 13.0 (3-35)


jU/L) p

0.725

Y= for0
/D ' ^r/DP (3-37)

0.220
8b=0365+ (3 38)
D/D7 "

Although the accuracy of the empirical correlation is good for a particular

set of data, it is difficult to apply this type of correlation with any

confidence to other conditions. Thus, using this type of correlation is not

practical because of the requirement of experimental data on same

condition. It was that reason which motivated numerous investigators to

study the question theoretically, in order to develop a mathematical

correlation of radial distribution of voidage based on the knowledge of

statistics and geometry.

A critical study of local bed porosity variation of spheres was carried out

by Haughey and Beveridge [1966], who developed the semi-theoretical

correlation of voidage distribution in a packed bed based on the

distribution of number of points of contact made by a reference sphere

with adjacent spheres (the coordination number). The coordination

number is determined by sphere center position and particle arrangement.

112
Characterisation of a Packed Bed

The coordination number is not an intensive property of a packed bed, but

is determined by method of packing, geometry of packing, and geometry

of the container [Haughey and Beveridge, 1966; German, 1989].

According to the local bed porosity prediction, Haughey and Beveridge

[1966] proposed a mathematical method only for spherical particles

bounded by a spherical wall container. Considering that the most

common type of packed bed system is bounded by cylindrical or

rectangular column [Perry and Green, 1984], this model is not useful

because its applicability is restricted to spherical wall container.

In a more recent study, Ridgway and Tarbuck [1968a] developed a

mathematical correlation of voidage variation over a cross section of

randomly packed beds of spheres in a cylindrical column. They used an

analytical derivation of bed porosity variation over the cross section of a

regular close-packed hexagonal array of spheres for general voidage

profile correlation of spherical particles. This generalisation is carried out

by introducing two empirical randomising parameters into the equation

that was derived from regular close-packed hexagonal packing.

In an array of close-packed spheres aligned on a flat wall, the bed

porosity profile is an oscillatory function as shown in Figure 3-8. If the bed

is cut by planes parallel to the wall, it can be divided into two types of

113
Characterisation of a Packed Bed

region, namely, A, B, C, etc. are occupied by one layer of spheres only,

whereas a, b, c, etc. are occupied by two interpenetrating layers.

By employing the analytical geometry technique, the local bed porosity for

any distance x of a regular close-packed hexagonal arrangement of

spheres may be expressed by the following general equation [Ridgway

and Tarbuck, 1968a]:

( ^ V
K X-..3PDP Dp-x + J-pDp
.3 J
=1-X- ^ ' - (3-39)
2
4 -P

Where p represents a particular layer, with p = 0 for the first layer, p = 1

for the second layer and so on. Accordingly, as the value of porosity is

between zero and unity, the bracketed terms must be positive or zero; if

one is negative, it is taken to be zero. In addition, similar results were also

proposed by Pillai [1977], Gotoh et ai [1978] and Kubie [1988] to

determine the wall effect on the bulk density variation.

In order to extend the applicability of equation (3-38) for a random packed

bed, Ridgway and Tarbuck [1968a] introduced two correction factors that

were called as randomising factors. The first randomising factor, F^ is to

allow for the voidage increase within a layer over that for a close-packed

array. The second randomising factor, F2 is to allow for the closer

approach of a given layer to the wall compared with a close-packed array.

114
Characterisation of a Packed Bed

0.00 1.00 2.00 3.00 4.00 5.00

Distance from wall in sphere diameters

Figure 3-8: Bed porosity variation of an array of close-packed uni-

sized spheres aligned on a flat wall [Ridgway and

Tarbuck, 1968a].

115
Characterisation of a Packed Bed

By introducing the randomising factors into equation (3-39), the Ridgway

and Tarbuck [1968a] relation for the random packed bed becomes:

Fill
(2 '
X_ PDp
p l
V3
D p - x + ^-pDpF2 (3-40)
1-1 —
^

Ridgway and Tarbuck [1968a] used the measurement data of Benenati

and Brosilow [1962] to fit the following correlation of randomising factors.

9 F\
F = -^(0.62 + 0.18e-0J6p) (3-41)

F2=0.991p (3-42)

The validation of the correlation, performed by making comparison of the

predicted results with the measurement data of Benenati and Brosilow

[1962] was reported by Ridgway and Tarbuck [1968a]. The root mean

square deviation between experimental results and prediction was 0.02

and the standard deviation of the experimental results was 0.01. It may be

thus concluded that the performance of the correlation is quite good;

however, it requires two empirical factors to be known from measured

data of porosity distribution.

A multi-channel model for estimating the local voidage profile in a

randomly packed bed of uniformly sized spheres in cylindrical column has

11.5
Characterisation of a Packed Bed

been proposed by Govindarao and Froment [1986]. The bed w a s divided

into a number of concentric cylindrical layers q of equal thickness. They

provided procedures for predicting the voidage variation up to distances

of five particle diameters from the wall. However, this model still requires

an empirical constant, but since they incorporated the aspect ratio

(a = D/Dp)into the model, therefore the applicability can be extended.

In order to simplify the mathematical manipulations, Govindarao and

Froment [1986] chose the thickness of cylindrical concentric layers, such

that m = DP/2Ar, as a suitable integer, and for realistic aspect ratios, the

effect of the curvature of a cylindrical concentric layer was neglected.

Consider the ith cylindrical concentric layer, the volume of spheres whose

center is in jth cylindrical concentric layer v^ may be calculated by

[Govindarao and Froment, 1986]:


. iAr

v , = — Jv(r)dr (3-43)
A I
(i-l)Ar

If Nj is the number of spheres whose center is in jth cylindrical concentric

layer, the total volume of solids in the ith cylindrical concentric layer is

given by [Govindarao and Froment, 1986]

V1=^NJviJ (3-44)
JI

117
Characterisation of a Packed Bed

where ji = 1 + m and j2 = i +rafor i < 2m,'^= 1 - m and j2 = i +rafor i > 2 771.

The voidage in the ith cylindrical concentric layer is then [Govindarao and

Froment 1986]

6i = 1 - (3-45)
TcAr'Lgj

where

g- = 2am-2i + l (3-46)

By solving equation (3-43) for different spherical slices and caps and then

substituting the results into equation (3-45) and defining x\, as the numbe

fraction of spheres with centers in the jth cylindrical concentric layer, gi

[Govindarao and Froment 1986]:

( 1 "\ i+m-I
_h_
e ; =_.-• n : j m —— + 3Xnjb, i < 2m (3-47)
Si V 4; j=m+l

and

\ i+m-l
6; = 1 — ( n i - ™ + n i + J m - - +3Xi-.b i > 2m (3-48)
Si j=m+l

where

b a =/n 2 -i 2 +j(2i-j)-- (3-49)

, N T Ar
h= —!— (3-50)
3L

118
Characterisation of a Packed Bed

The value of the number fraction {ni = N j / N T ) is an empirical parameter

[Govindarao and Froment, 1986]. Based upon the experimental data from

several investigators Govindarao and Froment [1986] have shown that

n, =n2...= nm = n„,+2 = nm+3 = ..= n3n(_1 =0 (3-51)

and Govindarao and Ramrao [1988] have described the following

correlations for predicting the two number fractions in terms of the aspect

ratio:

3 08
" (3-52) "m+l
a

2.60
niffl — (3-53)
a

Based upon the comparison of the local bed porosity, predicted by using

the procedure proposed by Govindarao and Froment [1986], to

measurement data has shown that the applicability of the procedure is

restricted up to distances of about two particle diameters from the wall

[Govindarao and Froment, 1986]. Additionally, this procedure also gives

an uncertain result for infinite value of the aspect ratio.

3.3 PERMEABILITY

Standish [1979] defined the bed permeability as the ability of a given

packing to allow a fluid to flow through it under given conditions. The

permeability is a function of four variables, namely particle size and

shape, bed voidage and geometric factor [Standish, 1979]. For packed

119
Characterisation of a Packed Bed

beds of multi-sized particles, the permeability is also affected by the

spread and size range of a particle size distribution [Yu and Zulli, 1994].

The permeability of the bed K is defined by Darcy's law [Bird et ai, 1960;

Fayed and Otten, 1984; Wilkinson, 1985; Kececioglu and Jiang, 1994]

uM=--(Vp + pg) (3-54)


M-

A large number of efforts has been expended on determining K for various

packed bed systems [Bo et ai, 1965; Standish and Leyshon, 1981;

Standish and Collins, 1983; Leitzelement et ai, 1985; MacDonald et ai,

1991; Yu and Zulli, 1994; Kececioglu and Jiang, 1994]. The following

semi-empirical expression has been found to accurately represent many

experimental data [Bird et ai, 1960; Kececioglu and Jiang, 1994; Poirier

and Geiger, 1994]. It is

where av is the surface area per unit volume of particles, and K' is an

experimentally determined constant and it has been found to equal 4.17

[Ergun, 1952].

The quantity av for spherical particles is defined by the following equation

[Bird etal., 1960; Poirier and Geiger, 1994]:


Characterisation of a Packed Bed

6
a v =D—P (3-56)

Substituting equation (3-56) into equation (3-55) and then inserting of the

value of the constant K" equal to 4.17 into the result, then gives

2.3
D/E
K= (3 57)
I_x7-_F "

The constant K" equal to 4.17 is not universally selected, s o m e believe the

value to be as high as 5.0 [Kececioglu and Jiang, 1994; Poirier and

Geiger, 1994].

The applicability of equations (3-56) and (3-57) is restricted to spherical

particles. As stated earlier, for non-spherical particles, equations (3-56)

and (3-57) may be used by introducing the sphericity concepts [Poirier

and Geiger, 1994].

According to Forchheimer's generalisation for pressure drop correlation of

fluid flowing in a packed bed [Kececioglu and Jiang, 1994], the energy

losses of fluid flow is due to inertia energy losses and kinetic energy

losses as shown in Ergun equation [Ergun, 1952]. The kinetic energy

losses and the inertia energy losses are expressed by bed permeability

and inertia parameter, respectively [Fayed and Otten, 1984]. According to

121
Characterisation of a Packed Bed

Ergun equation [Ergun, 1952], the inertia parameter appears to be

independent of the fluid properties and may be written as follows [Fayed

and Otten, 1984]:

Dp£3
I=
mTi) 0-58)

Comparing the permeabilities given by equations (3-57) and (3-58) with

equation (2-37), the Ergun equation, it is obvious that these equations

give the characteristics of the bed for viscous and inertial flow in the

Ergun equation, while the other terms, namely, p, \x and uM characterise

the flowing fluid.

An important assumption in the above definitions of permeability is that

the porosity is uniformly distributed throughout the bed. This condition is

equivalent to the geometric factor of the bed being equal to unity

[Standish, 1979]. However, as also pointed out by Standish [1979], this

condition is rather unusual and rarely met with in practice, where the

geometric factor is hardly ever unity due to the use of packings having a

size distribution and/or non spherical shape, causing particle segregation

of one kind or another. This definitely complicates prediction of

permeability distribution, at least by any model that seeks to achieve this

without any prescribed information.

[22
Characterisation of a Packed Bed

Recently, it w a s demonstrated [Yu and Zulli, 1994] that a good prediction

of radial permeability distribution in a blast furnace is possible if a

measured radial particle size distribution is given. Considering the

complexity of the system involved, namely coke and sinter of different size

distributions and absolute sizes (sinter: 5-20 mm and coke: 25-70 mm),

the reported good result may be regarded as an important achievement

that will undoubtedly be improved and extended in time to beds with

different geometric factors.

It is of interest to observe that in this regard, Yu and Zulli [1994] noted the

need for understanding the microstructure of packing of particles, giving

as an example a binary size system in a mixed state and in a segregated

state. The large difference in permeability in this example as given in the

paper [Yu and Zulli, 1994], is a direct result of a different geometric

arrangement of the same packings in the bed, ie. a different geometric

factor, viz \|/=1.0 f°r tne uniformly mixed bed and \j/<1.0 for the

segregated bed used.

It is also of interest to observe that Yu and Zulli [1994] were motivated in

their research "Because the radial permeability distribution is directly

related to the radial gas distribution in a blast furnace. It provides a more

quantitative and useful information for the process control then the radial

particle size distribution". Noting the above stated radial permeability-gas

123
Characterisation of a Packed Bed

distribution connection, it is suggested that a possibility of employing a

mathematical model of velocity distribution of a fluid in packed beds, as,

for example, the model proposed in the present work, be investigated

further.

124
CHAPTER FOUR

DEVELOPMENT OF A MATHEMATICAL
MODEL FOR VELOCITY PROFILE OF FLUID
FLOWING IN PACKED BEDS

Generally, there have been two main theoretical approaches for studying

flow conditions in packed bed systems. In the first approach the packed

bed is regarded as a bundle of tangled tubes; the theory is then

developed by applying the previous results for single straight tubes to the

collection of crooked tubes. In the second approach the packed beds is

visualised as a collection of submerged objects, and the point of view is

extended from conditions of submerged particles. For macroscopic level,

the first approach has been successful for bed porosities less than 0.5

[Bird etal., 1960].

In order to develop a general mathematical equation which may be used

to evaluate the velocity distribution of single phase fluid flow in the packed

bed, the fluid flow phenomena is treated by using the above first

approach. Considering the limitation of this approach that the applicability

is restricted to the bed voidage less than 0.5 [Bird et ai, 1960; Cohen and

Metzner, 1981; Foscolo etal., 1983] and the variation of the voidage over

the cross section of the bed as stated in the previous chapter, it has

125
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

seemed desirable to introduce s o m e modification to improve the

performance of the model. The improvement can be carried out by

considering the fluid flow phenomena in a packed bed as a continuous

system for the local voidage greater than 0.5, and as a discontinuous

system for the local voidage less than 0.5.

4.1 THE EQUATION OF FLOW THROUGH A SINGLE PIPE

The energy relationships of a fluid flowing through a pipe may be obtained

by an energy balance. Energy is carried with the flowing fluid and also is

transferred from the fluid to the surrounding, or vice versa [Brown et ai,

1950]. The energy carried with the fluid includes the internal energy, U,

and the energy carried by the fluid because of its condition of flow or

position, namely potential energy, kinetic energy and pressure energy.

The energy transferred between a fluid or system in flow and its

surrounding is the heat, q, [Brown etal., 1950; Foust et ai, 1960].

An energy balance around a flow system, such as between points 1 and

2 in Figure 4-1 and the surroundings, assuming steady state condition

(no accumulation of material or energy) at any point in the system is given

by equation (4-1) [Bird etal., 1960]:

Rate of Rate of Rate of


Energy > = < Energy >• — < Energy (4-1)
Accumulation Input Output

126
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

Point 2 Fluid Outlet


-i — z2

Zi
Point 2-
Fluid Inlet

Figure 4-1: Flow diagram of the fluid flow in a pipe.


Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

Assuming the direction of fluid flow is only in the z-direction, then equation

(4-1) gives:

AU + -muz2+A(mgz)+A(pV)=q (4-2)

The increase in internal energy is the s u m of the increases due to all

changes considered as taking place in the material in flow, including heat

effects, compression effects, surface effects, and chemical effects [Brown

etal., 1950].

AU = j*TdS + |2p(-dV)+|2odt+|2KAdmA + f KBdmB +etc. (4-3)

The pressure energy term, A(pV) is complete differential:

A(pV) = Ji2pdV + jVdp (4-4)

Combining equations (4-2), (4-3) and (4-4), including surface and

chemical effects in the etc. term, gives:

J] TdS + A -muz2 +A(mgz) + J Vdp + etc.= q (4-5)

The increase of the internal energy due to heat effects, TdS, as cited by
•M

Brown et al. [1950], is equal to the sum of the heat absorbed from the

surroundings and all other energy dissipated into heat effects within the

128
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

system due to irreversibility, such as overcoming friction occurring in the

process,

Ji2TdS = q + (lw) (4-6)

where the lost work (Iw) is energy that could have done work but was

dissipated in irreversibility within the flowing material.

Combining equations (4-5) and (4-6) gives,

r2 fl }
J Vdp + A - m u z 2 +A(mgz) + etc.= -(lw) (4-7)

Equation (4-7) is a general equation of fluid flow in a pipe and is

unrestricted in application to material flowing or transferred from state 1 to

state 2, except for unsteady state condition and the presence of shaft

work.

Assuming the fluid flowing through a pipe is free of chemical change,

surface effects, etc., equation (4-7) may be written, for a unit mass of

material as:

f - d p + ^ - A u z 2 + A Z = -F h (4-8)
g 2g

Where:

F h = ^ (4-9)
mg

129
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

The energy per unit mass lost as frictional conversion into heat, Fh, m a y

be calculated by the following equation [Brown et ai, 1950; Bird et ai,

1960; Foustefa/., 1960].

Fh=^^ (4-10)
2gd p

Substituting equation (3-10) into equation (4-9) gives:


2 2
r2 V Au fXu
k
—dp + —^- + Az+ / =0 4-11
Jl
g 2g 2gd p

Equation (4-11) is applicable for straight pipes with constant diameter and

with no heat or work transferred. Brown et a/.[1950] introduced the

concept of equivalent length, Le, for describing fluid flow through non-

regular piping sections such as bends or sections with changing cross

sectional area. The value of Le is the length of pipe itself plus an

equivalent length allowance for non-regular piping sections, the friction

energy losses equal to straight line pipe of length Le.

4.2 THE EQUATION OF FLOW THROUGH A PACKED BED


Considering fluid flowing through a packed bed with voidage less than

0.5, it can be assumed that the flow phenomena is similar to fluid flow

inside a bundle of tangled tubes [Bird et ai, 1960; Cohen and Metzner,

1981] with radius re. By using a definition of hydraulic radius, rH, [Bird et

130
Development of a Mathematical Model for Velocity Profde of Fluid Flowing in Packed Beds

ai, 1960; Foust, etal., 1960; Poirier and Geiger, 1994] then the quant

may be expressed in terms of the hydraulic radius, as follows:

re = 2rH
_d^ (4-12)
2

The hydraulic radius is defined by Bird etal. [1960] as follows:

cross section available for flow


H
wetted perimeter
(4-13)
volume available for flow
total wetted surface

By neglecting the wall effect, the hydraulic radius for a bed composed of

spherical particles can be shown [Bird et ai, 1960] to be:

EDD
r_
H = (4-14)
"6(l-e)

Mehta and Hawley [1969] have attempted to account for the wall effect by

modifying the expression for the hydraulic radius. Their modification takes

the form:

ED,
rHu = ... \.. (4-15)
6(1-e)M

where

4D
M = 1 + (4
f l & '16)

131
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

Although originally the approach of Mehta and Hawley [1969] w a s

proposed to apply over the entire cross-section of the column, Cohen and

Metzner [1981] suggested to apply this approach only at the wall region,

because the effect of the wall should be confined to the wall region

without affecting the nature of the hydraulic radius in the bulk of porous

bed.

W h e n the bed contains non-spherical particles and the material is

screened, and then the particle diameter, DPs, may be taken as the

arithmetic mean of the openings of the two screens. By introducing the

sphericity concept into equation (4-14), and then the hydraulic radius may

be calculated by the following equation [Poirier and Geiger, 1994]:

r =^^ (4-17)
H {
6(1-e) '

Considering the fluid flow in packed beds, the value of equivalent length,

Le, can be defined by the following equation:

Le=r;Az (4-18)

Interstitial velocities can be calculated using the following equation

[Cohen and Metzner, 1981] which relates the interstitial velocity with bed

voidage:

132
Development of a Mathematical Model for Velocity Profde of Fluid Flowing in Packed Beds

uM
uz=-fL (4-19)

By assuming the value of the correction factor, l%, is constant over the

cross section of the bed with local voidage less than 0.5 and replacing L

in equation (4-11) with Le, then substituting equation (4-12) into equation

(4-11) gives:

nV Au7 '1 + f_W^= 0 (4-20)


J —gd p + —_-^-
2g + Az
(^ 8grH /

The friction factor fk in equation (4-20) is calculated by using the Blake's

correlation [Ergun, 1952], as follows:

4 = 1.75+150^ (4-21)

Where

NRe="^L^ (4-22)
H

Because the applicability of equations (4-20) to (4-22) is restricted to bed

voidage less than 0.5, and considering the voidage higher than 0.5

usually occurs only at the wall region of the random packed beds,

equation (4-14) or equation (4-17) is taken into account for calculating the

hydraulic radius.

133
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

As discussed earlier, the local bed porosity has a limiting value of one at

the wall for the particles which have possibility to make point contact with

the wall of the container and then continuously decrease to reach a

minimum value at a distance about a half particle diameter from the wall.

In order to overcome the limitation of equations (4-20) to (4-22), which is

restricted to the bed voidage less than 0.5, then for the wall region the

mathematical equation for the velocity profile calculation may be

formulated by assuming the flow characteristics as a continuous system.

Consider the fluid flowing through a packed bed at the wall region, no-slip

condition is assumed at the wall or the velocity is equal to zero at r = R

and .is a radial position in the bed which has voidage equal to 0.5.

Employing the momentum balance [Bird et ai, 1960; Poirier and Geiger,

1994] in the wall region with the local voidage higher than 0.5, assuming

the flow of fluid only driven by the difference of momentum and that the

fluid is Newtonian, the following equation can be generated:

3u, z
re =0 (4-23)
3r 3r

The boundary conditions are;

1. at r = £, 6 = 0,5 uz = u, (4-24)

2. at r = R, e = 1.0 uz = 0 (4-25)

134
Development of a Mathematical Model for Velocity Profde of Fluid Flowing in Packed Beds

4.2.1 The Equation of Continuity

For a volume element (Figure 4-2), the material balance of the fluid flow is

as follows [Bird etal., 1960]:

Rate Rate Rate


of r-< of r = < of (4-26)
Input Output Accumulation

Which, with assumed steady state condition and no chemical reaction,

gives;

nre2(puz), )-[nre2{puz\ ' <) (4-27)


|z / \ U + AZ

and for Ar,Az -> 0, from equation (4-27) gives:

a(puz)
=0 (4-28)
3z

Or

dU, d
P rx
z
+ UM,-^ = 0
(4-29)
az az

4.2.2 The Incompressible Fluid

a
Considering that for the incompressible fluids, the value of —P is equal to

zero then equation (4-29) may be rearranged to give:

9UM
=0 (4-30)
3z

135
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Bed.

FLUID O U T L E T
_.Ar „

Z=L

r
R **—•

Z=Z+AZ
Z=Z

z=o

FLUID INLET

Figure 4-2: Diagram of the fluid flow in a packed bed.


Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

For an incompressible fluid the value of V is constant and equal to -


P
then by substituting equation (4-30) into equation (4-20) gives:

1 9p f&u2
~^T+ _ +1 = 0 4-31)
Pg dz 8grH

Substituting equations (4-14), (4-19), (4-21) and (4-22) into equation (4-

31) and then the result solved for positive value of uz is;

Uz = -^+A-2-4^ (4_32)
2A,,

Where

(1-e) pt
...=1.3125- £ -f*-
Dp (4-33)

(1-E) 2
x2 — 1
1 11Z.0
0. r,
^
(4-34)
£ Dp 2

dp
^ =-^-+pg (4-35)
dz

Considering the incompressible fluid, because the interstitial velocity is

independent of the axial position, equation (4-23) with the two boundary

conditions, equations (4-24) and (4-25), may be solved by using a

polynomial approximation [Burnett, 1987] and assuming that in the region

of concern porosity is a linear function of r, as follows:

(r-R)
8= 1
"2T7-R)' for£<r<R (4-36)

137
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

For the problem at hand, the following form of a complete linear

polynomial is chosen:

uz = k0 + k.r + k2r2 (4-37)

Where the coefficients k0, h and k3 are constants to be determined which

all boundary condition must be satisfied exactly. By using the collocatio

method [Burnett, 1987] to optimise the best value of k0, k: and k3, a

quantity called the residual, SH, is required. The residual is derived by

inserting equation (4-37) into equation (4-23) and gives:

9.= ^2i~R'kx +[2(21 -R)k2-k, }-3k2r2


2 (4-38)
= 0

By using the boundary conditions, two of the constants can be

determined. Then the rest of the constants may be determined by forcing

the residual to be exactly zero at a point, n in the domain. The followin

approximate solution of the constants for equation (4-37) is obtained by

employing this procedure.

/.0=-Rf>.+_-2R) (4-39)

5.+R
*,= ,, ^ 2u , (4-40)
3(.-R)

*a = U ' - * ' ( ' T R ) (4-41)


2 2 2
. -R

138
Development of a Mathematical Model for Velocity Profde of Fluid Flowing in Packed Beds

The value of the correction factor, \%, may be evaluated by using the

equations derived from a material balance of the system and assuming

steady state condition, using the following equations:

7.R2puM
m= P1- (4-42)
4
and
R

m = 27ipJ r£uzdr (4-43)


o

By minimising the deviations between the macroscopic result (equation

(4-42)) and the microscopic result (equation (4-43)) of the mass rate, the

value of the correction factor, £, can be determined implicitly.

Equations (4-32) to (4-35), (4-37) and equations (4-39) to (4-43) are a

complete set of the equations to predict the velocity profile of single-

phase incompressible fluid flow in packed beds. Use of these equations

requires only the knowledge of physical properties of the fluid, bed

characteristics and macroscopic data. Although these equations are

derived for uni-sized spherical particles, the applicability can be extended

for multi-sized spherical particles by employing the volume-surface mean

diameter for particle size or for non-spherical particles by employing the

sphericity concept.

139
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

4.2.3 T h e Compressible Fluid

Because the density or the specific volume of a compressible fluid is not a

r V
constant and is a function of the pressure, the term — d p in equation (4-
J
g
19) can not be directly solved. The solution of this term requires a

knowledge of the correlation between specific volume, V, and pressure.

This difficulty is complicated by the fact that the fluid velocity, uz, besides

being a function of bed radius, also depends upon the axial position in the

bed. These conditions lead to an increase in the complexity of the solution

of equations (4-20) and (4-23) to get the velocity profile of compressible

fluid flow in a packed bed.

On the other hand, based upon the measured data that were obtained at

both the inside of the bed [Stephenson and Stewart, 1986] and the outlet

of the bed [Price, 1968; Newell and Standish, 1973; Szekely and

Poveromo, 1975; Ziolkowska and Ziolkowski, 1993], for compressible and

incompressible fluid, it is shown that the velocity is more a function of the

bed character rather than that of the superficial velocity or the flow rate of

the fluid. This condition makes the assumption that a constant value of the

specific volume of compressible fluid over a small value of axial distance

of the bed becomes reasonable. This assumption is considered to simplify

the algebra considerably for the solution of the equations (4-20) and (4-

23).

140
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

The bed with large value of L/D is divided into finite number, N, of short

beds and the assumption of small pressure drop would be valid for every

short bed. The approximation is expected to approach the exact condition

as N^oo. For conditions of a small value of the pressure drop or the

value of the length increment (Az->0), equations (4-32) to (4-35), (4-37)

and equations (4-39) to (4-43) can be employed to calculate the velocity

distribution of compressible fluid flow in packed beds.

4.2.4 Pressure Drop Correlations for Packed Beds

A very large number of relations have been proposed for estimating the

pressure losses of a fluid flowing through a packed bed as discussed by

Brown, et ai [1950]; Bird et al. [1960]; Foust et al. [1960]; Perry and

Green [1984]; Agarwal and O'Neill [1988]; Leva [1992]; Poirier and Geiger

[1994] and Kececioglu and Jiang [1994], but only a notable few will be

described here. The principal reasons for not discussing the others were

their poor validity, limited applicability and complexity of the correlations.

However, because of the pressure drop correlation being an additional

equation for predicting the velocity distribution, the other correlations can

also be applied with regard to the availability of data and the validity of

correlations, if required.

The most widely used mathematical correlation for single phase fluid flow

in packed beds is that advanced by Ergun [1952], who proposed the

I4l
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

prediction of pressure drop due to friction losses based on mechanical

energy balance by using Blake equation of friction factor (equation (4-

21)). Although the theoretical explanation of Ergun correlation has been

made by Bird et ai [1960], actually this correlation is an extension of the

empirical correlation that was proposed by Forchheimer in 1901

[Kececioglu and Jiang, 1994]. Generalising Forchheimer's equation gives;

-^^afuM+bfuM2 (4-44)
dz

where a{ and bf are empirical constants.

In 1952, Ergun [1952] examined this general expression for gas flow

through crushed porous solids, based on its dependence upon the flow

rate, properties of fluid, porosity and character of the bed particles. He

obtained the following equation:

__-.15Q.k£^__f.M.751-E,'°PU"i (4-45)
L e'0J D p 2 e'0! Dp

When a bed contains a mixture of different-size particles and if the

material is screened to determine the diameter of each size faction,

Poirier and Geiger (1994) suggested to use the volume-surface mean

diameter, Dvs and then the sphericity, y, is introduced as a factor into

equation (4-45) to give:

142
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

1 2
^ = 150%^i^
2 + 1.75 "^^ , (4-46)
L e'0> Dv/ E V DVSV

As mentioned earlier, the applicability of the Ergun correlation is restricted

to the bed porosity of less than 0.5 [Bird et ai, 1960], and it is not

appropriate for systems containing particles of low sphericity [Gauvin and

Katta, 1973; MacDonald ef ai, 1979].

4.3 FLUID FLOW AT THE OUTLET OF THE BEDS


Although a considerable progress has been made in the development of

experimental techniques for investigating velocity distribution inside a

packed bed, they were limited to special situations. The optical technique

as employed by Stephenson and Stewart [1986] is restricted to

incompressible fluids. The use of a laser Doppler anemometry to measure

the velocity profile [McGreavy et ai, 1986] is only good for small values of

D/Dp ratio. The use of velocity sensing probes inside the bed could disturb

the packing arrangement and because the fluid velocity between the

packing particles is not uniform, as stated by Mickley et al. [1965], many

measurements of local velocity are necessary to give a true indication of

the mean axial velocity. The measurement of velocity profile by noting the

time taken for a step change in the electrical conductivity of the fluid to

travel between two fixed points in the bed [Cairns and Prausnitz, 1959] is

143
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

not satisfactory if radial mixing occurs in regions where velocity gradients

exist.

The measurement in the down stream of the bed is not directly

representative of the flow profile inside the bed because of the changing

of flow profiles. Based upon the result of previous investigators [Schwartz

and Smith, 1953; Newell and Standish, 1973; Cohen and Metzner, 1981;

Nield, 1983; Vortmeyer and Schuster, 1983; McGreavy et ai, 1986;

Tsotsas and Schluder, 1987, 1990; Stanek and Szekely, 1974; Szekely

and Poveromo, 1975; Ziolkowska and Ziolkowski, 1993], the difference in

the velocity profiles at the inside and the exit of the packed beds is clearly

shown. Use of the flow divider [Arthur et al. 1949; Price 1968; Newell

1971] is faced by problems in equalising the pressure losses through the

different passages. A schematic diagram of the fluid flow phenomena in

packed beds is shown in Figure 4-3. Clearly, the model of the velocity

distribution of the fluid flowing in packed beds needs to be corrected to

allow comparison with the data that were obtained at the outlet of beds.

A velocity profile at the outlet of the bed is a transition profile between the

inside bed profile and the fully developed flow of the fluid in an empty

pipe, which may be assumed as a developed flow profile in a pipe.

Numerous studies have been conducted for investigating the developed

flow profile [Langhaar, 1942; Foust et ai, 1960; Christiansen and

144
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

FULLY DEVELOPED FLOW

T T T TTTTTTTTT T T
EXIT OF BED FLOW

PACKED
BED
INSIDE BED FLOW

TTfTTTTTTTTTTTTTTT
FLUID INLET

Figure 4-3: Flow of a single-phase fluid in a packed bed.

145
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

Lemmon, 1965; Vrentas etal., 1966; Vrentas and Duda,1967; Atkinson et

ai, 1969; Chen, 1973; White, 1986].

In order to develop the velocity distribution model at the exit of the b

is assumed that there is no-slip at the wall, steady state conditions ap

the fluid is Newtonian with constant density and viscosity, no pressure

gradient for r-direction, and any angular motion is negligible (axisymme

flow). The following equations can be derived from momentum balance

and material balance [Bird et ai, 1960]:

f duz du z ^ 2
9p <___ a2„
u >
(4-47)
U +UZ + pg + H +• dz2
.'"37 "CJ7 ^3z Vf dr v dr

2
-„ A
f dur d I a a u (4-48)
u.1 dr U, = n (m +
dz L3r7a7 ^ "a?

1 d dUr
M+ =0 (4-49)
rd~r Tz

The boundary conditions are;

1. at z=0 u z = uzo(r) and ur = 0 (4-50)

2. at z=oo uz = uZL(r) and ur =0 (4-51)

3. at z>0 uz(R,z) = 0 and ur(0,z) = 0 (4-52)

146
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

The value of uzo(r) is the velocity distribution at the top of the bed, that

can be calculated from the velocity distribution inside the bed model

(Equations (4-32) to (4-35), (4-37) and equations (4-39) to (4-43)) times

the local bed voidage. The value of uZL(r) is the fully developed flow

distribution. The equations of motion for developed flow have been solved

numerically and analytically for many simplified cases by many

investigators, as discussed by Atkinson et ai [1969], Vrentas and Duda

[1967], Vrentas et ai [1966], Christiansen and Lemmon [1965], and

Langhaar [1942]. However, all of these solutions use the boundary

condition that the velocity at the entrance, uzo, is independent of the

radius, which is different for flow conditions at the exit of the bed. There is

a need to solve the equation of motion, equations (4-47) to (4-49) by

using different boundary conditions (equations (4-50) to (4-52)).

dimensionless
In order to simplify the problem at hand, the following

variables are introduced into equations (4-47) to (4-52);

u
z
(4-53)
<p = -^-

U
r
(4-54)
n= —

(4-55)

z
(4-56)

147
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed

R ap
v=U (4-57)
M
2
P az + p g

Then, the partial differential equations and the boundary conditions

become:

dry dv 2 1a f dfU
_*^ a2^
N R e ^ a c i C ac + ds2 (4-58)
ac as

dfU du 2 (a i a a2^ (4-59)


K— + V—--.—-ac C ac ^ ' as
ac as N R e
1
^ tr \ dP r. (4-60)

and the boundary conditions are;

1. at 5 = 0 V =% and «= 0 (4-61)

2. at 5 = oo f = # and « = 0 (4-62)

3. at s > 0 tf(l,s) = 0 and «(0,s) = 0 (4-63)

A complete procedure for numerical solution by employing the

dimensionless stream function and the vorticity vector of equations

(4-58) to (4-60) has been developed by Vrentas et ai [1966]. However,

the applicability of this procedure is restricted to a flat velocity profil

5 = 0, therefore it could not be applied to predict the velocity profile in

downstream of the packed bed.

148
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

After the fluid is far downstream from the outlet of the bed, one expects,

intuitively, that the flow profile will not undergo any further change of s

(fully developed flow profile condition). Hence, introducing the following

dimensionless parameter into equations (4-58) to (4-63) seems

reasonable for simplifying the problem:

<p = =- (4-64
%-Vw.

Because the parameter if is independent on C it is reasonable to assume

that the radial component of the equation of motion (equation (4-59)) is

negligible. Hence, the flow equation may be reduced to give the following

ordinary differential equation for if:

d
^-(A+E^)^-C = 0 (4-65)
ds ' ds

where

N If
A = ili*Lf__. (4_66)

B = N^V*0 (4.67)

N 7>
C = , *" x (4-68)

and the boundary conditions become;

1. at s = 0 if = \ (4-69)

2. at s = °o i/ = o (4-70)

149
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

By inspection, it is evident from equations (4-65) to (4-70) that the if is a

function of S and will be damped out exponentially with s. Equation (4-

65) with boundary conditions of equations (4-69) and (4-70) is a boundary

value type equation [Burden and Faires, 1993] and may be solved by

using the approximate methods of the weighted residuals [Burnett, 1987],

According to the method of weighted residuals, the solution of equation

(4-65) is approximated by functions. These functions are chosen so that

all boundary conditions are satisfied exactly, although the solution of

equation (4-65) itself is only approximate [Chow, 1979]. For the problem

at hand, the solution of equation (4-65) is approximated by the following

exponential equation;

it = €*s (4-71)

which automatically satisfies all the boundary conditions (equations (4-69)

and (4-70)). The parameterd is a constant to be determined. By using the

collocation method [Burnett, 1987] to optimise the best value of d, which

is carried out by choosing a point s in the domain and then forcing the

residual to be zero. The residual, Si, is derived by inserting equation (4-

71) into equation (4-65), and by choosing s = 1 to give:

ft = aV°+(de-(kXA + Ae-")-C ,^-jry,


=0

Solution of equation (4-72) for d, leads to a complete description of the

velocity field for the system under consideration. However, there arises a

150
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

mathematical difficulty in the derivation of the desired solution for

equation (4-72). This equation is a non-linear equation and hence a trial-

and-error technique is required to solve the problem. Besides, a time

consuming calculation is needed, and the convergence also cannot be

guaranteed [Burden and Faires,1993].

In order to overcome the above problem, the parameter d was

determined from experimental data of the entry length as given by

Atkinson ef ai [1969] and Bowlus and Brighton [1968]. The values of the

Reynolds number for the entry length data range from 1.0 to 3.88 x 105.

The correlation of d may be represented by the following equation:

-for NRe<2100

d = 2.41NRe-°'5 (4-73)

-for NRe>2100

d = 0.05 (4-74)

In order to perform a complete solution of the fluid velocity profile in the

entry region requires the knowledge of the fully developed velocity profile,

the pressure losses in an empty pipe and the entry length. Fortunately,

these had much attention from a theoretical and an experimental

viewpoint for exploring the physical phenomena; therefore, the availability

of the mathematical correlations is adequate. By using this equation, the

151
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

data that were measured at the exit of the bed may be used to evaluate

the model of the velocity profile inside the bed.

Consider a steady flow of a Newtonian fluid of constant density in a very

long tube of radius R, so the end effects are negligible. For the value of

the Reynolds number less than 2100, the equation for velocity profile is as

follows [Bird et ai, 1960]:

^=2(l-C2) (4-75)

For Reynolds number greater than 2100 the velocity profile may be

predicted by the following empirical correlation [Bird et ai, 1960]:

*.- 1.25(1 -,P (4-76)

The pressure gradient in the developing flow region is higher than in the

fully developed flow region due to the increasing friction and kinetic

energy losses [Langhaar, 1942; White, 1986]. Therefore, the pressure

drop in equation (4-68) may not be calculated by using the pressure drop

correlation that was formulated based upon measurement of fully

developed flow as proposed in Brown et al. [1950], Bird et al. [1960],

Foust etal. [1960], and Perry and Green [1984]. Considering the smooth

tube, the pressure drop in equation (4-68) can be calculated by the

following empirical correlation [Christiansen and Lemmon, 1965; White,

1986]:

152
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packe

2(-Ap) . L-M,
-pu..
— T ^ f k - MD^ + G- (4-77)
'M

where:

- For laminar flow:

64
fk = — , [Schmidt and Zeldin, 1969] (4-78)
IN'r Re

05 = 1.0 -1.41, [Christiansen and L e m m o n , 1965] (4-79)

For turbulent flow:

0.3164
f
k= 0.25 . [Bird et ai,1960] (4-80)
•^Re

05 = 1.314, [Schmidt and Zeldin, 1969] (4-81)

The entry length, which is defined as the distance along the axis of the

flow where the centre-line velocity reaches 99% of its fully developed

value [Chen, 1973], depends on the inlet profile and on the Reynolds

number [Foust et ai, 1960; Berman and Santos, 1969; Brady, 1984],

Numerous studies have been conducted for correlating the entry length to

its variables [Langhaar, 1942; Foust et ai, 1960; Christiansen and

Lemmon, 1965; Vrentas et ai, 1966; Bowlus and Brighton, 1968; Atkinson

et ai, 1969; Chen, 1973]. Based upon the survey of their results, it can be

shown that the following equation is satisfactory to predict the entry length

for uniform inlet profile:

153
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

- For 226 > NRe [Atkinson et ai, 1969]:

^L = 1.18 + 0.112NRe (4-82)


D

- For 104 < NRe < 107 [Bowlus and Brighton, 1968]

(4-83)

It is instructive to compare the calculation results of the present

correlation (equations (4-64), (4-71), (4-73) and (4-74)) with previous

models and experimental data. In order to facilitate comparisons, the

computed velocity profiles and the predicted entry length, together with

data from several previous solutions and experimental studies, are plotted

in Figures 4-4 and 4-5, respectively. Although the available experimental

data are inadequate for a very rigorous test of the present mathematical

model of velocity distribution in the entry region of an empty pipe, the

reasonable agreement of computed with experimental velocity profile and

entry length data indicates that a close approximation to reality for the

prescribed conditions has been achieved. Therefore, it may be concluded

that the present model of velocity profile under developing flow condition

can be used to fulfill the need of correction factor for study of flow profile

inside a packed bed which measurement is carried out at the downstream

of the bed.

154
Development of a Mathematical Model for Velocity Profile of Fluid Flowing in Packed Beds

1.5 -- 1.5 -

.1 8 8 Q 6 A A A
l . O . l Q Q D Q Q Q - I Q O i
I.0-- o „ o 8 |
if if f

0.5
s = o.oo 0.5 -
S = 0.49

0.0 I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' ' I ' ' ' ' 0.0 - 1 ' ' ' i ' ' ' ' i ' ' ' ' i ' ' ' ' i ' ' '
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 o.;
C c
1.5

" 2 8 _ _ G £ A A
A
1.0- ° o
if o

0.5 -
S =0.71
O Predicted A Data

0.0 I ' ' ' ' I ' ' ' ' I ' ' ' 'ft
0 0.2 0.4 0.6 0.8 1

c
1.5 -•

8 g fi A A A A A
° O A
1.0 O
if O

0.5
S = 0.92

0.0 -------- 1
' ' l ' ' ' ' l ' ' ' ' I ' ' ' '6
0 0.2 0.4 0.6 0.8 1

c
Figure 4-4: Comparison of the developing flow profile, at NRe = 47,
calculated by equation (4-71) with Berman and Santos

[1969] measurements.

155
-J
_
_
(J
o
a
a.
o
o
o
00 o
XO xo
Ox
xO o
Ox Ox Ox
o
o
is _ o
o
cc
c o
r. T3
tL o
c
C
C3 Ir
cd
1
<f3 o CO
_J

5)
Z
=:
c.
F c- o c
-C
u
_
CC -J
_
O o
o g
o "+-I

o ss.
<D
o i_

!_
o o
o
u
o _c
o c
+-I

cn cCD
o cu

o o
c
o o
o (fl
"SZ
CO
a
o E
o o
o
LO
I

a>
.. O !_.
3
D)

O
Q
o
CHAPTER FIVE

EXPERIMENTAL TECHNIQUE

In order to provide the data for validating the model of velocity profile of

single-phase fluid flowing in a packed bed, experimental work was

conducted to measure the velocity distribution. The measurements were

carried out in a vertical cylindrical tube, randomly packed with glass

spherical particles. The experimental equipment was designed and built to

allow the best possible observation of fluid flow profile in the random

packed beds of the spherical particles in which the particles are uni-sized

and multi-sized.

5.1 EXPERIMENTAL APPARATUS AND MATERIALS

A schematic diagram of experimental apparatus is shown in Figure 5-1.

The packed column that was used for the measurement of the velocity

distribution of the fluid flow in packed beds was made of clear perspex to

allow easy viewing of the packing material. The column dimensions were

3.17 mm wall thickness, 144.14 mm inside radius and 1000 mm length.

The cylindrical column was provided with an inlet pipe made from brass

with inside diameter of 4.28 mm, as a connector with a "AS 1335-996

157
Experimental Techniques

C O M W E L D " rubber hose with inside diameter 5.01 m m . The position of

the fluid inlet was in the center at the bottom of the column.

In the experiment, compressed air as the fluid was supplied from a bottle

(G size) of industrial grade compressed air. Air was chosen because of its

ready availability, its well known rheological and thermodynamic properties

[Bird etal., 1960; Reid et ai, 1977; Perry and Green, 1984], its Newtonian

fluid behaviour and its low cost. The inlet line of the fluid was completed

by a "CIGWELD COMET 500" gas regulator with two pressure gauges (0 -

27.58x106 kgm'V2 and 0 - 9.65x105 kgm"1s"2 of the gas bottle and outlet

pressures, respectively) and a "KEY INSTRUMENTS" air flow meter type

FR 4500 rotameter (2.4-16.8 m3/hr). The accuracy of this type flow meter,

as given by the manufacturer (KEY INSTRUMENTS, PA) is + 3 %. Air was

delivered from the bottle, at a pressure 6.89x104 - 41.37x104 kgnrfV2 and

a temperature of 26°C, to the inlet pipe via a "AS 1335-996 COMWELD"

rubber hose with inside diameter 5.01 mm.

Glass spherical particles were used to randomly pack the cylindrical

column to provide a packed bed system. Two different beds were used in

the column; namely the test bed and the inlet flow straightening bed, with

steel wire mesh with openings of 1.0 mm that separated them.

158
Data Processing Unit

350 m m Test Bed


1000 m m I
Wire Mesh

210 mm Straightening Bed

1 1 Wire Mesh

Rotameter

\^J Pressure Gauge

AIR

Figure 5-1: A schematic diagram of the experimental apparatus.


Experimental Techniques

Three sizes (15.69 + 0.19, 23.71 ± 0.59 and 34.33 ± 0.41 m m ) of spherical

diameter of "PHANTOMS" marbles and a 6.04 ± 0.05 mm of spherical

glass particles were used as the test bed and a mixture of sizes (4.0 - 7.0

mm) of spherical glass particles was used as the flow straightening bed.

The densities of spherical particles which used as the test bed were 2670,

2550, 2622 and 2453 kg/m3 for particles diameter of 6.04 ± 0.05, 15.69 ±

0.19, 23.71 ± 0.59 and 34.33 + 0.41 mm, respectively.

The height of the flow straightening bed was 210 mm in order to achieve

the H/D ratio greater than 1.0. The ratio of H/D greater than 1.0 was also

used for the test beds. This value is a minimum value to ensure a parallel

flow of the fluid inside the bed as stated by Poveromo and Szekely [1975].

The measurement of the air flow velocity was carried out by using glass

encapsulated thermistors with diameter of 1.4 mm, which is sufficiently

smaller than the bed particles to minimise flow disturbance. The accuracy

of this type of flow meter, as given by Bolton [1996], is ±1.0 %. These are

temperature dependent resistors that enable temperature variation, due to

the cooling effect on the thermistor, by the media flowing round it, to be

recorded as changes in the voltage across the thermistor [Benard, 1988].

A number of electronic data cables was used as connectors between the

thermistors and a "COMPAQ ProLinea 4/25S", IBM-compatible personal

160
Experimental Techniques

computer, 8 M B R A M and 540 M B H D . Glass tubes with an outside

diameter of 3.08 mm were used as sheaths for the connecting wires

between the thermistors and electronic data cables. Computer Board

"PPIOAI8" was used as an analog-digital interface to convert the voltage

changes from thermistors to digital numbers, and also to generate the heat

for the thermistors in order to keep the thermistors temperature higher

than ambient temperature.

5.2 EXPERIMENTAL PROCEDURE

In order to provide a correction factor for the rotameter reading, the total

equivalent length, Le, of line between the pressure indicator and the

rotameter was determined by using the mechanical energy balance

(equation (4-11)) and measured data of the inlet pressure and the gas flow

rate. The value of the total equivalent length is needed to determine the

actual pressure in the rotameter and then by employing the material

balance correlation, the gas flow rate at inside and at down stream of the

bed can be determined properly.

Before any velocity measurements were made the thermistors were

calibrated by using the flow of air in an empty pipe. The column

dimensions were 144.14 mm inside diameter and 2000 mm length, and a

mixture of sizes (4.0 - 7.0 mm) of spherical glass particles with 210 mm

height was used as the flow distributor. The thermistors were calibrated by

161
Experimental Techniques

placing them at 1750 m m above the flow distributor. Similar to the hot-wire

and hot-film anemometer, resistance-temperature relationship of a

thermistor is not linear [Bolton, 1996]. Hence, the similar type equation to

the hot-wire and hot-film anemometer for air flow measurement, as given

by Wasan and Baid [1971], was chosen to make a correlation of the

velocity and the digital computer data. Considering that the fluid properties

of air for calibrating the thermistor are similar to the air for experiment, th

Wasan and Baid [1971] equation may be simplified to give:

u0M = wi\oVe + T^ (5-1)

where the value of 6-\ and 4 is determined by using the experimental

relationship of the digital computer numbers, cMz and velocities.

The mean bed voidage was determined by weighing the column before

and after packing was added. From density measurements of the packing

and a knowledge of the total volume occupied by the packing plus voids,

the average volume fraction of voids was calculated. A good agreement

between measurement of the mean bed voidage was obtained by using

this technique, compared with the volume displacement technique as

reported by Newell [1971]. Additionally this technique is non destructive for

the packing arrangement.

162
Experimental Techniques

W h e n the air flowing in the packed bed had reached a steady state

condition, as indicated by constant position of the rotameter float, the

velocity measurement was started with a sampling rate 1 s"1. For each

position of thermistors, data logging was carried out with sampling time 60

seconds or 60 readings were taken in a single run.

The following parameters were investigated: fluid flow rate, D/DP ratio, and

particle size distribution. Test programs were also made to investigate the

changes of fluid distribution at the outlet of the bed. In addition, tests were

made to assess the reproducibility of the measurements with repacking of

the bed between tests for the same value of the mean bed voidage.

The thermistors position of 300 mm above the bed as a basis of

measurement was chosen to avoid the errors from the high turbulence

intensities of flow [Mickley et ai, 1965] and the axial component of flow

[Schwartz and Smith, 1953] at the exit of the bed. At a distance of 300 mm

above the bed, any gross changes in the flow distribution would be

expected. Although the flow distribution was changing from the inside of

the bed, validation of the mathematical model of fluid flow distribution at

inside the bed still can be done because the model includes developing

flow profile model, as discussed in the previous chapter. The

measurements for the thermistors positions of 250, 350, 400 and 450 mm

163
Experimental Techniques

above the bed also were carried out in order to provide data for validating

the developing flow profile mathematical model above the bed.

The study of the effects of the fluid flow rate and the D/Dp ratio on the

velocity profile was carried out by variation of air flow rate and particle

diameters. The air flow rate varied from 4.02 to 19.62 m3/hr where the uni-

sized particle diameters were 15.69 ±0.19 mm. The measurements of the

velocity profile for the uni-sized particles diameters of 23.71 ± 0.59 and

34.33 ± 0.41 mm also were performed to provide the validating data for

different value of the D/DP ratio. Binary sizes (23.71 ± 0.59 and 34.33 ±

0.41 mm, and 15.69 + 0.19 and 34.33 + 0.41 mm) and ternary sizes (6.04

± 0.05, 23.71 ± 0.59 and 34.33 ± 0.41 mm) mixtures of particles were also

used to investigate the effect of the mixture of packing particles on the

velocity profile of the fluid flowing in packed beds.

Ideal gas behaviour assumption for air was applied for the data analysis.

This is reasonable because of the low pressure and temperature of the

experimental conditions as stated by Reid and Sherwood [1966].

According to Bird et ai [1960], Newtonian fluid behaviour also can be

applied for the air under these conditions.

Since packed beds are random systems and cannot be exactly duplicated

[Schwartz and Smith, 1953; German, 1989], hence, the reproducibility is a

164
Experimental Techniques

key factor to verify the validity of the experiment. The reproducibility of the

measurements were calculated according to the following equation

[Davies and Goldsmith, 1977]:

R = - ^ — - — x 100% (5-2)
l
n.d

As stated earlier, the objective of the experiment is to provide the data for

validating the model of velocity profile of single phase fluid flowing in a

packed bed. The error probability of a good model is minimum [Reid et al,

1977]. In order to develop a good model there is a need to achieve

minimisation of errors by adjusting the form of the equation and the values

of constants that are included in the equation. In considering how to

guarantee the stability of error minimisation steps, Mickley et al. [1957]

suggested "the sum of square errors". The calculation of the sum of

square errors is made by using the following equation:

X/?2=Ife-^)2 (5-3)
i i=l

While the average deviation of calculated results compared with

measurements is determined as follows [Davies and Goldsmith, 1977]:

f
d,-c*

R =
H i=i
nd-l
-xl00% (5-4)

165
CHAPTER SIX

EXPERIMENTAL RESULTS AND


MATHEMATICAL MODEL VERIFICATION

In order to minimise the possibility of generation of unwieldy mathematical

equations, usually some assumptions are required in the development of a

mathematical model [Franks, 1972]. Of course, introducing any

assumptions has a consequence of carrying through an error or a

deviation from real situation, simultaneously. Therefore, verification of the

solution obtained from the mathematical model, by making a comparison

with measured data, is needed to check the validity and the consistency of

the mathematical model.

The requirement for experimental data involving an incompressible fluid for

validating the mathematical model has been fulfilled by the measured data

of Stephenson and Stewart [1986]. To provide the measured velocity

profile data of compressible fluid for validating the mathematical model, as

described at previous chapter, an experiment has been carried out by

measuring the velocity profile of the air at the downstream of a packed

bed.

166
Experimental Results and Mathematical Model Verification

6.1 E X P E R I M E N T A L R E S U L T S

Based on the data of the inlet pressure and the rotameter reading

relations, the total equivalent length, Le of the line between the pressure

indicator and the rotameter was determined. It was carried out by using

the mechanical energy balance (equation (4-11)), for which the average

value of Le was 5.152 m. Hence, the actual pressure of the fluid inside the

rotameter can be determined by using the Le value and then by employing

the material balance correlation, the gas flow rate at inside and at down

stream of the bed can be determined properly.

The thermistors were calibrated by placing them in a vertical column with

L/D ratio of 12.14. From the measured average velocity, obtained from the

rotameter reading and the cross-section of the pipe, the calibration of the

thermistors was made using equation (4-71) to determine the point

velocities of the thermistor's position. The calibration data of the

thermistors were fitted as a straight line by means of equation (5-1) and

the result demonstrates a good agreement, as illustrated in Figure 6-1.

For each position of the thermistor, digital computer number {cAQ is taken

as the average of 60 readings in a single run. This is to reduce the drift

errors of thermistor and high turbulence of fluid. Figure 6-2 represents an

example of the data read by a thermistor.

167
Experimental Results and Mathematical Model Verification

2.40 2.45 2.50 2.55 2.60 2.65 2.70 2.75


CfVsz

Figure 6-1: Typical calibration curve for thermistors.

168
Experimental Results and Mathematical Model Verification

2.8

2.7 -

:
2.6

cVc 2.5 -

2.4 -

2.3 -

0 10 20 30 40 50 60

Time, s

Figure 6-2: Typical reading data of the thermistor.


Experimental Results and Mathematical Model Verification

6.1.1 Reproducibility of Data

Five tests were made using three identical packed beds (34.33 ± 0.41 mm

diameter spherical particles, 390 mm bed length, 0.437 average voidage

and 0.1246 m/s superficial velocity). Measurements were carried out for

two of the packed beds twice and a measurement for the last packed bed,

once.

As shown in Figure 6-3, a good reproducibility of velocity measurement

was obtained. The absolute deviation from the mean of the five

replications was 12%, although maximum values as high as 28% were

observed. The reproducibility, which is on average 12%, is considered

very good for this type of experiment due to the nature of packed beds,

which are random systems and cannot be exactly duplicated [Schwartz

and Smith, 1953; Blum and Wilhelm, 1965; Fayed and Otten, 1984].

6.1.2 Measurement Results of Velocity Profile

Experimental work was conducted to investigate the effects of the

Reynolds number, the ratio D/DP, the particle size distribution and the

change of fluid distribution at the downstream of the packed bed. The

objective of this work was to provide the data to validate the mathematical

model of the velocity profile of a compressible fluid flowing in a packed

bed. For the case of an incompressible fluid, the model will be validated by

the commonly available data in the literature.

170
r-
8

no. 3
no. 2
ooa o

i
d
—M
c
-a -O "O c
- O ii
X)
-a T3 T3
s r-M r-. ---

_ o
•s O O <KX c,
ed
C-M
c_
OH
OI
MI II II
_
-J
K < oo-
-
a:
<© o << cc
+-*
s CO
c •_3
r
o 03 O 0)
-C
o
<D >
<o o
o
Q .3
i "O
<w ooo p_
o
i-
Q.
0)
oo _.
oo o < C "cc
1?
SO
r- co
Mr o
Tf — CO
<N co ^. Tt
0303 O "EL
' Tt Tt ON h-
O o CO — CO CO
d depth =

II II
O Gd<K> CD
J_

3
0 TT
LL.
o Br 1)
o o<] o
3 OJ QQm

<&0 oo
O «2> <
o
- r — d
Ox r- r- ro
d d d
d
_*
Experimental Results and Mathematical Model Verification

Figure 6-4 presents measurements of point velocities, uz, which is

normalised with respect to average velocity, uM, as function of wall

distance for mono-sized spherical particles. The velocity profile was

measured at 300 mm above the bed with superficial velocity 0.1246 m/s.

The bed particles in Figure 6-4 have diameters 15.69 mm, 23.71 mm and

34.33 mm which correspond to D/DP ratios 4.2, 6.1 and 9.2.

Figure 6-5 exhibits the velocity profile measured at 300 mm above the bed

with superficial velocity 0.1246 m/s for multi-sized spherical particles.

Figures 6-5a and 6-5b show the results for an equal volume binary-

mixture, whereas Figure 6-5c represents the result for an equal volume

ternary-mixture. In Figure 6-5a, the bed particles consist of diameters

34.33 mm and 15.69 mm. These give the following values: volume-surface

mean diameter, Dvs, of 21.42 mm, DP1/DP2 ratio of 2.2 and D/Dvs ratio of

6.7. The bed particles in Figure 6-5b comprise the diameters of 34.33 mm

and 23.71 mm, which correspond to Dvs of 28.15 mm, DPi/DP2 ratio of 1.4

and D/Dvs ratio of 5.1, while in Figure 6-5c, the bed particles comprise

diameters of 34.33 mm, 23.71 mm and 6.04 mm. These correspond to Dvs

of 12.43 mm, DP1:DP2:DP3 equal to 1.0:3.9:5.7 and D/Dvs ratio of 11.6.

The data in Figures 6-4 and 6-5 will be used to compare experiment with

prediction in the next section.

172
Experimental Results and Mathematical Model Verification

.

0
1.5
• 0 0
0 o
if 1 -:o 0 o o 0 o 0 o
p 0
0.5 -• oo D = 144.14 m m u M =0.1246 m/s
- Dp = 15.69 m m e'o = 0.43
Bed depth = 322.5 m m
• j -1—1—'— 1 1
' — l — ' ' — ' — ' — l — ' — L
- 1 1 1 s 1

0.0 1.0 2.0 3.0 4.0


(R-r)/Dp

1.5 0
-
o o o
O °
if i -
~: o o o o o o
-
o
0.5 --•• o D = 144.14 m m u M = 0.1246 m/s
- DP = 23.71 m m e'0 = 0.425
" , , .— 1 — L - Bed depth = 350 m m i—,—;—,—,—,—,—-]
—'—'—1—L- 1
• 1
— I — • — -I 1 1 1 1 L _

0.0 0.5 1.0 1.5 2.0 2.5 3.0


(R-r)/Dp
2.0

1.5 - o o
o o
if o o
1.0
o o o
0.5 -•
°°o ° D
Dp
= 144.14 m m
= 34.33 m m
u M = 0.1246 m/s
e'o = 0.437
Bed depth = 390 m m
0.0
0 a5 1 15 2
(R-r)/Dp
Figure 6-4: Velocity profile data at 300 mm above a bed of mono-

sized spherical particles of different diameter, DP.

173
Experimental Results and Mathematical Model Verification

z.u -
: a) Binary-mixture

1.5- :
0
0
if o o
1.0-
0
o o 0
0 ° o
0
0.5- 0 D = 144.14 m m u M = 0.1246 m/s
- Dvs = 21.42 m m e'0 = 0.383
- 1
Bed depth = 405 m m
0.0- ' 1 ' ' ' ' 1 '' ' ' 1 ' ' '1 I ' ' ' ' l — ' — • — ' — • — f — ' — ' — • -

0.0 0.5 1.0 1.5 2.0 2.5 3.0


(R-r)/Dp
2.0
; b) Binary-mixture
l

1.5
l

0 o
l
l

o o
-T'l

if
1.0 - o o o o
0.5 -
oo o
D =144.14 m m u M =0.1246 m/s
- Dvs =28.15 m m e'0 =0.428
Bed depth = 350 m m
, , , , , , , L _ _.—,—,—,—,—|—
0.0 _—,—,—,—,—,—,—,—,—,—__
0.0 0.5 1.0 1.5 2.0 2.5
(R-r)/Dp
2.0
c) Ternary-mixture
1.5 O
O
if
1.0
OO ° o 0
° 0 0 O O
OO o
0.5 oo D =144.14 m m u M =0.1246 m/s
Dvs = 12.42 m m e'0 = 0.306
Bed depth = 370 m m
0.0
0.0 1.0 2.0 3.0 4.0 5.0
(R-r)/Dp
Figure 6-5: Velocity profile data at 300 mm above a bed of

multi-sized spherical particles (see text for

details of a, b and c).

174
Experimental Results and Mathematical Model Verification

The measured data of the effects of the air flow rate and the distance

above the bed on the velocity distribution at the downstream of the bed

are presented in Figures 6-6 and 6-7, respectively. These effects will be

discussed later together with other pertinent effects.

6.2 MATHEMATICAL MODEL VERIFICATION

The purpose of this verification is to examine the performance of the

present mathematical model of the velocity distribution in a packed bed.

The verification begins with a comparison of the predicted results with

measured data, at the inside and at the downstream of the incompressible

and compressible fluid flowing in a packed bed, and then continuing with

the discussion of advantages and disadvantages of the present mode!

compared with other previous models. The criteria for a good

mathematical model as given by Reid and Sherwood [1966] is used in

verification of the mathematical model.

6.2.1 Validation of the Mathematical Model

The tests of the present mathematical model of velocity profile were

carried out by employing the measurement data taken from the literature

and from the present experimental work. The purpose of this procedure is

to test the present model, especially for the consistency of the model for

different experimental methods and materials.

175
Experimental Results and Mathematical Model Verification

2.0

o--C=0.5

1.5-

#1.0--
•o
o-
•o.
•o-

Column diameter 144.14 m m


0.5-
Spheres diameter 15.69 m m
Bed depth 322.5 m m
Average bed voidage 0.43

0.0 + + -1 I I I 1 1-

50 100 150 200 250 300 350


Air flow rate, L/min.

Figure 6-6: The effect of air flow rate upon the velocity profile at 300

m m above the bed.

176
Experimental Results and Mathematical Model Verification

2.0

o- - • £=0.5

1.5-

#1.0-- -o.

Column diameter 144.14 m m


0.5 Spheres diameter 15.69 m m
Bed depth 322.5 m m
Average bed voidage 0.43
Average velocity 0.1246 m/s

0.0
5 7
S

ure 6-7: T h e effect of thermistors positions o n the velocity profile

a b o v e the bed.

177
Experimental Results and Mathematical Model Verification

For beds packed with uniform and non-uniform size of spherical particles,

the model can be validated by the present experimental data. For beds

packed with uniform size of cylindrical particles, the model can be

validated by comparison with the measured data of Stephenson and

Stewart [1986].

6.2.1.1 Uni-Sized Particle Packed Beds

For the incompressible fluid flow in beds of uniform size cylindrical

particles, the model has been validated by using the data of Stephenson

and Stewart [1986]. Their velocity and porosity distribution data were

obtained by using optical measurements for particle Reynolds numbers

from 5 to 280 in the beds with D/Dp = 10.7 and velocity was measured

inside the beds. Table 6-1 shows the important parameters for

Stephenson and Stewart [1986] data, which were used for calculation by

the present model.

Because of the lack of mathematical correlation of porosity distribution for

cylindrical particles bed, the natural cubic spline method [Burden and

Faires, 1993] was used to fit a correlation for the measured data (Figure

3-5) of Stephenson and Stewart [1986]. Although the use of the natural

cubic spline method is cumbersome, this method is more accurate than

polynomial regression when the data contain regions of sharply different

178
Experimental Results and Mathematical Model Verification

behaviour [Tao, 1987a]. The value of local voidage m a y be approximated

by the following equations:

for r*j < r* < r*i+1:

e = -.i+fei(r*-r*i)-r-ci(r*-r*)2+c?i(r;t:-r*i)2 (6-1)

where

Table 6-2 contains a tabulation of the constants (a-,, h, cu d, and r*,) for the

natural cubic spline correlation of porosity distribution of equation (6-1).

Figure 6-8 shows the comparison of the measured velocity distribution

data inside the bed and those predicted by the present mathematical

model. The agreement between the experimental data with those

predicted by the model is good, with average deviation of 10% (30%

maximum) and the sum of square errors of 1.44 for the total number of the

data points of 120.

The deviation between the predicted and the measured data may be due

to the assumption of fitting the measured data (Figure 2-8) by Stephenson

and Stewart [1986], in which they assumed that the velocity profile,

normalised with respect to uM, is independent of the Reynolds number.

179
Experimental Results and Mathematical Model Verification

Table 6-1: T h e experimental conditions of Stephenson and Stewart

[1986] data.

Parameters Value
Cylindrical column
- Diameter, m m 75.5
- Bed height, m m 145.0
Cylindrical particle
- Diameter, m m 7.011
- Length, m m 7.085
- Volume-surface mean diameter, m m 7.035
Mean bed voidage 0.354
Fluids:
- Tetra ethylene glycol
- density, kg/m3 1,125
- viscosity, g/cm s 0.474
- superficial velocity, cm/s 3.1 and 12.4
- Tetra hydropyran-2-methanol
- density, kg/m3 1,027

- viscosity, g/cm s 0.114

- superficial velocity, cm/s 5.9 and 11.8

- Mixture of cyclo octane and cyclo octene


- density, kg/m3 834.3
0.0242
- viscosity, g/cm s
6.0 and 12.0
- superficial velocity, cm/s

180
Experimental Results and Mathematical Model Verification

Table 6-2: The value of the constants of equation (6-1).

i - i a\ bi Ci d\
0 0.00 1.00 -4.89 0.00 1.47
1 0.13 0.35 -0.21 0.59 -0.85
2 0.27 0.33 -0.06 0.25 -0.08
3 0.54 0.33 0.02 0.19 0.05
4 0.80 0.35 -0.10 0.23 -0.02
5 1.07 0.34 -0.06 0.21 -0.01
6 1.34 0.34 0.05 0.21 0.04
7 1.60 0.37 -0.10 0.24 -0.02
8 1.87 0.36 -0.10 0.22 -0.00
9 2.14 0.35 -0.10 0.22 -0.02
10 2.41 0.34 0.06 0.20 0.04
11 2.68 0.37 -0.06 0.24 0.00
12 2.94 0.37 -0.21 0.24 -0.05
13 3.21 0.33 -0.05 0.20 0.00
14 3.48 0.33 0.06 0.20 0.03
15 3.74 0.36 -0.06 0.23 0.01
16 4.01 0.36 -0.21 0.23 -0.07
17 4.28 0.32 0.17 0.18 0.07
18 4.55 0.38 -0.11 0.23 0.08
19 4.82 0.37 -0.12 0.30 0.51
20 4.95 0.36 -0.21 0.50 -0.33
21 5.08 0.34 -0.06 0.37 0.42
22 5.22 0.34 0.03 0.54 -1.35
23 5.35 0.35 0.00 0.00 0.00
Experimental Results and Mathematical Model Verification

2.5
2.5
N R e = 20
NRe = 5
2.0- 2.0

31.5 -: 1.5 -:

1.0 1.0 -;

0.5 Measured 0.5 -:


Predicted
0.0 i '' ' i 0.0 I i I I L_
I '' ' I

0 1 2 3 4 5 1
fe-i)ibp
(R-r)/Dp

0 1 2 3 4 5 0 1 2 3 4 5
(R-r)/Dp (R-r)/Dp
2.5 2.5
N R P = 145 N R e = 280
2.0-; 2.0 -

21.5
3

1.0

0.5

L
0.0 i ' ' ' i ' I ' '' I ' '' I 0.0 i ' '' i

0 1 2 3 4 5 0 1 2 3 4 5
(R-r)/Dp (R-r)/Dp
Figure 6-8: Comparison of velocity profile calculated by the

present model with the Stephenson and Stewart [1986]

measurements.

182
Experimental Results and Mathematical Model Verification

This assumption is not exactly true as demonstrated by measured data of

Morales et ai [1951], Schwartz and Smith [1953], Newell and Standish

[1973], Szekely and Poveromo [1975], and McGreavy etal. [1986].

For the compressible fluid, the model of the flow profile at inside and the

developed flow profile at the downstream of the bed has been validated

by using measurement data of air flow in a bed of spherical particles. The

flow profiles were taken from measurement at the downstream of the bed.

In accounting for the distribution of voidage over the cross section of the

bed, the empirical correlation proposed by Mueller in 1992 (equations (3-

33) to (3-37)), was used. This correlation was chosen because it is simple

and valid over a wide range of D/DP ratio as given in the papers by

Mueller [1990; 1992]. However, considering that the value of the bulk

porosity, eb, for a random packed bed is highly dependent on the method

of charging [Blum and Wilhelm, 1965; Cumberland and Crawford, 1987].

Therefore, It seems reasonable to use measurement value of the average

voidage rather than a value predicted by means of equation (3-38) for eb

[Mueller, 1991].

Figure 6-9 is a plot of the developed flow profile downstream of the bed

versus the position from the top of the bed. The solid curves represent the

calculated values by using the present model, whilst the symbols

183
Experimental Results and Mathematical Model Verification

represent the experimental values. It can be shown that there is

reasonable agreement between the experimental data and the predicted

values.

Figure 6-10 shows the effect of the flow rate variation on the velocity

profile of the fluid at 300 mm above the bed. The agreement between

measurements (symbols) and predictions (solid curves) is again quite

reasonable. It is seen, furthermore, that the fluid flow distribution,

normalised with respect to average velocity, is not independent of the flow

rate over the range tested.

Figure 6-11 shows the effect of the bed particle diameter variation on the

velocity profile of the fluid at 300 mm above the bed. The agreement

between measurements (symbols) and predictions (solid curves) is again

quite good.

It is seen, furthermore, that the oscillation pattern of flow profile has a

similar tendency for all bed particle diameters used. The velocity has a

zero value at the column wall, reaching the first peak value at about 0.2

particle diameter from the wall, and has a minimum at 0.5 particle

diameter. The velocity continues cycling with interval distance of the

peaks about 1.0 particle diameter.

184
Experimental Results and Mathematical Model Verification

2.0

O
1.5-
O O
C = 0.0
O

#1.0 --
£ = 0.5

0.5 Column diameter = 144.14 m m


Spheres diameter = 15.69 m m
Bed depth = 322.5 m m
Average bed voidage = 0.43
Average velocity = 0.1246 m/s

0.0 -J I l_ -J • •

0 8

Figure 6-9: Typical developed flow profile at the downstream of the

mono-sized particle bed.

185
Experimental Results and Mathematical Model Verification

3.0

C o l u m n diameter = 144.14 m m
Spheres diameter = 15.69 m m
B e d depth = 322.5 m m
Average bed voidage = 0.43

2.0 --

if

o ^o
V
1.0 --

\
£ = 0.5

0.0 I I • I • • t I • • I I I •
+ • J
— \ -

10 90 170 250 330 410 490


Airflow rate, L/min.

Figure 6-10: Comparison between measurements and predictions

for the effect of air flow rate upon the velocity profile at

300 m m above the bed.

186
Experimental Results and Mathematical Model Verification

It also can be seen that there is a rising tendency of the oscillation

patterns. This is because the flow conditions at 300 mm above the bed

represents the developing flow profile as discussed in section 2.1.

A complete comparison of the measured velocity distribution results and

those predicted by the present mathematical model is presented in Figure

6-12. The agreement between the experimental data with those predicted

by the model is good, with average and maximum deviation of 17% and

39%, respectively, and the sum of square errors of 1.750 for the total

number of the data points of 64.

Although the maximum deviation is 39%, it is still reasonable because the

difference between the maximum value of the calculation error and the

maximum reproducibility (28%) remains below the average error of

calculations. The deviation between the experimental data and the

predicted values may be due to a number of factors, particularly problems

caused by the use of Mueller's correlation [Mueller, 1992], which was

fitted from other packed bed systems to account for the radial voidage

profile. As stated by Blum and Wilhelm [1965], packed beds of spheres

are a random system, which almost certainly can not be duplicated, even

by repacking of the same particles.

187
Experimental Results and Mathematical Model Verification

2.0
Predicted
o O Measured
1.5 - •

O
O \^^/^~b
if 1.0 -
/ °\70 rV 0 0 ' 0

0.5 -\ \ol D = 144.14 m m


Dp = 15.69 m m
1
u M = 0.1246 m/s
0.0 '' ' ' ' 1
' '' ''' ' ' ' l
0.0 1.0 2.0 3.0 4.0
(R-r)/Dp
2.0
Predicted
O Measured
1.5

if 1.0 --

144.14 m m
0.5 --
23.71 m m
0.1246 m/s
_J I I L_
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
(R-r)/Dp
_. - -
- Predicted
O Measured
1.5 - / O \
0
/

<p 1.0-

0.5 - Ofc D = I44.l4mm


D p = 34.33 m m
u M = 0.l246m/s
0.0 - , — , — , — • _ — I 1 1 1 1 1 L _
— ' — ' — i — ' — ' — ' — ' — i — '

0.0 0.5 l.O 1.5 2.0


(R-r)/Dp
Figure 6-11: Comparison between measurements and predictions

for the effect of D/DP ratio on the velocity profile at 300

m m above the bed.

188
Experimental Results and Mathematical Model Verification

£..\J -

0 /

o /
1.5 - o o/
AA &Jf" X A +

TJ ^ > + A
&
1 1.0- AA
__ o
Q.
o
As
0.5 -

/ °
0.0- _L_, 1 , , 1 _. 1

—'—1—L— 1 — 1 — 1 — 1 — 1 — 1 — 1 — 1 —

0.0 0.5 1.0 1.5 2.0


#measured

Diagonal + Variation of flow rate


X Variation of s A Dp = 15.69 m m
O Dp = 23.71 m m O Dp = 34.33 m m

Figure 6-12: Comparison of predicted velocity distribution and

measurements at downstream of a bed of mono-sized

spherical particles.

189
Experimental Results and Mathematical Model Verification

6.2.1.2 Multi-Sized Particle Packed B e d s

In order to test the validity of the mathematical model for beds of multi-

sized particles, the predicted results were compared with experimental

data for beds of equal-volume of binary and ternary mixtures of spherical

particles. T h e experimental measurements of Goodling et ai [1983] were

used to account for the variation of local bed porosity over a cross section

of the multi-sized particle bed.

Because of the lack of mathematical correlation of porosity distribution for

multi-sized particle beds, the natural cubic spline method [Tao, 1987a;

Burden and Faires, 1993] w a s used to fit a correlation for measured data

(Figure 3-7) of Goodling et al. [1983]. However, because of the difference

of particles sizes and the relative motion between particles when pouring

into the cylindrical column to form a packed bed system, occurrence of a

size segregation is possible [Williams, 1976; Standish, 1990; Rhodes,

1990; Yu and Zulli, 1994]. Additionally, the value of local bed voidage is

also highly dependent on the method of charging [Blum and Wilhelm,

1965; Fayed and Otten, 1984; Cumberland and Crawford, 1987]. Hence,

the data of the radial voidage distribution should be chosen from the bed

with similar characteristics to that being considered. Table 6-2 represents

the comparison of bed characteristics between present work and Goodling

et al. [1983] measurement condition that is used for accounting of the

radial voidage distribution.


190
Experimental Results and Mathematical Model Verification

Table 6-2: Comparison of the bed characteristics between present

work and Goodling etal. [1983] measurement.

No. Parameters Present work Goodling etal., 1983

1 Binary-mixture A
- mixture basis equal volume equal volume
- Dpi, m m 15.69 4.76
- Dp2, m m 34.33 7.94

- DPI:DP2 1.0:2.2 1.0:1.7


- Dvs, m m 21.46 5.95

- D/DvS 6.7 8.8


-e'o 0.383 0.392

2 Binary-mixture B
- mixture basis equal volume equal volume

- Dpi, m m 23.71 6.35

- Dp2, m m 34.33 7.94

- DPI:DP2 1.0:1.4 1.0:1.3

- Dvs, m m 28.15 7.06

- D/Dv S 5.1 7.4


-e'o 0.428 0.426

3 Ternary-mixture
- mixture basis equal volume equal volume

- Dpi, m m 6.04 4.76

- Dp2, m m 23.71 6.35

- DP3, m m 34.33 7.94

- Dpi:Dp2:Dp3 1.0:3.9:5.7 1.0:1.3:1.7

- Dvs, m m 12.43 6.08

- D/Dvs 11.6 8.6


-e'o 0.306 0.415

191
Experimental Results and Mathematical Model Verification

As discussed in Chapter three, for studying fluid flow in a packed bed, the

volume-surface mean diameter is commonly used to represent the mean

diameter of multi-sized particles. Therefore, equation (2-74) can be

modified into the following form:

The values of the constants (a,, b\, c„ d\ and r*j) for binary-size mixtures

and ternary-size mixtures are given in Tables 6-3and 6-4, respectively.

A comparison of the measured velocity distribution results and those

predicted by the present mathematical model for binary-mixtures and

ternary-mixtures is presented in Figure 6-13. The agreement between the

experimental data and those predicted by the model is reasonably good.

For the total number of the data points of 43, the deviation between

experimental data and the calculated values is on average 21%, the

maximum deviation is 39%, and the sum of squares error is 1.564.

The deviation between the calculated results and observed values, as

shown in Figure 6-13, may be due to a combination of causes, which

include the method of charging, segregation of particles and the

distribution of particles. Moreover, because as noted earlier, a packed

bed is a random system which almost certainly can not be duplicated.

192
Experimental Results and Mathematical Model Verification

Table 6-3: The value of constants of equation (6-1) for beds of

binary-mixture of particles.

1 Binary-mixture A
i r*i a\ bi C\ ffj

0 0.000 1.000 -1.824 0.000 2.753


1 0.162 0.717 -2.463 1.334 -0.086
2 0.233 0.547 -1.746 1.316 -2.884
3 0.314 0.413 -0.812 0.617 0.226
4 0.422 0.333 -0.835 0.690 -1.008
5 0.485 0.283 0.224 0.500 0.049
6 0.601 0.316 0.289 0.517 1.669
7 0.673 0.340 1.376 0.876 -0.754
8 0.745 0.443 0.396 0.714 0.678
9 0.862 0.500 -0.901 0.951 -0.657
10 0.933 0.440 -0.763 0.810 -0.178
11 1.041 0.367 -0.886 0.752 2.728
12 1.086 0.329 -0.065 1.120 -2.998
13 1.158 0.329 0.120 0.474 0.362
14 1.274 0.350 0.185 0.600 1.637
15 1.346 0.367 -0.062 0.953 -1.179
16 1.418 0.367 0.112 0.699 0.412
17 1.508 0.383 -0.518 0.810 0.025
18 1.579 0.350 -0.479 0.815 -1.166
19 1.660 0.316 -0.217 0.533 0.213
20 1.759 0.300 0.472 0.596 0.086
21 1.831 0.337 0.043 0.615 -0.225
22 1.947 0.350 0.604 0.536 2.442
23 2.019 0.397 -0.264 1.062 -1.471
24 2.091 0.383 -0.334 0.745 -0.039
25 2.181 0.359 -0.360 0.735 0.124
26 2.253 0.337 -0.021 0.762 -0.493
27 2.333 0.340 -0.281 0.642 -0.527
28 2.432 0.318 0.271 0.486 0.504
29 2.531 0.350 -0.059 0.635 0.229
30 2.621 0.350 0.404 0.697 1.146
31 2.692 0.383 0.101 0.944 -0.381
32 2.764 0.395 0.304 0.862 -0.272
33 2.845 0.425 -0.197 0.796 -0.302
34 2.944 0.413 -0.399 0.706 0.201
35 3.033 0.383 -0.160 0.760 -0.056
36 3.114 0.375 -0.621 0.747 -0.334
37 3.204 0.325 0.087 0.657 -1.519
38 3.294 0.337 0.047 0.248 -0.098
39 3.662 0.383 -0.253 0.139 0.221
40 3.967 0.325 0.783 0.342 0.011
41 4.038 0.383 -0.232 0.344 -0.256
42 4.487 0.325
Experimental Results and Mathematical Model Verification

Table 6-3: The value of constants of equation (6-1) for beds of

binary-mixture of particles (Continued).

2 Binary-mixture B
i r*i fli b\ C\ d\
0 0.000 1.000 -1.025 0.000 10.346
1 0.078 0.925 -2.371 2.420 -3.880
2 0.166 0.733 -2.567 1.399 -2.863
3 0.244 0.540 -1.055 0.729 -0.332
4 0.400 0.392 -2.269 0.574 2.124
5 0.448 0.283 -0.697 0.885 -2.096
6 0.526 0.233 0.160 0.394 -0.058
7 0.614 0.250 0.506 0.379 0.384
8 0.731 0.315 1.386 0.514 1.650
9 0.809 0.427 1.035 0.900 0.416
10 0.897 0.525 0.654 1.010 1.792
11 0.975 0.583 -1.132 1.429 -2.897
12 1.053 0.502 -0.241 0.751 -0.252
13 1.180 0.483 -1.066 0.655 -0.202
14 1.297 0.367 -0.815 0.584 -0.011
15 1.384 0.300 0.143 0.582 0.579
16 1.462 0.315 0.050 0.717 -0.548
17 1.540 0.323 0.637 0.589 0.874
18 1.628 0.384 0.745 0.819 -0.166
19 1.706 0.447 -0.063 0.780 -0.731
20 1.833 0.450 0.207 0.502 1.159
21 1.950 0.483 -0.648 0.909 -0.056
22 2.028 0.438 -0.429 0.896 -0.363
23 2.115 0.407 -0.473 0.800 0.039
24 2.193 0.375 -0.165 0.810 -0.169
25 2.271 0.367 -0.654 0.770 -0.818
26 2.359 0.315 0.078 0.555 1.116
27 2.437 0.325 0.808 0.816 -1.992
28 2.515 0.392 0.014 0.350 -0.224
29 2.846 0.427 -0.226 0.127 0.012
30 3.412 0.342 0.068 0.148 -0.028
31 3.656 0.367 -0.090 0.128 -0.007
32 4.387 0.367 -0.036 0.112 -0.077
33 4.874 0.367
Experimental Results and Mathematical Model Verification

Table 6-4: The value of constants of equation (6-1) for beds of

ternary-mixture of particles.

i r*i a\ bi C\ d\
0 0.000 1.000 -1.009 0.000 10.589
1 0.070 0.933 -1.591 2.224 -4.140
2 0.190 0.767 -1.777 0.733 -0.361
3 0.350 0.500 -1.341 0.560 0.519
4 0.460 0.360 -1.146 0.731 -1.084
5 0.530 0.283 -0.104 0.503 -0.460
6 0.650 0.277 0.918 0.338 0.664
7 0.760 0.383 0.692 0.557 0.531
8 0.880 0.475 -0.149 0.748 -0.499
9 0.990 0.467 -0.328 0.583 0.194
10 1.130 0.433 -0.422 0.665 -0.586
11 1.220 0.400 -0.136 0.506 0.313
12 1.370 0.392 0.059 0.647 -1.742
13 1.450 0.400 -0.104 0.229 -0.073
14 2.000 0.400 -0.118 0.108 -0.010
15 2.500 0.367 -0.060 0.093 -0.019
16 3.500 0.380 -0.073 0.034 0.022
17 4.700 0.380 -0.023 0.113 -0.125
18 5.000 0.380

195
Experimental Results and Mathematical Model Verification

3.0
— Predicted Binary-mixture:
Dvs =21.46 m m
O Measured
D =144.14 m m
2.0 UM =0.1246 m/s
if Dpi:DP2 = 1.0:2.2

1.0

0.0 • M -

0.0 0.5 1.0 1.5 2.0 2.5 3.0


(R-r)/Dp
3.0
Binary-mixture:
O Measured
Dvs =28.15 m m
— Predicted D = 144.14 m m
2.0- uM =0.1246 m/s
Dpi:DP2 = 1.0:1.45
V O
1.0-

0.0
0.0 0.5 1.0 1.5 2.0 2.5
(R-r)/Dp
3.0
— Predicted Ternary-mixture:
Dvs = 12.43 m m
O Measured D = 144.14 m m
2.0 -
uM =0.1246 m/s
DPi:Dp2:Dp3 =1.0:3.9:5.7
O O
-CTQ
O

0.0 I ' ' ' ' l ' ' ' ' l ' ' ' ' I ' ' 'i i'' I' ''i ' '' ''' i'• I ' ' ' ' I '
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
(R-r)/Dp
Figure 6-13: Comparison of velocity distribution predicted by the

model with those measured at downstream of beds of

multi-sized spherical particles.

[96
Experimental Results and Mathematical Model Verification

Using porosity profile data of Goodling et ai [1983], measured under

different conditions to the present velocity profile measurements for this

calculation m a y also have introduced significant errors.

It should also be noted that unlike for mono-sized particles for which the

model predicts a smooth result (eg. Figure 6-11) for the case of binary

and ternary particle mixtures the model predicts a non-smooth line as in

Figure 6-13.

6.2.2 Comparison of the Mathematical Models of the Velocity Profile

in Packed beds

A velocity profile model of a single-phase fluid flow in packed beds w a s

developed in the present work by assuming that the flow characteristic is

a combination of a continuous and discontinuous system of fluid between

voids in the bed. Employing the proposed model of a single phase fluid

flow in the packed beds only requires the physical properties of the fluid

and those of the packed bed to be known, and does not require new

experimental variables.

The validity of the model has been checked by using previous and new

experimental data, and a reasonable agreement w a s obtained. S o m e

deviation of the model prediction from measured data is expected

because of the nature of packed beds, which are random systems and
197
Experimental Results and Mathematical Model Verification

cannot be exactly duplicated. However, based on the accuracy, simplicity,

and the requirement of empirical data, it is believed that the proposed

model provides a useful velocity distribution model.

It is instructive to compare the value of the present mathematical model

with previous mathematical models. As mentioned in the preceding

chapter, although the models that are based on phenomenological

approach are more accurate for a particular set of data, it is difficult to

apply this type of models with any confidence to other systems and

conditions. For these reasons, it would seem appropriate to make

comparison of the present model only with the previous models that were

based on a theoretical approach.

Of the many proposed mathematical models that have been developed

based on a theoretical approach, only Stanek and Szekely's model

[Stanek and Szekely, 1974; Szekely and Poveromo, 1975; Poveromo et

ai, 1975] and Vortmeyer and Schuster's model [Vortmeyer and Schuster,

1983] will be used to make comparison here. The principal reasons for not

using the other models for comparison were poor agreement between

their experimental and their predicted values, and the requirement of

empirical constants, which reduce the model's reliability, generality and

applicability.

198
Experimental Results and Mathematical Model Verification

All of these three models (Stanek and Szekely [1974], Vortmeyer and

Schuster [1983] and the present model) have almost similar condition of

model's simplicity and not require any new empirical constants. Hence, it

would seem reasonable to make comparison only of the accuracy of the

model in predicting velocity profiles. Because Stanek and Szekely's model

[1974] and Vortmeyer and Schuster's model [1983] only provide a

prediction for velocity profile inside a packed bed, the measured data of

Stephenson and Stewart [1986] were used as a basis in this model's

comparison.

The basis of Stanek and Szekely's [1974] model is the Ergun [1952]

macroscopic equation for pressure drop of fluids flowing in packed beds,

where the fluid pathways are regarded as bundles of tangled tubes

(discontinuous system) which are then treated in relation to individual

straight tubes. Originally, this model used the experimental measurements

of Benenati and Brosilow [1962], Figure 3-4, to account for the porosity

oscillation at the grid point adjacent to the wall [Szekely and Poveromo,

1975]. However, the experimental data of voidage profile measured by

Stephenson and Stewart [1986], Figure 3-5, are quite different from those

of Benenati and Brosilow [1962]. Hence, equation (6-1) with constant

values tabulated in Table 6-6 is used to predict the radial voidage

distribution.

199
Experimental Results and Mathematical Model Verification

The point superficial velocities, normalised with respect to average

superficial velocity predicted by Stanek and Szekely's model [1974],

compared with the measurements data of Stephenson and Stewart are

shown in Figure 6-14. Deviation between prediction and measurement is,

on average, 5 5 % , the maximum deviation is 3 1 9 % , and the s u m of

squares error is 59.9 for 120 data points.

The principal reason for this large deviation is probably the effect of the

bed voidage in the vicinity of the wall which has a value higher than 0.5.

This is a maximum value for which the Ergun [1952] equation is

appropriate [Bird et ai, 1960; Cohen and Metzner, 1983]. Another error

may also c o m e from the limitation of Ergun [1952] equation, as stated by

Gauvin and Katta [1973], which is not appropriate for systems containing

particles of low sphericity, whereas cylindrical particles were used by

Stephenson and Stewart [1986] as bed particles for their experiment.

Vortmeyer and Schuster [1983] have extended Brinkman's equation

[Brinkman, 1947] to develop a mathematical model of velocity distribution

in packed beds. The Brinkman equation [1947] is a macroscopic equation

for pressure drop of fluid flow in packed beds, in which the pressure drop

is obtained by summing the resistances of all the individual submerged

particles [Foscolo et ai, 1983], and it is an interpolation between Stokes

equation and Darcy's law [Durlofsky and Brady, 1987].


200
Experimental Results and Mathematical Model Verification

6.0—| 6.0
NRe = 5 N R e = 20

4.0- 4.0-
Measured
Predicted
I E
3
2.0- 2.0-

•• ' L I.

0.0 H '-H- -1—1—I


II — I' !__-.' ' I 0.0 I ' ' ' I- J — I — l — l I I l I 1 i ( i.

0 1 2 3 4 5 0 1
(R-r)/Dp fc-rvbp
6.0 6.0
N R e = 37 N R e = 75

4.0 4.0-

E E
3 3
2.0 4 2.0-
i.- *• • ... J t. . _ ..

0.0 i ' ' 'i '' 'i ''' i ' ' ' i 0.0 i ' ' ' i -1 \ 1^—1 L_

0 12 3 4 5 0 2 3 4 5
(R-rVDp (R-r)/Dp
6.0 f 6.0 i
N R e = 145

4.0

£
3
2.0-

0.0 • i ' ' ' i *

1 2 3 4 5 0 1 2 3 4
(R-r)/Dp (R-rVDp
Figure 6-14: Comparison of velocity profile predicted by Stanek and

Szekely's model [1974] with Stephenson and Stewart

[1986] measurement data inside the beds.

201
Experimental Results and Mathematical Model Verification

This model has better validity than Stanek and Szekely's [1974] model for

experimental data measured by Stephenson and Stewart [1986] as shown

in Figure 6-15. The deviation of predicted value from measurements is

16% on average and 94% maximum, with the sum of square errors being

59.9 for 120 data points. The deviation may be due to the exponential

profile assumption of the radial porosity profile (equation (2-52)) of

Vortmeyer and Schuster's [1983] model. Additional error is probably the

assumption in fitting the measured data by Stephenson and Stewart

[1986], in which they assumed that the uz/uM is independent of the

Reynolds numbers. This assumption is not exactly true as discussed

earlier in section 6.2.1.1.

Figure 6-16 demonstrates the validity comparison of the Stanek and

Szekely's model [1974] and Vortmeyer and Schuster's model [1983] with

the present model, which were tested by using the Stephenson and

Stewart [1986] data. It is clearly seen that the present model has better

validity than other models examined. This result may be expected to be

useful in furthering studies concerned with velocity distribution in packed

bed systems.

202
Experimental Results and Mathematical Model Verification

3.0 3.0
NRe = 5 N R e = 20
Measured
2.0 2.0-
Predicted

1.0 - - , m ,
1.0 •• - . >• •

0.0 _1 I I |_l I 1 L__l I i_


I '' ' I 0.0 i '' ' i
0 1 2 3 4 5 0 1 2 3 4 5
(R-rVDp (R-r)/Dp
3.0 3.0
N R e =-37 N R e = 75
- -

2.0

E E
3
1.0- --V __-..'
*• 1

_.
: .... 3
1.0 - ^ :• -
I _•- -'
i

_.
i . . , _ . . .

0.0 -H-1-1 1 ' ''1 ' ' '" i11 ' 0.0 - ' ' ' 1 ' ' ' i ' ' ' i ' ' ' i ' ' i i '1

0 1 2 3 4 5 0 1 2 3 4 5
(R-r)/Dp (R-rVDp
3.0
N Re = 280

2.0

E
T -. r -
3
1.0-

0.0 I I ' |
I II I I ' | I ' I | ' ' ' | ' ' ' |

0 1 2 3 4 5 0 12 3 4 5
(R-rVDp (R-r)/Dp
Figure 6-15: Comparison of velocity profile predicted by Vortmeyer

and Schuster's model [1983] with Stephenson and

Stewart [1986] measurement data inside the beds.

203
Experimental Results and Mathematical Model Verification

v50U 1

300 1

250 1

200

150 I

100 I
P
50-
i
r\ _
U
AE, %
I
ME, % SSE
• Stanek and Szekely, 1974 I Vortmeyer and Schuster, 1983
B Present work

Figure 6-16: Comparison of the velocity distribution models (AE =

average error, M E = maximum error, and SSE = the sum

of squares error).

204
CHAPTER SEVEN

DISCUSSION

A mathematical model of the velocity profile of a single-phase fluid flow in

packed beds and developing flow profile in the downstream of the bed has

been proposed. The model was tested against the experimental data

obtained from previous investigations and present measurements, and was

found to give an excellent fit. Verification of the model carried out by

comparing it with the previous models, showed an improvement in terms of

accuracy and simplicity, and the model does not require new empirical

constants. Based on these results, an attempt has been made to obtain a

more complete, comprehensive understanding of the fluid flow phenomena in

packed bed systems.

7.1 COMPUTED FLOW PROFILES


Figure 7-1 exhibits a velocity profile computed by the present model for a

cylindrical packed bed of spherical particles. It can be seen that the predicted

flow profile inside of the bed is a non-single peak, or oscillating profile. The

profile is in good agreement with the experimental data of McGreavy et al.

[1986] in Figure 2-9 and is qualitatively similar to the actual flow distributio

Figure 2-1. Moreover, the first maximum value is at about 0.2 particle

diameters, and this is consistent with the data of McGreavy et al. [1986] and

205
Discussion

Ziolkowska and Ziolkowski [1993], and is also consistent with the predicted

value of Vortmeyer and Schuster [1983].

Since the particles are mono-sized, then according to equation (3-57) the bed

permeability variation is represented by the local porosity. As would be

expected, the distribution of the bed permeability determines the velocity

profile, as shown by Figure 7-2. However, it also can be seen that, for

porosity above 0.5, the bed permeability has practically no effect on velocity

compares with its effect for porosity below 0.5. This is not surprising, because

equation (3-57) has been developed based upon an assumption that the

voids in the bed are not connected which each other, or as a discontinuous

system. This is almost true for small voidage beds in which the fluid in a void

is not interacting with the fluid inside the neighbouring voids. Hence, for the

beds with high voidage in which the interaction of the fluid between

neighbouring voids is possible to occur, the validity of the discontinuous

system assumption is no longer tenable. This confirms the findings in

macroscopic study of Foscolo et al. [1983]. In their study, for fluidised beds

system in which the bed voidage is higher than 0.5, the continuous approach

is more appropriate than the discontinuous approach. The kink in the velocity

curve at about K = 0.4, which is related to porosity about 0.5, represents the

transition between continuous and discontinuous behaviour of fluid between

voids.

206
r-
r.

(ft
0)
o
r
CO
a
CQ
O
'£-
CD
-C
a
to
•o
0)
N
'w
6T3O
c
JO
a CO
o
0)
9 TJ
E
CO
c
<D

a ^i,
> O
u o
o
o
CD II
0)
>
TJ cc
CO Z
•a
MM
3 c
a ca
E0 o
CM
u
CO II
o CJ.
n a
I- Q
..
r-
Wn^n N
CU
i_
3
O.
LL
Discussion

-T 2.5

K, mm

Figure 7-2: The effect of the bed permeability upon the fluid velocity

in a packed bed (D/DP = 12.0 and NRe = 1000).

208
Discussion

Although the effect of Reynolds number on the radial velocity profile in

packed beds has been studied by many investigators, a consensus has

not been reached. Therefore, it has seemed reasonable to investigate the

effect of NRe upon the velocity distribution by using the present model. The

results are presented in Figures 7-3 and 7-4.

As shown in Figure 7-3, the results clearly demonstrate that the

dimensionless flow profile is dependent on the Reynolds number if its

value is less than 500. This confirms the findings of Vortmeyer and

Schuster [1983], though their calculated distributions differ from that

shown in Figure 7-1. However, as shown in Figure 7-4, the interaction

between the flow profile with D/DP ratio and the Reynolds number is more

complicated than discussed by Vortmeyer and Schuster [1983]. This may

be due to the exponential porosity profile of Vortmeyer and Schuster

[1983] in which, as discussed in the preceding chapter, the porosity profile

is not exponential but follows an oscillation pattern. From Figures 7-3 and

7-4, it is also clearly shown that the flow profile is independent of the

Reynolds number greater than 500. The independency of the flow profile

on the Reynolds number, concluded by Price [1968], is discounted here

since this author measured the flow profiles only between NRe= 1470 and

4350.

209
Discussion

3.5

-3

2.5

2
•©-e-eee-o
1.5

*-* A A AM A 1

^-•-•<->-^ 0.5

_i i I I I I
0
10 100 1000 10000

NRe

Distance from wall, particle diameter :

—A—6.00 --•©•--1.38 --O--0.96 0.54 — a — 0.18

Figure 7-3: The effect of Reynolds number on the flow velocity for

D/Dp = 12.0.

210
Discussion

3.5
- e — D/Dp = 3

-a— D/Dp = 6

ZJ
1.
_:

"A -A

1.0 _i i I I I M

10 100 1000
N Re

Figure 7-4: The effect of Reynolds number on the flow velocity at 1.10

particle diameters from the wall.

2ll
Discussion

Computer simulations of fluid flow in a bed of mono-sized of spherical

particles were made using the present model to investigate the effect of

the LVD ratio on the flow profile. Both compressible and incompressible

fluids were used as a fluid for this investigation. As shown in Figure 7-5,

for a bed of mono-sized spherical particles with uniform axial properties of

the bed, the flow profiles is independent of the L/D ratio. This result agrees

with the finding of Price [1968], even though he measured the flow profile

at the downstream of the bed.

As discussed by Bird et al. [1960], for pressures up to critical pressure the

fluid viscosity is almost independent of pressure. Also uniform properties in

axial direction can be applied for mono-sized packed bed system. For an

incompressible fluid, if the parallel flow condition is achieved, it is clear tha

the Reynolds number of the flowing fluid is independent of the axial

direction in the bed, therefore, the flow distribution is independent of the

L/D ratio. To explain why the flow profile for a compressible fluid is also

independent of the L/D ratio, it is assumed that there is no significant

variation of bed properties in the axial direction and the fluid viscosity is

also independent of pressure. It is well known that the density of a

compressible fluid is a function of pressure and by assuming that the ideal

gas behaviour can be applied to correlate the density-pressure relation,

the increasing of the fluid density is linear with the increasing of the

pressure. On the other hand, based upon mass balance equation, the

212
Discussion

10 20 30 40 50 60
L/D

Reynolds number:
--•<>-• 10 ---O-- 100 ---A-- 1000

Figure 7-5: The effect of the L/D ratio upon the flow velocity at 1.0

particle diameters from the wall of a bed of mono-sized

spherical particles.

213
Discussion

average superficial velocity is linearly decrease with increasing pressure.

Therefore, the Reynolds number is independent of the pressure and as a

result the flow profile is also independent of L/D ratio for a compressible

fluid.

The effect of temperature on the fluid flow distribution also has been

studied by using the present mathematical model. Figure 7-6 exhibits this

temperature effect for various Reynolds numbers of a bed of mono-sized

spherical particles with D/DP ratio of 12.0. It is obvious from Figure 7-6 that

the flow profile is independent of temperature for a given Reynolds

number.

The dependency of the flow profile on temperature concluded by

Vortmeyer and Schuster [1983] is discounted here since these authors

investigated the effect of temperature by maintaining a constant average

superficial velocity, rather than a constant Reynolds number. Therefore,

the variation of the flow profile obtained in their investigation, is actually a

effect of the Reynolds number variation rather than of the temperature

variation.

Usually, in the study of mass and heat transfer in packed bed systems, it is

assumed that the effect of the column diameter can be neglected. It can

be seen that the average superficial velocity, particle diameter, bed

214
Discussion

E
3

Temperature, K

Reynolds number:
_g_ 1 _*_ 10 -&- ioo -e-1000

Figure 7-6: The effect of the temperature on the flow profile at 1.10

particle diameters from the wall of a bed of mono-sized

spherical particles.

215
Discussion

voidage and physical properties of the fluid, are usually used to predict the

mass and heat transfer coefficients [Perry and Green, 1984]. It may be

appropriate for packed bed systems with very big value of the D/DP ratio,

but probably is poor for systems having small D/DP ratio. For packed bed

systems with a small value of D/DP ratio, the velocity profile is far from the

flat profile condition, therefore, average superficial velocity can not

properly represent the local condition of fluid dynamics.

However, it is also a fact that the flat profile velocity assumption has been

very helpful, and much reducing the calculation time for analysing and

designing of the packed bed systems [Himmelblau and Bischoff, 1968].

For these reasons, it has seemed desirable to investigate the values of

D/Dp in which a flat profile assumption is applicable.

The present model is further used to investigate the effect of the D/DP ratio

on the deviation of local superficial velocity from average superficial

velocity (Figure 7-7). An inspection of Figure 7-7 reveals qualitatively the

deviation from flat profile flow condition as a function of the D/DP ratio for

mono-sized particles bed. At a Reynolds number below 500 the percent

deviation from the flat profile is found to decrease with corresponding

increase in the D/DP ratio and the Reynolds number. Moreover, for a

Reynolds number above 500, the deviation from the flat profile condition is

independent of the Reynolds number. It agrees with Figures 7-3 and 7-4 in

216
Discussion

100
Reynolds number
—B—1
—•—10

80-

60-

c
g
_
0)
a
40-

20-

_l J^ L.
0 - 1 —I- -I-
0 20 40 60 80 100
D/Dp

Figure 7-7: The effect of D/D P ratio on the deviation from the flat

profile condition.

217
Discussion

which there are no more significant changes in the dimensionless flow

profiles for Reynolds numbers above 500.

Considering that the recommended safety or over-design factor for packed

columns is about 15% [Peter and Timmerhaus, 1968], then for D/DP ratios

above 75, it is reasonable to apply the flat profile flow assumption. The

statement of Cairns and Prausnitz [1959] that the flat profile assumption

can be applied for systems with D/DP ratio above 10 is discounted here

since the deviation is far higher than 15%, and even for a Reynolds

number above 500 it is about 25%. However, when the condition of the

packed bed system is sensitive to the flow profile, for example hot spot

formation and nuclear reaction, the flat profile assumption should be

avoided.

Figure 7-8 illustrates the effect of particle diameter on flow profile inside

packed beds. It is clearly shown that the variation of flow profile is more

evident as an effect of the D/DP ratio rather than of particle diameter. The

significant effect of particle diameter on flow profile, reported by

Ziolkowska and Ziolkowski [1993], is argued here as being incorrect. This

is because these authors used the same column diameter to investigate

the effect of particle diameter, so their data is more representative of the

effect of the D/DP ratio rather than of DP. The significant effect of D/DP ratio

rather than of DP itself on the radial porosity in packed beds, as has been

218
Discussion

2.5
-©••• (R-r)/Dp=0.2;D/Dp=10
- D - (R-r)/Dp=1.0;D/Dp=10
2.4 - - © — (R-r)/Dp=0.2;D=160 m m
-_$_ — (R-r)/Dp=1.0;D=160 m m

2.3

2.2 -

5
.3
-£2.1

2.0 --

1.9 -

1.8 -

-a

1.7 -I ' H
3.0 5.0 7.0 9.0 11.0 13.0 15.0 17.0

Dp, m m

Figure 7-8: A typical effect of particle diameter on flow profile inside a

bed of mono-sized spherical particles (uM = 0.1 m/s).

219
Discussion

reported by Govindarao and Froment [1986] and Mueller [1992], is an

additional reason for the above conclusion.

The present model of fluid flow distribution, both inside and downstream of

packed beds, was developed by assuming the flow characteristic is a

combination of continuous and discontinuous systems. A good

performance of the model in predicting the flow phenomena in packed

beds also was clearly demonstrated. Therefore, it is suggested that this

mathematical model can then be directly incorporated into mass and

energy balance equations for the further mathematical simulation of

packed bed systems. Moreover, based upon investigation by the present

model, it is clearly shown that the Reynolds number and the bed

characteristics influence the flow distribution. However, when the Reynolds

number is higher that 500, the flow distribution is almost independent of

the Reynolds number.

7.2 SIMILARITY CRITERIA OF PACKED BED SYSTEMS

For economic and safety reasons, usually a physical modelling of systems

is required [Johnstone and Thring, 1957; Fleming, 1958; Peter and

Timmerhaus. 1968; Szucs, 1980; Euzen et ai, 1993]. Moreover, as stated

by Geldart [Rhodes, 1990], there is too little understanding of background

theory of packed bed characteristics, as an essential factor for fluid flow

distribution, in which the occurrence of a size segregation is possible.

220
Discussion

Therefore, experimental work becomes a key factor in studying the

system.

In order to achieve a greater confidence in studying of physical

phenomena by using a physical model, a knowledge of similarity criteria is

essential [Schuring, 1977; Szucs, 1980; Zlokarnik, 1991]. Therefore, it

seems reasonable to investigate the similarity criteria of a packed bed

system based on the fluid flow distribution point of view. In this

investigation, the two similar conditions of phenomena (systems) are

defined such that their corresponding characteristics (features,

parameters) are connected by bi-unique (one-to-one) mappings

(representations), as stated by Szucs [1980],

Studying flow phenomena in a packed bed system is usually based upon

dynamic and geometric similarity. If the flow distribution factor can be

neglected, then the Reynolds number, P"M p , and the Froude number,

—P-Y , are acceptable as similarity criteria [Hoftyzer, 1957]. if the fluid is


U
M

D2
pg p
liquid, the Bond number, , also should be considered [Hoftyzer,
o
1957]. However, as also reported by Hoftyzer [1957], the performance of a

model based on the above similarity to predict the prototype (actual size)

221
Discussion

behaviour is only about 2 5 % . Therefore, it is a continuing need for

investigation of similarity criteria on the basis of the fluid flow distribution

For fluid flow in packed beds, the characteristic of the bed can be

represented by the bed permeability. The bed permeability as noted in

Section 3.3, is influenced by the particle size and the particle size

distribution (the spread and the size range), as discussed by Yu and Zulli

[1994]. Therefore, the similarity of particle size and the particle size

distribution should be satisfied between the model and the prototype.

Considering that for the physical model with geometrical size smaller than

prototype, it is often impossible in practice to satisfy the requirement of

similar D/DP ratio together with the particle size distribution (the spread

and the size range). This is because, in engineering practice, the

components used to construct a particle mixture are themselves particles

mixtures [Yu and Standish, 1993a]. Accordingly, packed bed models are

easier to study by using actual (prototype) particle bed rather than

generate smaller particles with similar size distribution.

As discussed earlier, the fluid flow distribution is strongly dependent on

bed character and the Reynolds number. The Reynolds number has been

recognised by Hoftyzer [1957]; therefore, it seems reasonable to explore

110
Discussion

the bed character for similarity criteria and its possibility in developing of a

distorted model.

The similarity requirement of the D/DP ratio is still more difficult to satisfy

even for mono-sized particle beds since the pressure drop becomes very

high for small values of particle diameter. For these reasons, it has

seemed desirable to investigate the possibility of distortion for the similarity

criteria of the D/DP ratio.

To facilitate the investigation of the distorted model for D/DP ratio in

physical modelling, it is assumed that the deviation from flat profile

condition can represent the variation of flow profile. Hence, the minimum

D/DP ratio of the model is defined by the value of the D/DP ratio in which

the difference of the deviation from the flat flow profile condition with

prototype is equal to 10%. Figure 7-9 illustrates the effect of D/DP ratio and

Reynolds number on the minimum D/DP ratio of the model.

If a parallel flow condition and uniform condition in axial direction of the

bed is achieved, the flow distribution is independent of LVD ratio. However,

a L/D ratio higher than 1.0 is required to achieve a parallel flow distribution

[Szekely and Poveromo, 1975]. Therefore, in order to minimise the end

effects (at inlet and outlet), it seems reasonable to maintain the L/D ratio of

the model higher than 3.0.

223
Discussion

70

Reynolds number:

60
••--•IO
--—100
•-.--•500
-1000
50-

40--

o
A
Ow

Q
Q 30-

20-

10

_i i i

50 100 150 200

(D/Dp)prototype

Figure 7-9: Minimum D/D P ratio of a physical model as a function of

D/Dp ratio of prototype and Reynolds number.

224
Discussion

Although, the use of a physical model without any distortion of similarity

criteria is more preferable, it is often difficult to satisfy all requirements

unless it is by using the model with conditions exactly the s a m e as the

prototype conditions. Therefore, the results of this investigation are

expected to be useful in the development of physical models of packed

bed systems.

225
CHAPTER EI-3HT

CONCLUSIONS

A mathematical model of the flow distribution at inside and at downstream of

packed beds for single-phase fluid has been developed. The model is based

on the analysis of a packed bed viewed as a combination of continuous and

discontinuous systems of fluid between voids in the bed. The model is

applicable to both compressible and incompressible fluids. There is a

reasonable agreement between the prediction of the model and the

measured values. The new model compares favorably with previous models

in terms of accuracy and simplicity, and does not require new empirical

constants.

The disagreement of the previous investigators of the effect of the Reynolds

number and the particle diameter on fluid flow distribution in packed beds is

clearly explained by using the present mathematical model. It is demonstrated

that the Reynolds number has a significant effect on flow distribution only fo

NRe less than 500, and when the Reynolds number is higher than 500, the

flow profile is determined by bed characteristics. Moreover, that model result

226
Conclusions

has also demonstrated that it is the D/D P ratio that has a significant effect on

the flow profile rather than the particle diameter, as hitherto believed.

The results have also shown that models based on discontinuous systems

are only successful for local porosity of less than 0.5. When the local porosity

is higher than 0.5, especially at the vicinity of the wall, a model based on the

continuous systems approach, as used in the present case, is more accurate.

The results of the present model regarding the flat flow profile assumption for

packed bed systems have clearly shown that the deviation from the flat profile

condition does not only depend on the D/DP ratio, but is also depends on the

Reynolds number. This conclusion was also used to suggest some rules of

how a distorted physical model may be generated in term of D/DP ratio and

L/D ratio.

Finally, one practical conclusion that suggests itself is that the present

mathematical model of fluid flow distribution, together with energy and mass

balances equations and other rate processes equations, may be expected to

be useful in process design and process optimization of packed bed systems,

in general.

227
REFERENCES

Agarwal, P.K., 1988, Chem. Engng. Sci, 43, 2501-2510.

Agarwal, P.K., Mitchell, W.J. and Nauze, R.D.L., 1988, Chem. Engng.

Sci, 43, 2511-2521.

Agarwal, P.K. and O'Neill, B.K., 1988, Chem. Engng. Sci, 43, 2487-2499

Arthur, J.R., Linnett, J.W., Raynor, E.J. and Sington, E.P.C., 1950, Tr

Faraday Soc, 46, 270-281.

Atkinson, B., Brocklebank, M.P., Card, C.C.H. and Smith, J.M., 1969,

AlChEJ., 15, 548-553.

Beavers, G.S. and Joseph, D.D., 1967, J. Fluid Mech., 30, 197-207.

Benard, C.J., 1988, Handbook of Fluid Flow Metering, The Trade and

Technical Press Ltd., Surrey.

Benedict, R.P., 1977, Fundamentals of Temperature, Pressure, and Flow

Measurements, 2nd Ed., John Wiley and Sons, New York.

Benenati, R.F. and Brosilow, C.B., 1962, AlChEJ., 8, 359-361.

Berman, N.S. and Santos, V.A., 1969, AlChEJ., 15, 323-327.

Bey, O. and Eigenberger, G., 1997, Chem. Engng. Sci, 52, 365-1376.

Bird, R.B., 1965, Proceeding of the A.I.Ch.E.-l.Chem.E. Symposium

Series, No. 4, London, pp. 4.3-4.13.

Bird, R.B., Stewart, W.E. and Lightfoot, E.N., 1960, Transport

Phenomena, John Wiley and Sons, Inc. New York.

228
Bibliography

Blevins, R.D., 1984, Applied Fluid Dynamics Handbook, Van Nostrand

Reinhold Co., Melbourne.

Blum, E.H. and Wilhelm, R.H., 1965, Proceeding of the A.I.Ch.E.-

l.Chem.E. Symposium Series, No. 4, London, pp. 4.21-4.27.

Bo, M.K., Freshwater, D.C. and Scarlett, B., 1965, Trans. Instn. Chem.

Engr., 43, T228-T232.

Bolton, W., 1996, Instrumentation and Measurement, 2nd ed., Newnes,

Oxford.

Brady, J.F., 1984, Phys. Fluids, 27, 1061-1067.

Brinkman, H.C., 1947, Appl. Sci. Res. Sect. A1, 27-34.

Brown, G.G., Foust, A.S., Katz, D.L., Schneidewind, R., White, R.R.,

Wood, W.P., Brown, G.M., Brownell, L.E., Martin, J.J., Williams,

G.B., Banchero, J.T. and York, J.L., 1950, Unit Operations, John

Wiley and Sons, Inc., New York.

Buchlin, J.M., Riethmuller, M. and Ginoux, J.J., 1977, Chem. Engng. Sci

32, 1116-1119.

Burden, R.L., and Faires, J.D., 1993, Numerical Analysis, 5th. Ed., PWS

Publishing Co., Boston.

Burnett, D.S., 1987, Finite Element Analysis, from Concepts to

Applications, Addison-Wesley Publishing Co., Massachusetts.

Bowlus, D.A., and Brighton, J.A., 1968, J. Basic Engng., 90, 431-433.

Cairns, E.J. and Prausnizt, J.M., 1959, Ind. Engng. Chem., 51, 1441-

1444.

229
Bibliography

Chen, R.Y., 1973, J. Fluids Engng., 95, 153-158.

Cheng, P. and Hsu, C.T., 1986, Int. J. Heat Mass Transfer, 29, 1843-

1853.

Cheng, P. and Vortmeyer, D., 1988, Chem. Engng. Sci, 43, 2523-2532.

Choudhary, M., Szekely, J. and Weller, S.W., 1976a, AlChEJ., 22, 1021-

1027.

Choudhary, M., Szekely, J. and Weller, S.W., 1976b, AlChEJ., 22, 1027-

1032.

Chow, C.Y., 1979, An Introduction to Computational Fluid Mechanics,

John Wiley and Sons, New York.

Christiansen, E.B. and Lemmon, H.E., 1965, AlChEJ., 11, 995-999.

Christopher, R.H. and Middleman, S., 1965, Ind. Engng. Chem. Fund., 4

422-426.

Cohen, Y. and Metzner, A.B., 1981, AlChE J., 27, 705-715.

Courant, R. and Hilbert., D., 1953, Methods of Mathematical Physics,

Vol. 1, Interscience Publishers, Inc., New York.

Cumberland, D.J. and Crawford, R.J., 1987, The Packing of Particles,

Elsevier, Amsterdam.

Davies, O.L. and Goldsmith, P.L., 1977, Statistical Methods in Researc

and Production, 4th ed., Longman Group Limited, London.

Debbas, S. and Rumpf, H., 1966, Chem. Engng. Sci, 21, 583-607.

Delmas, H. and Froment, G.F., 1988, Chem. Engng. Sci, 43, 2281-2287.

230
Bibliography

Dorweiler, V.P. and Fahein, R.W., 1959, AlChEJ., 5, 139-144.

Durlofsky, L. and Brady, J.F., 1987, Phys. Fluids, 30, 3329-3341.

Durlofsky, L. and Brady, J.F., 1984, Phys. Fluids, 27, 3329-3341.

Ergun, S., 1952, Chem. Engng. Prog., 48, 89-94.

Ergun, S. and Orning, A.A., 1949, Ind. Engng. Chem., 41,1179-1184.

Euzen, J.P., Trambouze, P. and Wauquier, J.P., 1993, Scale-Up

Methodology for Chemical Processes, Gulf Publishing Co., Paris.

Fahien, R.W. and Stankovic, I.L., 1979, Chem. Engng. Sci, 34, 350-1354

Fayed, M.E. and Otten, L., 1984, Handbook of Powder Science and

Technology, Van Nostrand Reinhold Co., New York.

Fleming, R., 1958, Scale-Up in Practice, Chapman and Hall, LTD, London

Foscolo, P.U., Gibilaro, L.G. and Waldram, S.P., 1983, Chem. Engng.

Sci, 38, 1251-1260.

Foust, A.S., Wenzel, L.A., Clump, C.W., Maus, L. and Andersen, L.B.,

1960, Principles of Unit Operations, John Wiley and Sons, Inc., New

York.

Franks, R.G., 1972, Modeling and Simulation in Chemical Engineering,

John Wiley and Sons, Inc., New York.

Furnas, CC, 1931, Ind. Engng. Chem., 23, 1052-1058.

Gauvin, W.H. and Katta, S., 1973, AlChEJ., 19, 775-783

German, R.M., 1989, Particle Packing Characteristics, Metal Powder

Industries Federation, New Jersey.

231
Bibliography

German, R.M., 1981, Powder Technoi, 30, 81-86.

Givler, R.C. and Altobelli, S.A., 1994, J. Fluid Mech., 258, 355-370.

Goodling, J.S., Vachon, R.I., Stelpflug, W.S., Ying, S.J. and Khader,

1983, Powder Technoi, 35, 23-29.

Gotoh, K., Jodrey, W.S. and Tory, E.M., 1978, Powder Technoi, 20, 257-

260.

Govindarao, V.H. and Froment, G.F., 1986, Chem. Engng. Sci, 41, 533-

539.

Govindarao, V.H. and Ramrao, K.V.S., 1988, Chem. Engng. Sci, 43,

2544-2545.

Govindarao, V.H., Subbanna, M., Rao, A.V.S. and Ramrao, K.V.S., 1990,

Chem. Engng. Sci, 45, 362-364.

Haughey, D.P. and Beveridge, G. S. G., 1966, Chem. Engng. Sci, 21,

905-916.

Himmelblau, D.M. and Bischoff, K.B.,1968, Process Analysis and

Simulation Deterministic Systems, John Wiley and Sons, Inc.,

New York.

Hoftyzer, P.J., 1957, Proceeding of the Joint Symposium the Scaling-U

Chemical Plant and Processes, Instn. Chem. Engrs., London, pp.

S73-S77.

Jenson, V.G. and Jeffreys, G.V., 1963, Mathematical Methods in Chemic

Engineering, Academic Press, London.

232
Bibliography

Johnson, G.W. and Kapner, R.S., 1990, Chem. Engng. Sci, 45, 329-339.

Johnstone, R.E. and Thring, M.W., 1957, Proceeding of the Joint

Symposium the Scaling-Up of Chemical Plant and Processes, Instn.

Chem. Engrs., London, pp. S7-S8.

Kalthoff, 0. and Vortmeyer, D., 1980, Chem. Engng. Sci., 35,1637-164

Kececioglu, I. and Jiang, Y., 1994, J. of Fluids Engng., 116, 165-170

Keller, H.B., 1968, Numerical Methods for Two-Point Boundary Value

Problems, Blaisdell Publishing Co., Waltham.

Kondelik, P., Horak, J. and Tesarova, J., 1968, Ind. Engng. Chem. Pr

Des. Dev., 7, 250-252.

Kubie, J., 1988, Chem. Engng. Sci, 43, 1403-1405.

Kufner, R. and Hofmann, H, 1990, Chem. Engng. Sci, 45, 2141-2146.

Lamb, D.E. and Wilhelm, R.H., 1963, Ind. Engng. Chem. Fund., 2, 173-

182

Langhaar, H.L., 1942, J. Appl. Mech., 9, A55-58.

Larson, R.E. and Higdon, J.J.L., 1986, J. Fluid Mech., 166, 449-472.

Leitzelement, M., Lo, CS. and Dodds, J., 1985, Powder Technoi, 41,

159-164.

Lerou, JJ. and Froment, G.F., 1977, Chem. Engng. Sci, 32, 853-861.

Leva, M., 1992, Chem. Engng. Prog., 88, 65-72.

Levenspiel, O., 1984, Engineering Flow and Heat Exchange, Plenum

Press, New York.

233
Bibliography

Lundgren, T.S., 1972, J. Fluid Mech., 51, 273-299.

MacDonald, M.J., Chu, C.F., Guilloit, P.P. and Ng, K.M., 1991, AlChE

37,1583-1588.

MacDonald, I.F., El-Sayed, M.S., Mow, K. and Dullien, F.A.L., 1979,

Engng. Chem. Fund. 18, 199-208.

Macrae, J.C. and Gray, W.A., Brit. J. Appl. Phys., 12, 164-172.

Martin, H., 1978, Chem. Engng. Sci, 33, 913-919.

Masuoka, T. and Takatsu, Y., 1996, Int. J. Heat Mass Transfer, 39, 28

2809.

McGreavy, C. and Cresswell, D.L., 1968, Proceeding of the Fourth

European Symposium of Chemical Reaction Engineering, 9-11

Sept., Brussels, pp. 59-71.

McGreavy, C, Foumeny, E.A. and Javed, K.H., 1986, Chem. Engng. Sci,

41, 787-797.

Mehta, D. and Hawley, M.C, 1969, Ind. Engng. Chem. Proc. Des. Dei/.,

280-282.

Merwe, D.F.V.D. and Gauvin, W.H., 1971a, AlChEJ., 17, 402-409.

Merwe, D.F.V.D. and Gauvin, W.H., 1971b, AlChEJ., 17, 519-528.

Mickley, H.S., Smith, K.A. and Korchak, E.I., 1965, Chem. Engng. Sci

237-246.

Mickley, H.S.,Sherwood, T.K. and Reed, C.E., 1957, Applied Mathemati

in Chemical Engineering, 2nd Ed., McGraw-Hill Book Co., New York.

234
Bibliography

Miconnet, M., Guigon, P. and Large, J.F., 1982, Int. Chem. Engng., 22

133-141.

Morales, M., Spinn, C.W. and Smith, J.M., 1951, Ind. Engng. Chem., 43

225-232.

Morkel, W.S.T., and Dippenaar, R.J., 1992, 1(f PTD Conference

Proceedings, pp. 77-89.

Mueller, G.E., 1993, Powder Technoi, 77, 313-319.

Mueller, G.E., 1992, Powder Technoi, 72, 269-275.

Mueller, G.E., 1991, Chem. Engng. Sci, 46, 706-708.

Murphy, G., 1950, Similitude in Engineering, The Ronald Press Co.,

New York.

Murphy, D.E. and Sparks, R.E., 1968, l&EC Fund., 7, 642-645.

Newell, R.,1971, Velocity Distribution in Packed Beds of Rectangular

Geometry, M.Sc. Thesis, University of New South Wales, Australia.

Newell, R., and Standish, N., 1973, Met. Trans., 4B, 1851-1857.

Nield, D.A., 1983, AlChEJ., 29, 688-689.

Nield, D.A., Juqueira, S.L.M. and Lage, J.L., 1996, J. Fluid Mech., 3

201-214.

Ouchiyama, N. and Tanaka, T., 1989, Ind. Engng. Chem. Res., 20, 66-71

Ouchiyama, N. and Tanaka, T, 1981, Ind. Engng. Chem. Fund., 20, 66-

71.
Bibliography

Ouchiyama, N. and Tanaka, T., 1980, Ind. Engng. Chem. Fund., 19, 33

340.

Ouchiyama, N. and Tanaka, T., 1975, Ind. Engng. Chem. Proc. Des. De

14,286-289.

Ouchiyama, N. and Tanaka, T., 1974, Ind. Engng. Chem. Proc. Des. De

13, 383-389.

Perry, A.E., 1982, Hot-wire Anemometry, Clarendon Press, Oxford.

Perry, R.H. and Green, D., 1984, Perry's Chemical Engineers' Handboo

6th ed., McGraw-Hill Book Co., New York.

Peter, M.S. and Timmerhaus, K.D., 1968, Plant Design and Economics f

Chemical Engineers, 2nd Ed., McGraw-Hill Book Co., New York.

Pillai. K.K., 1977, Chem. Engng. Sci, 32, 59-61.

Poirier, D.R. and Geiger, G.H., 1994, Transport Phenomena in Materia

Processing, TMS, Pennsylvania.

Poveromo, J.J., Szekely, J. and Propster, M., 1975, Proceedings of B

Furnace Aerodynamics Symposium, Wollongong, pp. 1-8, 171-172.

Prausnitz, J.M. and Wilhelm, R.H., 1957, Ind. Engng. Chem., 49, 978-

Price, J., 1968,. Mech. Chem. Eng. Trans. Aust, MC4, 7-14.

Propster, M. and Szekely, 1977, Powder Technoi, 17, 123-138.

Puncochar, M. and Drahos, J., 1993, Chem. Engng. Sci, 48., 2173-2175

Ranz, W.E., 1952, Chem. Eng. Prog., 48, 247-253.


Bibliography

Reid, R.C. and Sherwood, T.K., 1966, The Properties of Gases and

Liquids, Their Estimation and Correlation, 2nd ed., McGraw-Hill

Book Co., New York.

Reid, R.C, Prausnizt, J.M. and Sherwood, T.K., 1977, The Properties o

Gases and Liquids, 3rd ed., Mc Graw-Hill Book Co., New York.

Rhodes, M., 1990, Principles of Powder Technology, John Wiley and

Sons, Chichester.

Ridgway, K., and Tarbuck, K.J., 1968a, Chem. Engng. Sci, 23,1147-1155.

Ridgway, K., and Tarbuck, K.J., 1968b, Chem. Proc. Engng., 49, 103-105

Ridgway, K., and Tarbuck, K.J., 1967, Brit. Chem. Engng., 12, 384-388.

Roblee, L.H.S., Baird, R.M. and Tierney, J.W.,1958, AlChEJ, 4, 460-464

Saleh, S., Thovert, J.F. and Adler, P.M., 1993a, Chem. Engng. Sci, 48,

2839-2858.

Saleh, S., Thovert, J.F. and Adler, P.M., 1993b, AlChEJ., 39, 1765-177

Saunders, O.A. and Ford, H., 1940, J. Iron Steellnst, CXLI, 291p-329p.

Schertz, W.W. and Bischoff, K.B., 1969, AlChEJ., 15, 597-604.

Schuring, D.J., 1977, Scale Models in Engineering Fundamentals and

Applications, Pergamon Press, Oxford.

Schmidt, F.W. and Zeldin, B., 1969, AlChEJ., 15, 612-614.

Schwartz, CE. and Smith, J.M., 1953, Ind. Engng. Chem., 45, 1209-1218.

Scott, G.D., 1962, Nature, 194, 4831-4832.

237
Bibliography

Scott, G.D. and Kovacs, G.J., 1973, J. Phys. D: Appl. Phys., 6, 1007-

1010.

Slattery, J.C, 1969, AlChEJ., 15, 866-872.

Slattery, J.C, 1968, AlChEJ., 14, 50-56.

Smith, J.M. and Van Ness, H.C., 1975, Introduction to Chemical

Engineering Thermodynamics, 3rd Ed., McGraw-Hill, New York.

Standish, N., 1990, Bulk Density of Coal, A Booklet of NERDDP Project,

University of Wollongong, Australia.

Standish, N., 1984, Chem. Engng. Sci, 39, 1530.

Standish, N., 1979, Principles in Burdening and Bell-Less Charging,

Nimaroo Publisers, Wollongong.

Standish, N. and Borger, D.E., 1979, Powder Technoi, 22, 121-125.

Standish, N. and Collins, D.N., 1983, Powder Technoi, 36, 55-60.

Standish, N. and Leyshon, P.J., 1981, Powder Technoi, 30, 119-121.

Standish, N. and Yu, A.B., 1987a,Powcter Technoi, 49, 249-253.

Standish, N. and Yu, A.B., 1987b,Powder Technoi, 53, 69-72.

Stanek, V. and Szekely, J., 1974, AlChEJ., 20, 974-980.

Stephenson, G., 1973, Mathematical Methods for Science Students, 2nd

Ed., Longman Scientific and Technical, London.

Stephenson, J.L. and Stewart, W.E., 1986, Chem. Engng. Sci, 41, 2161-

2170.

Szekely, J. and Kajiwara, Y., 1979, Met. Trans., 10B, 447-453.

23X
Bibliography

Szekely, J. and Poveromo, J.J., 1975, AlChEJ., 21, 769-775.

Szekely, J. and Propster, M., 1977, Ironmaking & Steelmaking, 1, 15-

Szucs, E.. 1980, Similitude and Modeling, Elsevier Sci. Publishing

Amsterdam.

Tao, B.Y., 1988a, Chem. Engng., 95, 107-110.

Tao, B.Y., 1988b, Chem. Engng., 95, 85-92.

Tao, B.Y., 1987a, Chem. Engng., 94, 109-113.

Tao, B.Y., 1987b, Chem. Engng., 94, 145-148.

Thadani, M.C. and Peebles, F.N., 1966, Ind. Engng. Chem. Proc. Des.

Dev., 5, 265-268.

Tsotsas, E. and Schlunder, E.U., 1990, Chem. Engng. Sci, 45, 819-83

Tsotsas, E. and Schlunder, E.U., 1988, Chem. Engng. Sci, 43, 1200-

1203.

Vortmeyer, D. and Haidegger, E., 1991, Chem. Engng. Sci, 46, 2651-

2660.

Vortmeyer, D. and Michael, K., 1985, Chem. Engng. Sci, 40, 2135-213

Vortmeyer, D. and Schuster, J., 1983,C..em. Engng. Sci, 38, 1691-16

Vortmeyer, D. and Winter, R.P.,1984, Chem. Engng. Sci, 39, 1430-143

Vrentas, J.S. and Duda, J.L., 1967, AlChEJ., 13, 97-101.

Vrentas, J.S., Duda, J.L and Bargeron, K.G., 1966, AlChE J., 12, 837

844.

Wasan, D.T. and Baid, K.M., 1971, AlChE J., 17, 729-731.

239
Bibliography

White, F.M., 1986, Fluid Mechanics, 2nd Ed., McGraw-Hill Book Company,

New York.

Wilkinson, D., 1985, Phys. Fluids, 28, 1015-1022.

Williams, J.C, 1976, Powder Technoi, 15, 245-251.

Wit, A. D., 1995, Phys. Fluids, 7, 2553-2562.

Yu, A.B. and Standish, N., 1993a, Powder Technoi, 76, 113-124.

Yu, A.B. and Standish, N., 1993b, Powder Technoi, 74, 205-213.

Yu, A.B. and Standish, N., 1993°, Ind. Eng. Chem. Res, 32, 2179-2182.

Yu, A.B. and Standish, N., 1991, Ind. Eng. Chem. Res, 30, 1372-1385.

Yu, A.B. and Standish, N., 1988, Powder Technoi, 55, 171-186.

Yu, A.B. and Standish, N., 1987, Powder Technoi, 49, 249-253.

Yu, A.B. and Zulli, 1994, Powder Handling Process., 6, 171-177.

Yu, A.B., Standish, N. and McLean, A., 1993, J. Am. Ceram. Soc.,76,

2813-2816.

Ziolkowska, I. And Ziolkowski, D., 1993, Chem. Engng. Sci, 48, 3283-

3292.

Ziolkowski, D. and Szustek, S., 1989, Chem. Engng. Sci, 44,1195-1204.

Zlokarnik, M., 1991, Dimensional Analysis and Scale-up in Chemical

Engineering, Springer-Verlag, Berlin.

Zou, R.P. and Yu, A.B., 1996, Powder Technoi, 88, 71-79.

240
APPENDIX

ALGORITHMS OF VELOCITY PROFILE CALCULATION

241
INPUT
bed properties
fluid properties
superficial velocity

COMPUTE GUESS
pressure drop

COMPUTE
radial porosity distribution

O
COMPUTE
radial velocity profile

CHECK
total mass-flow rate

UPDATE

Figure A-1: Flow diagram of the velocity profile calculation by the

present model.

242
INPUT
bed properties
fluid properties
superficial velocity

GUESS
uz at r = 0
COMPUTE
radial porosity distribution

O
COMPUTE
radial velocity profile

CHECK
total mass-flow rate

UPDATE
uz at r = 0

Figure A-2: Flow diagram of the velocity profile calculation by

Vortmeyer and Schuster [1983] model.

243
INPUT
bed properties
fluid properties
superficial velocity

COMPUTE GUESS
pressure drop Ks

COMPUTE
radial porosity distribution

o
COMPUTE
radial velocity profile

CHECK
total mass-flow rate

UPDATE

No

OUTPUT
uz=/(r)

ure A-3:Flow diagram of the velocity profile calculation by Stanek

and Szekely [1974] model.

You might also like