Seville, J. (1997) Processing of Particulate Solids
Seville, J. (1997) Processing of Particulate Solids
Particle Classification
K. Heiskanen
Hardback (0412493004), 330 pages
Jonathan Seville
Professor of Chemical Engineering
School of Chemical Engineering
University of Birmingham, UK
Ugur Tuzun
Professor of Process Engineering
Department of Chemical and Process Engineering
University of Surrey
Guildford, UK
and
Roland Clift
Professor of Environmental Technology
Centre for Environmental Strategy
University of Surrey
Guildford, UK
ISBN-13:978-94-010-7152-9 elSBN:978-94-009-1459-9
DOI:I0.I0071978-94-009-1459-9
Apart from any fair dealing for the purposes of research or private study,
or criticism or review, as permitted under the UK Copyright Designs and
Patents Act, 1988, this publication may not be reproduced, stored, or
transmitted, in any form or by any means, without the prior permission
in writing of the publishers, or in the case of reprographic reproduction
only in accordance with the terms of the licences issued by the Copyright
Licensing Agency in the UK, or in accordance with the terms of licences
issued by the appropriate Reproduction Rights Organization outside the
UK. Enquiries concerning reproduction outside the terms stated here
should be sent to the publishers at the London address printed on this
page.
The publisher makes no representation, express or implied, with
regard to the accuracy of the information contained in this book and
cannot accept any legal responsibility or liability for any errors or
omissions that may be made.
A catalogue record for this book is available from the British Library
Library of Congress Catalog Card Number: 96-83857
Preface xi
1 Particle characterisation 1
1.1 Particle size distributions 1
1.1.1 Definitions 1
1.1.2 Methods of presenting size data 3
1.1.3 Averages 7
1.1.4 Weighted distributions 9
1.1.5 Model distribution functions 16
1.2 Particle size measurement 18
1.2.1 Particle size 18
1.2.2 Particle shape 21
1.2.3 Sampling 26
1.2.4 Methods of particle size measurement 30
1.3 Nomenclature 48
A1 Derivation and properties of the log-normal distribution 49
References 52
2 Particles in fluids 53
2.1 Single particles 53
2.1.1 Fluid-particle drag 53
2.1.2 Fluid motion 56
2.1.3 Free fall or rise 59
2.1.4 Non-continuum effects 76
2.2 Unsteady motion of single particles 79
2.2.1 Drag in unsteady motion 79
2.2.2 Acceleration 80
2.2.3 Curvilinear motion 81
2.3 Assemblages of particles 85
2.3.1 Settling and particulate fluidisation 85
2.3.2 Flow through packed beds 86
2.4 Flow of particlelfluid mixtures in pipes 87
2.4.1 Continuity 88
2.4.2 Momentum balance 91
viii Contents
3 Particle mechanics 99
3.1 Interparticle forces 99
3.1.1 van der Waals forces 99
3.1.2 Liquid bridge forces 109
3.1.3 Electrostatic forces 120
3.1.4 Comparison of the magnitude of interparticle forces 123
3.2 Effects of interparticle forces at contacts 125
3.2.1 Contact mechanics 125
3.2.2 Assembly mechanics 127
3.3 Friction at a single contact 132
3.4 Impact and rebound of particles 135
3.4.1 The coefficient of restitution 137
3.4.2 The effect ofliquid layers 140
3.5 Nomenclature 142
References 143
Index 369
Preface
Well over half the products of the chemical and process industries are sold in
a particulate form. The range of such products is vast: from agrochemicals to
pigments, from detergents to foods, from plastics to pharmaceuticals.
However, surveys of the performance of processes designed to produce
particulate products have consistently shown inadequate design and poor
reliability. The science of fluid mechanics is usually dated back to Archimedes
and developed through the work of such as Galileo, Newton, D' Alembert,
Darcy and Euler. By the end of the last century, fluid handling had become a
scientifically based technological industry. In contrast, it was only in the 1950s
that the study of solids handling was seriously attempted and many solids
processing devices are still designed on a more-or-Iess empirical basis.
Particle technology is thus a very new subject and it faces new challenges. It
is apparent that chemical and process engineering is becoming less concerned
with the design of plants to produce generic simple chemicals (which are often
single phase fluids) and are more concerned with speciality 'effect' chemicals
which may often be multi-phase and multi-component. Chemical and process
engineers are also being recruited in increasing numbers into areas outside
their traditional fields, e.g. the food industry, pharmaceuticals and the
manufacture of consumer products of all kinds. We intend this book to meet
their needs.
We have aimed at a comprehensive coverage of the technology of
particulate solids in a form which is both sufficiently accessible and
sufficiently concise to be useful to engineering and science students in the
final year of an undergraduate degree and at Masters' level. Although it was
written with students of chemical engineering in mind, concern with granular
solids and powders is not limited to one engineering discipline. We therefore
hope that this book will be of use and interest to students of other disciplines.
The content and style follow a pattern which we have found useful in teaching
undergraduate and graduate students at the Universities of Surrey, Birmingham
and British Columbia.
The book is divided into two halves: fundamentals and applications. The
behaviour of single spheres in simple fluids is a standard part of any chemical
engineering degree; the challenge that we have attempted to meet here is to
widen that treatment to include non-spherical particles over a wide range of
xii Preface
J.P.K. Seville
U. Tiiziin
R. Clift
1
Particle characterisation
The most important characteristics of a particle are its size, its shape and its
density. At first sight these may seem simple unambiguous properties which
should be straightforward to determine. In practice, this is seldom the case,
particularly if the particle in question is smaller than, say, the diameter of a
human hair. This chapter is an introduction to the science of particle
characterisation; for a more detailed examination of this diverse subject the
reader is referred to Allen (1996). In practice, particles seldom occur singly,
and the first part of this chapter is concerned with the properties of particle
size distributions.
much below 0.55 [lm in diameter cannot be resolved using this method, which
has important implications for particle sizing, as will be discussed later. Most
particles show a distribution of particle size, i.e. they are polydisperse. Those
which do not are described as monodisperse, i.e. they are all of the same size.
In practice, it is only possible to approximate to the latter condition, although
in nature, plant pollen, and in industry, polymer latex spheres formed under
zero gravity are close approximations to monodisperse particles.
Fluid particles (i.e. bubbles or drops) are often spherical if they are small
enough (Clift et al., 1978) whereas solid particles are normally not. The term
'diameter' is therefore ambiguous and it is necessary to define an equivalent
diameter of a non-spherical particle as the diameter of a sphere having the
same value of a particular physical property as the particle of interest. Many
equivalent diameters are used. The choice, and the methods of measurement
are discussed later in this chapter.
ni
I".
Jt = - = h.!l.d.
I l
(1.1a)
N
*This section draws on the extremely useful chapter on this subject in w.e. Hinds' book Aerosol
Technology, Wiley (1982), to which the reader is referred for a more comprehensive account of
the subject.
4 Particle characterisation
-
300 -
- '-
100 -
fo-
I-
o Jr 111 1 1 1 1
~
1 1 I I
o 20 40 80 100 150 200
Particle diameter (microns)
I=l 0.03~-
8u
·s
§ 0.02 - e-
'.p
£ e-
0.01 -
o-
n
I I I I I I I I I I I
o 20 40 80 100 150 200
Particle diameter (microns)
Figure 1.3 Fraction per !Lm versus particle size, discrete number distribution.
line may be drawn through the tops of the frequency rectangles at the mean
point of each interval to form a continuous distribution, as in Figure 1.4. (For
the mean of an interval it is usual to take the geometric mean, i.e. (d i d i +1)'h. If
the interval width is small, this differs little from the arithmetic mean, V2(di +
d i +1 ).
For a continuous distribution, the fraction of the total number of particles
having diameters between a and b is the integral under the distribution curve
between these limits
(LIb)
where f(d p) is the continuous frequency distribution function. The total area
is again unity:
(1.2b)
0.035
0.030
::i.
5 0.025
.=-=..... 0.020
~
~ 0.015
0.010
0.005
0.000 I v=---.; • i • • i
Thus the frequency function at any point can be obtained from the slope of
the cumulative distribution function.
Because the cumulative distribution is the integral of the frequency
function, it is less sensitive to 'scatter' in the data. 'Smoothing' of
measurements and interpolation between measured points on the distribution
are therefore simple and reliable. For these reasons, it is common practice to .
work with the cumulative distribution function rather than the frequency
function.
1.1.3 Averages
To define a particle size distribution completely requires a large amount of
information, such as Table 1.1, for example. It is obviously advantageous to
be able to approximate the distribution by some form of mathematical
function, and many such functions are available. Each requires at least two
calculated parameters: one to define the location of the distribution and one
to define its width. The first of these parameters is usually some form of
'average', such as the mean (strictly arithmetic mean), geometric mean,
median or mode.
The arithmetic mean, dp , is given by the sum of all the particle diameters,
divided by the total number:
_ '2:.d '2:.n;d;
dp = - = - - (1.5)
N '2:.n;
= 1000
dJ( dp)dd p (1.6)
The geometric mean, d g , is defined as
dg = [dnldn2dn3
2
I 3 ...
dni]'/N
I (1.7)
The importance of the geometric mean will become clear in later sections.
The median is the diameter for which one half of the total number of
particles are larger, and one half smaller. It divides the frequency distribution
into equal areas and is the diameter which corresponds to F = 0.5 on the
cumulative distribution curve. The advantage of the median is that it is less
affected by skewness (lack of symmetry) of the distribution than the mean.
The mode is the most frequent size, i.e. the highest point on the frequency
curve. For monodisperse distributions all four of these averages are equal.
For distributions which are skewed towards larger sizes as in Figure 1.4,
which is frequently the case for particles,
·mode < median < mean.
Very often the particle sizing device which is used does not provide a direct
measurement of particle diameter but measures some property which is
100
-.
~
'-' 80
~
~
= 60
==~
~
~ 40
••
~
] 20
U=
o
o 20 40 60 80 100 120 140 160 180 200
Particle diameter (J.lm)
Figure 1.5 Cumulative number distribution.
Particle size distributions 9
Lm Jtd~
Iiz=--=--Q (1.8)
N 6 p
dm -_[6 ---Lm
QpJtN
]
'/3 [Ld
--
N
3
JY3 -_ [Ln;d~
--
N
] '/3 (1.9)
ds -
_ [Ln;df
--
JY2 (1.10)
N
The first moment average is simply the arithmetic mean as defined in equation
(1.5).
(1.11)
7a Particle characterisation
(1.12)
where m; is the mass of particles in the ith interval and M is the total mass.
The ratio (m;lM) is a form of weighting factor in the averaging process. If
particle shape is not a function of particle size for the distribution, i.e. we can
write
(1.13)
where k is a constant for all values of d;, then
"Lm·d· "Ln;d'f
d =_'_' __ _ (1.14)
mm M
(The mass mean diameter should not be confused with the diameter of average
mass.) In a similar way, the surface mean diameter (also known as the
'volume-surface mean' or the 'Sauter mean') is given by
"Ls;d; "Ln;dr
(1.15)
d sm = S = "Ln;dr
where s; is the surface area of particles in the ith interval and S is the total
surface area. The surface mean diameter is the appropriate diameter to use
when calculating the pressure drop through a packed bed of particles at low
Reynolds numbers (see Section 2.3.2).
It is also possible to obtain moment averages of weighted distributions,
analogous to the moment averages of the count distribution defined in
equations (1.9) and (1.10).
Solution
(Refer to Table 1.2)
Table 1.2 Solution to Example 1.1
_ 'Lnidi 24 085
(a) d p = N = 1000 = 24 f,lm
Solution
The fraction of the total number of particles in the ith interval is fin- The
weight of these particles (assuming spheres) is thus
ndr
6-
Nfin X - Qp
7,Nfin Cd')
-i i
-3
Qp 'L findi
1
Fim= ---- (1.16)
7Nfin
I (
-i
nd 3
)
Qp
I
-3
'L findi
1
where I is the total number of intervals. In fact it is not essential for the
0.050
Count Mode
0.045
0.040 Count Median Diameter
0.035
e Count Mean Diameter
%0.030
.9
.... 0.025
~
E 0.020 Diameter of Average Mass
~
0.015
Surface Mean Diameter
0.010 Mass Mean Diameter
0.005
0.000
0 20 40 60
I 80 100 120 140 160 180 200
Particle diameter (J.lm)
Figure 1.6 Averages of the distribution.
14 Particle characterisation
i
~
(f~:) I
Fin = (1.17)
(f~~)
I
~
Taking the data of Table 1.1 we now retabulate it as shown in Table 1.3.
The resulting cumulative mass distribution is plotted in Figure 1.7.
In converting from a number distribution to a mass distribution, or vice
versa, it is important to consider the accuracy of the extremes of the
distributions. A 10 i-!m particle has 1000 times the mass of a 1 i-!m particle.
Thus a small error in the upper size classes of a number distribution will have
a significant effect on the resulting mass distribution. Since the number of
particles in any upper size class is usually small it is essential to repeat the
count several times to avoid the usual problems of small-number statistics.
Conversely, when converting from a mass distribution to a number
distribution, small errors in the lower size classes of the mass distribution are
magnified by the conversion, producing large errors in the number distribu-
tion.
.9== 0.8
...
~
eu
==~ 0.6
eu
e
aI
~ 0.4
eu
-e=
= 0.2
U
0.0 I• .........-r
o 20 40 60 80 100 120 140 160 180 200
Particle diameter (f..lm)
Figure 1.7 Cumulative mass distribution.
76 Particle characterisation
Solution
Assuming that the particles are spheres,
S = Jt"Lnid7"
and
QpJt
M= - - "LnidT
6
Substituting in equation (1.15)
If the particles are not spheres, the constant (6/Qp) will change. This example
illustrates the ease with which it is sometimes possible to obtain weighted
mean diameters from macroscopic properties.
100
-.
e::::s..
'-' CMD
c..
"0
101 ~ I
50%
1
0.01 0.1 1 10 30 50 70 90 99 99.9
Percent less than indicated size
Figure 1.8 Log-normal plot-number distribution.
18 Particle characterisation
Og =--=--=
d so % d 16 %
[ Jd s4 %
d 16 %
Y,
(1.20)
So Og may also be obtained very simply from the graph. Note that unlike the
standard deviation of a normal distribution, Og has no units and must always
be greater than 1. A further important property of the log-normal
representation is that weighted distributions are parallel on a log-probability
plot. This means that the geometric standard deviation is the same for the
number and weighted distributions. This is shown for the data of Table 1.1 in
Figure 1.9. Transformation between different weighted distributions and
between the various averages discussed earlier in this chapter is made possible
by means of the Hatch-Choate equations (see Appendix AI), the transforma-
tion depending only on the value of GSD. Note however that mixtures (i.e.
sums) of log-normal distributions are not log-normal.
It is often the case that particle size distributions are only approximately
log-normal; even in these circumstances it may still be helpful to plot them on
log-probability paper, but it must be remembered that the special properties
described above apply only to true log-normal distributions.
dy
_ (6V)Y,
- - (1.21)
Jt
The value of d y is not dependent on the orientation of the particle.
1000 - - . - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
I
100
MMD •
,.-.,
~
'-'
Q.
I CMD
"0
lO
V _/ I
50%
1
0.01 0.1 1 10 30 50 70 90 99 99.9
Percent less than indicated size
Figure 1.9 Log-normal plot-number and mass distributions.
20 Particle characterisation
ds _(S)'h
- -
Jt
(1.22)
dA
_ (4A)'/2
- - (1.23)
Jt
The value of dA is dependent on orientation. However, it can be
shown (Vouk, 1948) that, averaged over convex particles in random
orientation,
S )Y2
dA = ds = ( --; (1.24)
than d y , but the difference is relatively small for compact particles, such
as cubes or spheroids.
v = uvAd~ =-i-
nd3
(1.34)
Here the extra subscript A emphasises that the shape coefficients are related
to the projected-area diameter; it is equally valid to derive expressions for the
shape coefficients which relate them to some other equivalent diameter.
Heywood now considered a particle resting on its plane of maximum
stability (Figure 1.10) and defined its dimensions in three orthogonal
directions: thickness, T (perpendicular to the plane of maximum stability),
length, L, and breadth, B. The shape is now described by
the elongation ratio, n = LIB
and the flakiness ratio, m = BIT.
The surface and volume shape factors can be combined with the shape ratios
as follows. Consider a rectangular parallelepiped of dimensions L X B X T
(Figure 1.10). The proj ected area of the particle,
ndi
A = - = rABL (1.35)
4
where rA is the 'area ratio', ndi/4BL. The volume of the particle is:
v = prABLT (1.36)
which is equal to uvAdl from the definition of UvA in equation (1.34). This
serves to define the 'prismoidal ratio' , p - another measure of particle shape -
as, after some algebraic manipulation, 4uvA d AlnT.
Combining equations (1.35) and (1.36) gives
nVn p
(1.37)
UvA =---
8 mvr;:n
If the particle is 'isometric', i.e. B =L= T and n =m = 1, then
nVn p
8 vr: (1.37a)
The term UvAI can therefore be used as an indication of particle form. When
the particle is not isometric, UvA is given by uvA1lmYn. In other words, the
form is modified by the relative proportions, represented by m and n.
Heywood tabulated values of rA and p for four general forms, as shown in
Particle size measurement 23
I
I
T I
Minimum
Dimension
~I"'--B --~~I
T T
.. . ..
.
~ . .. ..... . ...
. . . . ..
. .. . ..
........ . .
......•.•........
..
....•... ::.::::.:.: . :> ... ::" . ":-:: ... :.
. ..
~~ .... ~
~~L
Figure 1.10 Heywood's shape factors.
24 Particle characterisation
Shape 'A P
Angular
tetrahedral 0.5--0.8 0.4--0.53
prismoidal 0.5--0.9 0.53--0.9
Subangular 0.65--0.85 0.55--0.8
Rounded 0.72--0.82 0.62--0.75
R ,- ,-,
'0·- . '- ,_.0. .....
-e..
.... .•..•. •..•.. .-
e
Figure 1.11 A particle outline in polar coordinates (Hawkins, 1990).
Particle size measurement 25
(a) (b)
6.0
5.0 8= 1.17
4.0
P
3.0
2.0
~..
0.05 0.10 0.20 0.50
A
6.0
5.0 8=1.37
~ ~
4.0
P
3.0
2.0
0.05 0.10 0.20 0.50
A
6.0
5.0 0 8=1.42
~
4.0
P
3.0
2.0
0.05 0.10 0.20 0.50
A
establish them, the links between fractal dimensions and other measures of
particle properties, such as friction factors and flow behaviour, remain
somewhat tenuous.
The effect of particle shape on motion can be accounted for by means of the
dynamic shape jactor, X, which is defined as the ratio of the actual resistance
force acting on a particle, F o , to the resistance force acting on a sphere of the
same volume and velocity (in the Stokes law regime). Hence
x= F o /3JtflVd v (1.39)
Putting
Jtd~
Fo = - (Op - O)g,
6
we obtain the terminal settling velocity,
(Op - O)~g
Vt =------ (1.40)
18 flX
sphere 1.00
cube 1.08
cylinder (LID = 4)
axis horizontal 1.32
axis vertical 1.07
bituminous coal 1.05-1.11
quartz 1.36
sand 1.57
talc 2.04
1.2.3 Sampling
Only very rarely is it possible to perform a size analysis on the entirety of the
powder of interest. The aim of sampling is to obtain a fraction which is
representative of the whole.
Particle size measurement 27
Advantages Disadvantages
Extractive sampling
Equipment can be relatively simple and Analysis results are not immediately
cheap. available.
A portion of the sample removed can be Extractive equipment may disturb gas
retained. flow, producing an unrepresentative
sample.
Several forms of analysis can be Particulates can agglomerate in
performed and repeated if necessary. sampling and conversely agglomerates
can be dispersed during measurement,
producing an unrepresentative sample
for size analysis.
In situ sampling
Generally, sample extraction equipment Calibration can be difficult.
is not required.
Particles can be monitored as they exist in Sampled material not retained for further
the duct. analysis.
Rapid data return. Sensors may be exposed to 'hostile'
conditions and may need to be cleaned
periodically .
..
::;;;;;ZZ/7ZZZZZZtzZZ7ZtZZZ7Z7ZZlr
..
...
~
...
~/Z21//ZZZZZ;:7ZZZZZ/?ZZZZZ2/
----~......------------ ...
Uo=U
and Lundgren (1980) and Hinds (1982), among others. Figure 1.15 shows the
fractional error in measured particle concentration (CICo) as a function of the
fractional departure from isokinetic conditions (Uo/U), where the subscript
zero refers to the free-stream conditions and the unsubscripted variable to
conditions at the probe tip. The error in measured concentration is seen to
Particle size measurement 29
Vo ....
Gas streamlines
entering probe {
(a)
Vo
---$q?zzvvzzzzzzV/z/zz'jlzzvm
...
Gas streamlines { ..
~
----; ..
: -: I1 :
entering probe
~ ---.. ..
~ ~27ZZ7Z7VVVZ7ZZV)jVZZVZ
..
(b) ....
--------~-------~
--
--~~--~.----------- ..
----'u"-'Q'--------:-~?ZZZZZZZZ/zzzzzz;;zzzzzzzzz/
..--- .
{.--11..---------
---------
Gas streamlines
entering probe
. ..
---------1:-- .-_-----
U ..
..
..
(c)
--
_________:_-~ZZ7Z7ZZ7ZZZZZZ/~ZZZZZ7ZZZ7Z
1.0
0.5
0.2
0.1
~o
r:5
.~ 1.0 Stk = O.D1
___=:::;;;~~~:::..____-- 0.01
c:
o
.z; 0.1
e 0.2
~
"oc: 0.5
U
1.0
0.1 L-_.L---L_--L_...L..-L-L..L.l-L.L-_.L..--L_---'_-'---L...J
0.1 1.0
Velocity ratio, l'olU
Figure 1.15 Effect of velocity ratio on concentration ratio for several values of the
Stokes number (8 = 0°) (after Hinds, 1982).
Sieving 'sieve diameter' 20 !tm upwards mass slow; cheap, errors due to
blinding
Light interaction
extinction
single particle dA 2-1000 !tm number mainly used in liquids; limited
concentration range
multiparticle dA 2-1000 !tm average only total surface measurement
scattering 'equivalent scattering
single particle diameter' 0.2-20 !tm number fast; limited concentration
range; sensitive to refractive
index and shape
multi particle dA (for larger particles) 4--1000 !tm volume fast; has been used for on-line
measurement and control
j Mass
.1\ distribution
Equivalent
volume diameter
Equivalent
mesh size
Figure 1.16 The effect of shape on various types of distribution (adapted from
Scarlett, 1984).
1.2.4.3 Light scattering. There are three basic types of instrument III
which light scattering is employed (Figure 1.18).
(a)
(b)
/'
/'
1
/'
/'
light
Source '-': trap
" "- "" Collecting
Particle
at focus "" aperture
(a) 1
Aerosol
Pinhole Pinhole
apertu re aperture
;~t*lT: ~~
(c) ~L~
Figure 1.18 Instruments for particle sizing by light scattering. (a) forward-scatter
from a single particle; (b) diffraction sizing (reproduced by permission of Malvern
Instruments Ltd.); (c) extinction measurement (Hinds, 1982).
particles (nd/A < 0.3, where d is particle diameter and A is the wavelength
of incident light, i.e. d < 0.05 f!m in visible light) Rayleigh scattering theory
can be used, in which the relative angular distribution of scattered light, /(8),
is the same for all particle sizes, shapes and refractive indices. /(8) is then
proportional to d6 /'A 4. For large particles (nd/A > 3/(m - 1), d > = 2 f!m in
visible light) the forward scattered light intensity (8 < 90°) can be
approximated by Fraunhofer diffraction theory (see later).
Particle size measurement 37
scattering
plane
(1.43)
where i 1 and i2 are the Mie intensity parameters for scattered light polarised in
the perpendicular and parallel directions, i1 and i2 are complicated functions
of refractive index, m, size parameter red/A, and scattering angle 8. Two
general observations are of importance:
(a) As the scattering angle, 8, increases towards 90°, the scattered light
intensity becomes more sensitive to changes in refractive index, m,
(particularly whether or not the particle is absorbing) and to departures
from sphericity (Hodkinson, 1966).
(b) As the size parameter, a = red/A, increases, the angular scattering
pattern becomes more complicated. A plot of Mie intensity parameters
(i1 + i 2 ) against a will show variations in intensity of several orders of
magnitude over small changes in a for a > 10 (Figure 1.20) (Hinds,
1982).
For these two reasons most commercial single particle counters integrate the
scattered light over the forward direction (8 < 90°), ensuring the least
dependence of response on particle refractive index and shape. Integration
over a number of incident wavelengths (i.e. by using an incandescent white
light source) also serves to damp out oscillations in the Mie response curves.
Cooke and Kerker (1975) obtained the theoretical responses of a number of
counters by integrating the Mie intensities over the collecting aperture and
showed that, for those integrating the response in the forward direction, there
is comparatively little dependence on non-absorbing (real) particle refractive
index for d > 1 !lm. For absorbing (complex) refractive indices, however, the
38 Particle characterisation
1000
:"J<
+
'-e 100
OJ
E
~c:.
::-
'in
c
OJ
.~
OJ 10
~
.s'"
~r,;
OJ
>
.;::;
'"
0;
a: 1.0
0.1 L...I.._'--_'--..L-L.....l.._-'-_-.L...-'-....L-J
1.0 10 100
Size parameter, C(
0.17 1.7 17
Particle diameter for).. = 0.52 I'm (I'm)
Figure 1.20 Relative scattering (Mie intensity parameter) versus size parameter, for
water droplets (m = 1.33) at scattering angles of 30 and 90° (Hinds, 1982).
responses are very different, being in general much lower than for a real
refractive index of the same (real) magnitude. Instruments with less
convergent illumination (larger y; see Figure l.1Sa) and smaller collection
angles (smaller j3 - b; see Figure l.1Sa) show a multivalued response curve,
the principal uncertainty being in the range between indicated particle
diameters of 0.5 and 1 fLm.
For aerosols of industrial interest, which are in general both irregular and
absorbing (and also frequently of inhomogeneous refractive index), theoretical
light scattering predictions are too complicated to be used to predict the
Particle size measurement 39
(apart from the 'spreading' effect) providing that their upper concentration
limits are not exceeded. This means that, for accurate filtration
measurements, which require comparison of inlet and outlet concentra-
tion and size distribution, they are preferable to diffraction instruments
which only measure volume distribution. However, single particle
counters require special dilution systems for measuring particles at high
concentrations. Single particle counters are effectively subject to no
lower concentration limits, whereas diffraction instruments are subject to
signal-to-noise ratio problems at low concentrations.
(c) Diffraction instruments do not physically sample the suspension and
therefore cause no disturbance in the measured flow and their operation
is relatively insensitive to duct temperature and pressure. For these
reasons, they are preferable to the single particle counters as on-line
instruments. The requirement that the aerosol to be measured must pass
within the focal length of the collection lens is, however, a restriction.
In contrast to the devices which measure the intensity of scattered light, as
described above, extinction instruments measure the intensity of the light
which is not scattered out of the beam, i.e. the attenuation of the incident
light. According to the Lambert-Beer law (Hinds, 1982) the transmitted light
intensity is given by
(1.44)
Ad2
De = L N;--' (Qe); (1.45)
, 4
where N; is the number of particles per unit volume in size class i (m-3 ), d; is
the mean diameter of particles in class i (m) (which is to be interpreted as the
projected area diameter, dA ) and Qe is the 'particle extinction efficiency' for
particles in size class i, which equals the radiant power scattered and absorbed
by a particle divided by the radiant power geometrically incident on the
particle.
Qe is thus a measure of the ability of a particle to remove light from the
beam by scattering and absorption. In general, it is a complicated function of
both refractive index, m, and size parameter, a (Figure 1.21), but for a
greater than about 12 (particles in air of diameter greater than about 2 !-lm)
Qe approaches a constant value of 2. If this is the case
Qe
De = -4 L; N(JuE)
I I
= S/2 (1.46)
42 Particle characterisation
6,----,-----,----,----,-----,----,-----,----,----,----,
5
~
>-
"~ 4
'u
:;:
;:;
c
.'2 3
U
c
.;::;
x
w
2
Size parameter, 0:
o 2 3
Particle diameter for A = 0.52 11m (11m)
Figure 1.21 Extinction efficiency versus particle size for spheres (Hodkinson, 1966).
VOLUMETRIC I "' ..
METERING ",:::,:",." ,.""",«
DEVICE i "'.
OUTPUT
STIRRED SUSPENSION
OF PARTICLES IN
ELECTROLYTE SOLUTION
where f is the volume fraction occupied by the spheres and Q is the resistivity
of the carrier fluid. For a cylindrical sensing zone of length L and diameter D
filled with the fluid the resistance is
4QL
R1=-- (1.49)
nD2
When a sphere of diameter d is introduced, the value of f becomes
Vsphere 2d3
f= =- (1.50)
Vcylinder 3D2L
Hence the new resistance R2 is now
R 2 = 4QL
nD2
[1 + ~+
D2L
... J (1.51 )
Qp~UCst
St= - - - (1.53)
Impaction nozzle
or jet
Streamlines
Impaction plate
*The appropriate density is relatively easily measured for non-porous particles using a specific
gravity bottle or gas pyknometer. Measurement of the appropriate density for porous particles,
the so-called 'envelope' density, is discussed by Geldart (1990). The important area of powder
porosity and surface area determination, which is outside the scope of this book, is
comprehensively covered by Lowell and Shields (1991).
46 Particle characterisation
100,---------,----------r------~_,--------_,
80
20
o 2 3 4
Aerodynamic diameter (f.lm)
o 0.47 0.94
.J Stk
Figure 1.23 (b) Grade efficiency for inertial impactor.
where U is the average nozzle exit velocity and D j is the diameter of the
nozzle. Single impactors are easily designed and constructed; design
information is given by Marple and Willeke (1976). A good impactor
efficiency curve is shown in Figure 1.23b; this is a 'grade efficiency' curve,
showing the collection efficiency of particles as a function of their size. (A
theoretical approach to the prediction of impactor efficiency is given in
example 2.4 in section 2.2.3.) In practice, impactors rarely achieve as sharp a
'cut-off' as shown here because particles may bounce or break up on impact
or may deposit at various intermediate points between successive stages of a
cascade. The stages can be coated with a thin layer of sticky liquid to prevent
bouncing. A series of suitably sized miniature cyclones (see section 7.2) may
also be used to separate particles by their inertial behaviour.
Particle size measurement 47
Inner Nozzle
1 L/mln
Sheath Air Valve
4 'I.....
Pressure
Transducer
Figure 1.24 Aerodynamic particle sizer (reproduced by permission of Biral Ltd. and
TSI Inc.).
1.3 NOMENCLATURE
A Area (m 2 )
Cst Cunningham slip-correction factor
CMD Count median diameter (m)
c Hydrodynamic resistance (m) (equation (1.27))
D Dimension, diameter (m)
d Particle diameter (m) (equation (1.23))
dA Equivalent projected area diameter (m)
da Aerodynamic diameter (m)
dg Geometric mean diameter (m) (equation (1.7))
di Representative diameter of particles in interval i (m)
dIll Diameter of the particle of average mass (m) (equation (1.9))
dmm Mass mean diameter (m) (equation (1.12))
dp Arithmetic mean diameter (m) (equation (1.5))
ds Equivalent surface diameter (m) (equation (1.22))
dsm Sauter or surface-volume diameter (m) (equation (1.15))
d st Stokes diameter (m) (equation (1.25))
dv Equivalent volume diameter (m) (equation (1.21))
F Cumulative distribution function (equation (1.3))
FD Drag force (N)
f Frequency distribution function (equations (1.1a and b))
g Acceleration due to gravity (m S-2)
h Height of element of a frequency histogram (equation (1.1a))
I Number of intervals, radiation intensity (W m-2 )
Mie intensity parameter (equation (1.43))
M Total mass (kg)
MMD Mass median diameter (m) (equation (A1.7))
m Heywood flakiness ratio; refractive index (section 1.2.2)
Mass in interval i (kg)
Total number of particles
n Heywood elongation ratio (section 1.2.2)
Number in interval i
Perimeter (m)
Heywood prismoidal ratio (equation (1.36))
Particle extinction coefficient (equation (1.45))
Distance; electrical resistance (ohm)
Heywood area ratio (equation (1.35))
Total surface (m 2 )
Stokes number (equation (1.42))
Amount of surface in interval i
Gas velocity (m S-I)
Particle velocity (m S-I)
Particle terminal velocity (m S-I) (equation (1.25))
Particle volume (m 3 )
Appendix A 1 49
Greek letters
= nd/A, size parameter in light scattering
Heywood surface and volume coefficients (equations (1.33)
and (1.34»
Dynamic shape factor
Fractal dimension (equation (1.38»
Step length (m) (equation (1.38»
Angle (0)
Viscosity (N s m-2)
Fluid density (kg m-3 )
Particle density (kg m-3 )
Standard deviation
Extinction coefficient (m-l)
Geometric standard deviation
Wadell's sphericity
1
df= - - e x p
ov2n
[(d
-
p -
202
dp ?] dd p (ALI)
0.5 ""
/
-. "-
'0 0.4
-=
~
Q
..........= 0.3 \
(oJ \
c:
0.2
'=
-- 0.1
""- ..........
--,
a
1 10 100 1000
Particle diameter (I-lm)
Ln;ln d;
lndp = -N- =lndg (A1.2)
REFERENCES
Allen, T. (1996) Particle Size Measurement, 5th edn., Chapman and Hall, London.
Bayvel, L.P. and Jones, A.R. (1981) Electromagnetic Scattering and its Applications,
Applied Science Publishers, London.
Boeck, Th. and Leschonski, K. (1984) Proceedings of the 3rd European Symposium on
Particle Characterisation (PARTEC), Nilrnberg, 581.
Clift, R., Grace, J.R. and Weber, M.E. (1978) Bubbles, Drops and Particles,
Academic Press, New York.
Cooke, D.D. and Kerker, M. (1975) Appl. Optics 14, 734.
Davies, C.N. (1979) 1. Aerosol. Sci. 10, 477.
DeBlois, R.W. and Bean, C.P. (1970) Rev. Sci. Instr. 41 (7),909.
Durham, M.D. and Lundgren, D.A. (1980) 1. Aerosol. Sci. 11, 179.
Geldart, D. (1990) Powder Technol. 60, 1.
Ghadiri, M., Seville, J.P.K., Raper, J.A. and Clift, R. (1986) 1st World Congress of
Particle Technology, Nilrnberg 1, 203.
Hawkins, A.E. (1990) in Principles of Powder Technology (M.J. Rhodes, ed.), Wiley,
Chichester.
Heywood, H. (1938) Proc. Inst. Mech. Engrs 140, 257.
Hinds, W.C. (1982) Aerosol Technology, Wiley-Interscience, New York.
Hodkinson, J.R. (1966) in Aerosol Science (C.N. Davies, ed.), Academic Press,
London.
Jones, A.R. (1977) 1. Phys. D. 10, 163.
Kaye, B. (1989) A Random Walk Through Fractal Dimensions, VCH, Weinheim.
Kerker, M. (1969) The Scattering of Light and Other Electromagnetic Radiation,
Academic Press, New York.
Lowell, S. and Shields, J.E. (1991) Powder Surface Area and Porosity, 3rd edn.,
Chapman and Hall, London.
Marple, V.A. and Willeke, K. (1976) in Fine Particles (B.Y.H. Liu, ed.), Academic
Press, New York.
Pisani, J.F. and Thompson, G.H. (1971) 1. Phys. E. 4,359.
Pollack, J.B. and Cuzzi, J.N. (1979) in Light Scattering by Irregularly Shaped Particles
(D.W. Schuermann, ed.), Plenum Press, New York.
Scarlett, B. (1984) Proceedings of the 3rd European Symposium on Particle
Characterisation (PARTEC), Nilrnberg, 1.
Tate, A.H.J., Seville, J.P.K., Singh, A. and Clift, R. (1986) I. Chern. E. Symp. Ser.
99,89.
Van Santen, A. and Gannon, E.J. (1987) Filt. Sepn., Sept./Oct. 10, 328.
Vouk, V. (1948) Nature 162, 330.
Wadell, H. (1932) 1. Geol. 40, 443.
Williams, J.C. (1990) in Principles of Powder Technology (M.J. Rhodes, ed.), Wiley,
Chichester.
Wilson, J.C. and Liu, B.Y.H. (1980) 1. Aerosol. Sci. 11, 139.
Zerull, R.H., Giese, R.H., Schwill, S. and Weiss, K. (1979) in Light Scattering by
Irregularly Shaped Particles (D.W. Schuerman, ed.), Plenum Press, New York.
2
•
Particles In fluids
BUOYANCY
LIFT
DRAG
~
~ ___ Shear
~\ Stress
I I
- --
Normal
Stress
Figure 2.1 Forces on a stationary particle in a flowing fluid.
known respectively as form drag and skin friction drag and are the two
components of the hydrodynamic drag, F D , on the particle. Any component
of the total hydrodynamic force normal to U, F L , is known as lift.
The distinction between buoyancy and hydrodynamic force is useful for
several reasons beyond the basic point that only the buoyancy force exists
when the fluid is stationary. The few theoretical solutions available for fluid
motion and drag are obtained by eliminating the hydrostatic pressure so that
the equations describing fluid motion are written in terms of the reduced
pressure (see Batchelor, 1967). The drag force is frictional in the sense that it
corresponds to dissipation of mechanical energy within the fluid (although the
fluid can do no work directly on a rigid stationary particle). Buoyancy, on the
other hand, is not associated with mechanical energy dissipation.
There is no universal theoretical solution available for calculation of F D . It
is therefore necessary to proceed by applying dimensional analysis to
empirical results. In general, if the fluid is incompressible and Newtonian, FD
depends on U and Q, on the fluid viscosity I-l, and on some characteristic
dimension of the particle d, i.e.
FD = f[U, d, Q, I-l) (2.1)
for a particle of fixed geometric shape and orientation. Equation (2.1)
contains five parameters with three basic dimensions (mass, length and time)
so that it can be expressed in general in terms of two dimensionless groups.
These groups can be selected at will, provided they are independent.
Single particles 55
O.ll=-----j----+---+---+---+---+---t--;.oj
0.1 10 100
Re
Figure 2.2 Drag coefficient as a function of Reynolds number: ~~-, sphere; -'-'-,
cylinder in crossflow; - - - - -, disc with flow parallel to axis.
Conventionally the two groups chosen are the particle drag coefficient,
(2.2)
and the particle Reynolds number
Re = QdU/!J, (2.3)
In equation (2.2), A is the cross-sectional area of the particle, and therefore
proportional to tP. For a spherical particle, A is JttP/4 so that
Co = 8Fo/JtQU2tP (2.2a)
The particle Reynolds number is useful because it indicates the nature of flow
around the particle: a low value of Re indicates that the fluid is in creeping
flow, i.e. its motion is dominated by viscous effects in the fluid. At high Re
the inertia* of the fluid is dominant. Numerical ranges corresponding to 'low'
and 'high' are discussed below.
In dimensionless form, equation (2.1) now becomes
CD = f[Re] (2.4)
This type of relationship is summarised here as Co[Re]. Empirical drag
measurements can therefore be expressed in this form. Figure 2.2 shows the
*Some authors use the term 'turbulent' to describe high Re flows around particles. The term
appears to derive from a mistaken analogy with high-Re internal flows, such as in pipes. Its
application to external flows, such as those around particles, is misleading and is therefore to be
avoided.
56 Particles in fluids
2.1.2.1 Spheres. The shape of most interest for particle processing is the
sphere. The flow patterns around a spherical particle will therefore be
discussed with reference to the drag curve for the sphere. Figure 2.2 shows
that, at low Re, CD is inversely proportional to Re. This is the only range for
which a full theoretical solution is available. The resulting relationship is
Stokes' law,
CD = 24/Re (2.5)
or
FD = 3rq.,tdU (2.6)
so that, in this range, the drag force is directly proportional to U. Of the total
drag, two-thirds results from the pressure distribution over the sphere and
one-third from shear stresses, i.e. even in this creeping flow range, form drag
on a sphere is twice the skin friction drag.
Equations (2.5) and (2.6) are reasonable approximations for Re < 0.1: the
error in CD is about 2% at Re = 0.1 and nearly 20% at Re = 1. Outside this
range, it is necessary to estimate CD by some empirical result, for example
that due to Turton and Levenspiel (1986) given in Table 2.1. The changes in
drag coefficient shown by Figure 2.2 correspond to changes in the fluid flow
around the sphere. At Re = 20 flow separates from the surface, to form a
closed recirculatory wake which becomes larger as Re increases.
At higher Re the downstream tip of the wake oscillates until, for Re > 100,
fluid is shed from the wake. Throughout this range, the point at which flow
separates from the surface moves forward with increasing Re. For Re > 6000,
the separation point reaches a steady position somewhat forward of the
equator. In this range, skin friction drag is negligible and form drag depends
most strongly on the position of flow separation. Therefore the drag
coefficient varies very little with Re: over the range 750 ~ Re ~ 3.4 X 105 , CD
only varies by ± 13% about a value of 0.445 (or 4/9): this is usually known as
the Newton's law range, because Newton's concepts of fluid dynamic drag
implied a constant drag coefficient.
As Re increases above 6000, changes occur in the flow field which do not
immediately reflect in the drag coefficient. The fluid layers which have
detached from the surface of the sphere become turbulent, and the point at
Single particles 57
A. Spheres
(a) Drag
Range of Re Correlation
24 0.413
~ 3.38 X lOs Co = -(1 + 0.173 ReO. 6S7 ) + ---------,---:-:-
Re 1 + 1.63 X 104 Re-1.09
3.38 X lOs to 4 X 105 CD = 29.78 - 5.3 IOg1O Re
4 X lOs to 106 CD = 0.1 IOglO Re - 0.49
~ 106 CD = 0.19 - 8 X 104 Re- l
Range of Correlation
d*
2.1.2.2 Other shapes. Steady crossflow past a long cylinder shows much
the same features as flow past a sphere, but the transition Reynolds numbers
are different. Flow separation occurs for Re > 5. The closed recirculatory
wake oscillates for Re > 30, and wake shedding occurs for Re > 40. Over the
range 70 < Re < 2500, the wake sheds alternately from the two sides of the
cylinder to form a regular series of vortices alternating in their sense of
rotation; this is often termed a 'Von Karman vortex street'. Above Re = 105 ,
the critical region is entered, with flow transitions similar to those occurring
for a sphere. Table 2.1 gives some expressions for CD(Re) for long cylinders
in crossflow, from Clift et ai. (1978).
The comparison between discs and spheres illustrates some of the effects of
departure from a spherical shape. Because a disc in a fluid flowing parallel to
its axis has no surface parallel to the direction of flow, it can experience only
form drag. At very low Re, drag on a disc is lower than on a sphere because of
the absence of skin friction. Wake formation occurs much sooner for a disc,
around Re = 1, because the sharp edge encourages the flow to separate. Disc
wakes start to oscillate at about Re = 100 but, because the position of flow
separation is fixed, the drag coefficient is constant, at CD = 1.17, for
Re > 133. At high Re, flow patterns continue to be qualitatively similar to
those around a sphere except that, because the separation position is fixed by
the particle shape, there is nothing equivalent to the critical transition. Table
2.1 gives results due to Potter et ai. (1973) and Clift et al. (1978) for estimating
CD for discs fixed normal to the direction of flow.
For other particle shapes, it is usually helpful to compare with particles of
well-characterised shape, such as spheres, cylinders, discs, or spheroids (see
section 2.1.3.2). In the creeping flow range, the drag on a particle is always
less than or equal to that on a body which encloses it, and greater than or
equal to that on any body contained within it (Hill and Power, 1956). Drag in
this range is most influenced by surfaces on which shear and normal stresses can
act: for example, creeping flow parallel to the axis of a long fibre, needle or
extremely prolate spheroid causes half the drag arising in crossflow. At
intermediate and high Re, drag is most likely to be influenced by edges which
define the position of flow separation. Thus drag at high Re is reduced by
'streamlining', i.e. profiling the particle shape to prevent boundary layer
separation. Prolate (i.e. elongated) particles demonstrate the difference
Single particles 59
between these flow regimes: at low Re, flow parallel to the axis of a prolate
particle causes higher drag than on a sphere of the same volume, whereas at
high Re the drag on the prolate body is lower and can be very much lower if
flow separation is prevented. The effects of shape on free fall or rise of
particles are considered below.
where rtd3 /6 is the volume of the sphere, v. The drag coefficient then follows
from equations (2.2a) and (2.7) as
(2.8)
where Ret = QdVt/f.! is the particle Reynolds number at the terminal velocity.
For the creeping flow range, Ret < 0.1, in which Stokes' law applies, the
terminal velocity follows from equations (2.6) and (2.7) as
(2.9)
Buoyancy, pgv,
+ drag, FD
Weight, ppgv
In the Newton's law range (750 :;::; Re :;::; 3.4 X 105 ) where Co is close to 0.445,
the terminal velocity follows as
(2.10)
The interest in particle technology is usually in the intermediate range of Ret,
between creeping flow and Newton's law, say 0.1 < Ret < 750. There is no
explicit result for V t in this range. It is possible to calculate V t from equation
(2.8) using a correlation for Co[Re] such as that in Table 2.1; however, this is
inconvenient, because the equation is implicit and therefore an iterative
solution is needed. It is therefore preferable to express the dimensionless drag
relationship in terms of more convenient groups, remembering that the use of
Co and Re results solely from historical convention and that any two
independent dimensionless groups can be used.
From equation (2.8), a particle at its terminal velocity satisfies the
relationship
(2.11)
The group CoRe;, and various multiples of it, is known as the 'Archimedes
number', or the 'Best number'; it is a convenient dimensionless group, and an
independent variable for estimation of terminal velocities because it involves
the relevant particle and fluid properties but not the terminal velocity. Here,
rather than using the group in equation (2.11), we shall use the dimensionless
particle diameter which follows from it,
Retfd* (2.14)
The dimensionless drag correlation can then be expressed in the form Vi[ d*];
the resulting relationship is shown in Figure 2.4. Table 2.1 gives useful
expressions for Vi[ d*], developed by Grace (1986) from correlations for
Ret[Co Re;] given by Clift et at. (1978).
Single particles 61
10-2 L--...L...LJLLliJIlol.--L...Ll...L1JJlL---1---L.LLI.lill_L....l....L.L.llllL-....L-L..LJ...J...LW
0.1 10 10 2 103 104
Figure 2.4 also shows the simplifications for the creeping flow and
Newton's law ranges. In dimensionless form, with the ranges over which they
estimate V t within about 10%, they are:
Stokes'law
(d* < 2.5, Vi < 0.32, Ret < 0.8):
Vi = (d*?/18 (2.9a)
Newton's law
(60 < d* < 3500, 13 < Vi < 100, 750 < Ret < 3.4 X 105 ):
Vi = 1.73 v'd* (2.l0a)
Equation (2.9) shows that, in the creeping flow range, the terminal velocity
depends strongly on both particle size and fluid viscosity. For both liquids and
gases, viscosity is very sensitive to temperature (but is essentially independent
62 Particles in fluids
of pressure). On the other hand, the terminal velocity in the Newton's law
range is not strongly dependent on particle diameter, and depends on fluid
density rather than viscosity. Therefore measurements in which particles are
classified by terminal velocity are carried out at low Reynolds number, with
close temperature control. For the 'critical' range, roughly 2000 < d* < 3500,
Figure 2.4 shows that three terminal velocities are possible. The intermediate
value, portion AB of the curve corresponding to the range of Re in which CD
decreases sharply (see Figure 2.2), is unstable. Particles in this range have
been observed to fluctuate between the two stable terminal velocities, giving
a rather erratic motion. The curve beyond B represents supercritical motion.
The dimensionless diameter d* proves to be a convenient independent
variable in describing a range of fluid-particle operations, beyond just
estimation of terminal velocities. It is therefore used throughout this book.
Similarly, velocities made dimensionless in the form of equation (2.14) will be
used in a number of other places.
Solution
This type of calculation is particularly facilitated by using a correlation of the
form Vi[d*] because the property groups in equations (2.12) and (2.14) are
constants.
d*
d
[ Q(Q'"~ Q)g T
= [998.2 X (2650 - 998.2) X 9.81 J:;'
(1.002 X 10-3 )2
= 2.526 X 104
V *t
[ 2 ]11'
fA,(Qp ~ Q)g .
998.22 . ] Y3
[
1.002 X 10-3 X (2650 - 998.2) X 9.81
= 39.44
Single particles 63
d(m) d* Re
Stokes' Newton's
law law
10-5 0.2526 3.544 X 10-3 8.99 X 10-5 8.95 X 10-4 8.99 X 10-5
10-4 2.526 0.3251 8.24 X 10-3 0.821 88.99 X 10-3
10-3 25.26 6.328 0.160 160 0.899 0.220
10-2 252.6 28.99 0.735 7320 0.697
10-1 2526 82.69 2.10 2.09 X 105 2.20
Comments
This particle density is representative of many naturally and industrially
occurring materials. The inaccuracy in Stokes' law is already apparent for a
particle 10-4 m (i.e. 100 IJ.m) in diameter; Stokes' law is typically inapplicable
to particles bigger than about 100 IJ.m in liquids. Newton's law gives rough
estimates, but only rough estimates, for particles larger than a few millimetres
in diameter. Particles as large as 1 cm in diameter are in the critical range
where two terminal velocities are possible: the estimate above refers to the
lower, sub critical value.
Solution
The general procedure follows that used in Example 2.1, but the property
groups must be calculated for each set of conditions.
which leads to the following estimates for terminal velocity (in m S-l):
The values for 100 f,lm and 1 mm particles are plotted in Figure 2.5, together
with the intermediate diameter, 300 f,lm, for comparison.
Comments
Note how the terminal velocities of fine particles in the Stokes law range
decrease with increasing temperature, being inversely proportional to gas
Single particles 65
Vt
(m/5)
1 mm.
- - J-
--
o. 2 L....-~--'----L..----J'--..I....-.....L.-.......L..-----I
Figure 2.S Terminal velocities of spherical particles of density 2650 kg m-3 in air, as
functions of temperature: - - - , 1 bar; - - - - -, 10 bar.
IAXIS
I
IAXIS
I, 0 I--
I-O--l
J
=- -r
- - - PLANE OF
b SYMMETRY
b
1 J_
OBLATE
PROLATE
Figure 2.6 Oblate and prolate spheroids.
Single particles 67
particles includes shapes such as bars whose sections are regular polygons.
For more detailed accounts of the behaviour of particles in creeping flow,
including results for other shapes, see Clift et al. (1978) and Happel and
Brenner (1983).
In creeping flow, it is useful to define the principal resistances to translation
of the particle. For motion parallel to the axis of an axisymmetric particle at
velocity U1 , the drag force is parallel to the relative velocity and given by
(2.15)
where C1 is the resistance to axial motion. Similarly, for motion normal to the
axis at velocity U2 , drag and velocity are again parallel with
(2.16)
where C2 is the resistance to motion normal to the axis. For a sphere of
diameter d = 2a, C1 == C2 = 3nd = 6na. Expressions for the principal
resistances of spheroids, thin discs and cylinders are given in Table 2.2.
Table 2.2 Principal resistances of spheroids, discs and cylinders in creeping flow
Analytic
Spheroids: Oblate (E < 1) Prolate (E > 1)
8Jta(E2 - 1)
Disc: 32AI3
Mean, in
random 6JtaV(1 - E2) 6JtaV(E2 - 1)
orientation,
E:
cos-IE In(E + V(E2 - 1»
Disc: 12a
Empirical (from Heiss and Coull, 1952)
Cylinders E = Lid; 0.1 < E < 10
Axial, 5.82dE- Y' V(2E + 1) exp[0.367e-o.448 V(2E + 1){(3EI2)'h - I}]
1)]
Cl
Normal to 0.939E'i4(l.0145g-1I6 -
axis, C2: 6.57dE 1I6 V(2E + 1) exp [
V(2E + 1)
Note that, for a sphere, d = 2a and C1 = C, = 3nd or 6na.
68 Particles in fluids
Because drag and relative velocity are parallel when the motion is parallel
to one of the principal directions, axisymmetric particles with fore-and-aft
symmetry settle on a vertical path if the axis is either vertical or horizontal.
The corresponding terminal velocities, for a particle of volume v, are
(2.17)
Oblate bodies generally have C2 < Cl, so that they fall faster with the axis
horizontal. The converse is true for prolate shapes. As noted in section
2.1.2.2, a very prolate (i.e. needle- or fibre-like) axisymmetric particle has C2 =
2CI, so that it falls twice as fast when oriented vertically than when its axis is
horizontal.
For any other direction of motion, drag is not parallel to the relative
velocity. Thus, while an axisymmetric particle with fore-and-aft symmetry
moves on a straight trajectory without rotating, the trajectory is not in general
vertical. Figure 2.7 shows such a particle settling with its axis at some angle <\>
to the vertical. The path of the particle lies in the vertical plane through the
axis, but is at some angle J3 to the vertical. The components of relative
velocity parallel and normal to the axis are respectively Ucos( <\> - J3) and
Usin( <\> - J3). The two components of the drag force are then
acting in the directions shown in Figure 2.7. For steady motion, the total
horizontal and vertical forces on the particle must be zero, i.e.
v(pp-p)g
u
Figure 2.7 Axisymmetric particle moving at terminal velocity in creeping flow.
Single particles 69
(2.21)
(2.22)
(2.23)
Equation (2.21) shows that, for a prolate particle with Cl < C2, J3 is positive,
i.e. the particle tends to 'sideslip' on a trajectory between the vertical and the
axis. Conversely, an oblate body with Cl < C2 hasJ3 negative, i.e. its trajectory
lies between the vertical and the equatorial plane. Figure 2.7 represents both
these cases. For a large number of identical particles in random orientation,
the mean resistance and terminal velocity are:
(2.24)
(2.25)
Particles such as regular polyhedra, for which the principal resistances are
equal in three orthogonal directions, have the same resistance in every
direction and settle on a vertical trajectory whatever their orientation. The
terminal velocity can be estimated reasonably reliably from the empirical
correlation (see Clift et al., 1978):
V t = V ts (1 + 0.3671n '\jJ) (2.26)
where V ts is the terminal velocity in creeping flow of a sphere with the same
volume and density, and '\jJ is the sphericity of the particle.
One of the applications of these results is in estimating relationships
between volume-equivalent and Stokes' diameters of non-spherical particles
(see section 1.2.1). The Stokes diameter d st is, by definition, the diameter of
the sphere with the same density and terminal velocity as the particle, i.e.
from equation (2.9),
'12
dst = [18""V/(Qp - Q) g] (2.27)
d st = [
18v/c]
1/2
(2.28)
70 Particles in fluids
;' "-
;'
3 "-
;' '''-~
ds/de ;' "-
"-
:2' , "-
"-
"-
"-
0.5
Solution
A disc of diameter d and thickness kEd has volume (rr,d 3EI4) X Ed, i.e. v =
rr,d3E/4.
Similarly, a spheroid with equatorial diameter d and aspect ratio E' has
volume
v = rr,3E'/6
so that, if the disc and spheroid have the same volume,
E' = 1.5E
From equation (2.17) the terminal velocity of the particle is
Vt = v(Qp - Q)g/f.!C
where C is the hydrodynamic resistance in the appropriate orientation.
Therefore the ratio of the terminal velocity of the disc to that of the sphere
IS
V/V; = c'/c
where c is the resistance of the disc of aspect ratio E and c' is that of the
spheroid of aspect ratio E' (= 1.5E). Consider three aspect ratios, E = 0.2
(E' = 0.3), E = 0.1 (E' = 0.15) and E = 0.01 (E' = 0.015). For the first two
shapes, the Heiss and Coull (1952) empirical results can be used; the third
shape lies outside the range of these correlations, but can be approximated as
a vanishingly thin spheroid (E ~ 0) because the drag on the curved sides can
be neglected. For the oblate spheroids, the results in Table 2.2 are used in
each case:
Table Continued
Comments
These calculations further illustrate the general point that, in creeping flow,
terminal velocities are insensitive to particle shape. In fact, the differences
between the estimates obtained by treating lenticular particles as discs and as
oblate spheroids are within the sort of accuracy which can be reasonably
expected of an empirical correlation. The difference between the two shapes
increases, to pass through a maximum around E = 1.0, but is never very large
(see Clift et al., 1978).
of the particle is actually less than that of the sphere of diameter d A , while its
drag coefficient may be somewhat different from that for the sphere.
Generally the former influence is the stronger, which explains why KA is
always less than unity and probably helps to explain why this approach is the
most reliable of those available for estimating V t for irregular particles (see
Clift et al., 1978). The estimates are usually within ±20% and frequently
within about 10%. Given the impossibility of describing particle shape by a
single parameter, this is about the best which can be expected of any simple
empirical correlation.
2.1.3.2.3 Newton's law range. At higher Re, such that the wake behind
the particle sheds intermittently, secondary motion sets in. In addition to
shape, size, and (Qp - Q), the mean terminal velocity then depends on Qp
because the particle inertia determines how strongly it is influenced by wake
shedding. Even spheres show the influence of QpiQ in the Newton's law
range, and can follow zig-zag or spiral trajectories, although the effect on V t is
small. The effects are much stronger for other shapes.
Discs are probably the best understood. In addition to the dimensionless
diameter, conveniently defined for a disc of thickness Ed as
i
0,01
I,
c
15
<1>
I
~ 10' STEAOf
III<1> '
c:
,Q
'"c
<1>
E
o
Figure 2.10 Regimes of motion for discs in free fall or rise: _._._, contours of
constant Re; - - - , contours of constant Sr. Bold curves indicate approximate
boundaries between regimes of motion.
where d v is the volume-equivalent diameter and 1jJ the sphericity. For other
shapes in the Newton's law range, it is generally advisable to measure V t
rather than attempting to estimate it.
effects, arising when the relative velocity between the particle and the fluid is
comparable with the speed of sound in the fluid, can usually be ignored in
processing equipment, except in processes like atomisation which use high-
velocity gas jets.
The significance of non-continuum effects is indicated by the Knudsen
number, Kn, defined as the ratio of the molecular mean free path in the fluid
(/...) to some characteristic dimension of the particle, usually the diameter (d):
Kn =/.../d (2.39)
Equations using Kn are at best estimates, so that there is little point in
questioning in detail which measure of d should be used in equation (2.39) for
non-spherical particles; if in doubt, the volume-equivalent diameter should be
used. Clift et al. (1978) discuss some results for the effect of slip on non-
spherical particles.
The mean free path depends on the gas temperature and pressure, and can
be estimated as (Beard, 1976)
/. . = 3.2 X 10-4f.tYJt/2QP (2.40)
For air, equation (2.40) can be approximated as
/. . = 2.15 X lO-4f.tr'h/P (2.41)
Note that /... is in metres and P in bars in both these equations. As a general
approximation from equation (2.4), /... is roughly proportional to T/P,
showing that non-continuum effects become more significant at elevated
temperature and reduced pressure. To indicate typical values, /... for air is
0.069 f.tm at 300 K and 1 bar, 0.32 f.tm at llO0 K and 1 bar and 0.032 f.tm at
1100 K and lO bar.
Non-continuum effects reduce the fluid/particle drag, and for this reason are
sometimes termed 'slip effects'. The drag (always at low Reynolds number;
see above) can be estimated for near-continuum slip flows as
FD = 3Jtf.tdUlC (2.42)
where C is the slip correction factor or 'Cunningham coefficent' defined as
The value of C for a given particle and gas actually depends on the
'accommodation coefficient', i.e. the fraction of gas molecules undergoing
diffuse reflection from the particle surface. Clift et al. (1978) summarise
results which can be used to estimate C if the accommodation coefficent is
known. Usually it is not known and C must be estimated empirically from
Davies' (1945) modification of a form proposed by Knudsen and Weber
(1911):
78 Particles in fluids
\
10-3 10-9
\
\
\
\
\
'iU) ~
E 10"'" E 10-10 <..)
C!l
~ C
10-5 10
10-6 -------
c
0.1 1 10
dp (ll m)
Figure 2.11 Variation of terminal velocity, diffusion coefficient and Cunningham slip
correction factor with particle diameter (Clift et ai., 1981). For particles of density
2500 kg/m3 in air and 300 K, 1 bar (--); 1100 K, 1 bar (_._); and 1100 K, 10 bar
(----).
Unsteady motion of single particles 79
Figure 2.11 also shows values for the Brownian diffusivity , DAB, associated
with the random motion of the particle in the gas due to collisions with gas
molecules. The values are calculated from the Stokes-Einstein equation (see
Clift et at., 1981)
DAB = Ck B Tl3nfld (2.45)
where kB is Boltzmann's constant, 1.380622 X 10-23 JK-1 . Brownian
diffusivity is a significant parameter in some gas/solid separation processes, as
discussed in Chapter 7.
nd3 ]
[ __ _
dU + _d2-vJtQiA
3 t [dUJ
_ __ ds
6 dt 2 roo dt t=s-vt=S
(2.46)
where U(t) is the instantaneous particle velocity. Thus the total instantaneous
drag is made up of three terms, of which the first is the steady (Stokes) drag at
the instantaneous velocity. The second term on the right of equation (2.46) is
known as the 'added mass' or 'virtual mass' term. It arises from the entrained
fluid which accelerates with the particle; for a sphere, this 'added mass' of
fluid corresponds to a volume equal to half that of the sphere. The final term,
known as the 'Bassett term' or 'history term', involves an integral over all past
accelerations of the sphere. The form of this term arises from generation of
vorticity at the surface of the particle, to diffuse into the surrounding fluid.
The history term can be thought of as describing the state of development of
80 Particles in fluids
the flow around the particle: if it is small, then the relaxation time of the flow
is sufficiently small that the instantaneous flow field differs little from steady
flow at the instantaneous value of U.
Equation (2.46) can be developed to give an equation of motion of the
sphere, for example to describe the initial motion of a particle released from
rest. The form of equation (2.46) only applies in creeping flow. Nevertheless,
and slightly surprisingly, a modified form with empirically-determined
coefficients of drag, added mass and history gives a good description of
accelerated motion at higher Reynolds numbers; see Clift et al. (1978).
Solution is obviously complicated by the history integral, but some useful
general conclusions emerge:
(a) The history term is usually larger than the added mass term. Therefore it
is futile to ignore the inconvenient history term but to include the added
mass term.
(b) For particles in gases, with Qp » Qg, the added mass and history terms
are often small, so that the drag can be approximated by the steady-state
drag at the instantaneous velocity. However, this is by no means
universally true, even for particles in gases. It is therefore advisable to
estimate U(t) by making the quasi-steady approximation, and then
evaluate the two other terms to verify the validity of the approximation.
(c) For particles in liquids, the quasi-steady assumption is almost never
justified. Therefore the kind of calculations made for particles in gases,
for example in inertial devices for separating particles from gases (see
section 7.1), cannot be applied to particles in liquids.
2.2.2 Acceleration
As shown earlier (equation 2.9), the terminal velocity of an isolated particle
in free-fall under gravity, in the Stokes law range and in the absence of slip
effects, can be expressed as
(2.47)
where
(2.48)
'tis known as the 'relaxation time', so called because it characterises the time
taken for a particle to adjust to any changes in flow velocity or direction. This
approach can be generalised to give the terminal velocity of a particle
subjected to any constant external force, F,
F
V tF = 't- (2.49)
m
Unsteady motion of single particles 81
During acceleration in a gas (see section 2.2.1), the drag force may still be
approximated by the Stokes law, using the instantaneous velocity, so that for
acceleration from rest in a gravity field
dV(t)
m - = mg - 3Jt!lV(t)d (2.50)
dt
or
dV(t)
,; - - = ,;g - V(t) (2.51)
dt
o ------;--------- --->0-
X
y V
Figure 2.12 Particle trajectory for a settling particle with an initial horizontal
velocity.
'tracking' of the streamlines, while large values indicate that the particle will
'resist' changes in direction or velocity.
Nozzle
r
~~_r--_ _ _ _ _~U
-Impac';o" pi",
I
I
I
Solution
A full analysis of the problem would require numerical solution of the
complex flow field. Consider a rectangular impactor nozzle of half-width (d/
2) and infinite length (i.e. a two-dimensional problem) and assume:
(a) uniform flow across the entry;
(b) that the fluid streamlines are arcs of a circle centred on A;
(c) that departures from the streamlines are small, so that the radial velocity
can be taken as constant.
The terminal radial velocity can then be taken as
(2.61)
where ar is the radial acceleration, U2 /r, and r is the radius of curvature of the
streamline. The velocity, U, is assumed to remain constant throughout the
curved region, which is a gross assumption but adequate for the present
purpose. Hence,
(2.62)
Any particle in the flow which is closer to the centre line of the nozzle than ~
will be collected by the impaction plate. Now the critical Stokes number, for
which 50% of the particles in the flow are collected, will be given by
n 1
-St=- (2.66)
2 2
or
Stcrit = 0.32 (2.67)
Assemblages of particles 85
Comments
The assumptions made about the fluid and particle motion in this example are
gross and therefore good agreement with experimental values would not be
expected. Nevertheless, the experimental value of Stcrit for the rectangular
nozzles (0.53) is of the same order as that obtained from this first-order
calculation. Note that the nozzle-to-plate separation distance is relatively
unimportant in impactor design because the gas jet expands only slightly
before coming within one jet diameter of the impaction plate. The key
parameter is therefore the slot width, or diameter if circular holes are used.
(2.69)
where D..p is the pressure drop, S is the particle surface per unit volume of
the bed, U is the superficial flow velocity (the volumetric fluid flow divided by
the cross-sectional area of the bed) and L is the thickness of the bed in the
direction of flow. The constant k has been found to be about 5 in many
practical cases.
If the bed consists of spherical particles of diameter d, then the number of
particles per unit volume is
6 (1 - E)/nd3
so that
6(1 - E) 6(1 - E)
S=ncPX =--- (2.70)
nd3 d
Assemblages of particles 87
dp (1 - f!U Ef
- - = 180--- (2.72)
dl 103 d2
Thus, the pressure gradient is directly proportional to the flow rate, a result
first obtained by Darcy in 1846.
If the particles are not spherical or mono-sized, then (Kay and Nedderman,
1985)
6 (1 - E)
S=---- (2.73)
(2.74)
2.4.1 Continuity
Figure 2.14 illustrates schematically the flow being considered. Fluid and
particles flow at volumetric rates Qf and Qs along a pipe of cross-sectional
area A, inclined at e to the horizontal. The superficial velocities of the two
phases are then defined to be
Uf = QflA (2.75)
Us=QJA (2.76)
The delivered volume concentration of the solid is defined to be
(2.77)
Now consider the mixture flowing through BB', a section normal to the axis
of the pipe. Consider an element dA of the cross-sectional area, as shown in
Figure 2.14. The local volume fraction of the solid in the element is C\; this
can be envisaged as the average fraction of time for which the element dA is
occupied by the solid phase. The mean in situ concentration of solids across
BB' is then defined as
Cyt = -1 f CydA
A
(2.78)
A A
-
If (1 - Cy)dA
A
=1- If
- , CydA
A
=1- C Yt (2.79)
A A A A
Consider now the flows of the phases along the pipe. Suppose that the
average velocities of the two phases parallel to the pipe axis at the element are
Vs and Vf, where each average is evaluated over the times when the element is
occupied by the phase in question. The flows of the two phases through dA
are then C\VsdA and (1 - Cy)VfdA. The total volumetric flows across BB'
are then
(2.80)
(2.83)
The mean slip velocity between the two phases across BB' is defined as (Vf -
Vs) whereas the local slip velocity at dA is Vf - Vs. From equations (2.78) to
(2.83),
(2.84)
- - 1[ 1 f
Vf - Vs=-
A
- - ' (1- Cv)VdA- -I
1 - C Vt A
A A
C Vt
f CvVdA J
A
AA
(2.85)
90 Particles in fluids
where
Vf = Vs = V
Hence, from equation (2.85),
= 1 J (CYt -
..
Cv)VdA (2.86)
ACyt(l - Cyt) A
From equation (2.86), the mean slip velocity (Vf - Vs ) is still positive if V is
relatively large where (CYt - C\) is positive (i.e. Cy < Cyt ), and is relatively
small where (CYt - Cy ) is negative (i.e. Cy > Cyt ). To give a specific
example, applicable to both pneumatic and hydraulic conveying (see Chapter
6), there is mean slip, i.e. (Vf - Vs ) is positive, if the fluid flows at higher
velocity through a part of the pipe where the local solid concentration (Cy ) is
relatively small, even if there is no significant local slip between the phases.
At a different level, consider the case where Cy is essentially constant
across the section. In this case, it follows from equation (2.78) that Cy = Cyt .
This behaviour is approached by some vertical or high-velocity flows. Then
the above equations simplify to
Vf - Vs = Vf - Vs (2.88)
For flow in a vertical pipe, the local slip velocity, Vf - V" may sometimes be
related to the hindered settling velocity of the particles in the fluid. Equation
(2.88) is used below to describe some particle/fluid flows.
Equation (2.80) illustrates a further general point. A given solids flow (Qs)
can occur at high Cy and relatively low Vs, or low Cy and relatively high Vs.
The former case corresponds to dense phase flow and the latter to lean phase
flow. It may be possible for a flow to switch between these two states. This is
the condition known as choking in pneumatic conveying (see Chapter 6), and
is usually to be avoided.
Assemblages of particles 91
(2.93)
~ D\
af~
as --------~---------------------------.
Figure 2.15 Flow of a particlelfluid mixture in a pipe - momentum.
where the first term in equation (2.96) arises from shear at the wall, the
second represents the axial component of the weight of the mixture in the
element in question and the last term arises from the pressure gradient along
the pipe. The linear momentum balance along the pipe is then
nD2 dP
g sin 8[ CytQs + (1 - Cyt)Qd - --
4 4 dL
(2.97)
JM =JMs + JMf
= MsVs + MfVt (2.98)
Ms = f
A
QsCy VsdA = A QsCYt Vs
4
(2.99)
(2.100)
Assemblages of particles 93
Unless gas or solids are added to or removed from the pipe at the section in
question, continuity requires Ms and M f to be constant along the pipe.
Therefore, for a homogeneous flow,
Equation (2.102) shows that the pressure gradient along a straight pipe can be
expressed as the sum of three terms. The first arises from the shear stress at
the pipe wall; in this macroscopic formulation, it does not matter to which
phase the stress is imparted. The second term arises from the weight of the
mixture in the pipe, and can therefore be regarded as a hydrostatic term. The
last term arises from the acceleration of the two flowing phases.
We can now consider some further simplifications. For a fully-developed
flow, whose parameters do not change along the pipe, the acceleration term is
zero. Then
dP 41:
- = - - - g sin 8[CytQs + (1 - Cyt)Qr] (2.103)
dL D
This equation is used below to describe fully-developed flows with particles
transported by a liquid (hydraulic conveying) or a gas (pneumatic conveying).
For a vertical flow, the hydrostatic term commonly dominates. It should be
noted that it depends upon the in situ solids concentration, C yt , rather than
the delivered concentration, C yd .
As a further simplification, consider a horizontal flow [8 = 0; sin 8 = 0].
Equation (2.102) then takes the simpler form
(2.105)
For a lean phase flow, the pressure profile along the pipe therefore typically
takes the form shown in Figure 2.16. Upstream of the feed point, the pressure
gradient is small because 't arises solely from the shear on the gas and is
therefore relatively low. Downstream of the feed point, where the solids are
being accelerated, the term in Ms!1 Vs dominates and so the pressure drops
rapidly. As the solids reach their steady mean conveying velocity, the
acceleration term dies away and the pressure gradient is again dominated by
't, although the mean wall shear is larger than that for the gas alone and
transmitted primarily to the particles. By estimating the acceleration pressure
drop, !1PAccn in Figure 2.16, it may then be possible to estimate the mean
velocity of the conveyed solids.
In addition to pressure changes close to the feed point, this qualitative
aspect also explains why sharp bends have a strong effect on the pressure drop
in lean phase pneumatic conveying: the solids must accelerate, following the
bend, to a conveying velocity in a different axial direction.
In a vertically-upwards flow, the entrance effects are even larger. Low Vs in
the immediate vicinity of the feed point gives high solids concentration (CYt )
so that the hydrostatic term also contributes to the local high pressure
gradient.
L
i, -------------------_
,
,, L
~ __ 1Ms
---L-------------------------------~
Mf (Ms+ Mf )
IKE = f [CvQsVsCV;/2) + (1 -
A
Cy)QfVf(Vr/2)]dA
For a steady flow, the gradient of potential energy in the direction of flow is
g sin 8(Ms + Mf). The flowing mixture does 'displacement work' against the
pressure gradient given by (Qs + QD (dPldL). Putting these terms together,
the rate of mechanical energy dissipation from the flowing mixture per unit
length of pipe is
dP
- (Qs + Qf)- - g sin 8(Ms + Mf )
dL
(2.107)
dP1rr dP (Ms + MD
- dL = - dL - g sin 8 (Qs + Qf)
1 -d [Qs
- f'CyVsdA + -Qf f (1 -
'3 Cy)VfdA
"3 ] (2.108)
(Qs + Qf) dL 2 A 2 A
[ - dVs
-
1
MsVs- + MfV- f dVfJ
- (2.110)
(Qs + Qf) dL dL
For a fully-developed flow, with Vs and Vf constant along the pipe,
-dP1rr dP (Ms + Mf )
(2.111)
dL = - dL + g sin 8 (Qs + Qf)
Inserting the expression for the pressure gradient from equation (2.103),
(2.112)
The first term on the right of equation (2.112) corresponds to the frictional
(and therefore dissipative) wall shear stress, and is in the form familiar for
single-phase flow. The other term can best be investigated by noting that
[gsCYt + gf(1 - Cyt) ] is the mean in situ mixture density, gMt, whereas
the term (Ms + Mf)/(Qs + Qf) is the mean delivered mixture density g. In
these terms, equation (2.105) becomes
-dP~ ~ _
---= - + g sin 8[gMt - g] (2.113)
dL D
Note also that
gMt - g = (gs - gf)( C Vt - Cyd)
= (gs - gf)Cyd[A/(l - A)] (2.114)
from equation (2.92).
For an upflowing mixture, in the common case where gs > gh the
direction of slip is such that A is positive (see equation (2.91)): therefore the
density difference term in equation (2.112) is positive, representing a
contribution to mechanical energy dissipation. If the mixture is in downflow,
A is negative (see equation (2.91)) so that Vs > Vm. However sin 8 is now
negative, so that the term is again positive. Thus in both cases, slip of the
solids relative to the fluid leads to dissipation of mechanical energy - as
expected, given that particle/fluid drag is a dissipative frictional process.
2.5 NOMENCLATURE
A Area (m 2 )
C Cunningham slip correction factor (equation (2.43)
Nomenclature 97
Greek letters
a Shape coefficient in section (2.1.3.2.2)
f3 Angle between particle trajectory and vertical (Figure 2.7)
E Void fraction
Angle between axis of symmetry and vertical (Figure 2.7)
Mean free path (m)
Viscosity (Pa s)
Fluid density (kg m-3 )
98 Particles in fluids
REFERENCES
Batchelor, G.K. (1967) An Introduction to Fluid Dynamics, Cambridge University
Press, Cambridge.
Beard, K.V. (1976) 1. Atmos. Sci., 33, 851-64.
Clift, R., Grace, 1.R. and Weber, M.E. (1978) Bubbles, Drops and Particles,
Academic Press, New York.
Clift, R., Ghadiri, M. and Thambimuthn, K.V. (1981) in Progress in Filtration and
Separation, Vol. 2 (ed. R.l. Wakeman), Elsevier, Amsterdam, p. 75.
Davies, C.N. (1945) Proc. Phys. Soc, 57, 259-70.
Dullien, F.A.L. (1992) Porous Media, 2nd edn., Academic Press, London.
Ergun, S. (1952) Chem. Eng. Prog. 48, 89.
Ghadiri, M., Seville, 1.P.K., Raper, 1.A. and Clift, R. (1986) Estimation of the
Hydrodynamic Diameter of Fine Particles from Equivalent Area and Volume
Diameters, Proc. 1st World Congress of Particle Technology, Nurnberg, 36--54.
Govier, G.W. and Aziz, K. (1972) The Flow of Complex Mixtures in Pipes, Van
Nostrand Reinhold, New York.
Grace, 1.R. (1986) Can. 1. Chem. Eng. 64,353.
Happel, 1. and Brenner, H. (1983) Low Reynolds Numbers Hydrodynamics, Martinus
Nijhoff, The Hague.
Heiss, 1.F. and Coull, J. (1952) Chem. Eng. Prog. 48, 133.
Heywood, H. (1938) Proc. Inst. Mech. Engrs. 140,257.
Hill, R. and Power, G. (1956) Quart. 1. Mech. Appl. Math. 9, 313.
Hinds, W.C. (1982) Aerosol Technology, Wiley-Interscience, New York.
Kay, 1.M. and Nedderman, R.M. (1985) Fluid Mechanics and Transfer Processes,
Cambridge University Press, Cambridge.
Knudsen, M. and Weber, S. (1911) Ann. Phys, 36, 981-94.
Mayhew, Y.R. and Rogers, G.F.C. (1972) Thermodynamic and Transport Properties
of Fluids, 2nd edn., Basil Blackwell, Oxford.
Potter, R.L., Pruppacher, H.R. and Hamielec, AE. (1973) 1. Atmos. Sci. 30, 125.
Soo, S.L. (1989) Particulates and Continuum, Hemisphere Publishing, New York.
Stringham, G.E., Simons, D.B. and Guy, H.P. (1969) U.S. Geological Survey,
Professional Paper 562-C.
Turton, R. and Levenspiel, O. (1986) Powder Technol. 47, 83.
Wallis, G.B. (1969) One-dimensional Two-phase Flow, McGraw-Hill, New York.
Willmarth, W.W., Hawk, N.E. and Harvey, R.L. (1964) Phys. Fluids 7,197.
3
Particle mechanics
>-
0\
'-
QJ
C
QJ
d
....c
.l'! \ Repulsion
o \
CL
\
\. Separation r
0.111111111111111111111 1I11I1~III1IIIII1III1I1III1II1UIIIJIIIIIIIIIIII 111111111111111111111
_ro_7
I .
I Attraction
I
Figure 3.1 Potential energy as a function of the separation between two atoms
attracted by van der Waals forces. Similar curves apply to all types of interactions
though they may differ in scale and shape. (ro indicates equilibrium separation.)
(Tabor, 1987.)
There are essentially two theories for predicting the van der Waals
interaction between large bodies: the microscopic (Hamaker) and the
macroscopic (Lifshitz) theories. The Hamaker theory is easier to grasp and its
main results are summarised here; the Lifshitz theory is more rigorous but the
form of its predictions is very similar to those of the cruder Hamaker theory.
(3.4)
• z
Figure 3.2 Interaction between a molecule and a plane surface (after Israelachvili,
1991 ).
702 Particle mechanics
Note that
X=OO _ _ l 1
] (3.5)
n-2
x=o
Thus,
-2JtCg
W(a) = for n > 3 (3.6)
(n - 2)(n - 3)a n - 3
Therefore, for n = 6,
W(a) = -JtCg/6a3 (3.7)
and
-dW(a)
F(a) = - - = -JtCg/2a4 (3.8)
da
dz
a
z=R
2R-z
a+z
Figure 3.3 Interaction between a sphere and a plane (after Israelachvili, 1991).
Interparticle forces 103
so that the number of molecules in this section is n(2R - z)z dz Q. All these
molecules are at a distance (a + z) from the planar surface. Therefore, from
equation (3.6),
=f
z= 2R - 2nCQ nQ(2R - z)z
W(a)
z = 0 (n - 2)(n - 3)
. (a + Z)"-3
dz
W(a) =-
2n 2CQ2 f 00 2Rz dz
-- (3.12)
(n - 2)(n - 3) 0 (a + Zy-3
Therefore,
n 2CQ2R 00 z
(3.13)
W(a) =- (n _ 2)(n - 3)1'0 (a + Z)"-3 dz
Note that, by substitution y = (a + zt-3
00 z 1 1
(3.14)
fo (a + Zy-3 dz = (n - 4)(n - 5) . a"-s
Therefore,
-4n 2 Cg 2 R
W(a) = - - - - - - - - - - - - cS
(n - 2)(n - 3)(n - 4)(n - 5)a n -
which for n = 6 becomes
W(a) = -n2CQ2R/6a
=-AR/6a (3.15)
where A is the Hamaker constant, n 2 Cg 2 •
For a» R, however, equation (3.11) becomes
which is essentially the same as equation (3.6), since 4nR 3 g/3 is simply the
104 Particle mechanics
number of molecules in the sphere, i.e. at large separations the sphere now
behaves like a single point body located at its centre.
3.1.1.3 Other geometries. It can be shown that for two spheres of equal
radii, R, at small separation, a (<< R), the interaction energy is half of that in
equation (3.15), while at large separations W(a) is proportional to l/an, as for
two molecules.
In practice, the results at small separations are of most interest. In general,
for two unequal spheres of different materials,
-rr?0J1Q2R
W(a)=---- (3.18)
6a
where R = RIR2/(Rl + R 2) and R1 and ~h are the radius and molecular
number density, respectively, for sphere 1. Other interaction laws for
common geometries are given in Figure 3.4.
3. 1. 1. 5
Combining relations. Approximate relationships are available
which enable unknown Hamaker constants to be obtained from known ones.
If we define A 132 as the Hamaker constant for media 1 and 2 interacting
across medium 3, then
A132 = ± V(A l3l A 232 ) (3.19)
or, if medium 3 is a vacuum (or air)
A12 = V(A l1 A 22 ) (3.20)
Interparticle forces 705
Molecule-surface Sphere-surface
D p
AL (R'R' )112
W = - 12.,/20'12 R, + R,
R,
:
W = -A..JR:R, 160 W = -A I 12ltO' per unit area
Figure 3.4 Non-retarded van der Waals interaction free energies between bodies of
different geometries calculated on the basis of pairwise additivity. The Hamaker
constant A is defined by A = n 2 C!.hQ2, where Ql and Q2 are the number of atoms
per unit volume in the two bodies and C is the coefficient in the atom-atom pair
potential. (Israelachvili, 1991.)
106 Particle mechanics
Table 3.1 Hamaker constants for two media (1 and 2) acting across another medium
(3) (after Israelachvili, 1991)
1 3 2 Calculated Measured
Also,
A131 = A313 = (VAu - vA33)2 (3.21)
Combining equations (3.21) and (3.19),
Al32 = (vAll - VA 33 ) (VA22 - VA 33 ) (3.22)
A132 is negative if All < A33 < A22 or All > A33 > A 22 . Thus it is possible to
have a repulsive van der Waals force, as in the case of the quartz/water/air
system. These relationships apply only when dispersion forces dominate; in
particular, they are not reliable where water is concerned. Nevertheless,
some general rules may be derived:
(i) the force between any two bodies in vacuum (or air) is always attractive;
(ii) the force between two identical bodies in a medium is always attractive,
while that between different bodies in a medium can be attractive or
repulsive;
(iii) the Hamaker constant for two identical media acting across a second
remains unchanged if the media are interchanged, i.e. two water droplets
in air will attract with the same force as two air bubbles in water (all
other things being equal).
3.7.7.6 Retardation effects. The van der Waals force between two atoms
arises because of the temporary dipole set up in atom 2 in response to a field
produced by atom 1. If the time taken for the electric field of the first atom to
Interparticle forces 107
reach the second and return is comparable with the period of the fluctuating
dipole itself, the force arising from the interaction will be less than it would
otherwise be, i.e. it will be retarded. Hence the accurate value of the
Hamaker constant is reduced at larger separations, and in exceptional cases
can even change sign. This is accounted for by the full Lifshitz theory, but
obviously means that the Hamaker 'constant' is not really a constant at all,
but is itself a function of the surface separation.
3. ,. 1.8 Practical calculations. The expressions listed above give the van
der Waals interaction forces up to the point of contact between bodies, but
what is meant by 'contact' is unclear since if a is allowed to go to zero, the
interparticle forces become infinite. Many older references take a to be 4 X
10-10 m, which is close to the average interatomic distance for many common
solids. However, Israelachvili recommends the use of a 'cut-off distance', ao,
substantially less than the interatomic distance. He shows that a 'separation'
of 1.65 X 10-10 m gives values for surface energies within ±20% of
experimental values. From Figure 3.4 and equation (3.28),
y=A/24Jta5 (3.29)
Substituting for ao,
A = 2.1 X 1O-18y (3.30)
Real particle surfaces are, of course, rough and the radii of the asperities may
determine the van der Waals force. Fine particles may act as spacers between
the larger particles and therefore reduce the forces between them, or may fill
in gaps and thus increase the interaction force, depending on their relative
SIze.
Real particles are not rigid and will deform elastically and/or plastically at
the contact point, even under zero external load. The effect of elastic
deformation is considered in section 3.2. For plastic deformation, the total
interaction force must include a term for the extended contact area. From
Figure 3.5, for two spheres, Visser (1989) recommends
T
2r I
I
+
_l~
I I
.-.~1 ~_a -
AR A
F(a) =- + - . n,2 (3.31)
12a2 6na 3
where r is the radius of the extended contact area.
From the calculations leading to equation (3.15) it is apparent that the total
interaction energy for two bodies almost in contact is dominated by the
interactions between the surfaces of the bodies, the so-called 'screening'
effect. In fact, the interaction energy is determined almost entirely by the two
surface layers of depth equal to the surface separation. Thus thin layers of, for
example, oxide or adsorbed gases may have a strong effect on the observed
interparticle force.
PENDULAR FUNICULAR
CAPILLARY
3.1.2.1 The capillary force between two spheres. In general, the capillary
force between two particles is the sum of three terms: the axial component of
the surface tension force at the solid/liquid/gas interface, the force due to the
reduced hydrostatic pressure in the bridge itself, and the buoyancy force due
to the partial immersion of each particle. Princen (1968) and Picknett (1969)
demonstrated for spheres that both the buoyancy term and the distortion of
the shape of the bridge due to gravity can be neglected if the particle size is
small, say < 1 mm.
Consider two identical spheres of diameter 2R, joined by a liquid bridge of
half-angle p and separated by a distance 2a (Figure 3.7). In the arguments
which follow, zero contact angle (perfect wetting) will be assumed, although
the derivations can be readily extended to non-zero contact angles (Coughlin
et aI., 1982). According to the Laplace equation (Shaw, 1992), the reduced
hydrostatic pressure within the bridge, !1P, is given by
(3.32)
where rl and r2 are the principal radii of curvatur~f the bridge, as shown in
Figure 3.7, and y is the liquid surface tension. This equation requires that the
liquid surface must have constant total curvature, (ri1- r2:I), so that the surface
of the bridge in the plane of the paper is not a circular arc.
However, Fisher (1926) used a toroidal approximation for the shape of the
liquid bridge (i.e. assuming that the curvature of the bridge in the plane of the
paper is constant) and this has been shown theoretically (Orr et al., 1975) and
experimentally (Cross and Picknett, 1963; Mason and Clark, 1965) to lead to
an accurate estimate of bridge strength provided that the particles are smooth
and truly spherical. Consider first the case of zero sphere separation (a = 0),
Figure 3.7 Liquid bridge between two equal spheres (perfect wetting). a = half
particle separation.
Interparticle forces 111
SPHERICAL
PARTICLE
~--4-----------~ "
r,
Figure 3.8 Liquid bridge joining two spheres (only one shown; zero contact angle).
which is the situation examined by Fisher (1926). The axial surface tension
force acting at the dividing plane is given by
(3.33)
while the hydrostatic force evaluated from the reduced pressure in the liquid
bridge at the dividing plane is
(3.34)
(3.35)
By trigonometry,
r1 = R(secj3 - 1) (3.36)
r2 = R(1 + tanj3 - secj3) (3.37)
Putting t = tan (j312) , and substituting for r1 and r2 in (3.35) the following
simple result is obtained
2nRy
F =----- (3.38)
c 1 + tan(j312)
This result differs slightly from that obtained by Adams and Perchard (1985)
who argued that it is correct to evaluate the surface tension and hydrostatic
forces at the surface of the sphere, so that equations (3.33) and (3.34) become
(3.39)
112 Particle mechanics
and
(3.40)
i.e. the reduced hydrostatic pressure acts over the projected wetted area. The
main effect of this modification is to alter the proportion of the total force
which is attributed to each origin; the sum F1 + F2 remains almost the same as
that given by equation (3.38), as it should if the spheres are at equilibrium.
Adding (3.39) and (3.40), and substituting for '1 and '2 as before, we obtain
the result
Fe = 2nRy [
2f! - t + 1J (3.41)
(1 + f!)2
where t = tan(!3/2). This gives values for Fe within a few percent of those given
by equation (3.38) for values of j3 within the range of interest; the discrepancy
is attributable to the error in the toroidal approximation.
The variation of interparticle force with increase in the bridge half-angle j3
is shown in Figure 3.9, for the following values of the variables:
2R = 922!-tm
y = 0.072 N m- 1 (water at 25°C)
a = 0
Values of the surface tension and capillary forces and the total force are given
according to both approaches. It is clear from Figure 3.9 that the total force is
virtually independent of j3 at small values of j3; as predicted by equation
(3.38), Fe tends to a limiting value of 2nRy. At larger values of j3, Fe reduces
slightly. The range of interest of j3 extends· only to about 40°, since the
coalescence limits for liquid bridges between spheres are 30° and 45° for close
packed and cubic arrangements, respectively (Coughlin et al., 1982). It is
apparent that for values of j3 below about 10° the contribution of the surface
tension force is negligible.
Equation (3.38) predicts an increase in the magnitude of Fe for a decrease
in the size of the liquid bridge (represented by j3) until a maximum value is
reached at zero liquid content. Although careful laboratory experiments,
such as those of Cross and Picknett (1963) and Mason and Clark (1965),
reproduced this trend down to very low liquid volumes, this behaviour is
nevertheless the opposite of what one would intuitively expect, i.e. wetter
powders commonly appear stronger than drier ones (until the capillary state is
reached, at which point the strength is reduced). Pietsch (1968) attempted to
resolve this apparently paradoxical conflict between theory and experimental
result by suggesting that all real contacts are rough and that an effective
Interparticle forces 113
- - - Fisher coalescence
limits
Adams & Perchard
300 45°
~
z:
5x104·r---~r-~~~-r,,~----~---r~TT'
UJ
~
a
u..
UJ r-~~-==---=_==:--_:::::::-.!TO~TAL FORCE
---
-I
I..J
t= HYDROSTATIC
a::
Cf.
a:: 10-4 FORCE
UJ 2R = 922}1m
I-
z: If = 0·072 N/m
5x 10-5 Q= 0 /
/
/
/ SURFACE
/ TENSION
FORCE
Figure 3.9 Variation of hydrostatic, surface tension and total capillary forces with
liquid bridge half-angle, jJ.
3.1.2.2 The capillary force for cone and plate contact. Most particles of
practical interest cannot be considered spherical, especially when viewed on a
114 Particle mechanics
Figure 3.10 Schematic representation of a liquid bridge between two particles with
surface asperities (after Pietsch, 1968).
a./R= 0
F
2TTRa'
0·1
5 10 50
BRIDGE HALF-ANGLE 13 (")
Figure 3.11 The dependence of normalised adhesion force, F12Jtry, between two
equal spheres on the bridge half-angle, j3, as a function of dimensionless separation,
aiR.
CONICAL
r LIQUID BRIDGE
Figure 3.12 Liquid bridge at the contact between a cone and a plane.
angle and assuming the surface of the liquid describes an arc of a circle in the
plane of the paper as shown, the force acting perpendicular to the plane is
given by
Fe = 2n:yy cosa + I1Pn: y 2 (3.44)
where
(3.45)
cosa ]
r=! [ (3.47)
1 + sina
W
\..J
IX
o
1L
>-
IX
~
-J
0:::
«
\..J
10~~~~~~~~~~~~~~--~~~~
(}OO1 0·01 (}1
SEPARATION RATE (m/s)
Figure 3.13 Ratio of viscous to capillary force, FvfFc> as a function of separation rate.
-YVL
In(p/po)=-- (1 1)
--- =. ·k (3.51)
RT '1 '2
where the term in brackets is the total curvature of the bridge surface, k.
Equation (3.51) expresses the equilibrium between the liquid in the bridge
and that in the surrounding vapour; in other words, it provides a link between
the vapour pressure in the surrounding gas and the extent of capillary
condensation at the contact points between the particles in a powder.
Consider two spherical particles as shown in Figure 3.7, at contact (a ---c> 0).
By simple geometry,
?z + 2'1'2 - 2'1R = 0 (3.52)
Substituting for '2 in terms of the total curvature k, we have
rj(2k + 2k2 R) - '1(3 + 4kR) + 2R = 0 (3.53)
which is a quadratic in '1 with the following physically realistic solution:
118 Particle mechanics
3 + 4kR - (9 + SkR)'/2
'1 = 4k(1 + kR)
(3.54)
-----
3.5
4ynr
~ 3.0 3
00
-
<::>
~
0.>
2.5
....0 2.0
<S
0.>
U l.5
.~
e-
~ 1.0
Surface tension force
0.5
0
~--------- ---------------
0 10 20 30 40 50 60 70 80 90 100
Relative pressure, p/Po(%)
Figure 3.14 (a) Surface tension, pressure and total forces as a function of (P/Po);
contact angle = 0° (Coughlin et at., 1982).
Interparticle forces 179
i --
Pressure force
4.0-
____ 3.5-
2S 4yrrr
~ 3.0- 3
;<
~
.....
2.5-
<8
~ 2.0-
'i
e- 1.5-
OJ
:E 1.0 -
0.5-
Surface tension force
)
0 I I I I I I I I I
0 10 20 30 40 50 60 70 80 90 100
a = Cone half-angle
130
12
llO
100
a =80°
90
*...,
....
oJ 80
~
<E
.,
"0 70
<)
'"
"0 60-
~
50
40-
30
20
10
0
0 10 20 30 40 50 60 70 80 90 100
Relative pressure, p/Po(%)
Figure 3.15 Cone-on-plate contact. Effect of cone half-angle (a) on the interparticle
adhesion force, F*, for e = 0° (Coughlin et al., 1982). (F* = Jt"?vdRT).
-QA QB
FE = -----=-- (3.58)
4JtErEoa
2
Interparticle forces 121
(a) A B
Attraction
(0 0
0)
a
• ~
Repulsion
(0
(b)
Image
Charge
/~
(0 I
\
'"
'"
_/ /
\
a a
(c)
a r
. .
+Q -Q +Q
Figure 3.16 Electrostatic forces. (a) Coulombic force; (b) image charge force - semi-
infinite surface; (c) image charge force - confined surface.
J22 Particle mechanics
where QA and QB are the total charges on the two particles, separated by a
centre-to-centre distance a, and EO and Er are the permittivity of free space
and the relative permittivity, respectively. FE is thus positive in attraction.
(ii) Space charge forces. Each particle in a cloud of charged particles is
affected by its interactions with all other particles. If, as is often the case, all
particles have like charge, this will result in mutual repulsion. The effect is
generally only of importance at high charge levels and high mass loadings, as
in electrostatic precipitation. The maximum electrical field which can be
sustained in air due to space charging is about 3 X 106 V m- I , the breakdown
strength of air; this same limit effectively limits the maximum charge which
particles can carry.
(iii) Image-charge forces. When a charged particle approaches a surface,
it induces an 'image charge' in the surface, as shown in Figure 3.16(b). For a
charged particle and a semi-infinite surface, the attraction is the same as if
there were a second particle of opposite sign on the other side of the
interface. For a conducting surface,
Q2
F 1C = ------ (3.59)
4JtErEo(2a?
(3.60)
z:
w 10-6
w
0::
o
LL
W
-'
w
i=
0::
rt
7l
0::
~ 10-7 -
CAPILLARY
- - -
FORCE (MAX,)
--1---
a=1·65A-
VAN DER WAA
- - -0=4;"
10 100 10c0
PARTICLE DIAMETER (,urn)
arising from a chemical bond. Clearly, much larger interparticle forces than
those considered here can occur between particles if there is a chemical
reaction between the surfaces, or if sintering occurs, for example.
(3.63)
where
1- vT
k1=---
E1
aI-I-a l
Figure 3.18 Contact between two elastic spheres in the presence (contact radius al)
and absence (contact radius ao) of surface forces (after Johnson et ai., 1971).
W = normal load; [) = displacement.
and
1 - Y~
k2=---
E2
R1 and R2 are the radii of the spheres and Yj and Y2 are the Poisson's ratios
of the materials. For equal spheres of the same material and diameter d,
3 (1 - y2)
a3 = - Wd (3.64)
8 E
This analysis was first performed by Hertz (1882), after whom it is named.
Hertz also obtained the relationship between the displacement, 0, and the
normal load
(3.65)
dW
(3.68)
do
and the force necessary to separate the two spheres, the 'pull-off' force is
3nfd
W=- (3.69)
p 8
Johnson et al. (1971) were able to confirm these results by experiment using
optically-smooth rubber spheres.
Figure 3.19 Force transmission lines in a compressed bed of particles; the width of
the lines indicates the relative magnitude of the local contact force; principal stress
vertical (Thornton and Barnes, 1986).
128 Particle mechanics
3.2.2. 1 Elasticity. The link between the elastic nature of the interparticle
contact and the macroscopic elasticity of an assembly has been analysed by
Kendall et at. (1987). For uniaxial normal stress on the assembly * , 0*, the
mean load per contact is
W= o*ln (3.70)
where n is the number of particle-particle contacts per unit area across a
plane through the assembly. By considering a number of regular packings
with known void fractions, £, Kendall et at. (1987) derived results which are
equivalent to
n = 13.3 (1- £)4a2 (3.71)
where d is the diameter of spherical particles making up the assembly. (This
may be compared with an earlier result due to Rumpf (1962) which gives
similar values for n for values of £ of about 0.4, as shown by Abdel-Ghani et
at., 1991. The extensive subject of packing geometry is considered in detail by
Cumberland and Crawford, 1987.) An external stress causes deformation of
the contacting particles. If the average relative displacement of the particles
under load is 0, then the mean uniaxial strain is, by simple geometric
arguments,
s* = old (3.72)
The assembly elastic modulus is then
do* dW
E* = - = nd- (3.73)
ds* do
assuming that the strain causes no change in the contact density. Since W is a
non-linear function of 0, this equation is applicable only to small deformations
about the unloaded state. Substituting for (dWldo) from equation (3.68),
E* =n [
9Jtf E 2 d 5 J1jj
(3.74)
16(1 - y2)2
or, using equation (3.71) for n,
9JtfE2 ] 1"3
E* = 13.3(1 - £)4 [ (3.75)
16d(1 - y2?
* Asterisks are used to denote the properties of the assembly, so that, for example, E* refers to
the elastic modulus of the assembly and E is the modulus of an individual particle.
Effects of interparticle forces at contacts 129
Kendall et at. (1987) showed good agreement between this equation and
measured elastic moduli of compacts of submicron titanium dioxide powder,
where particle-particle contact might be expected to be good. However,
Abdel-Ghani et at. (1991), working with much coarser glass spheres,
measured elastic moduli which were much smaller than those predicted from
equation (3.75) and they proposed possible reasons for this disagreement. It
is likely that both the inhomogeneous distribution of stresses (see earlier) and
the imperfect nature of the interparticle contact in the case of larger particles
playa role.
(3.76)
t Anextremely entertaining layman's account of the subject is contained in J.E. Gordon's book,
The New Science of Strong Materials (Gordon. 1976).
130 Particle mechanics
Strain-
free
Figure 3.20 Elastic body containing a crack under tensile stress, showing strain
release around the crack. (B is the thickness, perpendicular to the plane of the paper).
1 aU (3.77)
- =2y*
B oA
=Gc
where a U/ aA is the 'rate of release of strain energy' with respect to crack
extension, y* is the free energy per unit area of fresh surface (i.e. the
'fracture surface energy'), and G c is known as the 'critical strain energy
release rate'. (Note that G c is not a 'rate' in the more conventional sense of
the term.)
Following Kendall et al. (1986), the fracture surface energy of a compact
might be expected to be related to the interactions between the constituent
particles as
2y* = nu (3.78)
where u is the energy required to separate two particles. For smooth spheres,
u follows from the JKR analysis as
SrSd4 ] v:,
u = [ 0.296 Jt E2 (1 - y2f (3.79)
In practice, equation (3.80) does not give a good prediction of the fracture
energy except in those rare cases where the particle interactions can be
considered entirely elastic. It is well known that there is a region of stress
concentration at any notch or flaw. When the local stress exceeds the elastic
limit, the compact deforms inelastically. Relative movement between the
particles may also dissipate mechanical energy by interparticle friction
(Adams et at., 1987). In macroscopic terms, a plastic 'process zone' develops
ahead of the notch in which energy is dissipated by plastic deformation rather
than being stored elastically (Mullier et at., 1987; Adams et at., 1989). In this
case, equation (3.77) becomes
(3.81)
where y; is the additional energy dissipated. It is common in many materials,
including polymers and metals, for this term to be considerably greater than
the interfacial energy. The formation of a process zone increases the effective
crack length due to perturbation of the stress field.
Adams et at. (1989) discuss methods of measuring G c and related
parameters. All tests involve loading a specimen in some way until failure
occurs; the commonest example is the 'three-point bend'test (Figure 3.21).
For 'stable fracture', in which the crack extends only as external work is done
on the specimen, G c can be obtained from a simple energy balance
U
G=--- (3.82)
c B(D -A)
t Applied force
where U is the work done on the specimen in the fracture test (the area under
the force-deflection curve) and B(D - A) is the fracture area. If fracture is
'unstable', sufficient energy is stored in the system at the critical point to
propagate the crack completely across the specimen and
U
(3.83)
Gc = BD¢>
where U is the work done on the specimen up to the fracture point and ¢> is a
calibration factor which accounts for the change in compliance (the rate of
change of displacement with change in load) of the specimen with notch
length; ¢> is tabulated for bars containing notches of various lengths by Plati
and Williams (1975).
Abdel-Ghani et al. (1991) used a modification of the three-point bend test
to measure the fracture energy of assemblies of glass spheres, and confirmed
the earlier conclusions of Adams et al. (1989) that the fracture of assemblies
or compacts of fine particles can be characterised using the principles of
fracture mechanics. However, the measured values of G c were much larger
than those calculated from equation (3.80), although this had been found
adequate to explain the fracture behaviour of submicron powders, albeit with
a fitted value for r (Kendall et aI., 1986). As discussed above, the explanation
is that most of the work done on the specimen is dissipated, rather than being
stored as elastic energy which can be released to create new surface.
F= !AwW (3.84)
0.30 ,--.,---,------,---,---,
(a)
0.20
0.10 .
• ••
-----~-.----.--------
•••••••
• ••
o ~~~==::==~~
0.60 I I 0.75 r : . : - - - - - , - - , - - - - , - - - - - , - - - ,
~ (b) (c)
0.64
+
I
I
0.54
0.40 ~ • -
\.
0.20
. '......
I-- ,
\
,
,
..
-
:l
0.43
0.32 \
..
'
•
I- • •- ... ----..--.\-- -
•• • 0.21 ~! ---
-
O~ __L_I__L_I__L_I__ L_I~
o 3 6 9 12 3 6 9 12 15
(Jw (kPa) (Jw (kPa)
Figure 3.22 Experimental wall friction coefficients: (a) mustard seeds; (b) glass
beads; (c) polyethylene beads (Tiiziin et al., 1988).
Highly polished metallic surfaces also exhibit very large effective coefficients
of friction at low normal loads. Tabor and his coworkers (Tabor, 1955;
Bowden and Tabor, 1955) explained such phenomena via an 'adhesion
model' requiring significant tangential force to break bonds formed between
the contacting surfaces. Briscoe et al. (1984, 1985) extended the same concept
to describe the interfacial shearing mechanism of thin polymeric films
wherein new 'junctions' continually form as old ones are ruptured.
The inclusion of an interfacial adhesive component in the friction model
results in a modification of Amonton's law as
F='t,A + aW (3.85)
where F is the shear force, W the applied normal load, A the interfacial
contact area, 'to the interfacial shear strength and a the pressure coefficient,
the last two being material properties. The effective coefficient of friction can
be obtained by dividing equation (3.85) by the normal load to give
(3.86)
734 Particle mechanics
!lw = (3.87)
W
E*=-
3
4
(1-V---
E1
T 1-V~)
+- --
E2
Here v and E are the Poisson ratio and Young's modulus of elasticity,
respectively, and subscripts 1 and 2 refer to the particle and the vessel wall.
Equation (3.87) has been used successfully to predict the load dependence of
friction at very low loads (Tiiziin et ai., 1988) and has the advantage of
relating the frictional force to measurable material and geometric quantities.
Adams et ai. (1987) show that it can be regarded as equivalent to the
empirical equation
F=kW' (3.88)
where the parameters k and n are load-dependent. It should be noted that the
load index, n, can vary between 213 and 1 even for a Hertzian (elastic) contact
when surface forces are taken into account.
The foregoing analysis applies to perfectly smooth particles; in practice it is
commonly the case that real particles show surface asperities. This case has
been considered in detail by Archard (1957) and his approach is summarised
by Adams et al. (1987). Archard's model is shown schematically in the second
column of Figure 3.23. The important feature is that the number of contacts
between the rough particle and the plane surface increases as the load
increases, so that the mean pressure at the asperity contacts is insensitive to
load. Hence, for a rough particle
Impact and rebound of particles 735
~ (Hertz) (Archard)
SURFACE SMOOTH ROUGH
EXAMPLES GLASS MUSTARD SEED
POLYETHYLENE
FRICTION LOAD-DEPENDENT NOT LOAD-
COEFFICIENT DEPENDENT
(3.89)
where P is a mean asperity contact stress, which is close to the flow stress of
the softer of the two contacting bodies, and a is a constant. Hence the
frictional force for rough particles is directly proportional to the load.
Vi
+--
0 Approach
0 Rebound
- V,
.......... ...... -
Initial surface
potential
-_ .. _----_ .. Rebound surface
potential
(3.90)
Upon rebound, the particle must exchange kinetic energy for potential energy
as it 'climbs out' of the potential well. The final kinetic energy of the particle,
beyond the influence of the surface potential, is
The particle will be captured when all the kinetic energy after rebound is used
up in trying to escape from the rebound surface potential; the limiting value
of Vi for which capture occurs is therefore given by
W2mVt 2 + Ei)e 2 - Er = 0 (3.92)
or
(3.93)
Impact and rebound of particles 137
vt = [AD 6am
1- e2 ] v,
e2
=~
d
[_A_
JtaQp
1- e2 ] v,
e2
(3.97)
+a b c d
--->o..G..-<-cx=_o-->G<---/o-->...8.-<-_8<--+:
cx",----
Separation Distance of
point d approach cx
Rebound
Figure 3.25 Schematic representation of the impact of a particle, including the effects
of adhesion (Reed, 1987).
Via> Vy = 2Jt]Z[
[ 3K
- _2 ]Y.> y'h (3.98)
5~?1
(3.99)
and
K=-
4 [1---.VI+1-- -
V~ ] -1
3 E1 Ez
(3.100)
....
c
o
·iii
<D ~
..
€';" 10- 1
_ E
co
CI)
o~
~~
EO
._ 0
(ij Qi 10-2
.. .
o >
."
<:5
.
10·3L---------~.----~--------~~~
1 10
Particle radius (Jlm)
Figure 3.26 Critical impact adhesion velocities for glass particles impacting onto steel
surfaces: A, stainless steel (elastic); B, mild steel (elastic-plastic). The broken line
shows the terminal velocity of the glass particles in air (Reed, 1987).
740 Particle mechanics
where
and
Co = (E2/Q2)'/2
Using the properties of glass, Aw = 0.025 so that for V io = 1 m S-l, e = 0.99.
dV
m =-=--- (3.103)
da V a
Impact and rebound of particles 141
Figure 3.27 A spherical particle approaching a plane surface covered by a thin liquid
layer.
but in this case we must also take into account the resistance of the liquid
layer after rebound from the surface, because the pressures generated in the
liquid are not low enough to cause cavitation. At the point of first contact
with the asperities, from equation (3.104)
(3.108)
Therefore, just after rebound
Vro = -eVio = -eVo [1 -In(aola2)ISt] (3.109)
Now integrating equation (3.103) with V = Vro at a = a2 and - V> 0 at a = ao
gives
= 2In(aola2)
if e = 1.
Barnocky and Davis (1988) obtained results for impact of particles on
smooth and rough surfaces, for a range of liquid layer thicknesses and
viscosities, showing good agreement with equations (3.106) and (3.110).
This approach to the prediction of the outcome of collision events opens
up the possibility of achieving a better understanding of 'capture' of particles
in some forms of filtration (see section 7.3.2.2) and in the various processes
which are used to manufacture agglomerated particles (see section 6.2.7).
3.5 NOMENCLATURE
A Hamaker constant (J), contact area (m 2)
a Separation between bodies (m); contact radius (m)
a' Dimensionless separation, aiR
d Particle diameter (m)
E Elastic modulus (Pa)
Ej,Er Surface potential energies (J)
E* Elastic modulus of assembly (Pa)
e Coefficient of restitution
F Force (N)
Gc Critical strain energy release rate (J m-2)
k Total curvature (m-I )
m Mass (kg)
n Index in pair potential; particle-particle contact density (m-2)
p Pressure (Pa)
Q Electric charge (C)
R Radius (m)
r Separation between molecules (m); radius (m)
St Stokes number
References 143
u Energy (J)
u Energy to separate two particles to infinity (J)
V Velocity (m S-1)
Liquid molar volume (m3 kmol-1)
Pair potential (J); normal load (N)
Pull-off force (N)
Greek letters
a Pressure coefficient
f3 Half-angle e)
y Surface energy (J m-2 )
r Adhesion energy (J m-2 )
o Displacement (m)
E Void fraction; permittivity (F m-1)
!l Viscosity (Pa s)
lAw Wall friction coefficient
Q Number density of molecules (m-3 )
o Stress (Pa)
v Poisson ratio
A Parameter in equation (3.100)
<j> Compliance correction in equation (3.83)
'to Shear stress
REFERENCES
Abdel-Ghani, M., Petrie, J.G., Seville, J.P.K., Clift, R. and Adams, M.J. (1991)
Powder Techno/. 65, 113.
Adams, M.J. and Perchard, V. (1985) 1. Chern. E. Symp. Ser. 91, 147.
Adams, M.J. and Edmondson, B. (1987) in Tribology in Particulate Technology (B.J.
Briscoe and M.J. Adams, eds), Adam Hilger, Bristol and Philadelphia.
Adams, M.J., Briscoe, B.J. and Pope, L. (1987) in Tribology in Particulate
Technology (B.J. Briscoe and M.J. Adams, eds) , Adam Hilger, Bristol and
Philadelphia.
Adams, M.J., Williams, D. and Williams, J.G. (1989) J. Mater. Sci. 24, 1772.
Archard, J.F. (1957) Proc. Roy. Soc. (London) A243, 190.
Barnocky, G. and Davis, R.H. (1988) Phys. Fluids 31, 1324.
Bowden, F.B. and Tabor, D. (1955) Friction and Lubrication of Solids, Cambridge
University Press, Cambridge.
Briscoe, B.J. and Adams, M.J. (eds) (1987) Tribology in Particulate Technology,
Adam Hilger, Bristol and Philadelphia.
Briscoe, B.J., Pope, L. and Adams, M.J. (1984) Powder Technol. 37, 169.
Briscoe, B.J., Fernando, M.S.D. and Smith, A.C. (1985) J. Phys. D: Appl. Phys. 18,
1069.
Cameron, A. (1981) Basic Lubrication Theory, Ellis Horwood, Chichester.
Capes, C.B. (1980) Particle Size Enlargement, Elsevier, Amsterdam.
Coughlin, R.W., Elbirli, B. and Vergara-Edwards, L. (1982) J. Colloid Interface Sci.
87,18.
144 Particle mechanics
Coury, J.R., Raper, J.A., Guang, D. and Clift, R. (1991) Trans. l. Chern. E. 69(B),
97.
Cross, N.L. and Picknett, R.G. (1963) International Conference on the Mechanism of
Corrosion by Fuel Impurities, Marchwood Engineering Laboratories, Butterworths,
London, p. 383.
Cumberland, D.J. and Crawford, R.J. (1987) The Packing of Particles, Elsevier,
Amsterdam.
Dahneke, B. (1971) J. Colloid Interface Sci. 37,342.
Drescher, A. and de Josselin de Jong, G. (1972) J. Mech. Phys. Solids 20, 337.
Fisher, R.A. (1926) J. Agric. Sci. 16, 492.
Gordon, J.E. (1976) The New Science of Strong Materials, 2nd edn., Penguin,
London.
Griffith, A.A. (1920) Phil. Trans. Roy. Soc. (London) A221, 163.
Hertz, H. (1882) in Miscellaneous Papers by H. Hertz (1896) (Jones and Schott, eds),
Macmillan, London.
Hinds, W.C. (1982) Aerosol Technology, Wiley-Interscience, New York.
Israelachvili, J.N. (1991) Intermolecular and Surface Forces, 2nd edn., Academic
Press, London.
Johnson, K.L. (1985) Contact Mechanics, Cambridge University Press, Cambridge.
Johnson, K.L., Kendall, K. and Roberts, A.D. (1971) Proc. Roy. Soc. (London)
A324,301.
Kendall, K. (1987) in Tribology in Particulate Technology (B.J. Briscoe and M.J.
Adams, eds), Adam Hilger, Bristol and Philadelphia, 110.
Kendall, K., Alford, N.McN. and Birchall, J.D. (1986) Inst. Ceram. Proc. Special
Ceramics No.8, Institute of Ceramics, Stoke on Trent, p. 255.
Kendall, K., Alford, N.McN. and Birchall, J.D. (1987) Proc. Roy. Soc. (London)
A412,269.
Lian, G., Thornton, C. and Adams, M.J. (1993) J. Colloid Interface Sci. 161,138.
Mason, G. and Clark, W.C. (1965) Chern. Engng. Sci. 20, 859.
Mullier, M.A., Seville, J.P.K. and Adams, M.J. (1987) Chern. Engng. Sci. 42,667.
Orr, F.M., Scriven, L.E. and Rivas, A.P. (1975) J. Fluid Mech. 67,723.
Palmer, K.N. (1990) in Powder Technology (M.J. Rhodes, ed.), Wiley, Chichester.
Parker, A.P. (1981) The Mechanics of Fracture and Fatigue, Spon, London.
Picknett, R.G. (1969) J. Colloid Interface Sci. 29, 173.
Pietsch, W.B. (1968) Nature 217,736.
Plati, E. and Williams, J.G. (1975) Polym. Eng. Sci. 15, 470.
Princen, H.M. (1968) 1. Colloid Interface Sci. 26, 249.
Reed, J. (1987) in Tribology in Particulate Technology (B.J. Briscoe and M.J. Adams,
eds), Adam Hilger, Bristol and Philadelphia, 123.
Rumpf, H. (1962) in Agglomeration (W.A. Knepper, ed.), Interscience, New York,
p.379.
Shaw, D.J. (1992) Colloid and Surface Chemistry, 4th edn., Butterworths, London.
Simons, S.J.R., Seville, J.P.K. and Adams, M.J. (1994) Chern. Eng. Sci. 49,2331.
Tabor, D. (1955) Proc. Roy. Soc. (London) A229, 198.
Tabor, D. (1987) in Tribology in Particulate Technology (B.J. Briscoe and M.J.
Adams, eds), Adam Hilger, Bristol and Philadelphia, 206.
Thornton, C. and Barnes, D.J. (1986) Acta Mechanica 64, 45.
Timoshenko, S. and Goodier, J.N. (1951) Theory of Elasticity, McGraw Hill, New
York.
Troadec, J.D., Bideau, D. and Dodds, J.A. (1991) Powder Technol. 65, 147.
Tiiziin, U., Adams, M.J. and Briscoe, B.J. (1988) Chern. Engng. Sci. 43, 1083.
Visser, J. (1989) Powder Technol. 58, 1.
Williams, J.G. (1984) Fracture Mechanics of Polymers, Ellis Horwood, Chichester.
4
Characterisation of bulk
mechanical properties
4.1 INTRODUCTION
The frictional behaviour of powder materials in all engineering applications is
described by two independent parameters: the coefficient of internal friction
and the coefficient of wall friction. The former determines the stress
distribution within a bed of powder undergoing strain and the latter describes
the magnitude of the stresses between the material bed and the walls of its
container.
Accurate knowledge of these two friction coefficients is essential to the
design of all bulk solids storage, handling and transportation equipment.
Both the structural strength of such equipment and the stability of the
material flow within them rely heavily on the accuracy of the values of the
friction coefficients used in design calculations (British Materials Handling
Board, 1985).
Values of friction coefficients are in general determined empirically, and
these empirical values are then used in all subsequent calculations to predict
stress and flow fields. The accuracy of such predictions will be affected
significantly by the bias and errors associated with a particular measuring
technique.
There have been few attempts to predict analytically the coefficients of
internal and/or wall friction. This scarcity is proof of the difficulty of such a
~
Lever
arm
v~ Top ring
Shear Base
Retaining force 5
_~====d:6d..pln ~ F
F I n put
orced
t rans Vibration
ucer
I
Sample
Figure 4.1 Shear testing equipment for granular solids: (a) Jenike shear box (from
Jenike, 1961); (b) annual shear cell (from Carr and Walker, 1968); (c) vibrating shear
cell (from Roberts et at., 1984).
Empirical measurements of coefficients of friction 147
materials at rest, and hence they can only be taken as a likely indication of the
'static' value of the internal angle of friction.
The angle of wall friction, <Pw, however, can be determined 'quasi-
statically' by accommodating a layer of the bulk material to be tested on an
inclined surface and measuring the angle of inclination of the surface at the
point of initiation of particle sliding (Augenstein and Hogg, 1974; Nedderman
and Laohakul, 1980), as shown in Figure 4.2(c).
A more direct determination of the wall friction is made possible by the
simultaneous measurement of the normal and shear loads at the walls of a
container using twin-axis cells (Tiiziin and Nedderman, 1983, 1985). In this
case, the wall friction corresponding to the static material fill can be measured
separately and then compared with the dynamic values resulting from a
flowing bed (Tiiziin et al., 1986).
However, such in situ measurements of the static and dynamic values of the
internal angle of friction are not possible. The shear cell tests described above
can at best be called quasi-static as the so-called yield loci obtained refer to
the stress distribution at incipient failure, which is analogous to the initiation
of motion in a static material bed. Unlike the dynamic wall friction
measurements described above, there is as yet no direct method of
measurement of the effect of the state of flow on the magnitude of the
internal friction. Such information would be invaluable to the successful
coupling of the stress and strain rate fields in particulate materials, as will be
demonstrated below.
Dead
-~
bead
~zone
))
Figure 4.2 Indirect measurements of angles of friction of granular solids: (a) angle of
repose in a flat-bottomed bin; (b) angle of repose of a static heap; (c) angle of friction
of an inclined wall.
10r----------------------.
8 4
~6
E
z
~4
\-'
2 4 6 8 10 12 14 2 10
(a) cr (kN m- 2) (b)
0.5~
I
0.4~
\
.'
0.3~~,
..+.'-.
.J. ",. -?.........
'.!'I.
..•.• - ' -
+'_ a=0.14
0.2 f-... + + ··'"-f.·····-+~-···T~-;;:0.20
... ... ...
...... ... ...... A ... A a = 0.15
0.1 I-
OL--~I~I---L-I-L-I-~I~I---L-I-
o 2 4 6 8 10 12 14
(e) crw (kN m- 2 )
Figure 4.3 Angles of friction of cohesionless materials. (a) Effective internal yield
loci of three cohesionless materials. Data from Tiiziin et at. (1986) with values of mean
particle size d (!im): 0, mustard seed (2290); *, table salt (360); 6, fine sand (160).
(b) Yield loci of cohesionless solids on various wall surfaces: 0, mustard seed on
aluminium alloy (Tiiziin et at., 1986); ., black coal on stainless steel (Roberts et at.,
1984); 6, polymer powder on glass (Cooker, 1980); ., polymer powder on steel
(Cooker, 1980). (c) Variation of wall angle of friction with normal stress on a smooth
wall (from Tiiziin et at., 1988). The curves are derived from theory using equation
(4.6). The data and curves are: ,;., full curve, mustard seeds; +, double broken curve,
glass ballotini; . , broken curve, polyethylene pellets.
Angles of friction of cohesive materials 151
The variation of the wall friction angle during filling and discharge of bulk
solid containers is best measured by the use of the double-axis load cells
described above. Tiiziin et al. (1988) have shown that with Coulomb-type
materials, the magnitude of the wall friction can in certain cases change
considerably once flow is initiated, and during flow, significant variation of
the wall friction angle is observed along the height of the container. Figure
4.3(c) shows the experimental values of the wall friction with mustard seeds,
glass ballotini and polyethylene beads as functions of the normal load acting
on the smooth aluminium side wall of a plane-strain silo. Clearly, the wall
friction behaviour is markedly different in each case; even though all three
materials behave as Coulomb-type, differences in the particle surface
roughness are shown to give rise to different functional dependence of !!w on
the wall stress Ow (see Tiiziin et al., 1988).
Conventionally, the continuum mechanics approach is used to explain the
variation of the internal angle of friction with consolidating load at the point
of initiation of failure within the material. The continuum mechanics
approach is therefore also the underlying theory of the shear-testing devices
described in the previous section. In practice, however, this approach cannot
fully explain the frictional behaviour of Coulomb materials during sustained
flow. Its apparent deficiencies result from neglect of effects of the key particle
properties such as those described above. Where the wall friction is
concerned, these effects can be investigated to a certain extent by carrying out
single-particle friction experiments. A description of one type of apparatus
for performing these experiments is given by Briscoe et al. (1973). The
apparatus is designed to slide a single particle against a well-defined surface
under the action of a given normal load. The theory of single contact friction
is summarised in section 3.3, while section 5.2 describes some recent attempts
at assembly models of friction.
8
.. '
a
(a) (b)
't
o a
(c)
Figure 4.4 Angles of friction of cohesive materials. (a) Incipient yield loci of a
polymer powder « 1400 !lm). (1) Ql = 380 kg m-3 , aLl = 0.928 kN m-2 ; (2) Q2 =
444 kg m-3 , aL2 = 3.32 kN m-2 ; (3) Q3 = 484 kg m-3 , aL3 = 8.101 kN m-2 • Here Q
is the initial bulk density and aL is the termination point of the incipient yield locus.
(b) Mohr-Coulomb failure of cohesive materials: IYL, incipient yield locus; EYL,
effective yield locus; CVL, critical voids line. (c) Termination locus of incipient yield
and consolidation surfaces: YL, yield line; CL, consolidation line; CVL, critical voids
line (= termination locus).
three loci have a non-zero intercept on the shear axis and all are distinctly
curved. An equation of the form
(4.7)
°
can be used to fit the data in Figure 4.4(a) for < 0L, i.e. for stress values
lower than the termination point of the yield locus. Here T is termed the
tension or tensile strength of the material and is given simply by the intercept
ofthe yield locus on the normal load axis. Equation (4.7) is well known as the
Warren-Spring equation in recognition of the pioneering work on granular
Angles of friction of cohesive materials 153
(JA (J
Dilation prior to
Consolidation prior ~Iastic failure
to plastic failure 1:2
I:
--.
1:4
Incipient failure
at £5
y y
(a) (b)
Figure 4.5 Flow initiation in a static bed of cohesive solids: (a) plastic failure of an
underconsolidated fill; (b) plastic failure of an overconsolidated fill.
appear to be cohesionless even though individual slip planes will occur within
the flow field given by the point of tangency with the individual incipient yield
loci; these points are illustrated in Figures 4.4 and 4.5.
The result of the critical state analysis, as put forward by Roscoe et al.
(1958), Roscoe (1970) and Schofield and Wroth (1967), is that it allows us to
define the frictional behaviour of a cohesive powder in steady flow as being
independent of the value of the interstitial voidage. However, the material
properties cohesion, c, and tensile strength, T, are both strong functions of
the voidage and are proportional to the termination point, 0L, of the
individual yield loci, as seen in Figure 4.4(a):
c = feE,
ad (4.9a)
T= feE, ad (4.9b)
Roberts et al. (see under Roberts, 1984) have developed a vibrating shear cell
in which the material yield loci can be obtained as a function of the amplitude
and frequency of the vibration applied to the cell; the device is illustrated in
Figure 4.1(c). This work is significant in that it demonstrates a 'dynamic' way
of altering and controlling the interstitial void age distribution in a bed of
material, thereby significantly altering the material yield strength under
shear. Their experimental results show a strong correlation between the
cohesive strength, c, of the material samples and the frequency of the applied
vibration.
The void age distribution that results from the filling of the material into a
container will decide the magnitude of the cohesive forces at different
sections in the static fill. For failure under compression, flow will only occur if
the magnitude of the compressive stresses acting on a failure plane are larger
than the cohesive strength of the material. Once flow is initiated, the critical
state analysis described above suggests that it can be sustained indefinitely
since the bulk will act like a cohesionless material at its critical state (see
Figure 4.4(b».
In dealing which cohesive materials, the continuum mechanical approach
based on the assumption of a critical voidage state suffers from two major
drawbacks:
10
--
0..
..:>C
14
Figure 4.6 Wall yield loci of cohesive solids from Roberts et al. (1984). The full
curves are for brown coal with 65% moisture content and the broken curves are for
black coal with 18.7% moisture content. The curves are for the following materials: A,
D, stainless steel; B, C, polyethylene.
Characterisation of 'f/owability' of cohesive powders 157
the wall yield locus of a cohesive material in the conventional shear cell
equipment described above, as the magnitude of the so-called adhesive (<Pw)
forces will depend strongly on the local voidage adjacent to the wall surface,
which is very difficult to measure. Figure 4.6 shows the experimental wall
yield loci of moist pulverised coals measured by Roberts et al. (1984), in
comparison with those of a Coulomb-type material given by equation (4.10)
above. Some of the experimental wall yield loci in Figure 4.6 are significantly
curved, indicating that the wall friction angle, <Pw, is a function of the normal
stress acting on the wall. This behaviour is unlike that described by equation
(4.10) which gives rise to a constant value of the wall friction angle over the
entire normal load range. In general, the Coulomb failure criterion given by
equation (4.10) fails to explain the wall frictional behaviour of a cohesive
powder material. Moreover, as the cohesive strength increases, as in very wet
materials for example, the adhesion term Cw in equation (4.10) can become
much more significant that the term <Jwtan<pw. Wet materials handled in
industry have a strong tendency to stick to the surfaces of storage vessels and
the mechanical conveying equipment, thereby causing numerous material
handling problems, as described by Bransby and Heywood (1984). Yet very
little is currently understood about the wall frictional behaviour of wet
powder beds. The effects of the variation of the values of the stress normal to
the wall and the bed voidage adjacent to the wall have been observed in
practical applications but are not as yet accounted for by a theoretical
analysis.
Active
state of Vertical
stress part
Discontinuity
Transition
Passive
state of Hopper
stress
wall
Virtual "-'"--Outlet
apex
(a)
Shear Stress 1:
-
C ['------!:----::------!c=_____
o (Ix. Normal Stress 0
(b)
therefore gives the yield strength of the material confined between two
retaining walls under a stress-free top surface, as shown in Figure 4.7(b).
The major principal compressive stress, ar, corresponding to the flowing
state of the material is obtained by drawing a Mohr's circle through the
termination point of the internal yield locus; see Figure 4.4. The procedure is
repeated for the family of yield loci obtained at different levels of
preconsolidation during the Jenike shear cell tests (refer to section 4.2
above).
a = Qbg Yr
X-1
[a _
rr
Qbg YR ]
X-1 .
[~] x
R
(4.13)
where the variables (X, Y) are functions of the hopper half-angle, a, and the
internal and wall angles of friction, and arr is the radial component of
compressive stress, as shown in Figure 4.7, which also definesj3, the direction
angle of the principal compressive stress
2m sin<l>e [sin(2j3 + a) ]
X= +1 (4.14)
1 - sin<l>e sina
critical
No
Flow
Flow
o
Figure 4.8 Jenike's flow/no-flow criterion.
aH(a)
B=--- (4.16)
Qbg
where the empirical function H( a) is provided in graphical form in Figure
4.9. The major limitation of this approach is that it often predicts a minimum
outlet width of the order of a metre or more depending upon the cohesive
strength of the powder. This then means that the gravity discharge of the
powder from such a wide opening is controlled by the design parameters of
the mechanical conveyor system placed underneath. While being very
effective in the design of large-scale storage installations, the lenike method
is of limited use when designing pneumatic dosing and flow metering units
often used in the chemical, pharmaceutical and food industries where quite
small discharge rates are required coupled with strict requirements for the
solids fraction discharged.
3.0
2.5 V
--
1 -I-"'"
.,./
...- - ~ ...;.--
...;.--
E..
n
H(a) 2.0
1.5
-- -I-"'"
':J -
10' 20· 30' 40' 50' 60'
(l'
CURVE i_Circular
2_ Square
3_ Rectangular
(L ~3B)
Figure 4.9 lenike's empirical function H(a) to determine minimum critical outlet
size for flow for different orifice shapes: B = orifice width, L = orifice length.
z
~x
(jyy
~'tyx
'tyz .c t . (jzz
(jyy
(jyy
't:--±
'txy
/
/ O(jxx 0 +Xo
(jxx ~ I (jxx + ax x (j--H
~(jxx
xx
Oy )-
"0+
~
'tXyV~
l 'tyx
(jyy
o
1,,0 >-
X
+
"0'" r4'
6"';;'
'C/o'\, >,
X \:)~I rt>
rt> (b)
6"'' ' +
;f
(a)
+ = 0 (4.19)
(4.20)
Equations (4.19) and (4.20) together define the state of stress equilibrium in
planar geometry.
,
I
I
,
I
I
I
:0:
0: :,7l
7~ Element of bulk solid
t,Ojl OJ
(a)
(b)
material and the equation of material failure is given by equation (4.4). The
construction in Figure 4.11 is often called the Mohr-Coulomb failure
criterion. This is the most popular theoretical model used to describe the
onset of plastic failure (i.e. flow) in granular solids. The model is essentially
two-dimensional and it does not take into account the influence of the
intermediate normal stress (Jzz) component in a three-dimensional element
(see Figure 4.10). In dealing with axially-symmetric geometries, the Mohr-
Coulomb failure criterion can be applied by assuming that
or
where (Jxx and Oyy are the principal normal stresses. This assumption is often
referred to as Harr and Von Karman's hypothesis; see for example
Nedderman (1992). More sophisticated failure criteria (e.g. Von Mises)
which take into account the variations in the intermediate principal stress and
the 'conical' nature of the yield surfaces in the three-dimensional flow have
been applied to predictions of granular flow fields in mass flow. Cleaver and
Description of bulk solids stress states 165
Figure 4.13 shows the Mohr's stress circles corresponding to 'active' and
'passive' states of failure. Note that in the 'active' state, the direction of the
major principal stress, Or, is near to the vertical (i.e. Oyy> oxx), while in the
'passive' state, Or is close to the horizontal (i.e. Oyy < oxx).
In practice, the filling of a bunker results in the bunker walls bulging
slightly outwards and consequently, the initial material fill is generally
associated with an active state of stress. However, the material in the
converging section of a bin-hopper system is forced to contract horizontally
and thus expand vertically during discharge. This implies that in such a
geometry, the direction of the major principal compressive stress is near to
the horizontal, suggesting a passive state of failure.
On the basis of the above arguments, a discontinuous change from one
stress state to the other is to be expected in the bunker during flow initiation.
The experiments of Handley and Perry (1965,1967), Perry and Jangda (1970)
and Rao and Venkateswarlu (1974, 1975) confirm the above hypothesis.
These workers were able to measure point values of vertical and horizontal
stress components at various points along and across a bunker by tracing the
signals of a radio-pill. Their results show that after the initial filling, the
material in the bunker is in the active state, i.e. the direction of the larger
stress component is nearly vertical. However, during steady flow, they found
that the material in the parallel-sided bin section remains in the active state
while in the converging hopper section, the major principal stress rotates to
(a) (c)
Passive
Active
the horizontal direction giving a passive stress field. Near the bottom corners
of the bin section, known as the 'transition corners' (see Figure 4.13), their
results also show 'peak stress' values due to transition from the 'active' to the
'passive' stress state, as predicted by lenike (1961, 1964), and later by Walters
(1973a, b).
CJ
b-slip line
Figure 4.14 The direction of the slip planes in cohesionless granular material at
incipient flow.
768 Characterisation of bulk mechanical properties
material 'slip planes' during flow in relation to the direction of the major and
minor principal stress planes. Unlike the experimental procedures described
above for measuring internal and wall friction angles of bulk materials, no
established experimental technique currently exists which can be used to
quantify the so-called 'angle of dilation'. It is this angle and its prediction
which holds the key to the definition of a characteristic flow rule for a given
granular material.
Om = W/A
t 0,
~
°2=°3
T """ ..,r
o
-- "-
(a)
t
(b)
Qbo = (4.22)
AHbo
where Wb and Hbo are the weight and the initial height of the bulk sample and
A is the cross-sectional area. The mean value of the bed void age is then given
by
Qbo
1--
= 1 - (4.23)
QsAHbo
The values of bed void age corresponding to different uniaxial consolidation
loads are measured by monitoring the reduction in the sample height Hb in
the cylinder, i.e.
(4.24)
which allows a plot of the mean bed voidage as a function of the uniaxial
consolidation load. Figure 4.13 shows some typical results obtained with
different granular materials. In general, over a relatively small range of
consolidation loads, a linear relationship of the form
(4.25)
can be used as a best fit to experimental data, where the empirical coefficient,
a, is a material property and Oc is the applied load divided by the cross-
sectional area. However, over a larger range of consolidation loads, a
logarithmic relationship of the form
(4.26)
is more appropriate. The distribution of the applied load across the bulk
material sample will undoubtedly be affected by the angle of the material top
surface as well as by the wall friction. To minimise these effects, it is
important to use cylinders with a height-to-diameter aspect ratio >2 and a
ratio of cylinder diameter to average particle size >30.
Equation (4.20) approaches a rather 'ill-defined' asymptotic limit at the
large values of the applied load, so that it can be rewritten as
1
--- = Vo - AlnP (4.27)
1- E
where Vo is an empirical constant and the coefficient A is often referred to as
the compressibility coefficient in the soil mechanics literature (Roscoe et at.,
1958). An analogous expression in terms of the form is also possible
770 Characterisation of bulk mechanical properties
1 Vo
- =- - (A/Qs)lnP (4.28)
Qb Qs
where Qb and Qs are the bulk and solid densities, respectively, and P is the
applied pressure. With cohesionless bulk materials such as glass ballotini,
coarse sand and table salt (see Figure 4.16), typical values of the compress-
ibility coefficient are reported to be in the range 0.1-0.01 over a range of
consolidation loads up to 10 bar; see for example Crewdson et al. (1977) and
Thorpe (1992). With fine powders «100 !-lm) of irregular shape, and bulk
materials which exhibit multimodal size distributions (sauce mix and icing
sugar in Figure 4.16), the compressibility of the bed of bulk material is much
higher; as much as 30-40% change in bulk density is obtained at consolidation
pressures as small as 0.02-0.05 bar. With cohesionless materials, the change
in bulk density under such small pressures is often <1 or 2%. Hence with
cohesionless materials, the coefficients of friction measured in the shear cell
experiments described earlier will be independent of the applied consolidating
pressure. When the bulk material is significantly compressible, the values of
bulk cohesion and of the angles of bulk friction are strongly affected by the
degree of consolidation.
The compressibility of a bulk material bed also affects its permeability to
C
0
U
jg
en
:!2
(5
en
c
nl
Q)
:::i:
Figure 4.16 Typical compressibility curves for various granular materials. 1. Table
salt, granulated sugar, fine sand (100 !-tm < d < 500 !-tm). 2. Sauce mix (d < 50 !-tm). 3.
Flour, corn flour, silica (d < 20 !-tm). 4. Icing sugar (d < 50 !-tm).
Nomenclature 171
4.11 NOMENCLATURE
A Cross-sectional area
B Orifice width, critical outlet width
C Cohesion
Cw Adhesion
D Diameter
d Particle diameter
Hb Initial height of the bulk
H (a) lenike's empirical function
L Length
MYL Material yield locus
P Pressure = W/A
r Radial distance
T Tension/tensile strength of the material
W Normal load
Wb Weight of the bulk
X, Y Empirical constants
X o, Yo, Zo Components of gravity in 3D space
Greek letters
a Hopper half-angle
j3 Angle defining principal compressive stress direction, pressure
coefficient
y Weight density = Qg
E Interstitial voidage
Voidage corresponding to zero consolidating stress
Compressibility coefficient
Coefficient of internal friction
Coefficient of wall friction
Empirical constant
Q Bulk density
Qb Initial bulk density
Qs Solid density
a Normal stress
OC Unconfined yield stress
0L, 'tL Termination point of the incipient yield locus
am Mean compressive stress
Ow Normal stress at wall
172 Characterisation of bulk mechanical properties
REFERENCES
Augenstein, D.A. and Hogg, R. (1974) Powder Technol. 10,43.
Bransby, P.L. and Heywood, N.J. (1984) Needs for research on the handling of wet
particulate solids. Warren-Spring Laboratories, UK Department of Trade and
Industry (CHISA '94 Prague, Czechoslovakia).
Briscoe, B.J., Scruton, B. and Willis, F. R. (1973) Proc. Roy. Soc. (London) A333,
99.
Briscoe, B.J., Pope, L. and Adams, M.J. (1984) Powder Technol. 37, 169.
British Materials Handling Board (1985) Draft Code of Practice for the Design of
Silos, Bins, Bunkers and Hoppers.
Carr, J.F. and Walker, D.M. (1968) Powder Technol. 1,370.
Cleaver, J.A.S. and Nedderman, R.M. (1993) Chem. Engng. Sci. 48, 3693.
Cooker, B. (1980) PhD Thesis. Cambridge University.
Coulomb, C.A. (1776) Memoires de Mathematique de I'Academie Royal des Sciences,
vol. 7, Paris, p. 343.
Crewdson, B.J., Ormond, A.L. and Nedderman, R.M. (1977) Powder Technol.
16, 197.
Drescher, A. (1991) Analytical Methods in Bin-Load Analysis, Elsevier Science,
Amsterdam.
Enstad, G. (1975) Chem. Engng. Sci. 30, 1273.
Haaker, G. and Rademacher, F.J.C. (1982) in European Symposium on Storage and
Flow of Particulate Solids, EFCE, Braunschweig, Germany.
Handley, M.F. and Perry, M.G. (1965) Rheologica Acta 4, 225.
Handley, M.F. and Perry, M.G. (1967) Powder Technol.l, 245.
Hvorslev, M.J. (1937) Ingenior Vidensk Skr. 45.
Jenike, A.W. (1961) Gravity Flow of Bulk Solids, Utah Engineering Experimental
Station, University of Utah, Bulletin 108.
Jenike, A.W. (1964) Storage and Flow of Solids, Utah Engineering Experimental
Station, University of Utah, Bulletin 123.
Jenike, A.W. (1967) Powder Technol. 1,237.
Jenike, A.W., Johanson, J.R. and Carson, J.W. (1973) Trans ASME. B95, 6.
Nedderman, R.M. (1978) Powder Techno!. 19, 287
Nedderman, R.M. (1992) Statics and Kinematics of Granular Materials, Cambridge
University Press, Cambridge.
Nedderman, R.M. and Laohakul, C. (1980) Powder Technol. 25, 91.
Perry, M.G. and Janda, H.A.S. (1970) Powder Techno!. 4,89.
Rankine, C. (1857) Phil. Trans. Royal Soc. (London), 147,9.
Rao, V.L. and Venkateswarlu, D. (1974) Powder Technol. 10, 143.
Rao, V.L. and Venkateswarlu, D. (1975) Powder Technol. 11, 133.
Roberts, A.W. (1984) in Handbook of Powder Science and Technology (M.E. Fayed
and L. Otten, eds), Chapter 6, Van Nostrand Reinhold, New York.
Roberts, A.W. Ooms, M. and Scott, O.J. (1984) Trans. Aust. Institute Mech. Engrs
C(355) , 123.
References 773
5.1 INTRODUCTION
Study of the mechanical behaviour of an assembly of solid particles in contact
with each other or with a wall surface is essential to a significant number of
process engineering applications involving bulk solids. The effects of single
particle properties on the bulk mechanical behaviour of granular materials is
an area of significant interest to process and materials engineers which has,
however, received little attention in traditional civil engineering and soil
mechanics literature (see, for example, Jenike, 1964; Schofield and Wroth,
1967). Historically, this is a direct consequence of the overwhelming
popularity of a continuum mechanics approach in mathematical modelling of
the granular flow and stress fields (Spencer, 1964; Pariseau, 1969a,b). In civil
and mechanical engineering applications, the internal and wall friction
coefficients and cohesion are traditionally measured as 'bulk' properties in
planar shear test equipment such as the apparatus described in Chapter 4.
The past two decades have also seen the development of more sophisticated
bi-axial and tri-axial bulk testers in which it is possible to exercise more
accurate control of the stress distribution and the interstitial void age during
both consolidation and shear of the material samples. However, the data
from such equipment are more difficult to generate and invariably require
more complex theoretical analysis than the results of the simple planar shear
cell tests. The major shortcoming of the bulk mechanical failure tests is that
they cannot be used to explain the effects of the single particle frictional
properties on bulk flow.
tThis chapter was written in collaboration with Dr Paul Langston, Department of Chemical
Engineering, Nottingham University, UK.
WALL WALL
• ••
•••••
•
•
•••••
•••••
• ••
.....
.:.:. ~
••
•••••
S>2R
Load Cell
Active Surface
Figure 5.1 Packing at the silo wall. (a) Close packed (static); (b) dilated (flowing).
contacts on the wall surface. It is possible to define the interstitial void space
at a distance R from the flat wall plane in terms of S
mrcR2
E=l---- (5.1)
SZ
where m = 1 for square packing and m = 1.155 for hexagonal close packing of
touching spheres, i.e. S = 2R. For the dilated assembly, S > 2R and the
voidage will scale as
1- El S~
----- (5.2)
l- E2 •• Si
where the subscripts 1 and 2 refer to the different states of interstitial voidage,
as shown in Figure 5.1(b). The maximum value of the flowing voidage
attainable by a bed of coarse spheres is given by the voidage at minimum
fluidisation, Emf (see section 6.3). Measurements of E made in two-
dimensional beds by numerical integration of the particle velocity profiles
next to a flat transparent wall seldom result in values in excess of 0.45 (see,
for example, Laohakul, 1978, and Tiiziin and Nedderman, 1979). Substitu-
tion of this maximum value for E in equation (5.1) yields a maximum value of
S = 2.4R for square packing and S = 2.6R for hexagonal packing. In general,
factors like the particle shape and the width of the size distribution are likely
to influence the limiting values of the voidage obtained. Furthermore, when
flow is initiated in a static bed, there is no reason to believe that the packing
geometry will remain unaltered. However, it is difficult to envisage
circumstances which will give S > 3R.
176 Assembly mechanics
The number of points of contact with the surface varies inversely with the
square of the spacing, S. Hence, for an area of surface A * (the area of a load
cell for stress measurement, for example)
N=A*/S2 (5.3)
Therefore, for a closely-packed bed N = A */4R2.
We can now modify the single particle friction equation resulting from the
adhesion model of friction for elastic point contacts (see section 3.3) to
include the effect of the number density of particles in contact with the silo
wall.
For each contact point, say
of= :rt'to(E*R) 'hOWl) + aoW (5.4)
For N point contacts, F = NoF and W = NoW and substituting for N from
equation (5.3) gives
(5.5)
Thus
E*R )2/1 1
( -S- - '11 +a (5.6)
~
- - Mustard
seeds
- - - - - - Polythene
beads
.
.
2
-- 3
0
I
...
··
4 ;.
·
Static Static Flowing
50 6
0
Glass Ballotini
--
0
I
H 3
4
Static Flowing
5
60
-t 12
E = Eo - C In [:0 J (5.8)
will predict dilation as the vertical loads during flow, 0, are significantly less
than those given by the static bed values 0 0 • Equation (5.8) is very similar in
form to equation (4.26) except here the reference state (Eo, 0 0 ) is well
defined. Since 0/0 0 < 1 then E > Eo where Eo is the value appropriate for
a static bed of closely-packed spheres. Here C is an empirical coefficient
whose values lie in the range 0.1-0.01. A similar expression can be used to
explain the variation of voidage with depth inside the silo. However, in this
case, the experimental stress profiles presented in Figure 5.2 indicate an
almost negligible variation of bed voidage with height except (i) immediately
below the top surface and (ii) close to the silo outlet where the stress levels
are found to change significantly with depth. This is typical of 'cohesionless'
materials whose internal failure loci remain independent of the interstitial
voidage. In line with Janssen's analysis described in Chapter 8, the
relationship between the vertical stress (Oyy)w at the wall and its normal
component is given by
(5.9)
where Kw is known as Janssen's pressure coefficient at the wall plane and can
be calculated for the active, K wA , and passive, K wp , stress states, as described
in detail in section 8.5. We can therefore rewrite equation (5.8) as
(5.10)
where 0wo is the static bed wall stress and Ow is the normal stress measured
during flow. For most cohesionless materials, the internal angle of friction is
about 30° and this would give values of the order of 0.1 for the ratio KWA/ Kwp.
It is conventional in soil consolidation tests to relate the mean value of the
The distinct element method 779
bulk void age to the mean isotropic stress so that effects due to anisotropic
loading could be averaged over the sample volume (see, for example,
Schofield and Wroth, 1967 and Roscoe et al., 1958). In equations (5.8) and
(5.10) above, the vertical component of the wall stress (Oyy)w is used rather
than the mean stress Om, where Om = 0.5(1 + Kw)ow' If the mean isotropic
stress is used then the term KWA1Kwp in equation (5.9) is replaced by (1 +
KwA/1 + Kwp) which has now a value of about 0.33 for the same material.
Equation (5.9) can be rewritten as
C = Co + C' - C In ( Ow ) (5.11)
Owo
where the new empirical constant C' is decided by the assumption about the
controlling stress component. If vertical stress is assumed to be controlling
then C' is about 0.23 with C lying close to the limiting value of 0.1 (refer to
equation (5.8». If however, the isotropic mean stress is controlling then C'
= 0.11. The experimental wall stress profiles presented in Figure 5.2 indicate
a three- to fourfold increase in the wall normal load due to flow initiation in a
static bed. Under these circumstances, equation (5.10) will predict c = co,
i.e. negligible or no dilation at the bed wall interface if the voidage is assumed
to be controlled by the mean isotropic stress at the wall. If, on the other hand,
the vertical stress component at the wall is assumed to be controlling then c
= 1.5-2.0co with a three- to fourfold increase in the wall normal load (see
Figure 5.2). It is well known that a dilated shear zone of some two to five
particles across is created when a bed of coarse spheres is allowed to flow next
to a vertical flat wall (see Chapter 8). The above results appear to suggest that
this observed dilation is predicted correctly if the vertical stress at the wall is
assumed to be controlling the interstitial void age at the bed wall interface.
However it must also be born in mind that equation (5.11) is an empirical
form with two adjustable parameters, C and C', which makes it difficult to
state unequivocally the precise nature of the relationship between the bulk
stresses and the interstitial voidage at the bed wall interface. Clearly, more.
systematic measurements of the interface voidage are required as a function
of the state of flow, and hence the value of interparticle separation distance S.
The above analysis is essentially two dimensional; a more complex analysis
is required to model assembly wall friction against a curvilinear surface in an
axially symmetric container. This would also require a more specific model of
the particle packing structure in three dimensions.
Wall Particles
(a)
(b)
(c)
(d)
Figure 5.3 Principle of the distinct element Newtonian method - see equations (5.11)
and (5.12). (a) Position at time t; (b) interaction forces; (c) net acceleration (including
gravity); (d) position at time t + nl'o.t (several time steps later).
182 Assembly mechanics
5.3.3 Methodology
The basic methodology of DE models is shown in Figure 5.3 for a dynamic
Newtonian analysis. First, the particle positions are recorded. From these the
particle interactions are determined and then the subsequent dynamics are
evaluated. The particle positions are determined at certain time intervals.
These are often regular (t, t + I:1t, t + 21:1t, etc.) but some analyses are event-
based, e.g. conditions are assumed constant until the next collision. Constant
or linearly varying conditions are assumed over the time interval, and hence
the time interval must be small enough to maintain sufficient accuracy. The
effect of the magnitude of the time step is discussed later.
In Newtonian dynamics, all the forces acting on each particle are
determined (e.g. friction at contacts, gravity) and then the net acceleration of
the particle is determined, both linear and angular. The positions and
orientations at the end of the next time step are then determined using an
explicit numerical time integration such as
(5.12)
(c)
Figure 5.6 Stochastic simulation of a vibrating assembly. Thc original pile (a) and
after a large shake (b) and after a small shake ( c). Insets show contact network of
highlighted cluster. Reproduced with permission from T.A.J. Duke, G.c. Barker and
A. Mehta. A Monte Carlo Study of Granular Relaxation, published by Institute of
Physics, 1990.
The distinct element method 185
N
~
NW~2*1 O-rNE
SW
~3
4
5,.. SE
'f
S
Request process Exchange process
(a) (b)
Gravity
action
W Kinetic
~
Interactions energy
W ! W
~ ~
Probability
distribution
Exchange
(d)
(c)
Figure 5.7 Use of Cellular Automata. (a) Orientation and numbering around a site.
(b) Request-exchange process governing grain displacements. (c) Transition rule
diagram. (d) Silo flow. Reproduced with permission from J. Martinez, S. Masson and
D. Deserable. Flow Patterns and Velocity Fields during Silo Discharge Simulation with
a Lattice Grain Model. PARTEe '95, published by Niirnbergmesse GmbH, D-90471
Nuremberg, 1995.
186 Assembly mechanics
L/. • • ~ .~
•
{j
•
• • ••
• •
• • • • •••
• • •
• • • • •
•• •
significant and the results are inaccurate or the model becomes unstable
(because particles overlap too much and they undergo increasingly large and
unrealistic accelerations). If the time step is too small the computer run time
will be unnecessarily long. Rounding errors can propagate over the time steps
causing errors for time steps which are too small. Another effect of a small
time step is spurious oscillations. Disturbances can be propagated faster than
the wave speed through the medium. In terms of the Rayleigh wave speed the
corresponding time step is given by (Johnson, 1985)
M = (nR/a)(Q/G)'/2 (5.14)
where R is the particle radius, a is a constant (0.9-0.95), Qis the particle
density and G is the shear modulus. Some researchers relate the time step to
the natural period of oscillation of the system. Thompson and Grest (1991)
state that the time step should be < 0.12Y(mlk) for particle mass m and
stiffness k. Zhang and Campbell (1992) propose that it should be
< 0.075Y(mlk). These estimates should give a good idea of the required time
step, but selection should be followed by trial runs with different I'3.t values to
check that the results are stable and reasonably invariant over this range of
I'3.t.
5.3.4 Forces
Chapter 3 summarises models for forces between particles. This section looks
at the application of these forces in DE models. Langston et al. (1995) have
developed DE granular models for non-cohesive particles using three
different functional forms for the normal interaction. These are illustrated in
Figure 5.9. The continuous interaction curve is widely used in molecular and
colloidal simulations and is fairly stable. The Hertz curve is based on the
contact mechanics of spheres (see section 3.2). The simple linear spring is the
simplest mathematical form and is fairly widely used. In deciding how to
model particle normal interaction, the analyst needs to consider the eventual
application of the simulation. For example, since the characteristics of flow in
silos are dominated by geometrical effects, the velocity and stress profiles
may be realistically modelled using softened interactions which allow some
particle 'overlap'. The precise details of the contact mechanics may not be
necessary, and since the time step would have to be smaller for a more
realistic particle stiffness, this enables more particles to be simulated. Bearing
in mind that most real granular systems are irregular and rough, it seems
unnecessary to model granular flow with ideal Hertzian contacts, although
the interaction must not be so soft as to allow excessive overlap, or the
analysis will be inaccurate. If the analyst is interested in modelling a few
particles in a shear cell to obtain 'continuum' constitutive properties, then it
may be necessary to model the contact physics more closely. Frictional forces
are usually modelled using the Mindlin equation (Mindlin and Deresiewiez,
188 Assembly mechanics
o Overlap, r
Figure 5.9 Modelling normal forces. Non-cohesive normal force options shown.
(a) Continuous interacti?n (molecular dynamics), Fn oc r-n; (b) Hertz (contact
mechanics), Fn oc (1 - r) 12; (c) Hooke (spring contact), Fn oc (1 - r).
1953), as shown in Figure 5.10. Some other analysts use a linear spring for
friction, usually with a cut-off at the maximum force given by the friction
coefficient. At the other end of the scale, Thornton and Yin (1991) model the
normal and tangential contact mechanics in considerable detail in simulations
of elastic sphere impact.
Contact damping is an important consideration in granular dynamics. This
is most generally modelled by treating the contacts as second-order systems as
shown in Figure 5.11. The damping force is most generally characterised by
relating it to the ratio of critical damping. This ratio is typically in the range
0.1 to 0.5. Some analysts work in terms of coefficient of restitution. Zhang
and Campbell (1992) relate the coefficient of restitution to critical damping
for a second-order system.
The simulations can be extended to model cohesive forces such as those
due to electrostatic charge, using the interaction force equations in Chapter 3.
Gravitational forces can be defined in DE models using constant vectors.
Fluid forces such as those encountered in pneumatic conveying, fluidised
beds or air-assisted hopper discharge present other problems which can be
handled in a number of ways. Tsuji et al. (1993) has modelled fluidised beds
using a DE model for the particle dynamics and a fixed (Euler) grid for the air
flow. Mass and momentum balances are carried out on the gas for each zone,
using a finite element technique. The drag force on each particle is calculated
using the Ergun equation (see section 2.3.2), the region around each particle
being modelled as a localised packed bed with a voidage dependent on the
positions of the neighbouring particles. The drag force effect on the gas
The distinct element method 189
dp
o
Figure 5.11 Modelling contact damping forces. Includes normal and tangential.
Retarding force, ex. relative velocity; mx +ex
+ kx = F; damping ratio ex. critical
damping; e ex. v' (mlk).
Figure 5.12 Simulation of a f1uidised bed. Shows particle velocity vectors for
different jet velocities. Reproduced with permission from Y. Tsuji, T. Kawaguchi and
T. Tanaka. Discrete Particle Simulation of Two-Dimensional Fluidized Bed, published
by Elsevier, 1993.
particle within the hopper the superficial air velocity is known. Then from the
local voidage at each particle, the interstitial air velocity is calculated and
hence the relative air-particle velocity is known. This gives a local drag force.
This is added to the other forces acting on the particle and its dynamic
response is modelled in the same way as in Figure 5.13. If the assumption of
radial flow is valid the FE method is not required. This approach is probably
reasonable for hopper flow and some pneumatic conveying applications
where spatial and temporal gradients of voidage are relatively small, but it
will not be sufficient for more complex applications such as fluidised beds.
An advantage of DE methods is that they can handle any container
geometry, for example a screw feeder, drum mixer or conveyor. These do not
present any geometrical difficulty (although their size may be a problem in
terms of the number of particles in the simulation). Particle---container
interactions can be handled in a manner similar to particle-particle contacts.
Frequently, the analysis is simplified by the presence of a fixed container
boundary.
(a)
(b)
Figure S.lS Two-dimensional non-disc simulation of shale flow. Shows position and
velocity vectors. Reproduced with permission from O. Walton. Particle-Dynamics of
Shear Flow, published by Elsevier, 1983.
The internal distributions of the normal and tangential bulk stresses have
been calculated by adopting a local averaging technique that works on the
values of the normal and tangential contact force vectors at various 'probe'
points within the particle assembly. The information thus generated is
analogous to the experimental measurements obtained by placing force
The distinct element method 795
transducer pills at different positions within the bulk (see, for example,
Handley and Perry, 1967). Normal and shear stresses for each probe point in
two dimensions are calculated from the following equations:
(5.15)
V 2DO y = L FnYnuYn + LFtYtuYn (5.16)
V 2D Oyx = L FtXtuYn + LFnxnuYn (5.17)
V 2D Oxy = L FtYtuxn + LFnYnuxn (5.18)
°
where V 2D is the sample volume, is the stress, Fis the force vector, Xn is the
contact x component vector (from particle-to-particle centre) and subscript u
indicates unit vector, n normal vector and t tangential vector. The three-
dimensional equations for the horizontal stress, On and the vertical stress an
are similar. This is essentially a numerical integration of the force vectors
within a small radius (e.g. two particle diameters) of the probe point to obtain
an average stress. Figure 5.19 shows an example of the internal normal stress
196 Assembly mechanics
CLLLL
lLL.LL.
LLLlL
LLLu
LLLL
LL'--Ll
5.4 NOMENCLATURE
A Area
c Damping constant
C Compressibility coefficient
D Elasticity constant; = E* in equation (3.88a)
E* Effective contact modulus (see equation (3.88a))
F Contact force
FD Air drag on particle
g Gravitational acceleration
G Shear modulus
h Height of packing
k Particle stiffness
K Janssen's coefficient
m Particle mass
N Number of particles
n Index in normal interaction curve
Appendix A5 197
jj.p Overpressure
R Particle radius
r Separation of particle centres
S Separation distance
Sn Normal displacement at contact
t Time
jj.t Time interval
Vo Superficial air velocity
Vj Interstitial air velocity
vp Particle velocity
V 2D Volume of 'sensor' search area in two-dimensional analysis
x Particle position
Greek letters
Pressure coefficient
Tangential displacement at contact, maximum before sliding
Local void age
Friction coefficient
Poisson's ratio
Q Density
o Normal stress
Subscripts
j For each particle
n Normal
r Radial
Tangential
u Unit vector
w wall
x,Y,Z Axes
N
F = 'toNbA + a I Wi (A7)
;= 1
but
N
I Wi = NbW
;=1
REFERENCES
Aizawa, T. (1993) Granular modeler of two and three dimensional powder dynamics
in powder forming and metallurgy, in The First Nisshin Engineering Particle
Technology International Seminar, January 18-20, Nisshin Engineering Co. Ltd.,
p.27.
Barker, G.C. and Mehta, A. (1992) Physical Review A 45(6), 3435.
Campbell, C.S. and Potapov, A.V. (1993) Recent applications of computer simulation
to granular systems, in The First Nisshin Engineering Particle Technology
International Seminar, January 18-20, Nisshin Engineering Co. Ltd., p. 27.
Cundall, P.A. and Strack, O.D.L. (1979) Geotechnique 29(1),47.
Duke, T.A.J., Barker, G.C. and Mehta, A. (1990) Europhys. Letters 13(1),19.
Fincham, D. (1992) Molecular Simulations 8, 165.
Haile, J.M. (1992) Molecular Dynamic Simulation, Wiley, New York.
Handley, M.F. and Perry, M.G. (1967) Powder Technol. 1,245.
Hogue, C. (1993) PhD Thesis, University of Cambridge.
Jenike, A.W. (1964) Storage and flow of solids, Bulletin 123, Utah Enging. Exper.
Station, University of Utah.
Johnson, K.L. (1985) Contact Mechanics, Cambridge University Press, Cambridge.
Kafui, K.D. and Thornton, C. (1993) in Powders and Grains '93 (C. Thornton, ed.),
Balkema, Rotterdam, p. 401.
Langston, P.A., Heyes, D.M. and Tiiziin, U. (1994) Chem. Engng. Sci. 49, 1259.
Langston, P.A., Heyes, D.M. and Tiiziin, U. (1995) Chem. Engng. Sci. 50,967.
Laohakul, C. (1978) Velocity distributions of flowing granular materials, PhD Thesis,
University of Cambridge.
Martinez, J., Masson, S. and Dcserablc, D. (1995) in Proceedings of the Third
European Symposium on Storage and Flow of Particulate Solids, Niirnberg, March
1995. Niirnbergmesse GmbH, p. 367.
Mindlin, R.D. and Deresiewicz, H. (1953) J. Appl. Mech. 20,327.
Mustoe, G.G.W. and DePoorter, G. (1993) in Powders and Grains '93 (C. Thornton,
ed.), Balkema, Rotterdam, p. 357.
Nedderman, R.M. (1992) Statics and Kinematics of Granular Materials, Cambridge
University Press, Cambridge.
Ning, z. (1995) PhD Thesis, University of Aston.
Ouwerkerk, C.E.D. (1991) Powder Technol. 65, 131.
Pariseau, W.G. (1969a) Gravity flows of ideally plastic materials through slots, Trans.
Am. Soc. Mech. Engings., J. Eng. Ind. May.
Pariseau, W.G. (1969b) Discontinuous velocity fields in gravity flows of granular
materials through slots. Powder Technol. 3, 218.
Roscoe, K.H., Schofield, A.N. and Wroth, c.P. (1958) On the yielding of soils.
Geotechnique 8, 22.
Satak, M. (1993) in Powders and Grains '93 (C. Thornton, ed.), Balkema,
Rotterdam, p. 3.
Schofield, A.N. and Wroth, c.P. (1967) Critical State Soil Mechanics, McGraw Hill,
London.
Smith, I.M. and Griffiths, D.V. (1982) Programming the Finite Element Method,
Wiley, Chichester.
Spencer, A.J.M. (1964) A theory of the kinematics of ideal soils under plain-strain
conditions. J. Mech. Phys. Solids 12, 337.
Taguchi, Y. (1993) Convection, turbulence and anomalous diffusion in vibrated bed,
in The First Nisshin Engineering Particle Technology International Seminar, January
18-20, Nisshin Engineering Co. Ltd., p. 84.
200 Assembly mechanics
Tanaka, T. and Tsuji, Y. (1993) Cluster formation in gas-solid flows, in The First
Nisshin Engineering Particle Technology International Seminar, January 18-20,
Nisshin Engineering Co. Ltd., p. 84.
Thompson, P.A. and Grest, G.S. (1991) Phys. Rev. Lett. 67(13), 1751.
Thornton, C. (ed.) (1993) Proceedings of the Second International Conference on
Micromechanics of Granular Media. Published as Powders and Grains '93,
Balkema, Rotterdam.
Thornton, C. and Yin, K.K. (1991) Impact of elastic spheres with and without
adhesion. Powder Technol. 65, 153.
Tsuji, Y., Kawaguchi, T. and Tanaka, T. (1993) Powder Technol. 77,79.
Tiiziin, U. and Nedderman, R.M. (1979) Experimental evidence supporting kinematic
model of the flow of granular media in the absence of air-drag. Powder Technol. 24,
257.
Walton, O. (1983) Particle-dynamics calculations of shear flow, in Micromechanics of
Granular Materials, Elsevier, Amsterdam.
Zhang, Y. and Campbell, C.S. (1992) J. Fluid Mech. 237,541.
6
Fluid-particle systems
The behaviour of mixtures of more than one phase is important in many areas
of process engineering, particularly where particulate material is transported
by a fluid, in reaction engineering or where energy or mass transfer occurs
across phase boundaries:
Compared with single phase systems, multi-phase systems present inherent
experimental and theoretical difficulties, for example in the need to
characterise particle shapes, describe size distributions and their evolution,
treat surface layer effects including contamination, consider various
regimes of motion, and deal with a host of secondary factors such as
Brownian motion and electrostatic charges. (Grace, 1986)
Moreover, structures can develop in the flow with their own characteristic
length scales, different from those of either the particles or the equipment,
and this often presents further complications (Clift, 1996). The range of fluid-
particle contacting devices in use in industry is very extensive. The subject
matter of this book is restricted largely to gas/solid systems, but even here the
variety of devices is extremely wide. One selection, again due to Grace
(1986), is presented in Table 6.1. No uniform view of the fluid and particle
processes in these devices yet exists. (Molerus, 1993, gives one of the best
accounts of those currently available.) In this chapter, therefore, a small
selection of fluid-particle contacting devices is presented, illustrating the
applications of the theoretical developments presented in Chapters 1 to 5.
Solids
Slurry
preparation
Booster
Feed pump
(a)
pump
Solids
Conveying gas
(air)
(b)
substantial drop in the pressure of the conveying gas, but the energy
transferred to the particles is not recoverable. Therefore the 'entry loss' is
usually significant in a pneumatic system (see sections 2.4 and 6.1.3.1), and
may even be the dominant contribution to the overall system pressure drop.
Similarly, there is usually a significant pressure drop following a bend in a
pneumatic system, because the particles must be accelerated up to the
transport velocity in the new direction.
On the other hand, because the solids pass through the pumps in a
hydraulic system, while the relative inertia of the particles is in any case small,
any 'entry loss' can usually be neglected. Similarly, the pressure drop
associated with flow through bends and fittings can be estimated to acceptable
accuracy from standard results for single-phase fluids. Therefore, estimation
of pressure loss in hydraulic systems concentrates on the effect of particles on
the shear stresses between the pipe wall and the flowing mixture.
Flow regimes. In both hydraulic and pneumatic systems, two general types
of flow are possible (see section 2.4.1):
Lean phase: low particle concentration and relatively high velocity.
Hvdraulic and pneumatic conveying 205
The distinction between settling and non-settling slurries has already been
introduced. In a non-settling slurry, the particles are sufficiently fine and
Hydraulic and pneumatic conveying 207
concentrated that their tendency to settle out from the conveying liquid can
be ignored in designing the conveying system; an extreme example is a paste.
The slurry can then be treated as if it were a single-phase fluid. The pressure
drop for a non-settling slurry in a horizontal pipe then varies with mean
velocity as shown qualitatively by curve NS in Figure 6.2, with the gradient
discontinuity around point T corresponding to the transition from laminar
flow at lower velocities to turbulent flow at higher velocities.
Settling slurries, by contrast, contain particles sufficiently large or dense
that their tendency to settle out from the vehicle cannot be ignored. As a rule
of thumb, this case arises when the hindered settling velocity is greater than
about 1.5 mm S-l. For particles with the density of sand (2650 kg m-3 ) in
water, this means that the slurry will normally show settling behaviour if the
particles are larger than about 75 !-tm. The system characteristic for a settling
slurry flowing through a horizontal pipe at constant delivered concentration
(see below) typically takes the form of curve S in Figure 6.2. Point D
corresponds to the deposit velocity, V" below which the pipe contains a
stationary deposit of solids which have 'settled out' from the slurry. At
velocities above V., the solids are all in motion, but the pressure gradient
typically decreases with increasing mixture velocity, to pass through a
minimum and then to approach the pressure gradient for the vehicle alone at
high velocities.
Systematic design of a hydraulic conveying installation depends on
calculating system curves like those shown schematically in Figure 6.2. This
PRESSURE GRADIENT
NS
CARRIER LlaUID
Vs MEAN VELOCITY, Vm
Figure 6.2 Velocity/pressure drop characteristics for flow of slurries in pipes at fixed
delivered concentration (schematic): NS = non-settling slurry; S = settling slurry; T =
laminar/turbulent transition; Vs = deposit velocity_
208 F/uid~particfe systems
approach is developed in detail, with case studies, by Wilson et al. (1992 and
1996). An outline of their approach is given here.
Sp(Sm - 1)
Sm(Sp - 1)
( IlP) 1 (6.3)
if = - ilL f gQf
where (- IlP/ IlL)f is the pressure gradient along the pipe. For a slurry, two
measures of friction gradient are then possible depending on the density used
to convert pressure to head. In terms of the vehicle density
(6.4)
(6.5)
so that
(6.6)
Both im and jrn will be used here, to illustrate different points.
Even for relatively short-distance conveying, the dominant operating cost
can be the power required to drive the pumps. A simple measure of the
efficiency of a conveying system is then the specific energy consumption, i.e.
210 Fluid-particle systems
the energy consumed per unit mass of solids transported per unit distance
travelled. In terms of the friction gradient, i m , the specific energy consumption
IS
6.1.2.2 Slurry flow in a horizontal pipe. In terms of slurry head, jm, the
friction gradient for a non-settling slurry varies with concentration and
velocity as shown schematically in Figure 6.3. To a good approximation, the
friction gradient in turbulent flow is independent of slurry concentration (and
close to the value for the vehicle alone). In the laminar flow regime, to the left
of the turbulent flow curve, the slurry friction loss depends on solids
concentration. Non-settling slurries almost always show non-Newtonian
properties, most commonly with a significant yield stress and with shear-
thinning properties (i.e. mean viscosity decreasing with increasing shear
rate). However, some slurries can show dilatant behaviour (mean viscosity
increasing with shear rate), and this behaviour is often desirable for drilling
muds. The rheological properties of a slurry depend on the solids concentration
and size distribution, and on the interaction forces between the particles
which depend inter alia on the pH and ionic strength of the carrier liquid and
on trace concentrations of surface active agents. Because of the complexity,
jrn
TURBULENT FLOW
VELOCITY, Vm
Figure 6.3 Friction gradient for a non-settling slurry: C1 < C2 < C3 < C4 .
Hydraulic and pneumatic conveying 211
(a) Suspended load. These are particles which are carried by turbulent
suspension in the conveying liquid. The immersed weight (i.e. weight minus
212 Fluid-particle svstems
im
VELOCITY. Vm
jm
VELOCITY. Vm
Figure 6.4 Friction gradient expressed as head of vehicle Om) and head of slurry Urn)
for a settling slurry: EF = equivalent-fluid point; C I < C2 < C3 .
o
o 009~
:...... :...... N I\)
01 00001
Velocity at limit of stationary deposition
:-" :--'"
Vsm (m/s) S 2.65
:--"':-" -t.....L -t. f\)
= I\) UJ UJ ~ c.n -..JO
.....
..... I\) w~o,Cn 000 0, 00,00 00
(b) Contact load or bed load. These are particles whose immersed weight is
supported by stresses transmitteci directly between particles and wall. This
214 Fluid-particle systems
includes particles which move through the pipe as a sliding bed, but also
extends to other types of particle motion, notably 'saltation', in which
particles are supported intermittently by collisions either with the wall of the
pipe or with other particles which in turn collide with other particles or with
the wall. The essential feature of the contact load is that, because it is
supported by solid/wall contacts which can be represented by a local mean
solid/wall normal stress, it gives rise to direct momentum transfer between
the particles and the wall which can be embodied in a mean solid/wall
tangential stress. The ratio of the mean tangential stress to the mean normal
stress can be regarded as a coefficient of dynamic friction between the solids
and the wall. The development of this model by Wilson and his coworkers has
simplified the contact load as a sliding bed but the general approach is not
limited to this type of motion: it is an approximation to any hydraulic
transport in which the immersed solids weight is supported by direct contact
between particles and wall, whether continuous or intermittent. Representa-
tion on a more detailed or sophisticated level - for example by single-element
modelling or by treating collisions as a kinematic process - is a very long way
off because, by definition, hydraulic transport is a case for which the
interstitial fluid cannot be neglected and the unsteady fluid drag terms (see
section 2.2.1) are usually significant.
The distinction between the modes of transport is illustrated in Figure 6.6,
which shows schematic sections through a horizontal pipe carrying a settling
slurry at different velocities corresponding to different ranges in Figure 6.4.
At or below the deposit point (Figure 6.6a), the pipe contains a stationary bed
with particles transported primarily by saltation along the top of the bed. If
the slurry contains fine particles, then these may be conveyed in turbulent
suspension even in the presence of a stationary bed. At some velocity V m >
Vs (Figure 6. 6b), the slurry is typically in heterogeneous flow, with part of the
solids transported as contact load and part as suspended load. The pressure
gradient is greater than that for the vehicle alone, due primarily to mechanical
friction with the contact load. As the velocity is increased, the proportion of
solids transported as contact load decreases so that, for very high velocities
(Figure 6.6c), virtually all the solids are in turbulent suspension. Because the
suspended solids do not contribute to the pressure gradient (see above),
im ~ iw at high V m, and hence, from equation (6.6), jm ~ iw/Sm < iw.
This representation of the flow of a settling slurry illustrates the importance
of the stratification ratio, i.e. the proportion of the conveyed solids which are
transported as contact load. For a slurry in which the suspended solids do not
contribute to wall shear, the friction gradient can be written
im = if + (Sm - 1) fn ( V m) (6.8)
Vso
where the (Sm - 1) term is proportional to delivered solids concentration
(see Table 6.2) and therefore accounts for variations in solids throughput.
Hydraulic and pneumatic conveying 215
~-----~---------
Cv
------
(a)
------r--c-------
------~
(b)
------1---\-------
------~
(c)
Figure 6.6 Particle concentration profiles within a pipe for settling slurry (schematic):
(a) below deposit velocity (Vrn < Vs); (b) in heterogeneous flow (Vrn > V s); (c) high-
velocity pseudo-homogeneous flow.
Vso = w ~
':yI[;
cosh[60d/D] (6.9)
where ff is the (Moody) friction factor for the flow of the vehicle alone in the
pipe, and w depends on the terminal velocity of the particle, V t ,
w = 0.9Vt + 2.7[(Qp - Qf)g~/Q~V3 (6.10)
and ~ is the shear viscosity of the vehicle. Equation (6.9) embodies the idea
that, for a particle to pass into turbulent suspension, it must be enveloped by
an eddy whose characteristic velocity exceeds the terminal velocity; the
wV(8/ff ) term provides a rough estimate for velocity in an eddy (see Wilson et
at., 1996). It also contains the idea that the eddy must be large enough to
envelop the particle, and unless diD is very small only a proportion of the
eddies will be sufficiently large; the term cosh [60d/D] is a semi-empirical
expression to account for this effect.
216 FIUld-particle systems
The index M depends on the width of the particle size distribution. For a
narrow size distribution with a geometric standard deviation less than about
1.22, M can be taken as 1.7. For broader size distributions, M is smaller, and
can be estimated from results given in Wilson et al. (1996). In this case, din
equation (6.9) and V t in equation (6.10) should be taken as the values for the
particle size such that 84% of the particles are smaller and 16% are larger (i.e.
the particle diameter which is one geometric standard deviation times the
geometric mean size; see Chapter 1) because stratification is dominated by
the coarsest particles present.
6.1.2.3 Slurry flow in a vertical pipe. In vertical slurry flow of both settling
and non-settling slurries, there is nothing corresponding to the limiting
velocity, Vs , in horizontal flow of a settling slurry - in principle, any
superficial fluid velocity above minimum fluidisation (see section 6.2) can be
used in vertical upwards flow, while conditions in downward flow are limited
by bridging or blocking. The solids distribute themselves fairly uniformly
across the pipe section, so that equations (2.87) to (2.92) apply. The slip
velocity (Vm - Vs ), can then be approximated by the hindered settling
velocity, V;, at the in situ concentration. From equations (2.92) and (6.2), the
in situ concentration is then related to the delivered concentration for a slurry
in up flow by
Cvd = ( + V;)
1 Vm Cvd (6.13)
provided that V; < < V m' The in situ concentration in vertical downflow is
Cvt = -Vm)
( Vm Cvd = (1 - --V:) Cvd (6.14)
+ V; Vm
--
For a settling slurry, V: is normally very much less than the slip velocity in
horizontal flow, so that in either case the in situ and delivered concentrations
differ less in vertical than in horizontal flow.
Because the solids are uniformly distributed across the pipe, they are
Hydraulic and pneumatic conveying 277
(6.15)
where Smt is the in situ slurry specific gravity and 1'0 is the shear stress
between the wall and the conveying fluid (or the vehicle in the case of a slurry
with a significant concentration of particles in the non-settling size range).
The first term on the right of equation (6.15) is the 'hydrostatic' term, and the
equation can be written in terms of the in situ concentration (see Table 6.2) as
(6.16)
(6.17)
---;;:z:
( -IJ.P ] D = -g
[
Qf + (Qp - Qf)Cvd
( V ; )]
1 - V rn + D
41'0
(6.18)
because the 'hydrostatic term' now acts in the direction of the flow.
Equations (6.17) and (6.18) can be used to estimate pressure drop in
vertical conveying. They also provide a relatively simple means of measuring
the density (and hence concentration) of a slurry directly on-line. Subtracting
equation (6.18) from equation (6.17) gives
-IJ.P)
( - - - ( - IJ.P ) = 2g[Qf + (Qp - Qf)Cvd] = 2gQrnd
IJ.L u IJ.L D
(6.19)
Thus averaging the pressure gradient up a riser and down a downcomer gives
the mean delivered slurry density and hence the delivered solids concentration.
The result is used in practice by passing the slurry through a vertical U-Ioop,
with the pressure gradient measured in the two legs (see Wilson et al., 1996).
A more rigorous analysis, avoiding the approximation in equations (6.17) and
(6.18) and allowing for the dependence of hindered settling velocity on in situ
concentration, has been given by Clift and Clift (1981); it delineates the
conditions under which this measurement technique becomes inaccurate.
218 Fluid-particfe systems
SLURRY
Pump
Head of Vehicle
---- - - - - -----~
'"
., ---
'"
....
-........ " At
- -- --
Discharge
Head of Slurry
• f
I
Discharge
Figure 6.8 System and pump characteristics for settling slurry conveyed with
centrifugal pumps: f = vehicle alone; 1 = concentration Cj ; 2 = concentration C2 (>
C j ) ; A = pump A; B = pump B (larger or faster than A).
relatively low-speed devices (see below), the pump characteristics for fixed
speed operation have the form shown schematically in Figure 6.8.
Consider first operation with Pump A, selected to ensure that the operating
point (defined by the intersection of the pump and system characteristics) is at
modest discharge rate or slurry velocity. If the slurry concentration is
controlled at levell, then the system will be operable. However, if the
concentration goes up to level 2, then the pressure loss through the system
goes up to characteristic 2. The head generated by the pump (characteristic
220 Fluid-particle systems
A 2 ) is now insufficient to maintain flow through the system, and the discharge
will reduce back into the deposit range. Thus, if the concentration is variable,
it will be necessary to use a pump like B, selected to give operation around
the 'equivalent fluid' point (jrn = if; see above). Variations in concentration
and throughput can now be accommodated with little variation in the steady-
state velocity or total discharge rate. A more detailed analysis, along these
lines but covering transients and changes in particle size, is given by Wilson et
al. (1996).
Figure 6.8 illustrates why the operating condition for a settling slurry with
centrifugal pumps is usually limited by operating stability rather than by the
deposit velocity. Positive displacement pumps, by contrast, deliver a fixed
discharge rate, so that their characteristics can be represented by vertical lines
on the coordinates of Figure 6.8. It is then possible to select the most
economic velocity - as is required for long-distance transport - and the
deposit velocity may be limiting.
difference between the inside and outside of a 900 bend (see Wilson et al.,
1996). Magnetic and Doppler meters are widely used for more precise
measurement of flow rate.
6.1.3.1 Flow regimes and flow transitions. It has already been noted that
both lean- and dense-phase flow are used in pneumatic conveying, a(
distinction from hydraulic conveying where genuinely dense conveying is
rare. A further distinction lies in the difference between the two: the
transition from lean to dense flow with decreasing conveying velocity is sharp
in a pneumatic system, whereas there is rarely a sharp transition in a slurry. In
general terms, this difference mirrors the much-studied case of the stability of
fluidised systems (see for example Jackson, 1985) and arises from the
different solid/fluid density ratios in the two cases. A gas-fluidised bed shows
sharp transitions; for example, an expanded bed is usually unstable (but see
section 6.2.3.1), and it collapses into a bubbling bed with particle-free voids
(i.e. 'bubbles') distinct from the dense fluidised 'phase'. By contrast, a liquid-
fluidised bed can expand, but the two phases separate only gradually so that
'bubbles' only develop towards the top of deep beds.
In horizontal flow, the transition from lean- to dense-phase flow is usually
termed the saltation velocity (although the term is somewhat misleading,
because part of the transport in the lean phase may be by the process of
saltation, as for settling slurries). The transition is shown schematically in
Figure 6.9. At constant solids flow rate, M s , and starting with a high
superficial gas velocity, Uf, the mixture is in lean-phase flow. If the gas flow is
reduced, the pressure gradient at first decreases, much as for lean-phase flow
of a settling slurry. When the saltation velocity (Usa1t ) is reached, the pressure
gradient increases sharply, as an external indication of transition to dense-
phase flow. Further reduction in gas flow causes the pressure gradient to
increase. The reverse transition occurs on increasing flow, without appreciable
hysteresis if the flow is allowed to reach steady state. The saltation velocity
increases when the solids flow rate is increased. Therefore, the transition
222 Fluid-particle systems
...
.J::
'"c:
..9:!
H
.'1::
c: F
::I
:;;Q. ~ G MS2
Q. AL
0 E
~
MSl
~
::I
II>
II>
Q)
D: Ms=O
o
from lean to dense flow may be triggered by an increase in solids flow rate, if
the line is operating close to the saltation point.
Figure 6.10 shows schematically the corresponding flow transition in flow
vertically upwards. In this case the transition is somewhat less sharp. As the
gas velocity is reduced with the solids flow rate kept constant, the in situ
particle concentration goes up (see equations (2.77) and (6.13» so that the
'hydrostatic' contribution to the pressure gradient goes up (see equations
(2.103) and (6.15». This contribution eventually outweighs the decrease in the
frictional pressure gradient, so that the total gradient passes through a
minimum and then increases. Eventually it rises more sharply, to a condition
in which the riser pipe contains a slugging (or, more rarely, bubbling)
fluidised bed. This condition is called choking. A 'choked' vertical conveying
line is typically characterised by pressure fluctuations associated with the rise
and eruption of slugs. As for horizontal conveying, the transition velocity
depends on the solids flow rate. At higher solids flow and therefore higher
choking velocity, the transition can be to a fast-fluidised system, characterised
by 'clusters' or 'strings' of particles (i.e. with the dense regions dispersed in a
lean phase, rather than a slugging or bubbling bed in which lean regions are
dispersed through a dense fluidised phase). The pressure fluctuations in the
'choked' system are then less violent.
.EC>
"
..9!
'2
":;; Ap
a.
a. AL
~
."
2!
""'
"'2!
0..
therefore it will not be considered further here. Rizk (1973) proposed a semi-
empirical correlation which has become widely used. In SI units, the Rizk
correlation can be written
Ms Ms 10 (-1.96 - 1440d)Fr (2.5 + l1ood)
------- salt (6.20)
Mf QfUsaltA
where Frsal t is the pipe Froude number at the saltation point
(6.21)
The pressure loss in a horizontal pneumatic conveying line is obtained from
the macroscopic linear momentum equation. The pressure drop along the
same length L of pipe downstream of the solids feed point is made up of the
following terms, introduced in general terms in section 2.4.2 and illustrated
schematically in Figure 6.11:
(a) Acceleration or 'entry loss'. The mean gas velocity increases from Uf
to the mean value allowing for the fraction of the pipe occupied by the solids,
Uf /(l - Cvi ). This contributes a total pressure drop corresponding to the fluid
momentum term in equation (2.105):
Solids
--+--~~~---------------
Gas • ~
:... l ------~.:
Pressure, p
• ••••
- !1 p •
..t· .
- !1 p,
..f .. ......... I!.. ...... _ ...... __ .................. __ ......... ..
Distance
More significant is the pressure drop required to accelerate the solids to their
mean conveying velocity, V p , i.e. the solids momentum term in equation
(2.105)
(6.23)
These two terms together represent the entry loss (or acceleration term)
(6.24)
(b) Friction loss. The pressure gradient due to friction arises from shear
stresses between the wall and both the gas and the conveyed solids, as in
equations (2.104) and (2.105). Because the entry length is usually relatively
short, while the accelerational and frictional components cannot readily be
distinguished in this zone, the friction gradient is usually taken to be the value
for fully-developed flow even in the entry region. As for hydraulic conveying,
the contributions to friction from the solids and the conveying fluid are
evaluated separately and taken to be additive as in equation (6.8). At least for
lean-phase conveying, the contribution from the gas is evaluated as for single-
phase flow at the mean mixture velocity. The solids contribution can be
estimated from a correlation originally proposed by Hinkle (1953)
(-/},PsF)IL = 2JsMsVpIDA (6.25)
where (in SI units)
Hydraulic and pneumatic conveying 225
6.2 FLUIDISATION
When a gas is passed upwards through a settled bed of particles, several types
of behaviour are possible. Figure 6.12 (Grace, 1986) shows the six flow
regimes which are generally distinguished, although not all these regimes are
observed for all particles/gas systems, and some types of behaviour (such as
'spouting' and 'slugging' - see later) are dependent on the dimensions of the
apparatus. In general, as the gas flow increases, the bed passes through three
types of 'aggregative fluidisation', in which the particulate 'phase' is more or
less continuous, before entering a so-called 'fast fluidisation' type, in which
the fluid phase is continuous and the particles are suspended as clusters. In
this type of behaviour, 'elutriation' (the removal of particles from the
apparatus by the gas flow) is very great, so that the solids must be collected
(usually in a cyclone - see section 7.2) and returned to the bed as shown. This
gives rise to the so-called 'circulating fluidised bed' which is in wide use as a
reactor and in combustion processes. Despite extensive research work since
the early 1980s, a comprehensive understanding of the fluid mechanics of
circulating fluidised beds is not yet available. At higher gas velocities still, the
pneumatic conveying regime is reached; this type of behaviour has already
been discussed in section 6.1.3. In this section, attention is confined to the
bubbling and slug-flow regimes of fluidisation.
For a settled bed, there comes a point at which, as the fluid velocity is
increased, the drag force on the particles becomes equal to the buoyant
weight of the bed. The bed is then supported by the fluid flow and itself
possesses fluid-like properties: flowing easily, maintaining a horizontal level
and exhibiting an apparent viscosity, for example, so that this regime of
multiphase behaviour has become known as 'fluidisation'. By virtue of their
highly mobile character, fluidised beds are very good mixers and possess good
[5.9.'7;
;~;o
~.,{b6P
fiXED BED BUBBLING SLUG flOW TURBULENT FAST PNEUMATIC
OR ~IR_EG_I_M_E____~Vr-____R_EG_IM_E~1 FLUIDIZATION CONVEYING
DELAYED
BUBBLING AGGREGATIVE FLUIDIZATION
INCREASING U, t
Figure 6.12 Principal flow regimes of upward flow of gas through solid particulate
materials (Grace, 1986).
Fluidisation 227
(6.33)
= 150
+ (6.34)
228 F/uid~partjc/e systems
(6.36)
Figure 6.13 shows schematically how the pressure drop across the bed
increases with increasing superficial gas velocity (linearly if the viscous term
in equation (6.33) is the dominant one) up to the minimum fluidisation
velocity. A simple force balance shows that it is not possible for this pressure
drop to exceed the buoyant weight of the particles. At higher fluid velocities,
therefore, either the bed voidage must increase so as to maintain the pressure
drop at or below this level, or not all the fluid can flow interstitially. The
following types of behaviour are now possible:
• uniform expansion
• 'jetting', 'bubbling' or 'slugging'
liP
Possible "overshoot"
Bed weight
per unit area
Figure 6.13 Dependence of the pressure drop liP on thc fluid velocity U through a
packed and f1uidised bed.
Fluidisation 229
• 'spouting', in which the gas enters only through a single central orifice
(Mathur and Epstein, 1974)
• nucleation and growth of stationary cavities
• channelling ('rat-holing')
All of these types of behaviour, with the exceptions of spouting and
channelling, can be described as fluidisation, because both the bed and the
individual particles within it are wholly supported by the pressure drop.
Spouting and channelling cannot, because, in general, the pressure drop
during these types of behaviour is less than that required to support the bed.
There have been several attempts to devise theoretical and empirical
classifications of these behavioural types, most of which relate only to gas-
solid fluidisation. Of these, the most widely used is the empirical classification
of Geldart (1973), who divides fluidisation behaviour according to mean
particle size and density difference between the solids and the fluidising gas
(Figure 6.14). Geldart recognises four behavioural groups, designated A, B,
C and D. Typical fluidisation behaviour of groups A, Band C is illustrated in
Figure 6.15:
"! 104
CI
.lO:
/
re' /
I
,/ B SPOUTABLE
cf"
,/
,/
UJ
,/
/
SAND- LIKE 0
u
z /
UJ /
0::
UJ
u.
u. 103
/
/ A
/
15 AERATABLE
>- "
/
!::
VI
"
/
Z ""
UJ
Cl ( "
COHESIVE
""""
1~~~~~~~--~~~~~~~~
10 50 100 500 100J
MEAN PARTICLE DIAMETER (,urn)
Group C. These particles are cohesive and tend to lift as a plug or channel
badly; conventional fluidisation is usually difficult or impossible to achieve.
Group A. These particles are intermediate in particle size and in behaviour
between groups Band C, and are distinguished from group B by the fact that
appreciable (apparently homogeneous) bed expansion occurs above the
minimum fluidisation velocity but before bubbling is observed. There is now
much experimental evidence (Seville, 1987) that group A particles are also
intermediate in cohesiveness between groups Band C, their interparticle
cohesive forces being of the same order as the single particle weight.
Group D. These particles are those which are 'large' and/or abnormally
dense. Such particles show a tendency to 'spout', rather than fluidise.
Other properties of the groups are summarised in Table 6.3 and are discussed
below.
It should be emphasised that the 'Geldart diagram' (Figure 6.15) is
applicable only to particles fluidised by air under ambient conditions, and in
the absence of artificially-enhanced interparticle forces, due to the presence
of liquid layers on the particles, for example. However, the grouping of
fluidised systems in terms of experimentally observed behaviour is of wider
application, even in cases where interparticle forces are artificially enhanced
(Seville and Clift, 1984). A later and more comprehensive classification due
Uf uf 6.P uf
H H H BEHAVIOUR ERRATIC
AND IRREPRODU5.!BLE
I'"
'" '" BUBBLING
HMF I------!"" - - - - - - 1--_ _--"1 : ~U~L.!..NG_
1 1
EXPANDING
1
u u
Figure 6.15 Typical fluidisation behaviour of particles in Geldart's (1973) groups A,
Band C (note that scales are different for each group) (Seville, 1987). H = bed height;
W/A = bed weight per unit area.
Table 6.3 Characteristic features of Geldart's (1973) classification of fluidisation behaviour (after Geldart, 1986)
Group
C A B D
Most obvious characteristic Cohesive, difficult to Bubble-free range of Starts bubbling at Umf Coarse solids
fluidise fluidisation
Typical examples Flour, cement Cracking catalyst Building sand, table Crushed limestone,
salt coffee beans
Property
1. Bed expansion Low when bed High Moderate Low
channels; can be high
when fluidised
2. Deaeration rate Can be very slow Slow Fast Fast
3. Bubble properties Channels Splitting and coales- Stable size large, and Size limited by vessel
cence predominate. may not be reached Small wake fraction
Limiting bubble size.
Large wake fraction.
4. Solids mixing Very low High Moderate Low
5. Gas back-mixing Very low High Moderate Low
6. Slug properties Solid slugs Axisymmetric; break- Asymmetric Horizontal voids.
down to turbulent Solid slugs.
fluidisation Wall slugs
7. Spouting No, except in very Shallow beds only Shallow beds only Yes, even in deep beds
shallow beds
232 Fluid-particle systems
to Grace (1986) is shown in Figure 6.16. This uses the dimensionless velocity,
U*, and particle diameter, d*, which were introduced in section 2.1.3, to
define the behavioural groups.
Figure 6.16 also shows the various processing options which might be
considered for particles of various sizes and gases of different properties.
Grace's classification successfully accounts for the effects of variation in gas
properties due to operation at elevated temperature and pressure (see later),
DILUTE CONVEYING
•
:J
T /' I
/ i GROUP 0
1 ;' I i PARTICLES
CONVENTIONAL FLUIDIZED BEDS
I I /GROUP B I
1' ,PARTICLES k
1 GROUP A / ' / '1 Proposed B- 0
1 PARTICLES Boundary
, I 1
, / II
I~
I~
: I.i~
<1;/ i
ly<>1
I I
Is ~I I
'" : ;1;'
,~ .~ ;; to.Approximote A-B 3
1 CI, I .t! 1 3
:~ Boundory lor 6(>· 10 - 2x 10
:0 L
'I fI
, .~ ~I
PACKED BEDS
'ff
Q. "
1>-
, .... I~
10-3~~~'~~__~~~~__~__~~~~~~~
10 10 2
but there is, as yet, no satisfactory classification which also takes into account
interparticle forces, which in many practical situations may be of considerable
importance.
1 Time
Figure 6.17 Typical de aeration curves (bed height, H, versus time) for powders in
groups A, Band C (Geldart, 1986). Fluidizing gas flow cut off sharply at time zero.
Uc = bed collapse rate.
234 Fluid-particle systems
Reynolds number is low, so that n is usually about 4.65; see section 2.3) and
the fact that it can also be used to characterise the bubble-free expansion of
gas fluidised beds has been used to suggest that the expansion mechanism is
the same in both cases. However, as Martin (1983) points out, the comparison
is misleading. In liquid fluidisation, expansion is uniform and particle-particle
contacts are transient. In group A fluidisation, bed expansion is thought to
occur by nucleation and growth of cavities whose sizes range from a few to
about ten particle diameters (Massimilla et al., 1972; Donsi and Massimilla,
1973). The surrounding particles maintain the surface contacts which are
essential for the stability of the structure. Further evidence for the non-
uniform mode of expansion of group A beds is given by Geldart and Wong
(1984) who showed that the Richardson-Zaki index, n, increases with
decreasing particle size below about 100 !-lm, as the interparticle forces are
able to stabilise the expanded structure more effectively. Other aspects of the
stability of expanded non-bubbling beds are reviewed by Seville (1987).
As the superficial gas velocity exceeds Umb , the passage of bubbles breaks
up the expanded structure, causing a decrease in bed height (Figure 6.15) as
the dense-phase voidage is reduced to somewhere between Emf and Emb.
When the gas supply is suddenly cut off, the bed initially collapses rapidly as
the bubbles leave (Figure 6.17) and then continues much more slowly. This
property of slow de aeration is responsible for the ease with which fluidised
group A solids are maintained in circulation around catalytic cracking plants,
for example, but is also responsible for their tendency to 'flood' on discharge
from hoppers (Geldart and Williams, 1985).
In bubbling group A beds, all bubbles travel faster than the interstitial gas,
but a tendency towards bubble splitting (see section 6.2.6.1.3) limits the size
to which they can grow by coalescence. Circulation and mixing are rapid, bed-
to-surface heat transfer is favourable, and gas exchange between the bubbles
and the dense phase is high. All of these factors, together with a larger solid
surface area per bed volume than for groups Band D, favour the use of group
A particles in fluidised bed reactors.
where F is the mass fraction of the powder <45 flm in size (another
illustration of the effect of fines on interparticle forces). The constant is
dimensional, so SI units must be used. For air at ambient conditions and
F = 0.1, this reduces to
(6.40)
(Again, SI units must be used.) Geldart (1986) used his own correlation for
U rnf and substituting this and equation (6.39) into equation (6.38) obtained
the boundary expression for group A type behaviour
2300Q~·126flo.523 exp(0.716F)
--~~~------------ > 1 (6.41)
cfi,.8 gO.934(Qp _ Qg)O.934
m '"0
E GROUP 0
en Group A properties
.><: 3
m x reported
0
~ V):, ~ e
x '(//\ GROUP B
2'
CI
"
c£" 'J 'J ~ ~1jI IjI
() ()OO . xx I!I
10..
0- x x xl Il
~.~~ ~
ell
\/
u
C
ell
0
. 1!II!II!II!II!It!I •
... e
~
()
~
~ 0,5 Group B properties
-0 GROUpe GROUP A/' reported
>.
iii
-
c
••
~ 0,2 Group 0 properhes
reported
x gas/solid
.. liquid/solid
20 50 100 200 500 1000 2000
-------------= K (6.43)
(6.44)
= 10-3 (6.45)
6.2.3.3 BID boundary. As mentioned earlier, there are two criteria for a
particle system to be in group D:
(i) that the bubble velocity, Ub, is less than the interstitial gas velocity, i.e.
(6.47)
the gas density and viscosity terms in the Wen and Yu (1966) correlation
(equation 6.35). For fine particles, Umf decreases with increasing temperature
and is hardly affected by pressure, reflecting the change in gas viscosity, while
for coarse particles Urnf increases with increasing temperature and decreases
with increasing pressure in response to changes in gas density. However, it
must be noted that equation (6.35) must be used with caution if there are
appreciable interparticle forces, because these may affect Emf, which is
subsumed within the constants of that equation (see later).
The effect of temperature and pressure on Umb are predicted by equation
(6.39), which is in reasonable agreement with the results of King and
Harrison (1980). However, Piepers et ai. (1984) showed that Umb and other
measured characteristics of a fine catalyst under pressure depend strongly on
the type of gas which is used; independent measurements suggested that this
was due to different levels of gas adsorption on the solid surface, which is
known to influence the van der Waals forces. It is therefore to be expected
that the position of the AlB boundary will depend not only on temperature
and pressure but also on gas adsorption onto the particles.
The effect of temperature on the AIC boundary will depend on whether the
sintering temperature has been exceeded. This is considered briefly in the
next section.
There is some evidence (King and Harrison, 1980) that the minimum
spouting velocity decreases with increasing pressure; hence, the BID
boundary should move to the right as the pressure is increased, bringing
larger particles into group B. Grace (1986) also considers the BID boundary,
and proposes another correlation to locate it.
OPERATING
DIAGRAM
>-
~
u
o
...J
UJ
FLUIDISED
>
VI
«
l!J
...J
:;!;
~
LL
0:::
UJ
c... DEFLUIDISED
::>
VI
TEMPERATURE
DlLATOMETER
TRACE
:z:
t-
l!J
Z
UJ
...J
~
UJ
l!J
Z
«
:z:
u
Minimum
sintering
temperature
group B copper, polymer and glass particles. In all cases, it was observed that
the temperature at which the measured defluidisation velocity departed from
the predicted behaviour in the absence of interparticle forces coincided with
the 'initial sintering temperature', as measured using a dilatometer. This
approach was developed further and applied by Tardos et al. (1984, 1985). It
was also observed that normal fluidisation could be recovered by increasing
the gas velocity, provided that the bed had not been allowed to remain
defluidised for too long.
It may be noted that the cohesive effects of, for example, liquid layers and
sintering are qualitatively different. The presence of a liquid layer may
influence whether the outcome of a particle collision is sticking or rebound
(see sections 3.4.2 and 6.2.7); this is an almost instantaneous effect. Sintering,
however, is time-dependent and in order to predict whether particles brought
into contact will fuse by sintering, it is necessary to consider the time they are
242 Fluid-particle systems
able to spend in contact, which will depend on the local bubble frequency and
therefore on the fluidising gas velocity.
There is experimental evidence (see Seville, 1987) that a small increase in
interparticle forces in a group B material can cause it to exhibit group A type
behaviour, i.e. Umb/Umf > 1. Such behaviour would be consistent with
Molerus' (1982) boundary expressions, and suggests that it is possible to move
from group B to A to C simply by an increase in interparticle forces alone.
from the nature of the bubble boundary. For a bubble in a liquid, the
boundary is a true interface dividing two phases. It has a measurable and
recognisable surface tension and, while material can diffuse across the
interface, there can be no bulk convection across the bubble boundary. By
contrast, a bubble boundary in a fluidised bed is not a sharp interface and has
no property really analogous to surface tension. It is just a boundary between
a region in which the particle concentration is very low, and a region in which
the particle concentration is of the same order as that in the whole bed at
incipient fluidisation. This boundary is permeable, so that the gas can, and
does, flow through it. Thus interphase transfer in a fluidised bed can result
both from diffusion and from bulk convection.
6.2.6. 1 Bubbles
Bubble
volume Vb
Wake
volume Vw
(a) (b)
f
(dv)max
V lOOOK
300K
10m
$
'<l
f/
"-
1m 3OOK
,/
1~
0'1m
1em I
h
1mm L -_ _--'-_ _ _-L._ _-.J
Figure 6.22 Predicted maximum stable bubble size as a function of the mean size of
the powder and gas pressure or temperature (Clift, 1986).
246 Fluid-particle systems
6.2.6.2 Slugs
Provided that d v is less than about D18, where D is the bed diameter, the
shape and rise velocity of a bubble are unaffected by the walls. For larger
bubbles, wall effects reduce the rise velocity and change the bubble shape to
reduce the wake friction. For d)D greater than about 0.6, the bubble takes up
the characteristic shape shown in Figure 6.23(a) and is known as a slug.
Increasing slug volume simply adds further to the cylindrical body of the slug,
without changing the shape of the nose. Slugs sometimes adhere to the walls
of the bed, as shown in Figure 6.23(b).
These slug shapes are found typically in beds of groups A and B. For
coarse, angular or cohesive particles, or for groups A and B in very narrow
tubes, plugs rather than slugs tend to be formed. Plugs are slow-moving voids
which fill the tube so that the bed particles 'rain through' them rather than
flowing around. 'Plug flow' is usually undesirable and can be suppressed by
roughening the bed walls (Geld art et al., 1978).
Wall effects retard bubbles. For 0.125 < d)D < 0.6, the rise velocity can be
estimated roughly (Wallis, 1969) as
Figure 6.23 (a) Axisymmetric slug. (b) Asymmetric slug. (c) Square-nosed slug or
'plug' (Clift, 1986).
Fluidisation 247
(a)
iii i i
0
a
(i)
C) (ii)
(b) 6
Figure 6.24 (a) Bubble and solids flow pattern and (b) coalescence models:
(i) bubble overtake; (ii) bubble sideways motion (Clift, 1986).
6.2.6.3.2 Bubble and slug flow rates. The bubble flow rate (also known
as the visible flow rate) in a fluidised bed, Qb, is defined as the rate at which
bubble volume crosses any level in the bed. A first estimate for Qb is given by
the 'two-phase theory of fluidisation' (Davidson and Harrison, 1963), which
conceives the bed as consisting of two phases: (a) a dense phase in which the
gas flow rate is equal to the flow rate for incipient fluidisation', i.e. the
voidage is constant at the minimum fluidisation value; (b) a bubble phase
which carries the additional flow of fluidising fluid.
The visible flow rate is then estimated as
(6.56)
where V is the superficial fluidising velocity and Vmf the value at minimum
fluidisation. In words, the simple two-phase theory can be stated as:
Ffuidisation 249
The excess gas flow above that necessary for minimum fluidisation passes
through the bed in the form of bubbles.
Extensive experimental work since this theory was first proposed has
revealed certain shortcomings. As a first approximation, however, equation
(6.56) can be used as stated, noting that Qb is actually rather lower in
practice. The discrepancy is typically of order 10% for group A, 20-30% for
group Band 40-50% for group D (Clift, 1986; Clift and Grace, 1986). There
is some evidence that the discrepancy is less in slugging beds.
bubbling beds rise faster than isolated bubbles of the same size. Unfortu-
nately, the theoretical basis for equation (6.59) is at best questionable (see
Clift and Grace, 1986). In fact, VA> Vb primarily as a result of the
increase in bubble velocity caused by coalescence. There is some evidence
that equation (6.59) accounts approximately for the average effect of
coalescence. In this case, it can be applied as a first estimate to a bed overall,
although instantaneous bubble velocities vary widely about this result and the
mean velocity for a given bubble diameter varies with position in the bed,
being highest in the 'active' bubbling zones where coalescence is frequent.
(6.62)
H - Hmf V - Vmf
(6.63)
Hmf Vb
This result, first proposed by Davidson and Harrison (1963), has been very
widely used to predict bed expansion. For more precise estimates, it is
probably best to use equation (6.62) with V A based on estimates of bubble
size given, for example, by the correlation of Darton et al. (1977) (see above).
6.2.7 Agglomeration
Although this area is largely neglected by researchers into fluidisation,
agglomeration is both a major problem and a major application of
Fluidisation 251
fluidisation. If particles are sticky, often due to the presence of free surface
liquid, they will tend to stick together. This can lead to defluidisation (see
section 6.2.5) but if controlled can be used as a method of manufacture of
agglomerated products.
Agglomeration of a powder can be used to change the shape, porosity and
bulk density of a material as well as the particle size. This can bring
improvements in the appearance, flow properties and solubility of the
material and reduce the amount of dust present. Agglomeration can be used
to ensure a well-dispersed non-segregating blend of, say, a high-cost material
with a low-cost filler or an active material with an inert filter. Comprehensive
reviews of particle-size enlargement processes including agglomeration have
been produced by Capes (1980), Sherrington and Oliver (1981) and Pietsch
(1991).
All methods of agglomeration have in common the mobilisation of
adhesive or 'binding' forces and disruptive or 'break-up' forces, and it is the
relationship between these which gives rise to the particular properties of the
products which are associated with the different agglomeration processes. In
most cases, the binding forces arise from the deliberate addition of a binding
agent (Krycer et al., 1983), usually in the form of a liquid spray. The binding
agent may remain in liquid form, or solidify/evaporate to leave solid layers or
bridges between particles. The binder can be sprayed into the bed through an
atomising nozzle which can be placed above, in or below the bed (see Figure
6.25). Circulation patterns in the bed cause particles to enter the spray region
and collide with atomised binder droplets. They then experience particle-
particle impacts as they circulate through the bed and return after some
circulation time to the spray region.
An attractive feature of fluidised bed agglomeration is that it can combine
the agglomeration and drying steps in one piece of equipment. Drying in
fluidised beds is extremely rapid so that the bed usually contains a large
proportion of dried agglomerates which periodically recycle through the
binder feed zone in order to pick up more binder.
During the agglomeration process, single particles or wet or dry agglomer-
ates are brought into contact with each other at a certain relative velocity.
Their kinetic energy may be converted to elastic or plastic deformation
(viscous dissipation) or, in the case of break-up, the creation of new surface.
As a result, the particles may coalesce, stick to each other, rebound and/or
break-up, depending on the kinetic energy of each partner and their
cohesional and adhesional strengths. Under different circumstances, and in
different types of agglomerator, agglomeration behaviour may be controlled
by the wet or dry agglomerate strength, and breakage may occur due to gross
fracture, particle attrition or some combination of these.
The agglomeration rate and the final size of agglomerates are both affected
by parameters which can change the balance of forces in the agglomerator
(Nienow and Rowe, 1985). When the binding forces (which depend on the
252 Fluid-particle systems
(b)
Figure 6.25 Fluidised bed agglomerator (Glatt Protech Ltd.). (a) top spray;
(b) bottom spray ('Wurster' column).
Fluidisation 253
binder properties as well as the particle size) are initially large compared with
the break-up forces, the embryonic agglomerate grows. This will be the case
with coarse binder atomisation, high concentration binder and low operating
gas velocities. If the binding strength is low compared with the break-up
forces, the embryonic agglomerate will break down before it can re-enter the
feed zone, often leaving the constituent particles with a coating or a patch of
dried binder. In time, this may lead to successive layers of coating, known as
'onion-ring' growth. One disadvantage of fluidised bed agglomeration by
comparison with other methods is this sensitivity to the formulation/operation
combination, although their relatively gentle action does mean that fluidised
beds can be used to produce agglomerates of relative low bulk density with
the open structure suitable for 'instant' products. Figure 6.26 illustrates the
effect of operating variables on growth rate for a strongly-agglomerating
system. Here, the use of a higher operating gas velocity or a larger starter
particle size leads to slower growth and a maximum stable agglomerate size.
600 (a)
------;0::---°
500 o
400
300
3·00 (b)
2·60
2·20
0- d po = 245 fJm
o
~"
-00.. '.80
'·40
,·00
Figure 6.26 Granulation in a strongly agglomerating system: (a) the effect of excess
gas velocity (glass powder; 5% Carbowax; constant U - Umf = 0.65 n S-l); (b) the
effect of starter particle size, d po (glass powder; 5% Carbowax; constant U - Umf =
0.525 m S-1) (Nienow and Rowe, 1985).
St 8t* 8t 8t
8t*
X x
8t*
----
X
(a) (b) (e)
decreases in granule kinetic energy will increase the overall rate of growth as
intuitively expected. Further increases in St (Figure 6.27c) so that all values
exceed St* and all collisions are 'unsuccessful' will lead to a change of
behaviour to coating rather than coalescence. All of these qualitative aspects
of agglomeration behaviour are well-documented experimentally.
It should be emphasised that in general not only the agglomerate size but
also the binder properties change with time, due to evaporation. Other things
being equal, evaporation of the binder will make coalescence more likely,
through increase in the binder viscosity. A further factor is that the liquid
loading and liquid properties determine the rate at which an agglomerate is
able to deform and consolidate (Ennis et al., 1991) - if the binder becomes
viscous very rapidly the agglomerates may have insufficient time to
consolidate and may therefore remain as loose open structures (Figure
6.28(a)), which may have good 'instant' properties but poor resistance to
attrition (Mullier et al., 1991). This, of course, opens up the possibility of
controlling granule morphology by a combination of binder selection and
careful control of the evaporation rate.
There are several problems with the approach outlined above. One is that
the Barnocky and Davis (1988) analysis (section 3.4.2), on which the analysis
of Ennis et al. (1991) is based, considers only the energy loss on impact due to
viscous dissipation in the liquid layer between two approaching spheres. In
reality, there are other dissipation mechanisms if the 'particles' are them-
selves agglomerates. Impact may lead to extensive plastic deformation
throughout the agglomerate and/or to fracture and break-up. The liquid
content and properties within the agglomerates are crucial here, because they
determine the 'deformability' of the two bodies; small liquid bridges lead to
brittle behaviour while large ones allow larger extension of particle-particle
bonds before rupture occurs, leading to a more ductile behaviour. Kristensen
et al. (1985) have attempted to relate the strength and deform ability of
agglomerates to their composition, and have established experimentally a link
between these properties and their granulation behaviour which is consistent
with the reasoning above.
A further problem is that one does not know, in general, the distribution of
256 Fluid-particle systems
(a)
(b)
Figure 6.28 Products of fluidised bed agglomeration: (a) a rapidly solidifying binder;
(b) a slowly solidifying binder (Ennis et at., 1989).
Nomenclature 257
6.3 NOMENCLATURE
A Pipe cross-sectional area (m 2 )
Ao Distributor area associated with one gas inlet point (m 2 )
Cv Volumetric concentration of solids
Cvd Delivered volume concentration
Cvt In situ volume concentration
Cw Solids concentration by weight
D Internal diameter of pipe (m), bed diameter (m)
d Particle diameter (m)
dp Mean bed particle diameter (m)
dy Volume-equivalent bubble diameter (m); Vb = Jtd~/6
(dv)max Maximum value of d v limited by stability (m)
FH Interparticle force (N)
If Moody friction factor
g Gravitational acceleration (m S-2)
H Bed depth
Hmax Maximum bed depth (m)
Hmf Bed depth at minimum fluidisation point (m)
if Hydraulic gradient for vehicle alone (m m-I )
im Hydraulic gradient for slurry in terms of vehicle head (m m-1)
jrn Hydraulic gradient for slurry in terms of head of delivered
slurry (m m- 1)
Length along pipe (m)
Mass flow rate of conveying fluid (kg S-1)
Mass flow rate of solids (kg S-1)
Pressure (Pa)
r Radius of curvature of bubble surface (m)
Sm Relative density of slurry
Ss Relative density of solids
SEC Specific energy consumption (equation (6.7»
V Superficial gas velocity (m S-I)
Vf Superficial fluid velocity (m S-1)
Vb Rise velocity of single isolated bubble (m S-1)
Vch Value of V f at choking point (m S-1)
VA Bubble velocity in freely bubbling bed (m S-1)
V rnb Gas velocity at minimum bubbling (m S-I)
258 Fluid-particle systems
Greek letters
E Void fraction
Shear viscosity of fluid (kg m-1 S-l)
Particle density (kg m-3 )
Density of fluid (kg m-3 )
Gas density (kg m-3 )
Mean slurry density (kg m-3 )
Shear stress at pipe wall (Pa)
REFERENCES
Abrahamsen, A.R. and Geldart, D. (1980) Powder Technol., 26, 35.
Bagnold, R.A. (1941) The Physics of Blown Sand and Desert Dunes, Methuen,
London.
Barnocky, G. and Davis, R.H. (1988) Phys. Fluids 31 (6), 1324.
Brown, N.P. and Heywood, N.!. (eds.) (1991) Slurry Handling: Design of Solid-
Liquid Systems, Elsevier Applied Science, London.
Capes, C.E. (1980) Particle Size Enlargement, Elsevier, Amsterdam.
Chandnani, P.P. and Epstein, N. (1986) Fluidization. V, 0stergaard, K. and
S0rensen, A. (eds.), Engineering Foundation, New York, p. 233.
Chaouki, J., Chavarie, C., Klvana, D. and Pajonk, G. (1985) Powder Technol. 43,
117.
Clift, R. (1986) Hydrodynamics of Bubbling Fluidized Beds, in Gas Fluidization
Technology, (ed. D. Geldart), Wiley, Chichester.
Clift, R. (1996) Powder Technol., 88, 335.
Clift, R. and Clift, D.H.M. (1981) Int. f. Multiphase Flow 7, 555.
Clift, R. and Grace, J.R. (1986) in Fluidization, Davidson, J.F., Clift, R. and
Harrison, D. (eds.), Ch. 3, Academic Press, New York.
Clift, R., Grace, J.R. and Weber, M.E. (1978) Bubbles, Drops and Particles,
Academic Press, New York.
Cranfield, R.R. and Geldart, D. (1974) Chern. Eng. Sci. 27,2309.
Darton, R.C., LaNauze, R.D., Davidson, J.R. and Harrison, D. (1977) Trans. I.
Chern. E. 55,274.
References 259
Potter, O.E. and Nicklin, D.J. (eds.) (1992) Fluidization VII, Engineering Foundation,
New York.
Punwani, D.V., Modi, M.V. and Tarman, P.B. (1976) International Powder and Bulk
Solids Handling and Processing Conference, Chicago.
Richardson, J.F. and Zaki, W.N. (1954) Trans. I. Chern. E. 32, 35.
Rietema, K. and Mutsers, S.M.P. (1973) Proceedings of the International Symposium
on Fluidisation, Toulouse, 28.
Rizk, F. (1973) Doktor-Ing Diss., Universitat Karlsruhe, Germany.
Seville, J.P.K. (1987) in Tribology in Particle Technology, Briscoe, B.J. and Adams,
M.J. (eds.), Adam Hilger, Bristol.
Seville, J.P.K. and Clift, R. (1984) Powder Technol. 37, 117.
Sherrington, P.J. and Oliver, R. (1981) Granulation, Heyden, London.
Shook, e.A. and Roco, M.e. (1991) Slurry Flow: Principles and Practice,
Butterworth-Heinemann, New York.
Stein, M., Martin, T.W., Seville, J.P.K., McNeil, P.A. and Parker, D.J. (1997) in
Non-invasive Monitoring of Multiphase Flows, Chaouki, J., Larachi, F. and
Dudukovic, M.P. (eds.), Elsevier, Amsterdam, p. 309.
Tardos, G., Mazzone, D. and Pfeffer, R. (1984) Can. 1. Chern. Eng. 62,884.
Tardos, G., Mazzone, D. and Pfeffer, R. (1985) Can. J. Chern. Eng. 63,377,384.
Wallis, G.B. (1969) One-Dimensional Two-Phase Flow, McGraw-Hill, New York.
Wen, e.Y. and Yu, Y.H. (1966) A. I. Chern. E. J., 12, 160.
Wilson, K.e. (1979) Proc. Hydrotransport 6, BHRA Fluid Engineering, 1.
Wilson, K.C., Addie, G.R. and Clift, R. (1992) Slurry Transport using Centrifugal
Pumps, Elsevier Applied Science, London.
Wilson, K.C., Addie, G.R., Sellgren, A. and Clift, A . .(1996) Hydraulic Conveying of
Solids with Centrifugal Pumps, 2nd ed., Chapman and Hall, London.
Wong, A.C.Y. (1983), PhD Dissertation, University of Bradford.
Woodcock, e.R. and Mason, J.S. (1988) Bulk Solids Handling, Chapman and Hall,
London.
Zenz, F.A. (1964) Ind. Eng. Chern. Fundamentals 3, 65.
7
Gas/solid separation
=-=-- (7.1)
v RT
where M is the mean relative molecular mass of the gas, v the molar volume
and R the universal gas constant. With P in bars, T in Kelvin, and g in
kg m-3 , R is 8.3143 kJ kmol- 1 K-1 .
Elementary kinetic theory gives a first approximation for the effect of
temperature and pressure on gas viscosity. Viscosity is predicted to be
independent of pressure and proportional to the square root of absolute
temperature. In practice, the temperature dependence is usually stronger but,
at least over the pressure range of interest here, it can be assumed that
pressure has no significant effect on gas viscosity (Reid et al., 1977).
These general conclusions are useful for quick estimates of the pressure
drop associated with gas cleaning equipment. For filters in which the
Reynolds number of the gas flow is small, e.g. for most barrier filters (see
later sections), a form of the Carman-Kozeny equation applies (see section
2.3.2):
t!.P ex: [.tU (7.2)
where U is the actual face velocity through the filter. Therefore, for constant
volumetric throughput of gas,
(7.3)
centripetal, and the particles therefore move centrifugally towards the outside
of the cyclone. Almost all practical cyclones induce the vortex motion
'passively', by appropriate design of the gas flow channel so that the device
has no moving parts. Some rotating devices - perhaps more properly called
'gas centrifuges' - have been proposed but have seen very limited application.
The principal types of cyclone design are illustrated in Figure 7.1. Figures
7.1(a) and 7.1(b) show axial-flow cyclones, where the vortex motion is
induced by stationary vanes. Axial-flow cyclones are used for fairly
specialised duties, typically with many such cyclones mounted in parallel (for
reasons developed below) to give a relatively compact device to handle a
large gas flow; a widespread use is in the air intakes of diesel engines to
operate in dusty environments such as deserts. However, the commonest
cyclone in industrial use in the reverse-flow cyclone, shown schematically in
Figure 7.1(c), where the vortex motion is induced by introducing the gas
tangentially to a cylindrical section called the cyclone barrel. The gas exits
through an axial pipe sometimes called the vortex finder. The end of the
vortex finder extends beyond the gas outlet, so that the gas executes a helical
outer vortex in the barrel and tapered section or cone, before moving into the
much narrower inner vortex and leaving through the vortex finder. The
particles are flung out to the wall while, in a properly designed cyclone, the
helical motion in the outer vortex pushes them down towards the apex of the
cone. Most commonly, cyclones are mounted vertically for easier discharge of
particles from the cone, but they can be (and sometimes are) used in different
orientations. A disengagement hopper is sometimes provided to control
particle discharge, as shown in Figure 7.1(c). However, in some applications
where the particles are retained within the process, a dip-leg is used instead,
in the form of a pipe down which the disengaged particles move in dense-
phase flow. This is common practice in collecting particles from the gases
leaving fluidised bed reactors or crackers or combustors, and the dip-leg then
usually extends down below the surface of the bed.
The gas entry to the cyclone is critical. It is preferable to ensure a sufficient
length of straight pipe - at least several pipe diameters - before entering the
cyclone. The different gas entry configurations are exemplified by the
archetypal designs of Stairmand (see Stairmand, 1951; Stairmand and
Kelsey, 1955) shown in Figure 7.2, which are almost 'industry standards'.
A simple tangential entry, shown in Figure 7.2(a), is cheaper to construct but
may give higher pressure drop across the cyclone (see below); Stairmand used
the tangential entry for his 'high efficiency, medium throughput' design. For
lower pressure drop (but increased construction cost), a volute (or 'scroll' or
'wrap-around') entry can be used, as in Stairmand's 'high throughput,
medium efficiency' design shown in Figure 7 .2(b).
Other types of inertial separator have not seriously impinged on the
popularity of the cyclone, which remains the 'workhorse' of gas-particle
separation. The Cardiff Cyclone (Syred et al., 1986), which exists in a range of
Inertial separators 265
(a)
-
-
Cleaned
gas stream
Ta secondary
circuit
Oust out
Figure 7.1 Cyclone designs. (a) Fixed impeller through-flow (Strauss, 1975);
(b) axial entry reverse flow (Strauss, 1975); (c) tangential entry reverse flow
(Institution of Chemical Engineers, 1985).
266 Gas/solid separation
0'50
h1
Cl
In
N
,
(a) (b)
Figure 7.2 Standard cyclone designs (Strauss, 1975, after Stairmand, 1951). (a) High-
efficiency, medium-throughput pattern. Nominal flow rate = 1.5D 2 m3 S-1.
(b) Medium-efficiency, high-throughput pattern. Nominal flow rate = 4.5D2 m 3 S-1.
Entrance velocity at these flows is approximately 15.2 m S-1 in both types.
7.2.2. 1 Pressure drop. Starting with the pressure drop across the cyclone,
I1P depends on the cyclone dimensions, the gas volumetric flow rate, Q, and
the gas density and viscosity, Q and !.l
I1P = f(D, Q, Q, !.l) (7.12)
where D is a characteristic cyclone dimension, typically the barrel diameter.
Equation (7.12) applies for the gas flow alone; particle loading effects are
discussed below. Given the five parameters in equation (7.12) and that the
system has the usual three dimensions (mass, length and time), the problem
can be expressed in terms of two (i.e. 5 minus 3) dimensionless groups. These
are, most conveniently, the pressure coefficient
1t = I1PI(QW2)
1t2 D 4 11P
(7.13)
268 Gas/solid separation
(7.16)
and is used here as a measure of gas velocity. The inlet velocity is typically of
order lOUe, the precise value depending on the geometric design of the
cyclone.
In practice, cyclones operate at very high Reynolds number so that, as for
many devices which operate in the range where flow is fully turbulent and
dominated by inertia, the effect of changes in Reynolds number can be
neglected and the pressure coefficient taken as constant.
The effective value of Jt depends on the cyclone design; it is affected by
both inlet and outlet designs, and the relative importance of inlet and outlet is
still not clear. The equation of Shepherd and Lapple (1940) is as good as any
(Hoffmann et aI., 1992)
Jt 2 D 4
Jt = k - - (7.17)
16abD~
where a and b are the dimensions of the rectangular gas entry and Dc is the
diameter of the vortex finder (see Figure 7.3). The parameter k depends on
the cyclone geometry, but is approximately 16 for tangential and volute
entries. Equation (7.17) should only be used for estimates, in the absence of
test data on any particular design being considered.
T
s
h
large that 'slip' effects can be ignored, the collection efficiency for a particle
of diameter d depends on the following variables,
'l1 = f[d, Qp' Q, [l, D, Q] (7.18)
where Qp is the particle density. Proceeding as above, the problem can be
generalised by writing it in terms of (six minus three), i.e. three dimensionless
groups. (Collection efficiency is already dimensionless.) Hence we can write
'l1 = f[St, Re, dID] (7.19)
where St is the Stokes number (see section 2.2) derived from the ratio of the
stopping distance to cyclone diameter
Qp d 2 Ue
St=--
18[lD
2 Qp~Q
(7.20)
9n [lD 3
However, equation (7.19) can be simplified. As noted above, variations in Re
have negligible effect. Similarly the size ratio, d/ D, is usually so small that its
effect can be ignored. This leaves
270 Gas/solid separation
100
90 -' l -
80 I
~ 70
g
>-
Q)
60
~ 50
~ 40
~
<!J
30
20
(a)
-
10
100
90
V
/
80
0 II
0 I
0 l
0
0'
20
(b)
0
0' I I
10 20 30 40 50 60 70
Particle size (11m)
Figure 7.4 Fractional efficiency curves for 200 mm diameter cyclones, inlet velocity
15.2 m S-l, solid density 2000 kg m-3 in air at 20o e. Strauss (1975) after Stairmand
(1951). (a) High-efficiency cyclone. (b) High-throughput cyclone.
or
(7.23)
Equation (7.23) shows that the cut size decreases (i.e. the cyclone efficiency
improves) if the throughput increases, and also that the performance of small
cyclones is better than that of large cyclones, all other things being equal.
(7.24)
d so ex (D 3fQ)'h
ex D'h/j"p-Y4 (7.25)
(7.30)
where Jtc is the pressure coefficient at inlet loading C (kg m-3 ) and Jt o is the
pressure coefficient for gas flow alone.
(7.34)
so that the cut size increases (i.e. collection efficiency deteriorates) with
increasing temperature. The available experimental data generally support
this simple result (e.g. Abrahamson and Allen, 1986). However, the
temperature effect revealed by equation (7.34) is relatively slight; generally,
use of cyclones at elevated temperatures has been straightforward, except
sometimes for mechanical problems or problems in discharging particles from
the cyclone (see below).
The effects on pressure drop can be estimated in a similar way. From
equation (7.13),
!!.P oc QQ2fD4 (7.35)
while, from equation (7.34), keeping cut size constant implies
Q oc T'D 3 (7.36)
Therefore, if a cyclone of the same size is used at the extreme temperature
and pressure,
!!.P oc PT(2n-l)
(7.37)
On the other hand, if the same velocity is used (i.e. QfD2 constant) so that
smaller cyclones are needed, then from equation (7.34)
!!.pocPT- 1 (7.38)
On either basis, increasing pressure increases the pressure drop needed across
the cyclone. If cyclone size is kept constant, then increasing temperature
increases !!.P modestly, from equation (7.37). Equation (7.38) shows that if
smaller cyclones can be used for elevated temperature, then !!.P can actually
be reduced. However, this again raises the problem of distribution between
cyclones in parallel, noted above.
274 Gas/solid separation
The results for the effect of particle loading, discussed above, are untested
at conditions remote from ambient. The qualitative effects are almost
certainly the same at elevated temperature (and pressure), because the nature
of the physical processes should be the same. However, it is not likely that
Smolik's purely empirical equation (7.30) for the effect of loading on !1P will
be applicable at high temperature and pressure. At least there is some
comfort in the expectation that a cyclone designed for low particle loading
should perform better at high loading.
7 .3 FILTRATION
7.3.1 General features of filtration behaviour
Filtration is the commonest method of particle removal from a fluid, and
refers to the process by which the particle laden gas passes through a
permeable filter 'medium', which consists commonly of a membrane or an
array of fibres or collector particles. The particles to be separated from the
gas - often known as the 'aerosol' - make contact with the collecting surfaces,
or other aerosol particles which have already been deposited thereon, and are
removed. Examples of filter media are shown in Figure 7.5(a). Most
commonly, industrial filters consist of arrays of vertically-hung cylindrical
elements ('bags' in the case of flexible fabric filters; 'candles' in the case of
rigid media), as shown in Figure 7.5(b). The flow is from the outside inwards,
and exits via a plenum chamber at the top. Many other geometries are
known.
A distinction can be drawn between two main types of filtration behaviour:
'depth' filtration and 'surface' or 'barrier' filtration (Figure 7.6). In depth
Filtration 275
(a)
Clean gas
(b) /\~:~....
Nozzle for
reverse pulse
cleaning
Dirty
gas
Figure 7.6 Cross-section of a dust cake built up on a fibrous filter (surface filtration).
The scanning electron microscope image is taken from a resin block, in which the holes
represent where the particles originally were. The larger shapes at the bottom of the
picture are the filter fibres: (Reproduced by permission of Dr.-Ing. E. Schmidt,
University of Karlsruhe.)
filtration, collection of particles from the gas occurs throughout the filter
medium. In surface filtration, the medium acts as a barrier to the solids so
that a dust 'cake' is built up on the upstream surface, with no penetration into
the medium itself. In practice, filtration behaviour depends on the properties
of both the dust and the medium: not only the relative size of the dust
particles and the pores in the medium but also the surface properties (such as
adhesion) of both. In general, gas cleaning using membranes and the finer
grades of fibrous and granular filtration media approximates to surface
filtration. Perfect surface filtration is rare, and with a virgin 'surface filter' it is
usual for there to be a short period of penetration into the surface layers of
the medium before cake formation begins. During this short period, the
filtration efficiency may be slightly lower than in steady operation.
A further important practical difference between depth filtration and
barrier filtration is that in general it is not possible to remove particles
effectively from a depth filter after they have been captured, certainly not
'on-line' (i.e. without shutting down the filter and removing the filtration
medium). Barrier filters, however, are usually cleaned in situ, sometimes by
shaking or other mechanical action but more commonly by administering a
Filtration 277
short pulse of pressurised gas in the opposite direction to that of the filtration
flow. This detaches the cake in relatively large pieces, which drop into a
collection vessel for disposal or recycle.
Since the considerations which govern filter design for depth and barrier
filtration are rather different, they will be dealt with separately in the sections
which follow.
*'Collector particles' here includes fibres. The derivations of the equations which follow are
slightly different for fibres, as are the correlations for the single-particle collection mechanisms
given in Table 7.1, but since the general approach is very similar, fibrous collectors are not
considered separately. For a comprehensive review of this subject, the reader is referred to
Brown (1993).
278 Gas/solid separation
(a)
(b) TRAJECTORY OF
PARTICLE WITH
Figure 7.7 Particle collection (a) on a single fibre (scanning electron microscope);
(b) on a single collector particle (schematic).
bed. Unit volume of bed contains 6(1 - £)/Jtd~ collector particles. The
collection rate per unit bed volume, R y , is then given by
(7.40)
3EU(1 - £)
Ky = ----- (7.41)
2de
is the collection rate constant per unit bed volume. Equation (7.40) shows
that stationary filtration is a first-order process, i.e. the rate of filtration is
proportional to the dust concentration. For non-spherical collector particles,
particle shape will modify this result, but the effect will be slight provided that
the collector particles are roughly equi-axed, i.e. no one collector dimension
is very much greater than the others. If the collector particles cover a range of
sizes, the correct value for de is the surface-volume mean (also known as the
Sauter mean; see section 1.1.4).
We now consider the filter as a whole, and define a filter penetration, f
aerosol concentration in exit gas
(7.42)
f = aerosol concentration in entering gas
The overall filter efficiency, 'Y], is then given by
'Y]=l-f (7.43)
In order to obtain penetration in terms of E, we consider the case of deep bed
filtration in the stationary regime, for which E is constant throughout the
filter. We also consider a monodisperse aerosol or one size 'cut' in a
polydisperse dust.
Consider a fixed bed of height H and cross-sectional area A (Figure 7.8).
Let the challenging aerosol concentration be Co and the penetrating aerosol
concentration be C 1 . The rate of collection in a thin slice, of thickness dh, is
given by
3EU(1 - £)
RyAdh= CAdh (7.44)
2de
Now, aerosol entering per unit time = CUA and aerosol leaving per unit time
= (C + dC)UA so that
dC -3E(1 - £)C
------=--- (7.45)
dh U
Integrating over height H with C = Co at 'Y] = 0, and C = C 1 at 'Y] = H
C1
f= C = exp[-Ky H/U] (7.46)
o
280 Gas/solid separation
D cross-
section,
A
H
h
D
Figure 7.8 Filtration in a fixed bed (schematic).
It is often the case that additional collection occurs at the entrance to and exit
from the filter, for example on retaining screens, so that equation (7.46) is
modified as
/=f exp[-KvHIU] (7.47)
where f accounts for entrance and exit effects.
It remains to calculate E. The four purely mechanical processes which can
cause collection are shown in Figure 7.9 and reliable correlations for these are
listed in Table 7.1. When several processes occur together, it is common
practice simply to sum the single-mechanism values for E. This is valid if all
the individual efficiencies are much less than unity or if one collection
mechanism dominates.
Figure 7.10 shows individual and overall collection efficiencies for typical
values of bed and aerosol properties at ambient temperature plotted as
functions of aerosol size. For any filter operating conditions, there is a 'most
penetrating particle size' at which the collection efficiency is a minimum. For
Filtration 287
(a)
, I
I (b)
,,, ,,
I
,,
, , PARTICLE
TRAJECTORY ,
\
\ \
\ \
\ \
\ \
\ \
\
,
\
,
, ,
, ;-
I
t ;
I
(d)
,
I
, I
\
\
\ , ,
\ , \
I
I ,
I
t
I
:. t ;
Figure 7.9 Mechanical mechanisms of collection: (a) diffusion; (b) inertia; (c) direct
interception; (d) gravity.
aerosol particles below this size, diffusional collection dominates, while, for
large aerosols, inertial impaction usually dominates. The effects of temperature
and pressure follow from the effects on gas properties and gas-particle
interaction discussed in sections 2.1.4 and 7.1. Increase in gas pressure has
relatively little effect on the combined efficiency, but increase in temperature
enhances collection of sub micron aerosols (because it increases Brownian
diffusivity) and reduces the collection of larger aerosols (because of the
increase in gas viscosity). The combined effect is to increase the most
penetrating particle size at higher temperatures as shown in Figure 7.11.
In addition to these purely mechanical processes, electrophoretic collection
282 Gas/solid separation
Note: numerous other correlations exist; those used here have been found reliable for granular
bed filters at ambient and high temperatures in air.
de = 500J.,lm
Pa= 2650 kg/m3
air
>-
w 15°(
:z 10-1
l1.J
w
1 bar a.
u:::
u..
U = 0·05 m/s
l1.J
:z
0
;::: TOTAL E
W
l1.J
--.J
--.J
0
W
l1.J
--.J
W
;:::
ex:
ct.
l1.J
-'
\!J
:z 10-3
Vl
104~~~--wu~--~--~---,
0·1 1 10
AfROSOL DIA. (J.,Im)
Figure 7.10 Typical single-particle collection efficiencies.
>-
u
z U= 0·08 rnls
LU
dc= 461 ,.urn
u
u.... E = 0·41
u....
LU
Pa= 1800 kg/rn 3
z:
0
~
nitrogen
u
LU
-I
1 bar a.
-I
0
10-1
u
LU
TEMPERATURE:
-I
U
~
0::
ct I
LU
-I
l!J
z 10-2
VI
-I
~
0
~
10
AEROSOL DIA. (,.urn)
granular filters for larger particles and at higher gas velocities, i.e. in the
region in which inertial collection is most effective. For a particle which
adheres on impact, the adhesion force is normally sufficiently strong for fluid
drag alone not to cause re-entrainment (Stenhouse and Freshwater, 1976), so
that the problem is generally one of instantaneous rebound. However, in
some types of granular filters, subsequent attrition of collected particles may
also occur.
A theory of rebound has been described in section 3.4. Even though the
analysis remains to be developed further, some useful qualitative conclusions
can be drawn. Most obviously, any effects which dissipate energy, such as
Filtration 285
GRADE
EFFICIENCY
1·0 ,-
/
I
/
I
I
I •
0·5 ••
t
I
I
0 0
0
0
0 Z 3 4 5 6 7 8 9
EQUIVALENT - VOLUME
DIAMETER (,.urn)
Figure 7.12 Grade efficiency for 922 (.tm silica sand filtering gasifier fines at 800°C
and superficial gas velocity 0.4 m S-1 (Ghadiri et at., 1993), d s ~ 0.89 d v . • , Sand
coated with a sticky 'retention aid'; D, clean sand.
Cohesive
and/or
low energy
particles
Uncohesive
and/or
low energy
particles
Figure 7.13 Filter cake structure. The larger circles indicate the filter medium
(schematic) (after Houi and Lenormand, 1986).
Once a cake forms, there is, as yet, no satisfactory method for predicting its
voidage and hence the pressure drop, which will increase rapidly as further
dust is collected. In general, retention of an aerosol particle at the point of
first contact gives a cake of higher void fraction (Figure 7.13) and therefore
lower resistance, because particles rest where they first contact the particles
deposited previously. Thus a high-voidage dendritic cake generally forms
when the dust is fine or cohesive or deposited at low face velocity, while dense
high-resistance cakes are typical of dense or uncohesive particles and high
approach velocities. Quantitative predictions are not yet possible so that
experiment is invariably required.
Filtration 287
"p
/
/ Baseline
/
/
/
qj/
/
/
/
/
'---------- - - - - - -----..:..:.:.:-=-.~
No. of Cycles
Figure 7.14 Filter conditioning behaviour (schematic). See text (Seville, 1993).
(7.48)
where (-dP/dz) is the pressure gradient in the direction of flow and U is the
superficial fluid velocity, i.e. the actual volumetric flow rate divided by the
area available for flow. In the case of the media considered here, the
Reynolds number (Ugdp/[!) is much less than unity, so that the second term
in equation (7.48) can be neglected and k j can be replaced by the Carman-
Kozeny expression (equation 2.69) provided that the void fraction E is not
too high (Kyan et aI., 1970),
(7.49)
where E is the void fraction, So the specific surface area of the medium and
kKis the Kozeny parameter, which depends on the geometrical structure. For
example, Seville et al. (1989b) investigated the resistance to flow of samples
of ceramic filter media prepared from fibrous and granular constituents mixed in
various proportions. Values of the Darcy's law resistance, k j , varied from 4 X
1010 to 5 X 10 11 m~2 but the value of kK remained approximately constant at
6.1 with a standard deviation of about 0.7, suggesting that equation (7.49) can
be used to 'design' a medium with the desired resistance characteristics. In
cases where E > 0.95, such as in many fibrous media, the prediction of
pressure drop is much more complex; reviews are presented by Strauss (1975)
and Brown (1993).
If Darcy's law applies, the flow resistance of an arbitrarily shaped filter
element can be predicted by solution of Laplace's equation (1). 2p = 0) with
appropriate boundary conditions, provided that the pressure drop across the
filter is small compared with the system pressure so that the gas density can be
regarded as constant.
As noted earlier, it is the conditioned rather than the virgin flow resistance
which is of most importance in practical applications. This depends on the
residual dust layer and is a great deal more difficult to predict. In principle,
the flow resistance attributable to the filter cake could be predicted by
application of equation (7.49) but in practice the void fraction is extremely
Filtration 289
(7.50)
290 Gas/solid separation
Figure 7.15 Patchy cake detachment from rigid ceramic filters (Seville et aI., 1991).
Filtration 291
gas flow
-
medium cake
-1-
position
Figure 7.16 Pressure distribution in medium and cake during reverse flow (Seville et
at., 1989a).
The gas viscosity is effectively constant and the cake and medium thicknesses
can be incorporated into modified resistances, ke and km, so that
(7.51)
and
APe = keU (7.52)
Combining equations (7.50), (7.51) and (7.52)
APe=APT[~J
ke + k m
(7.53)
This is the pressure drop across the cake itself and also, as shown in Figure
7.16, the tensile stress acting at the cake/medium interface. It is therefore this
quantity which is of prime interest when investigating the cake removal
characteristics of a given dust/medium combination.
Equation (7.53) was developed for a uniform cake. The analysis above is
equally qpplicable to a partially cleaned filter (Seville et al., 1989a). Because
of its inhomogeneous resistance to flow, a patchily cleaned filter will show
regions of preferential gas flow (Figure 7.17). However, in the uncleaned
292 Gas/solid separation
gas flow
..
cake
medium
T T
Figure 7.17 Patchily cleaned filter during reverse flow (Seville et at., 1989a).
areas the total pressure drop across the filter must still be distributed across
the medium plus cake as shown in equation (7.53), provided that the flow is
rectilinear, i.e. in the uncleaned areas the gas velocity is the same in the
medium and in the cake. This approximation is valid provided that the
undetached cake patches are large compared with the cake thickness, which is
often but not always the case.
From the analysis above it is clear that when comparing the cleaning
behaviour of different dust/medium combinations it is I!!.Po the pressure drop
across the cake alone, which should be considered and not I!!.PT , the total
pressure drop. Indeed, comparison of values for I!!.PT necessary to detach the
cake may be misleading, because they depend on the cake loading, whereas
I!!.Pc does not. It is sometimes asserted that thick cakes are easier to clean
from filter media than thin ones. The foregoing analysis shows why this
appears to be so. For the total pressure drop across the cake to remain at its
critical value for detachment, a thicker cake requires a smaller pressure
gradient for removal and therefore less cleaning gas flow and less pressure
drop across the medium itself.
The approach outlined above considers only steady reverse-flow cleaning,
but it also applies to pulse cleaning, since the maximum stress to which the
cake is subjected corresponds to the steady flow cleaning value from equation
Filtration 293
100 -l
I
J
% cake
J
remaining
I 'ideal' cake
_ . - - - detachment
-I
50 I.
J
-
I -
I
I
I
I -
Stress applied to cake (Pa)
(7.53). In many industrial applications, the pressure rise associated with the
cleaning pulse is, in any case, so slow that the process can be considered in the
same way as steady reverse-flow cleaning.
Because such factors as the cohesion of the collected dust and the adhesion
of the dust to the medium cannot be predicted a priori, it is essential to carry
out experimental work to investigate the filter cleaning behaviour. Because of
spatial variations in both the filter medium and the cake, the critical stress at
which the cake detaches is not uniform across the surface of a filter so that
detachment occurs progressively as the cleaning pressure is increased.
It is possible to determine the range of tensile stresses over which the cake
detaches from the medium by progressively increasing llPT , and measuring
the fractional cake removal at each value of llPT by collecting and weighing
the dust removed from the filter. This is most conveniently carried out on a
small flat 'coupon' of filter medium, as described by Koch et al. (1993) or as
specified in the German VDI standard 3926 (1994). Not only are the
quantities of dust required for conditioning more manageable but since the
coupon is small it is comparatively easy to ensure that the imposed cleaning
stress is uniform across it; this is not usually the case for a candle or bag
(Berbner and Loffler, 1993). Results from the coupon test can be plotted in
the form of 'percentage cake remaining' versus 'applied stress' (Figure 7.18),
where the applied stress is the appropriate value of llPc for each point,
calculated from equation (7.53). (Values of kc and k m are obtained from
294 Gas/solid separation
pressure drop measurements before the test and after complete cake
removal.) This 'cake detachment curve' provides the information needed for
rational selection of cleaning pressure. For fabrics, the cake detachment
stress has been determined in a similar way but by using 'jerk tests', in which
progressively larger accelerations are applied and the proportion of cake
detached is measured at each acceleration level (Leith and Allen, 1986;
Sievert, 1988).
7.4 NOMENCLATURE
A Filter cross-sectional area (m 2 )
a, b Dimensions of gas inlet to cyclone (m)
C Cunningham slip-correction factor; Inlet particle
loading (kg m- 3 )
D Cyclone barrel diameter (m)
DAB Brownian diffusivity of particle in a gas (m 2 S-l)
da Aerosol particle diameter (m)
de Collector particle diameter (m)
De Diameter of vortex finder (m)
d Particle diameter (m)
d 50 'Cut size' (m)
References 295
Suffices
c Particle-laden flow
o Gas flow alone
Greek letters
Collection efficiency
Gas viscosity (N s m-2 )
Pressure coefficient (equation (7.13))
Gas density (kg m-3 )
Particle density (kg m-3 )
Void fraction
REFERENCES
Abrahamson, J. and Allen, R.W.K. (1986) First International Symposium on Gas
Cleaning at High Temperatures, Inst. Chem. Engrs. Symp. Ser. no. 99, 31.
Berbner, S. and Laffler, F. (1993) in Gas Cleaning at High Temperatures (R. Clift and
J.P.K. Seville, eds) Blackie, Glasgow, p. 225.
Boericke, R.R., Giles, W.G., Dietz, P.W., Kallio, G. and Kuo, J.J. (1981) US DoE
Meeting on 'High Temperature, High Pressure Particulate and Alkali Control in
Coal Combustion Process Streams', Morgantown, February 3-5, CONF-810249.
296 Gas/solid separation
Boyson, F., Ayers, W.H. and Swithenbank, J. (1982) Trans. Inst. Chem. Engrs 60,
222.
Brown, R. (1993) Air Filtration, Pergamon, Oxford.
Clift, R., Ghadiri, M. and Hoffmann, A.C. (1991) A.I. Chem. Eng. J. 37,285.
Coury, J.R., Thambimuthu, K.V. and Clift, R. (1987) Powder Techno!. 50,257.
Coury, J.R., Raper, J.A., Guang, D. and Clift, R. (1991) Chem. Eng. Res. Des.
69(B), 97.
Dietz, P.W. (1981) A. I. Chem. Eng. 1. 27,888.
Ghadiri, M. (1980) PhD Dissertation, University of Cambridge.
Ghadiri, M., Seville, J.P.K. and Clift, R. (1993) Chem. Eng. Res. Des. 71(A), 371.
Giles, W.B. (1982) 4th Symposium on the Transfer and Utilization of Particulate
Control Technology, Houston, October 11-15.
Gokoglu, S.A. and Rosner, D.E. (1984) Int. 1. Heat Mass Trans. 27,639.
Hoffmann, A.C., van Santen, A., Allen, R.W.K. and Clift, R. (1992) Powder
Techno!. 70, 83.
Hoflinger, W., Stocklmayer, Ch. and Hackl, A. (1993) Proceedings 'Filtech' 1993,
Karlsruhe, p. 563.
Houi, D. and Lenormand, R. (1986) Filtrat. Separ. (Proc. Filt. Soc.) (July/Aug. 1986),
238.
Institution of Chemical Engineers (1985) User Guide to Dust and Fine Control, 2nd
ed.,L Chern. E., Rugby.
Koch, D., Schulz, K., Seville, J.P.K. and Clift, R. (1993) in Gas Cleaning at High
Temperatures (R. Clift and J.P.K. Seville, eds.), Blackie, Glasgow, p. 244.
Kyan, C.P., Wasan, D.T. and Kinter, R.c. (1970) Ind. Eng. Chem. Fundam. 9,596.
Leith, D. and Allen, R.W.K. (1986) in Progress in Filtration and Separation (R.J.
Wakeman, ed.) Elsevier, Amsterdam, p. 1.
Leith, D. and Licht, W. (1972) A. I. Chem. Eng. Symp. Ser. 68, 196.
Linhardt, H.D. (1981) US DoE Meeting on 'High Temperature, High Pressure
Particulate and Alkali Control in Coal Combustion Process Streams', Morgantown,
February 3-5, CONF-810249.
Mothes, H. and Loffler, F. (1984) Chem. Eng. Process 18, 323.
Muschelknautz, E. (1980) VDI Berichte no. 363,49.
Nienow, A.W. and Killick, R.c. (1987) Powder Technol. 50,267.
Paretsky, L.C. (1972) PhD Thesis, The City University of New York.
Reid, R.C., Prausnitz, J.M. and Sherwood, T.K. (1977) The Properties of Gases and
Liquids, McGraw-Hill, New York.
Schmidt, E. (1993) Proceedings 'Filtech' 1993, Karlsruhe, p. 553.
Seville, J.P.K., Cheung, W. and Clift, R. (1989a) Filtrat. Separ. (May/June 1989),187.
Seville, J.P.K., Clift, R., Withers, C.J. and Keidel, W. (1989b) Filtrat. Separ. (July/
Aug. 1989), 265.
Seville, J.P.K., Legros, R., Brereton, C.M.H., Lim, C.J. and Grace, J.R. (1991)
Proceedings of the 11th International Conference on Fluidised Bed Combustion,
ASME, New York.
Seville, J.P.K. (1993) KaNA Powder and Particle (11), 41.
Shepherd, C.B. and Lapple, C.E. (1940) Ind. Eng. Chem. 32, 1246.
Sievert, J. (1988) Dissertation Universitat Karlsruhe Fortschr. Ber. VDI Reihe 3, Nr.
161.
Stairmand, C.J. (1951) Trans. 1nst. Chem. Engrs 29,356.
Stairmand, c.J. and Kelsey, R.N. (1955) Chem. Industry 1324.
Stenhouse, J.LT. and Freshwater, D.C. (1976) Trans. Inst. Chem. Engrs 54,95.
Strauss, W. (1975) Industrial Gas Cleaning, 2nd edn., Pergamon, Oxford.
Svarovsky, L. (1981) Solid Gas Separation, Elsevier, Amsterdam.
References 297
8.1 INTRODUCTION
Particulate materials are stored in vessels which are referred to by different
names depending on their geometry; Figure 8.1 shows illustrations of some of
the most commonly used. Those containers with a flat bottom are commonly
known as 'bins' or 'bunkers'; if the vessel side walls are inclined then the term
'hopper' is used. Bins and hoppers can have either axially-symmetric or
rectangular cross-section as illustrated in Figure 8.1. With rectangular
containers, if the vessel thickness, I, is much less than the vessel half-width, a
(usually 1 < aI2), then the container may be referred to as 'planar' or 'plane'
type. The implication is that when the vessel thickness is quite small, then the
geometry is essentially two-dimensional and hence can be approximated by a
single plane.
The term 'silo' is also used especially in European literature to describe
bulk solids storage vessels irrespective of their geometry. Here, we will use
the word 'silo' to describe a storage vessel which comprises a bin section
placed on top of a hopper, a configuration often used for industrial containers
and illustrated in Figure 8.1 (d). Bulk solids storage vessels can be used for a
variety of purposes in different industrial sectors; a classification is presented
in Figure 8.2.
The vessels used in the 'primary' bulk solids handling sector, which includes
agricultural, mining, cement and refractory products, are essentially long-
term storage units for batches of material between 50 and 1000 tons. On the
other hand, when secondary processing or manufacture are involved, such as
in pharmaceuticals, chemicals, foodstuffs, plastics, dyes and pigment industries,
then bins and hoppers are more often used for intermediate storage of
materials in between other process units. Figure 8.3 shows two process flow
sheets where the storage container serves an additional 'functional' purpose
as well as holding a batch of particulate material for a given length of time. In
Figure 8.3(a), several hoppers are used to regulate the feeding of the
components of a bulk mixture which is continuously blended to form a paste.
In this case, the hoppers are used to feed different components of the mixture
(a)
(b)
_2a-·- I -
(c)
(d)
Figure 8.1 Typical bulk solids storage vessel geometries used in industry. (a) Conical
or axisymmetric hopper; (b) plane-flow wedge hopper; (c) plane-flow chisel hopper;
(d) pyramid hopper.
Major bulk
solids
producing
11310 30286.3 15658.6 521.2 14785.2 2514.7
Industries
Major
process
Industries 9007
using 29913.9 18978.9 669.9 9810.5 1990.0
granular
solids
Energy
industries
using bulk
454 21950.3 15935.0 494.1 6214.5 635.0
solids
Bulk solids
machinery, 4896.8 2256.5 209.3 2444.3 341.9
plant 2113
equipment
All UK
production 109595 193288.0 108955.4 5436.9 80680.0 12492.4
industries
% % %
100 100 100
75 75 75
"if.
"!
50 ;!!.
0
50 :;;:
0>
0
25 C\I 25
0 0
No. of establishments Net annual output Annual industrial costs
The process silos illustrated in Figure 8.3 are expected to operate at steady-
state flow conditions appropriate to the process in question. For reliable
operation, it is important to predict the steady-state flow fields set up by
different materials and their relation to the silo geometry as well as to the
physical properties of the particles. Furthermore, transient analyses may
become necessary in order to predict the dynamic changes in hopper flow
conditions as a function of certain step changes in various process variables
such as temperature, pressure, humidity and bulk composition.
Introduction 307
(a) (b)
Figure 8.3 Process silos. (a) Feed forward mode: 1, feeder; 2, reactor/contactor; 3,
non-solid phase; 4, conveyor; 5, collector. (b) Recycle mode.
Figure 8.4 Silo collapse due to structural failure upon initiation of discharge.
is reduced and the cohesive forces between the particles are increased. The
reader is referred to Chapter 4 on the failure properties of particulate
materials for an illustration of consolidation effects with cohesive materials. If
the yield strength, 0c> of the material above the silo outlet reaches a critical
value during storage which is larger than the mean compressive stress, am,
acting in the bulk at the same point, then the material will not flow out when
the outlet is opened. Instead a 'stable arch' will form at the point where Oc >
Om and discharge will be prevented, as shown in Figure 8.5(a). The prediction
of the conditions which will give rise to stable arch formation is the basis for
the major design manuals used by industry and the equipment manufacturers;
see, for example, the Draft Code of Practice of the British Materials Handling
Board (1985). The design procedures outlined in such manuals are all variants
of a semi-empirical analysis due to Jenike (1961,1964,1967), Johanson (1965,
1968) and Jenike et ai. (1973) which is discussed in detail in Chapter 4. Table
8.1 compares the design aspects which are important to large-scale batch
storage silos and small-scale continuous process silos.
(a) (b)
Mass Flow Funnel flow
Table 8.1 Factors relating to the design of bulk solids storage vessels
(ii) funnel or core-flow conditions. Figure 8.5(a) and 8.6(b) illustrate these
two types. In the so-called 'mass-flow' condition, particulate material is in
motion everywhere within the container during discharge. The flow region
boundary in this case coincides with the vessel walls and the entire cross-
sectional area of the vessel is made available for flow. As illustrated in Figure
8.6(a), such conditions are met in general when discharging free-flowing
materials from conical or wedge-shaped hoppers. To a first approximation,
the (included) hopper half-angle, u, measured clockwise from the vertical
has to be less than U e ,
304 Storage and discharge of particulate bulk solids
Jt
Uc =-- <Pw (8.1)
2
where <Pw is the angle of wall friction (see Chapter 4) so that particles next to
the wall are kept in motion. The limiting value of U c in equation (8.1) is given
by the so-called fully-rough wall condition, where <Pw = <p, i.e. the internal
and wall friction angles are identical. The case <Pw = <p becomes analogous
to the so-called 'core' or funnel flow illustrated in Figure 8.5(b) when the
material forms its own flow region boundary such as during discharge from a
flat-bottomed bin or a shallow hopper. In this case, we would expect a funnel
to form if
Jt
U
c
> -2 - <Pd (8.2)
where <Pd is the angle between the flow region boundary and the horizontal at
the bin outlet. If we can treat the flow region boundary as a fully-rough wall
then we would expect <Pd = <PR where <PR is the 'angle of repose' (see
Chapter 4) of the material. It is important to recall here that the angle of
repose will be equal to the internal friction angle <P when the material is
cohesionless, as discussed in Chapter 4. Hence with materials of negligible
cohesion, we would expect to see funnel flow inside a vessel if
(8.3)
Figure 8.6 Mass flow, funnel flow and other typical flow regimes. (a) Mass-flow, (b) Semi-mass flow, (c) Funnel flow,
(d), (e) Expanded flow.
306 Storage and discharge of particulate bulk solids
It is worth noting that in industrial practice, the same vessel is often used to
discharge a number of different material batches with varying particle and
frictional properties. The changes in the physical properties of the materials
handled will often result in different flow regimes within the same container
which may have been in the first instance designed to provide mass flow
assuming a set of nominal values of material properties.
In Chapter 4, the frictional properties of powder beds and the experimental
techniques used for their determination were considered. The accuracy of
such measurements relies largely on the matching of particle-packing
conditions and the compressive load levels to those experienced in the silo.
Hence it is not difficult to over- or underestimate the 'effective' coefficients of
internal and wall friction used in the mass-flow design criterion.
In general, if the particulate material is not significantly cohesive then the
size of the stagnant zone in funnel flow will be quite small and hence the
volume of material trapped within this region will not be significant. This
disadvantage is easily compensated for by factors like the reduced wall
stresses, smaller vessel height to hold the same volume and reduced wall wear
and particle attrition. With cohesive materials, however, the funnel flow
Table 8.2 Comparison of mass flow and funnel flow of particulate materials
Characteristics
No stagnant zones Stagnant zone formation
Uses full cross-section of vessel Flow occurs within a portion of vessel
cross-section
First-in, first-out flow First-in, last-out flow
Advantages
Minimises segregation, agglomeration Small stresses on vessel walls during flow
of materials during discharge due to the 'buffer effect' of stagnant zones
Very low particle velocities close to vessel
walls; reduced particle attrition and wall
wear
Disadvantages
Large stresses on vessel walls during Promotes segregation and agglomeration
flow during flow
Attrition of particles and erosion and Discharge rate less predictable as flow
wear of vessel wall surface due to high region boundary can alter with time
particle velocities
Small storage volume to vessel height
ratio
308 Storage and discharge of particulate bulk solids
PLUG FLOW
12
• - Height No: 2
t::
<lJ o - Height No: 3
> x
VI
VI
•
<lJ
...J
c
.2
VI
c
<lJ
E
o
o 0.5 1.0
Dimensionless distance across the bin
the flow retardation of the particles in contact with the vessel wall surface. It
is analogous to the boundary layer set up next to a wall surface in laminar flow
of fluids.
Experimental investigations by Takahashi and Yanai (1973) in axially
symmetric bins, and by Nedderman and Laohakul (1981) in planar bins, have
however shown significant differences between the boundary layers in
laminar flow and the shear zones observed in plug flow of particulate solids.
In laminar flow of Newtonian fluids, the boundary layer thickness is a
function both of the distance from the leading edge, x, and of the fluid
velocity, Uf
(8.4)
where v is the fluid kinematic viscosity and 0 is the boundary layer thickness
(Kay and Nedderman, 1974). In particulate flow, the experimental evidence
suggests that the shear zone thickness is independent of the magnitude of the
plug-flow velocity and does not vary with height below the material top
surface. Instead, it is found to be a strong function of the particle size.
Davidson (1981) suggests a correlation of the form
v = vP - (v P - v W )f3 L -x/d (8.5)
where vp is the plug flow velocity and Vw is the particle velocity at the wall. f3L
is an empirical constant whose value is found to be between 1.5 and 2.0 (see
Laohakul, 1978). Equation (8.5) will predict the shear zone thickness to be of
the order of five to six particle diameters adjacent to the wall; see Figure 8.8
and Example 8.2.
Nedderman and Laohakul (1981) predict the particle velocities in the shear
regions using a hyperbolic tangent profile (see Figure 8.8). This is expressed
in the form
v = Vw + (vp - vw)tanh(x/o) (8.6)
where vp , and Vw are the plug-flow and wall velocities, respectively, and 0 is
the characteristic length closely related to the displacement thickness of the
conventional boundary layer theory of Newtonian fluids, as used in equation
(8.4). The particle velocity Vw at the wall is found to be a strong function of
the coefficient of wall friction. When the walls are smooth (i.e. [tw < 0.2)
then there is little retardation of the particles next to the wall surface and Vw
approaches vp ' the plug-flow velocity in the bulk. With fully-rough walls (i.e.
<Pw = <p), however, the particles adjacent to the wall are severely retarded
and the particle velocity at the wall, Vw, approaches zero, as shown in Figure
8.8.
The shear strain rate y across the shear region can also be related to the
velocity at the wall
Velocity distribution in bins and hoppers 311
head of material inside the silo should not fall below the top of the conical
section. If it is decided to fill the silo completely and then discharge its
contents in a batch manner, what would be the residence time of solids in the
cylindrical section of the silo?
Solution. Using equation (8.9)
tR = 5/vp
where vp can be calculated using equation (8.8). However, we need to guess
the interstitial voidage E. A good guess is E = 0.4 ± 0.05. Thus we get
5000 X 4
v = =3.10 mm S-1
p 3600 X Jt X 950 X 0.6
Hence tR = 5/3.10 X 10-3 = 1613 s = 0.45 h.
Comments. The magnitude of the residence time for the process in this
case is rather sensitive to the value of the voidage E used in the calculations.
tR = 0.52 h if E = 0.3 and tR = 0.37 h if E = 0.5.
where x is the horizontal distance from the wall. The velocity profile vex) is
tabulated below for different values of the empirical constant f3L.
f3L = 1.5
f3L = 2.0
Comments. The shear zone thickness calculated here is of the order of six
to seven average particle diameters depending on the chosen value of the
empirical constant fh. Hence, in proportion to the bin cross-section, the size
of the shear region next to the wall is found to be insignificantly small.
Furthermore, according to the above table, it is only the first two particles
next to the wall whose velocities are severely retarded. This leads to the
conclusion that the shear zone phenomenon in plug flow of particulate solids
is purely a geometric effect resulting from the contact friction between the
wall surface and the particle layer next to the wall. The wall friction is almost
completely dissipated in this thin region. If a contact friction argument is
used, then it must simply follow that the ratio vw/v p will be independent of the
bed height so long as the wall friction angle <Pw does not vary significantly
with position along the wall. Furthermore, if vwlvp is a function of <Pw only
then it will be independent of the plug-flow velocity, vp.
using the Mohr's circle construction in Figure S.9(b). (Refer to Chapter 4 for
the definition of the stress and strain states in granular materials undergoing
continuous deformation.) If the material is assumed to be essentially
incompressible during flow, then the equation of continuity in polar
coordinates (r, 8) is given by
df(8)
- + (2 + m)f(8)tan21jJ = 0 (8.14)
d8
(J
1jI= 90° at e = 0
90 + ro+<I>w at e = a
o 1jI =
2
(a) (b)
where A is the integration constant. This is the radial velocity field solution
first introduced by lenike (1964). The limiting value of the hopper half-angle
a for which the radial velocity field is possible is given by the conditions 'ljJ1
= rt/4 and 'ljJ2 = 3rt/4. In either case, equation (8.16) will produce Vr = O. It
can also be shown quite easily that in these limiting cases, aV/ a8 will either
be zero or infinity (see Jenike, 1961) and the resultant radial velocity profiles
are of the type seen in Figure 8.10. As noted earlier, the condition Vr = 0 at
the wall is met only when the hopper walls are fully rough (i.e. <l>w = <1».
With smooth hopper walls there will always be some slip at the wall, therefore
Vr > O. Hence the rays at which 'ljJ = 'ljJ1 and 'ljJ = 'ljJ2 are not likely in
hoppers whose walls are sufficiently smooth. This also implies that the radial
velocity field is likely in narrow-angled hoppers with half-angle a « 45°.
Furthermore, in his more recent work, lenike (1991), using a different form
of the mechanical failure function (known as Von Mises criterion) instead of
the Mohr-Coulomb criterion given by equation (8.11), has calculated much
smaller values of the hopper half-angle over which the radial velocity field will
apply; Nedderman (1992) gives a more detailed description of these
calculations.
8.3.7.4 Kinematic flow model. The radial velocity field solution due to
Jenike (1964) relies on the assumption of a radial stress field immediately
above the hopper outlet. In addition it also assumes that the particle
streamlines are radial and the flow region extends as far as the hopper walls.
316 Storage and discharge of particulate bulk solids
(a) (b)
Figure 8.11 (a) Flow region boundaries of coarse 'cohesionless' materials in funnel-flow bins (Tiiziin, 1979). i. Flat-
bottom plane-strain bin (bin thickness = 2 cm) ii. Plane-strain bin; hopper half-angle = 60°. iii. Plane-strain bin; hopper
half-angle = 45°. iv. Mid-plane of a cylindrical bin showing tracers marking the stagnant zones. v, vi. Deformation
patterns of coloured particle layers in a cylindrical bin.
318 Storage and discharge of particulate bulk solids
(b)
'0
'T
r TTT"''l7 1\ ~''''TTrrl'l 0..9
80
70
'
,1
.'/ ,,"
r-
0 0 "\
v.:"'.Jrom§.,;:) n.;:'\ ~
('I.
'
\',
('I.
' Height: 1200>
d p - 1.5mm
1
....... B ::: O.25cm
_.------B = O.35cm
0..8
0.7
6.0
ciJ ._ B ::: O.45cm 0.6
,r ~
~
'0 5 0.5
E
<>
>
4D 0..4
>
,~:
3.0 0..3
2D 7: "\ 0..2
'\
0..1
1.0
, "-
"\
~"".LUJ..U..L.UJJ.C"w.'J..'1w.,J..'~"w.1~
•• LLL-,-LLL.LU~L·LL.·
00
- , - 6. - 4, - '. 0. 1. 2. 3 " 5 6 7
Xem rem
(i) (ii)
Figure 8.11 (b) Vertical velocity profiles in planar and axially-symmetric bins
(Tiiziin, 1979), (i) Plane-strain bin with flat bottom, Bin thickness: 2 cm. Orifice
width: 2 cm. (ii) Cylindrical bin. Orifice diameter: 2 cm.
(e)
10 2O~~mmmm~'~I'I~'I~,m"~I'''I~''I'~'~I'~1
d p =1.5mm
Height: Scm
no
"
dp = 1.0mm
Height: 6.0cm
1.0 -
... ....
~ 00 ~ 0.5 -
~
.'.
E E
u -o..t <>
:J :J -'
.. •
-0.5 -
-no
-1.0 -
Figure 8.11 (c) Horizontal velocity profiles in planar and axially-symmetric bins
(Tiiziin, 1979). (i) Plane-strain bin with flat bottom. Bin thickness: 2 cm, Orifice
width: 2 cm, (ii) Cylindrical bin, Orifice diameter: 2 cm.
Velocity distribution in bins and hoppers 319
OV
- +
ov
-=0 (8.17a)
ox oy
in plane strain where (x,y) are Cartesian coordinates and
OV OU U
+ (8.17b)
oz or
+-=0
r
X-++ve
in axial symmetry where (r,2,8) are the cylindrical polar coordinates. Motion
can be considered to be propagated by particles moving into the voids vacated
by others in the layer below. As seen in Figure S.12, if the two particles, B
and C, in the lower layer have different velocities, then there will be a
tendency for the particle A to move sideways in the direction of the faster
moving particle. Hence we expect the horizontal velocity, u, to be a function
of the gradient of the vertical velocity. Nedderman and Tiiziin (1979) chose
the simplest non-trivial form by allowing the horizontal velocity to vary
linearly with the vertical velocity gradient. This results in
u = -B-
ov in plane strain; u = -B -
ov in axial symmetry (S.lS)
ax Or
Substitution of equations (S.lS) into (S.17) results in
ov oZv
-=B- (S.19a)
oy oxz
in plane strain and
(S.l9b )
v = v o e-x2/4By (S.20a)
in plane strain and
v = vo e-r2/4Bz (S.20b)
in axial symmetry. Here Vo is the centreline velocity which is given by
(S.21a)
in plane strain and
Vo = QI4rtBz (S.21b)
in axial symmetry. Hence we find that in axial symmetry, the centreline
velocity varies with liz as opposed to (ltVy) in the two-dimensional case for
the same volumetric flow Q.
2. Product solution: The product solution is obtained by considering a single
central orifice of finite size at the bottom of a vessel of finite cross-section.
These conditions result in a solution of the form
;) (S.22a)
The kinematic constant, B. The scaling of the flow fields predicted by the
model depends entirely on the chosen value of B. It has already been noted
that B might be expected to be a small multiple of particle size. Indeed, the
experimental profiles measured in planar containers (see Figure 8.11(b» do
seem to indicate B = 2d when funnel flow occurs (Tiiziin, 1979; Dosekun,
1981). In two-dimensional flow fields, both the shapes of the velocity profiles,
as well as the magnitudes of the velocities are predicted with reasonable
accuracy when values of B of the order of 1-2d are chosen. However, in
axially-symmetric flow, the shape of the vertical velocity profile is no longer
bell-shaped (Figure 8.11). The velocities in the fast-moving central core are
predictable with values of B ::::; 2d; however in the surrounding plateau region,
Velocity distribution in bins and hoppers 323
40~====~==~=?2;a:=~30n.~5~em~=+===+====~
-12a =7.6 em\--
35 2a= 12.7em
t<--t- 2a = 20.3 em -l-~
30
25
E
~
1: 20
0>
'w
I
15
10
o~~-~~~~~
-15.0 -10.0 -5.0 0 5.0 10.0 15.0
X (em)
the profile flattens considerably which would require much larger values of B
(usually of the order of 5-lOd), in order to satisfy equations (S.20a) and
(S.20b). Moreover, an accurate prediction of the stagnant zone boundary
does also require B to be quite large (Figure S.U(b». This result cannot be
explained on the basis of the physical arguments advanced above. Hence we
must conclude that the proposed linear relationship between the horizontal
a a
velocity u and the gradient vi x (see equation S.lS) is far too simple and
restrictive to account for the velocity fields observed in axial symmetry. One
possible physical explanation for the observed variations in the value of the
kinematic constant could be the variation of local void age within the flow
field which is not included in the model. In an axially-symmetric flow field,
there is evidence to suggest the existence of a dilated core region separated by
a more compacted slow-moving annulus (McCabe, 1974).
324 Storage and discharge of particulate bulk solids
v 1
- ex in axial symmetry (S.26a)
vp Bz
and
v 1
- ex - - in plane strain (S.26b)
vp VBy
where v is the particle velocity at a given height within the converging flow
zone and vp is the plug-flow velocity.
Equations (S.26a) and (S.26b) above also show quite clearly that the size of
Velocitv distribution in bins and hoppers 325
1.0
./
". .--- ---- ~~-
0.9 ~
/ ~
~
., ,
~
/
/ .,
0.8 ,,
/
, ,,
! 2,
,,
0.7
/
e-
I
I
'S / ,
I
"il 0.6 I
> / I
I
~ I I
~
I
I I
0.5
e; I I
I
I
.~ I I
I
0
"il 0.4 I I
I
> I I
, I
~ 0.3
I
I
I
I I
I
I I
I
I I
0.2 , I
I
I ,,
0.1
,/
I ,'
0.0
.(.r'I" I I I I I
0 10 20 30 40 50 60 70 80
Height (em)
the transition zone depends on the particle diameter. As the particle diameter
is increased in the same vessel, v tends to vp much faster with height above the
orifice. This is also consistent with the measurements discussed earlier of the
size of the stagnant regions in funnel flow of coarse granular materials in
plane-strain and axially-symmetric containers (section 8.3.1.4).
Nedderman (1992) points out that there are no radial flow solutions to
equations (8.19a) and (8.19b). This can be seen by expressing equation
(8.19a) in terms of the stream function in equation (8.23)
-v = -
as and hence u =
as (8.27)
ox oy
which results in
(8.28)
326 Storage and discharge of particulate bulk solids
In radial flow, the stream function 'i; is a function of the ratio x/y only.
Denoting a = x/y and replacing derivatives with respect to x and y in terms of
a gives
which shows that the stream function 'i; is not a function of a only and radial
flow does not apply. However, it is worthwhile to note here that the
magnitude of the kinematic scalar B/x in equation (8.30) is a function of the
cross-sectional distance within the flow field which is entirely consistent with
the observations of flowing voidage change across the bin cited in section
8.3.1.4.
Figure 8.15 depicts the movements of the material top surface in a silo during
batch discharge and illustrates the resulting flow fields during each successive
transient stage of the batch discharge; see also Tuzun et at. (1982).
8.3.2. 1 Initial transient stage. Following the start of discharge from the
outlet, there exists a small time interval during which the 'opening up' of the
Velocity distribution in bins and hoppers 327
2Bz 2 rt!'l.Q
T = - - - - exp (
QQf
~
4Bz
) (8.31)
where Qf is the flowing bulk density and !'l.Q is the difference between the
flowing Qf and static bulk densities, Qo (Figure 8.16).
According to these calculations the duration of the initial transient with
free flowing coarse materials (i.e. d ;?: 1 mm), is predicted to be a strong
function of the particle size, d, as well as of the vessel geometry.
Throughout the initial transient stage of flow, the material top surface in
the bin will remain stationary. The material head will start to drop only after
the flow region boundary meets the vessel walls (Figure 8.15(c) and 8.15(d)).
8.3.2.2 Pseudo steady-state. When the dilation wave has passed through
the material, the initial transient phase of batch discharge is complete and
thereafter the velocity fields will be the same as those observed during steady-
state discharge. However, if the materials are discharged as a batch, then the
material top surface will continue to decline for some time within the so-
called 'plug-flow zone' without any change of shape (Figure 8.15(c)). Pseudo
steady-state will continue until the top surface descends to a level comparable
with the top of the stagnant zone (Figure 8.15(f)).
8.3.2.3 Final transient stage. As the material top surface approaches the
converging flow zone, the centreline velocity becomes greater than the
velocities to either side and a depression is formed in the top surface. This
depression grows as the surface descends until finally the slope exceeds the
angle of repose, whereupon material cascades down the top surface towards
the centre (Figures 8.15(g) and 8.15(h)).
These observations point to a time-dependent velocity field within the
converging flow zone as the velocity profiles below the material top surface
will now be a strong function of the changing shape of the top surface. If the
328 Storage and discharge of particulate bulk solids
Figure 8.15 Transient stages of batch discharge from bins and hoppers.
Velocity distribution in bins and hoppers 329
.............
",...... • •••••~.~ s
135,25
,,
: " ,
'\\ ,
,
,
: d p =2.07 mm I '"
.
" <Pw =1 8 ° ' *
122.00
.·
··
106.75
··
91.50
E
.s. 76.25
1:
OJ
·iii
I
61.00
23s
,, ,
.,.
45.75 ,,
..~~
30.50
52s
85s
150 s
15.25
,
L-____~~~~I~~_____ L_ _~
-10.16 -5.08 o 5.08 10.16
X (em)
Figure 8.16 Calculation of the transient flow region boundary in axially-symmetric
discharge. v = Particle velocity; v1 = propagation velocity of flow region boundary;
00 = static fill bulk density; Of = flowing bulk density. Do = 0.65 cm; D =20.3 em.
330 Storage and discharge of particulate bulk solids
material is in funnel flow, then the stagnant region extends up to the top
surface and therefore the continuous descent of the surface must also result in
the partial erosion of the stagnant region (Figures 8.15(f) and 8.15(g)). In
funnel flow, the completion of discharge from a vessel will not always ensure
complete emptying of the vessel contents. In wide-angled hopper sections and
flat-bottom containers, pockets of stagnant or 'dead' material may remain
inside the vessel on either side the discharge orifice (Figure 8.15(h)). The
amount of material remaining inside the vessel will be a function of the
internal and wall angles of friction and of the hopper half-angle (see Chapter
4). The observations cited above relating to the transient stages of batch
discharge from a tall vessel show quite clearly that it is the aspect ratio of the
container which determines to a large extent the relative durations of initial
and final transient stages. If the vessel aspect ratio is small (i.e. H < 1-2D)
then the so-called final transient will immediately follow the initial transient
as the material top surface will start to deform as soon as the flow region
boundary meets the vessel walls. In shallow bins and hoppers, therefore,
there will not be a pseudo steady-state.
In contrast, when the aspect ratio is high (i.e H > 3-4D) then we should
expect to see a significant pseudo steady-state as the material top surface
descends more or less horizontally in the plug-flow region of the container
(Figures 8.15(c) and 8.15(d)). In vessels of high aspect ratio, the duration of
the initial transient during which a dilation wave moves up the bunker will
also be substantially prolonged, although this period will also be affected
significantly by the ratio of the orifice width to the vessel width. In general, if
the aspect ratio is high and the vessel width is an order of magnitude or more
larger than the orifice width, then the initial transient stage becomes
significant. The model tests on tall concrete silos indicate initial flow
transients lasting as long as 0.5-1.0 h of discharge time (Nielsen, 1983a,b). If
the vessel is short and narrow, then the initial transient is reduced to a few
minutes (Tiiziin, 1979).
r+-- D
-- t
I
I
I,
\
j
--
Figure 8.17 Schematic diagram defining the notation in cylindrical and conical
hoppers. See also Nomenclature.
332 Storage and discharge of particulate bulk solids
these low values of H, the top surface of the material almost always shows a
central depression. Thus there seems to be substantial agreement that W is
independent of H until the hopper is almost empty though the precise value
of the critical height is not clear.
Similarly, there seems to be general agreement that the mass flow rate is
independent of the vessel diameter D provided this is not too small. Both
Ketchum (1929) and Brown and Richards (1960) reported that W is constant
provided D > 2.5Do and the latter gave correction factors for the smaller
values of D. Franklin and Johanson (1955), on the other hand, gave the
criterion as (D - Do) > 30d where d is the particle diameter. For smaller
values of D, larger flow rates are found and in the limit as D ~ Do, the whole
mass of material accelerates indefinitely under gravity.
Ignoring the possible effect of the particle diameter, d, dimensional
•
analysIs suggests that W ex: Qb g Y2 Do5/2. However, many earIy work ers
reported that W ex: D~ where n ;:,; 2.5. Under these circumstances, the
correlation must contain dimensional empirical constants or d to the power
(2.5 - n). Ketchum (1929) gives n = 3.0, Franklin and Johanson (1955) 2.93,
Rausch (1948) 2.8 and Brown and Richards (1959) 3.1. Here the dimensions
of the correlation are balanced by the insertion of a term in d(2.5 - n); this
rarely predicts the effect of particle size correctly.
Beverloo et ai. (1961) plotted W 2/, against Do and obtained a linear
relationship as shown in Figure 8.18. The interesting feature of this
presentation is the intercept Z. They correlated this with the particle
diameter, finding Z = kd. For spherical particles k is about 1.5 but somewhat
larger values were found for angular particles. In the latter case the value of k
depends crucially on the choice of typical dimension of the particle. The
constant k was found to be a function solely of the particle shape and no
dependence on other properties could be found. Beverloo therefore
correlated his results in the form
(8.32)
and this correlation has been applied successfully by many subsequent
workers. Though normally attributed to Berverloo, similar ideas had been
current prior to Beverloo's paper and indeed a correlation of this type had
been proposed by Hagen as early as 1852. Weighard (1952) also gave
W ex: (Do - Z)'h, and Brown and Richards (1960) devoted considerable
attention to the concept of the 'empty annulus'. No particle centre can
approach within a distance d/2 of the orifice edge and therefore all particle
centres must pass through a circle of diameter (D - d). This does not in itself
explain why k > 1 but Brown and Richards (1960) also found that there was a
decrease in the number of particles flowing per unit time in the zone adjacent
to the orifice edge. The measurements of Laohakul (1978) show that adjacent
to a solid boundary, there is a region of retarded flow a few particles in
thickness. This region clearly has much in common with the shear zone
Discharge rates from bins and hoppers 333
described by Roscoe (1970) and the term kd may simply be the displacement
thickness of this zone.
Alternative formulations for the effects of particle and orifice diameters
include the correlations of Fowler and Glastonbury (1959),
\12
D )0.185
Wex: ADh ( - -
h
(8.33)
d
Rose and Tanaka (1959),
W ex: D'/~[:o -3 r· 3
(8.34)
All the above correlations break down for fine materials, i.e. for d <
500 !lm. According to Crewdson et at. (1977), this is due to the effects of
interstitial pressure gradients and this aspect is discussed in greater detail in
section 8.4.4.
For orifices that are less than about six particle diameters across, the flow is
intermittent and irreproducible. Equation (8.32) should therefore not be used
if Do < 6d, nor may it be assumed that putting the group (Do - kd) equal to
zero correctly predicts the orifice-to-particle diameter ratio at which flow
ceases. This is given by Brown and Richards (1959) as 2.5 for slots and 4.0 for
circular orifices. These results of Brown and Richards also support the earlier
experimental work by Langmaid and Rose (1966). However, for fine
powders, a different mechanism known as arching is found. A stable arch can
be formed over orifices of considerable size and this prevents flow taking
place. The prediction of such arches is considered in Chapter 4 when
discussing lenike hopper design procedure.
Clearly, the Beverloo correlation requires a density for dimensional
consistency but there is dispute as to the appropriate value to use since the
density within a hopper varies with respect to both time and position. Many
authors are not at all clear in their definition of density, but Beverloo's
original paper in 1961 specifies the use of the initial density Qi resulting from
the filling process. However, as described in section 8.3.2 above, on initiation
of flow the material dilates to some voidage characteristic of the flowing
material. Hence for more compacted beds, there is a longer period between
the initiation of flow and the time at which the top surface starts to descend.
There is therefore a need to define a 'flowing density', Qr, as the ratio of the
mass flow rate to the volumetric flow rate calculated from the observed rate
of descent of the top surface. This definition of Qf is found to be substantially
independent of the initial voidage or the mass flow rate. Using this density in
the Beverloo correlation, Nedderman (1992) reports a much smaller range of
values of C. In the original formulation,
W = CQig'h (Do - kd)'/2 (8.36)
and C was found to be in the range 0.55 < C < 0.65. In the modified
formulation
W= C'Qfg 'h ( Do - kd) 'I 2 (8.37)
and C' lies in the range 0.575 < C' < 0.595.
Note that C is dimensionless and thus does not depend upon choice of units
so long as they are consistent. The above values of C therefore differ from
those given in the paper of Beverloo et at. which uses mixed units.
Both X-ray investigations (Bosley et at., 1969; Bransby and Blair-Fish,
1979; Lee et at., 1974) and observations by transmitted light through narrow
bins (Tiiziin, 1979) clearly show considerable changes in voidage both on
initiation of flow and from point to point within the hopper. Tiiziin (1979)
Discharge rates from bins and hoppers 335
gives an initial voidage of 0.32 and flowing voidage of 0.37 for glass ballotini.
Thus a potential error of up to 20% exists if the wrong density is used.
Clearly, some density characteristic of the flowing material is more
appropriate than the initial density. In more recent work by Hosseini-Ashrafi
and Tiiziin (1992), voidage profiles at the orifice were measured during
discharge from conical hoppers using a y-ray tomographic scanner. They
report a 10-20% reduction in flowing density in the vicinity of the orifice
plane due to the dilation of the flowing material.
The effects of the angle of friction <I> and the particle shape on the flow rate
have been investigated by Demming and Mehring (1929), Franklin and
Johanson (1955), Rose and Tanaka (1959), Harmens (1963) and Kotchanova
(1970). The effects however seem to be small. Rose and Tanaka recommend
the inclusion of a multiplicative factor of
exp (-7.7 X 10-6 c! d 3 Qg'l2)
to account for the effects of cohesion c. However C is a strong function of the
voidage and it is not specified at which voidage or stress level c should be
measured. It should also be noted that the group above is not dimensionally
consistent.
Orifice shape is clearly an important parameter and has been studied by
Fowler and Glastonbury (1959). They report that
W = A(Dh)V2 (D hld)o.185 (8.38)
where A is the area of the orifice and Dh is the hydraulic mean diameter. This
correlation preceded that of Beverloo and includes no allowance for the
empty annulus. Following the Beverloo approach it would seem reasonable to
work in terms of the area A * and hydraulic mean diameter Dt of the space
remaining after a zone of width kdl2 has been removed from the perimeter of
the orifice. The Beverloo correlation can be expressed in these terms as
4C 4C
W= Q0 *(gDt)'12 where = 0.75 (8.39)
Jt Jt
inclination will have little effect. Following these ideas, Rose and Tanaka
(1959) proposed the use of a multiplicative factor F( a, X) where
F(a, X) = (tana tanx)-D·35 a < 90° - X (8.41)
for an orifice adjacent to a wall and up to a 35% increase for an orifice in the
corner of a rectangular hopper.
All the results discussed in this section have been otained in laboratory-
scale hoppers. There seem to be no published results for large orifices. One
reason for this may be that most hoppers discharge directly onto a conveyor
of some sort and the flow rate is controlled by the conveyor characteristics.
The reader is therefore referred to specialist literature in national design
codes for bulk solids transport equipment.
(8.43)
(8.44)
Many earlier theories started from the assumption that the flow was
independent of H and was determined solely by the nature of the free-fall
arch. The work of Harmens (1963) is typical of this approach. He assumes a
shape for the free-fall arch which is independent of scale so that the height
above the orifice plane is necessarily of order Do. Particles are assumed to
detach from the arch with considerable velocity and to accelerate freely under
gravity. Their velocity on passing through the plane of the orifice is therefore
of order (gDo)'Iz and the flow rate is thus proportional to g'lzDos/2, whatever
shape of arch is assumed. The constant of proportionality is determined by
the assumed shape of the arch and the unknown rate of detachment of
particles from the arch.
The 'minimum energy theorem' of Brown and Richards (1965) is likewise
crucially dependent on the details of the free-fall arch. They consider an
equation of the Bernoulli type for the total energy content, T, of unit mass of
338 Storage and discharge of particulate bulk solids
material as it flows rapidly towards a point sink somewhat below the orifice,
as shown in Figure 8.19.
v2 a
T = gr cosO + + (8.45)
2 g
they argue with conviction that the energy content must decrease with time
and hence that (dT/dr) must be positive. The stress term (alg) is dismissed
from the analysis using the argument of Janssen (1895) that, for a cylindrical
bunker, the stresses become invariant with depth and hence (daldr) is zero
(see section 8.5). This part of the argument is not sound, as the Janssen
analysis cannot hold right down to the free-fall arch where a must fall to zero.
The dismissal of the a term ensures that H does not enter the analysis and in
the absence of any other quantity with the dimensions of length, a result of
the form W ex gY2 D 5/2 is inevitable.
For radial flow, the velocity v is given by (Alr2) in axisymmetry (or (Air) in
plane strain) where A is a function of angular position only. Thus
dT
g cosO - (8.46)
dr ,s
While it is entirely reasonable to assume that (dT/dr) is positive throughout
the region above the free-fall arch, Brown and Richards make the additional
assumption that (dT/dr) falls to zero on the arch. Besides this unconfirmed
assumption, Brown and Richards also assume that the arch is spherical and
centred on the virtual apex as shown in Figure 8.19. Hence
A2 = ~ cosO (8.47)
and the radial acceleration at the arch will be
Carleton (1972) came to the same conclusion in saying that the particles are in
free-fall below the arch and hence their radial acceleration is g cosO. He
assumes that acceleration is a continuous function of position and hence the
acceleration immediately above the arch must also be g cosO. This argument
is open to question since there is no physical requirement that acceleration is
a continuous function of position. Sokolovskii (1965) has shown that in
granular materials discontinuous forces and accelerations are to be expected.
The radial velocity Vr is given by
Discharge rates from bins and hoppers 339
Figure 8.19 Schematic diagram of 'free-fall arch' and 'virtual apex' at the outlet of a
cylindrical bin.
A
v=-
r ?
V2 5/
g r o"COS
V,
e
=------------ (8.49)
where 13 is the included half-angle of the flowing zone, i.e. (900 - X).
Since the orifice diameter, Do = 2ro sin13 (Figure 8.19), hence
V2Jt
Q= g V, Do5/2 (1 - cos3/213)/sin5/213 (8.51)
6
340 Storage and discharge of particulate bulk solids
This theory provides a not unreasonable prediction of the mass flow rate,
commonly overpredicting by a factor of about two. However it must be asked
how much of this is due to coincidence. The suppression of the term in a
makes the correct dependence on Do inevitable and the remaining assump-
tions merely affect the dependence on /3, a quantity that is not often known
with certainty. However the prediction that the velocity distribution is given
by
is not in agreement with experiment. Tiiziin (1979), for example, found a very
much stronger dependence on 8 (see Figure 8.10). Like all the other theories
considered in this section, the Brown and Richards analysis is based on
continuum mechanics and can therefore give no information on the effect of
particle diameter.
Three very similar analyses have been presented by Savage (1965), Sullivan
(1972) and Davidson and Nedderman (1973), to explain the lack of
dependence of the mass flow rate on the height H. Unlike the analyses of
Harmens (1963) and Brown and Richards (1965), which assume that W is
independent of H, these analyses include H and show that it has negligible
effect on the flow rate.
The analysis of Davidson and Nedderman (1973) considers a narrow-
angled, smooth-walled conical hopper as shown in Figure 8.20. Since the
shear stress is zero on the smooth wall and on the axis of symmetry, it is
assumed to be zero throughout, and the radial, tangential and circumferential
stresses are therefore principal stresses. As the flow is converging towards the
apex of the cone, a passive state of stress is assumed, and invoking the Haar-
Von Karman hypothesis and the Coulomb failure criterion (see Chapter 4) it
is shown that
= a
(1 +
'1''1'
= sin<j))
Orr (8.52)
1 - sin<j)
1 + K ] Y2 [ 1 _ 3-2K] Y2
W = 2:rtgg Y2 (1 - cosa) [ - - - . /~ y (8.53)
2K - 3 1 - y-2-2K
It must, however, be borne in mind that this analysis is only valid for small
values of a, since it has been assumed that the material is subjected to a
radial body force of gg.
342 Storage and discharge of particulate bulk solids
w = _Q_g'-,-V21-;-b h
3
_ [ 1 +K ] v,
(8.55)
sin v2 a 2(K - 2)
The achievement of these analyses is the demonstration that the flow rate is
independent of y and hence H. The dependence on the orifice dimensions
follows automatically. The effect of angle of friction, via the term in K, is
shown to be small, but this does not seem to have been tested experimentally
in any satisfactory way. On the other hand, the analysis can give no
information on the effect of particle size and the dependence on the hopper
half-angle a is not in accord with experiment. In general this analysis
overpredicts the flow by a factor of 1.5-2. This method cannot predict a
velocity profile since it is necessary to assume that the velocities are
independent of angular position.
Savage (1967) attempted an alternative approach to the solution of the
equations of motion based upon the method of integral relations. By fitting
the stress and velocity distributions using polynomials and by integrating the
equations of motion across the width of the hopper, a set of ordinary
differential equations was obtained. Applying the appropriate boundary
conditions, these equations were integrated to yield closed form solutions for
the stress and velocity distributions as well as the flow rate. The predicted
flow rates in wedge-shaped hoppers using realistic values for the wall and
internal friction angles were lower than those for zero wall friction but still
considerably higher than the measured flow rates.
Williams (1977) proposed upper and lower limits to the flow rate which he
obtained by solving the equations of motion approximately along the
centreline and along the hopper walls. His analysis suggests that wall friction
has only a small effect on the velocity distribution and mass flow rate.
Savage and Sayed (1979) used a development of the method of integral
relations approach of Savage (1967) to investigate flows in wedge-shaped
hoppers. They found that small density variations near the aperture could
produce flow rates lower than those obtained when the bulk was assumed to
be incompressible. The analysis showed that roughening the walls could
increase the flow rate at large hopper half-angles and this prediction was
confirmed by experiments performed to investigate the effect. Savage and
Sayed (1981) more recently extended the method of integral relations
approach to solve for the stress fields and flow rates in rough-walled conical
hoppers.
Thorpe (1984) presented a somewhat simpler analysis to account for the
effect of wall friction, by relaxing the assumption of zero shear stress within
the flowing material. He incorporated momentum terms which take into
Discharge rates from bins and hoppers 343
account the radial variation of particle velocities into Walker's (1966) stress
analysis for the conical hopper, as discussed in section 8.5 below. An equation
similar to that found by Davidson and Nedderman (1973) results, but the
constant K is replaced by a function of internal and wall angles of friction and
also the hopper half-angle u. The resulting predictions of the mass flow rate
are only slightly smaller than those of Davidson and Nedderman (1973).
(8.57)
2F
CD = --- (8.58)
ApQsgv;
344 Storage and discharge of particulate bulk solids
3 (1 - £)
grad p= 4 --d- QsgCov; (8.59)
Vr
grad pex - (8.60)
cf
and the same result can be obtained from equation (8.59) at low Reynolds
numbers where CD ex 1/Re.
It is clear from equation (8.60) that interstitial pressure effects will be more
significant with fine powders. Moreover, in a conical hopper, both the gas and
solid velocities will vary as 1/,2 and therefore the pressure gradient will be
greatest near the orifice. This, coupled with the result of the previous section,
where it was shown that it is the features within the immediate vicinity of the
orifice that control the flow, suggests that any analysis of the effect of
interstitial pressures must concentrate on the behaviour near the orifice.
The adverse pressure gradients in the vicinity of the orifice are believed to
have a direct effect on the discharge rate and are claimed by Crewdson et al.
(1977) to be the cause of the well-known reduction of flow rate with
decreasing particle size.
By making sweeping assumptions about the voidage distribution within the
material and by assuming that the Carman-Kozeny equation was applicable,
Crewdson et ai. (1977) were able to obtain analytical expressions for the
interstitial pressure profile and the discharge rate from a conical hopper.
These showed substantial agreement with experiment. Spink (1976), working
with a wedge-shaped hopper, considered the problem in more detail and also
measured the voidage profile by a capacitance method.
Crewdson et al. (1977) simply postulate that there is an extra body force
equal to the pressure gradient at the orifice (dpldr)lr and therefore modified
the Beverloo correlation to the form 0
(8.61)
W = CQb [ g + -C' ~
P ] y,
(Do - kd)'/2 (8.62)
Qb dr ro
For most granular materials, K is large and C' does not differ greatly from
unity. In contrast, Carleton (1972) does not consider the relative motion
caused by the changing void age but adopts the alternative assumption of a
stationary fluid medium and uses the well-known relationship between
Reynolds number and drag coefficient, as outlined above, to deduce the
retardation of fine powders. Holland et al. (1969) used a similar approach to
modify the Minimum Energy Theorem of Brown and Richards (1965) by
incorporating such a fluid drag term in the velocity profile across the orifice.
They showed that
where Up is the particle velocity and the pressure gradient term is given by
equation (8.56) above. This approach also seems to predict trends which are
in qualitative agreement with experimental discharge rate measurements.
While the argument of Crewdson et al. (1977) that the flow rate depends on
the pressure gradient near the orifice seems well founded, this result is rarely
directly useful in practice since it is most unlikely that the pressure gradient
can be measured. A much more common approach has been to relate the flow
rate to some pressure difference. This is measured between tappings just
below the orifice and at some arbitrarily selected point on the hopper wall.
Clearly any resulting theory or correlation will depend on the position of the
tappings but, fortunately, in many cases the bulk of the pressure change takes
place near the orifice and the precise position of the upper tapping is
relatively unimportant. This approach has been adopted by Bulsara et al.
(1964), McDougall and Evans (1966), McDougall and Knowles (1969),
Resnick et al. (1966) and many others. Commonly this method is coupled with
the use of Bernoulli's equation, either explicitly or implicitly (by the use of an
energy balance) and inevitably gives rise to a relationship of the form
W cc (/).p Y2 ).
Bulsara et al. give
W= C D (Qb/).P)Y2 (Do - kd? (8.64)
a result that yields a good correlation of experimental results for large values
of /)'P but incorrectly suggests that W tends to zero as /)'P ~ O.
346 Storage and discharge of particulate bulk solids
written as
(8.66)
where K2 and K3 are empirical functions of the material properties.
Leung et ai. (1978), on the other hand, do not add the driving forces but
add the flow rates resulting from each separately. They therefore obtained a
relationship of the form
W= [C1Qb(gDo ) '/2 + C2 11PVz] Do2 (8.67)
Though different in principle from Resnick's form, this equation gives very
similar results for the limiting cases when I1P is either small or large. The
experimental results of Yuasa and Kuno (1972), however, suggest that the
Resnick/McDougall approach is the better of the two.
Resnick's result, equation (8.66), is very similar in form to that obtained by
Crewdson et at. (1977), equation (8.61), and all that is required to reconcile
the two approaches is a satisfactory relationship between the pressure
gradient at the orifice and the overall pressure difference. For a conical
hopper, the analysis is trivial provided one can assume that the Carman-
Kozeny equation holds. The velocities of both gas and particles are
proportional to l/r and on integration it can be seen that
dPI
-
dr Yo
[1 1]
- I1P -
r0
- -
r1
(8.68)
where ro is the radius from the virtual apex to the orifice and r1 is the radial
coordinate of the upper tapping, as shown in Figure 8.20. Normally r1 » ro
so that (dp/dr)lr = (I1P/ro) and the equation of Crewdson et at. becomes
(8.69)
i1P
dPI f(Re) (8.70)
dr r"
(8.71)
28
~
0 1
:'A 24 '--
Oil .-----:::--::". 1. Beverloo e/ ars correlation (1961)
C 2/0 - - 0"'" -3. Crewdson e/ ars theory (1977)
t;;--
--
'1 20
0 B :----
I
o Open
><
Ei:
" 16
___ - - ia
I
"tij
~
"~oS 12
* ,/ ~Carrnan-KOZeny's packed bed model
..c:
~
"6 8
I B/ . .o~
3/ (P~",0 o D p~latm
p~O.latm
4
/ p~ Total pressure
/
0
0' 0.3
Average Particle Size a(mm)
Figure 8_21 Variation of discharge rate with average particle size in a conical hopper.
348 Storage and discharge of particulate bulk solids
W2= W 2
o
[ 1 -C"W]
--
Qbgro
(8.73)
where
180!lg(1 - E)2 [2K - 3]
C" = -2-j[-~-Q-scf2-(-1---c-o-sa-)-E3- 2K - 1
The correction factor C" is inversely proportional to the square of particle
size, which for discharge in air results in significant reduction in discharge rate
when the particle size is significantly smaller than 500 !lm (see Figure 8.21).
With a cylindrical bin, the relationship between the orifice pressure
gradient dp/drl ro and the pressure drop I:!..P within the bin requires the
numerical solution of Laplace's equation; hence no simple analytical forms
are available for the prediction of interstitial air effects on bulk solids
discharge rates in flat bottom containers.
8.5.1 Introduction
Design and construction of bulk solids storage and handling equipment
require a knowledge of the stress distributions prevailing in static and moving
beds of granular material. Structural stability of the storage vessels can only
be ensured if the limiting values of the stresses on vessel walls are predicted
correctly. Furthermore, the changes that occur in bulk stress distributions at
the onset of discharge from a storage vessel can in many cases impose further
constraints with regard to the choice of critical vessel dimensions such as
vessel height-to-diameter ratio, shape and size of vessel orifice, and hopper
half-angle. Frequently, flow-promoting inserts are placed inside storage
vessels to facilitate easy discharge and prevent blockages. Such inserts have to
be designed so that they can withstand the stresses exerted by the material
without any structural failure. Shapes and relative sizes of these inserts and
their positioning inside the vessels are quite critical in achieving the desired
modifications in the bulk stress distributions and the material flow behaviour.
Bulk stress distributions also affect the magnitude of wall friction between the
material bed and vessel walls as well as the friction along the external surfaces
of the flow-promoting inserts. Wall friction is in many cases responsible for
the abrasion and wear of the mechanical surfaces and in some cases the
attrition and size degradation of the particulate material. The coupling of
storage vessels with mechanical conveying equipment such as belt, screw or
bucket type feeders will also require reliable predictions of static and moving
Stress distnbutions in bins and hoppers 349
~-------- D --------~
cr zz + d crzz 8 z
dz
I+---------D --------~
or
dozz 4'tw
--+--=y (8.75)
dz D
Defining the Janssen constant K as the ratio of the principal stresses (refer to
Mohr's stress circle in Figure 8.24),
Orr OIl Or
K = -- = -- or = -- (8.76)
Figure 8.24 Mohr's circle of stress depicting the major and the minor principal
compressive stresses. YL = yield locus.
(8.78)
This is inconsistent with assumption (i) which specifies Orr to be a principal
stress and is considered in more detail elsewhere (Drescher, 1991; Nedder-
man, 1992).
Equation (8.77) is a first-order differential equation of standard form, with
solution
Ozz = --
yD
+A ( 4!-lWKZ)
exp - - - - (8.79)
4!-lwK D
where A is an arbitrary constant.
If the top surface of the fill is stress-free, we can use the boundary condition
Ozz = 0 on Z = 0
giving
Ozz = ~(1
4!-lwK
_ ex p (-4!-lW KZ )]
D
(8.80)
Tw = yD [
-4- 1 - exp
(-4!-lWKZ)]
D (8.82)
354 Storage and discharge of particulate bulk solids
(8.83)
yD
(8.84)
00
au
00 yD
Tw=-- (8.85)
4
This last result, equation (8.85), can be derived solely by a force balance and
the assumption that a steady-state stress situation is approached. It has
general validity, not only for granular materials but also for fluids in motion.
The apparent anomaly of hydrostatics results solely from the rate of approach
to this asymptote being infinitely slow. Equation (8.85), which can be derived
from equation (8.84) and the Coulomb failure criterion, can be taken as
correct for cohesionless granular materials, but equation (8.83), involving the
Janssen constant K, has less general validity.
Referring to the Mohr's circle construction in Figure 8.24, there are two
limiting values of the stress ratio,
1 - sin<l>
KA =----
1 + sin <I>
and
1 + sin<l>
Kp=----
1 - sin <I>
Thus in the passive state the value of K is the reciprocal of its value in the
active state. It is therefore convenient to differentiate between these values of
the Janssen constant with the symbols KA in the active state and Kp in the
passive state. However, since most equations involving the Janssen constant
are equally applicable for active and passive states, the undifferentiated
symbol K is used above, and this takes the values Ka and Kp as appropriate.
For typical materials, i.e. 30° < <I> < 50°, the constant Ka lies in the range
0.33-0.13 and the corresponding values of Kp are therefore in the range
3-7.5. It will be readily seen that the analysis leading to equations (8.80) to
(8.85) is equally applicable for both the active and the passive states provided
the appropriate value of K is used. Thus the same asymptotic values of Orr
and Tw are obtained but the asymptotic value of Ozz is changed by a factor of
K2. The rate of approach to these asymptotes is seen to be a function of K and
is therefore more rapid in the passive state. Figure 8.25 shows the variation of
Stress distributions in bins and hoppers 355
'tw
,---------------------------------------------r------
0.50 50 z
Figure 8.25 Variation of wall shear stress with depth in a cylindrical bunker.
the wall shear stress with depth for a material with <I> = 30° and Ilw = 0 for
both states. It is seen that the stress reaches 90% of its asymptotic value at a
depth of 4.8 bin diameters in the active state and of only 0.53 diameters in the
passive state. It is seen that even in the active state the approach to the
asymptote is comparatively rapid and that the asymptotic stresses will occur
for the larger part of all but the shallowest bunkers.
= -yD
- - [ 1 - exp (4!lWKZ)]
- --- + Qo exp [4!lWKZ]
- --- (8.86)
OZZ
4!lwKz D D
This differs from equation (8.80) solely by the addition of the extra term in Qo
which, since it contains the exponential factor, will die away with increasing
depth. Thus the surcharge only affects the stress distribution in a compar-
atively narrow region near the top of a bunker. This is an important result for
the understanding of the flow of powders from an orifice in the base of a
hopper. Unless the hopper is very shallow, the stress state near the orifice is
356 Storage and discharge of particulate bulk solids
Z
Figure 8.26 Effect of surcharge on stress distribution below the material top surface.
where (ozz)w and (orr)W are the vertical and horizontal components of the
compressive stress on the wall. Referring to the Mohr's stress circle
construction in Figure 8.27 we get
(J
Figure 8.27 Mohr's stress circle depicting 'active' and 'passive' failure states at the
vessel wall. MYL = material yield locus; WYL = wall yield locus.
358 Storage and discharge of particulate bulk solids
and
(8.91b)
(8.91c)
!iJ = (ozz)w
(8.92)
(8.94)
however, presupposes that !iJ is not a function of z. This is true only when the
stresses are no longer varying with depth, i.e. at great depth. Detailed stress
analysis shows that !iJ is far from constant, and this is the main deficiency in
Walker's analysis. Nevertheless, Walker's method predicts the wall stresses
adequately in many cases and is to be recommended if a rapid estimate is
required. Figure 8.28 gives a plot of the great depth values of the distribution
factor!iJoo as a function of the internal and wall angles of friction.
3.0,-------------------------------~~=F__.
2.8
2.6
2.4
,0
~ 2.2
"3
""""iJ
..2- 2.0
II
tA
o 1.8
N
§ 1.6
:;:0
::J
.0
·c
.~ 1.4
""0
(/J
"Qj
~ 1.2
:s:
1.0
0.8
0.6
0.4L---~----~--~----~--~----~--~-----L----J
o 10 20 30 40 50 60 70 80 90
<p (0)
Figure 8.28 Great depth values of Walkers' distribution factor (1966), as a function
of the internal friction angle <p for different values of the wall friction angle <Pw in the
active and passive stress states.
(Ozz)w = ~[1
IlwKw .
- exp ( - IlwKwZ)]
a
(8.95)
where 2a is the separation distance between the parallel walls (see Tiiziin and
Nedderman, 1985 (a) and (b)). In conical (three-dimensional) and wedge-
360 Storage and discharge of particulate bulk solids
(a)
(b)
where h is the vertical distance from the virtual apex, as shown in Figure
Stress distributions in bins and hoppers 361
20r-~r------------------------------.
o m= 4.0
• m=2.0
o m= 1.5
'l'
E
z • m=0.8
C ¢ m=O.O
lfl10
~ • m = -0.5
Cii
O~~~------------~--------------~.
o 1 2
Height (m)
8.29(b), and ho is the distance to the orifice plane. The value of the parameter
m is given by
+ x<J>w - 2xa)
nsin<J> sin( w
m= (8.97)
tana[1 - xsin<J> cos(w + x<J>w + 2xa)]
where x = -1 in the active case and x = + 1 in the passive case (see Figure
8.29(b)). Here n = 2 in conical hoppers and n = 1 in wedge-shaped hoppers.
Figure 8.30 shows the wall stress distribution in conical hoppers, for different
values of m, see Nedderman (1992) for a detailed discussion of the
relationship between 0hh and the stress at the wall which calculates the stress
at A Band C in Figure 8.29(a).
8.5.5 Walters' switch stress analysis
Walters (1973 (a) and (b)) presented an analysis for the transient switch stress
that occurs during flow initiation within a bunker. His argument is that the
material filled into the bunker is in an active state prior to discharge. On
initiation of flow, a dilation wave will pass upwards, separating the static
material from the dilated flowing material. Since the flowing material is being
pushed inwards towards the central orifice, a passive stress state will occur
within the flowing region.
Walters (1973a) assumes that at any moment the boundary between the
static and dynamic zones is a horizontal plane at depth Z, as seen in Figure
8.31. Using the Janssen-Walker analysis outlined above, we obtain
(8.98)
362 Storage and discharge of particulate bulk solids
T
cr
z z
1 "
,,
T''''''''' -.-.-..-.,-...-......-..........-..-
.,,~ ....... '"
,
O'rr
..
,#### I
,
, f
"
,f
: yO
;, : 4 Jlw ~iJ KwA
:~
,,, , 4 Jlw (j)
z ' yO
as the value of the mean vertical stress at the active zone boundary where
z = Z. Below the zone boundary seen in Figure 8.31, the passive form of
equation (8.98) applies, with the surcharge Q being given by (ozz)z. Thus, in
this zone, ozz decreases from (ozz)z to its great depth asymptote (yDI
4~wKwAg). This variation is shown by the full line in Figure 8.3l.
At all points within the static zone above the boundary Orr = GzzIKwA' and
within the dynamic zone below the boundary, Orr = 0zzIKwp. Hence, we can
also plot the variation of the normal stress 0xx as shown by the dotted line in
Figure 8.31.
Walters' (1973b) analysis would, therefore, predict a so-called wall stress
jump at the zone boundary of magnitude equivalent to KwplKwA . For the
model granular material with the internal angle of friction <I> = 30°, the
magnitude of the stress jump could be as large as 9-1O-fold (refer to section
8.5.3 above). Details of Walters' analysis are far from perfect and no reliance
should be placed on the exact numerical values because horizontal switch
planes do not occur in real granular materials. However, localised large
stresses are frequently measured especially after initiation of discharge
(Tiiziin and Nedderman, 1985b). Such high values of wall stresses resulting
from the switch phenomena are also often recorded as being responsible for
industrial silo collapses, as discussed in section 8.1.
Nomenclature 363
8.6 NOMENCLATURE
a Hopper half-width (m)
Ap Cross-sectional area of particle; constant of integration (m 2 )
b Breadth of slot outlet; orifice half-width (m)
B Kinematic constant
Co Drag coefficient
c Cohesion
C, C', C" Empirical constants
d, dp Particle diameter (m)
D Diameter (m)
Dh Hydraulic mean diameter (m)
Do Outlet diameter (m)
!if Distribution factor
gradp Pressure gradient (bar m-1)
h,H Height (m)
10 Bessel f]Jnction
kp Permeability
KA Janssen constant in the active state
Kp Janssen constant in the passive state
I Thickness (m)
p Pore pressure (Pa)
Q Volumetric flow (m 3 S-1)
Qo Surcharge (uniform applied compressive stress at the top
surface) (Nm-2)
Radial co-ordinate (m)
Reynolds number
Residence time (s)
Time/total energy content (s/Joules)
Particle velocity (m S-1)
Fluid velocity (m S-1)
Horizontal velocity (m S-1)
Vertical velocity (m S-1)
Fluid velocity (m S-1)
Centreline velocity (m S-1)
Plug-flow velocity (m S-1)
Radial velocity/relative velocity (m S-1)
Particle velocity at wall (m S-1)
Gravity-induced mass discharge rate (kg S-1)
Mass flow rate (kg S-1)
x Horizontal coordinate (m)
y Vertical coordinate (m)
z Axial coordinate (m)
Z Depth (m)
11 Po The adverse pressure difference at the hopper orifice (Nm-2 )
364 Storage and discharge of particulate bulk solids
Greek letters
0. Hopper half-angle (0)
Critical hopper half-angle for mass-flow (0)
The included half-angle of the flowing zone (0)
Empirical constant (1.5-2.0)
The inclination to the horizontal of the stagnant zone boundary
when core flow occurs (0)
Boundary layer thickness/shear zone thickness (m)
E Interstial void age/angle of slip plane CO)
Angle of internal friction CO)
Angle between the flow region boundary and the horizontal at
bin outlet CO)
<PR Angle of repose CO)
<PW Angle of wall friction CO)
y. Weight density (kg m-2 S-2)
y Shear strain rate (S-1)
A Eigenvalue
iJ, Coefficient of internal frictionlfluid viscosity (Nm-2 S-l)
iJ,g Fluid viscosity (Nm-2 S-1)
f-lw Coefficient of wall friction C)
e Polar co-ordinate (0)
Qb Bulk density (kg m-3 )
Qf Flowing bulk density (kg m-3 )
Qo Initial density (kg m-3 )
Qs Solid density (kg m-3 )
~Q Qf - Qo (kg m-3 )
a Normal stress (Nm-2)
Yield strength (Nm-2)
Mean compressive stress (Nm-2)
Normal stress at wall (Nm-2)
Angle between the major principal stress direction and the r-
direction C)
Angle between the major principal strain rate direction and the
r-direction (0)
Shear stress (Nm-2)
Shear stress at wall (Nm-2)
Stream function
REFERENCES
Arnold, P.e., McLean, A. G. and Roberts, A.W. (1982) Bulk Solids: Storage, Flow
and Handling, Tunra Research Associates, NSW, Australia.
References 365
Australian Institute of Engineers (1983) National Code of Practice for Bulk Solids
Handling.
Beverloo, W.A., Leniger, H.A. and Van de Velde, J. (1961) Chern. Engng ScLI5, 260.
Bosley, J., Schofield, C. and Shook, e.A. (1969) Trans. Inst. Chern. Engrs. 47,147.
Bransby, P.L. and Blair-Fish, P.M. (1979) Powder Technol. 8, 197.
British Materials Handling Board (1985) Draft Code of Practice for the Design of Silos,
Bins, Bunkers and Hoppers.
Brown, R.L. and Richards, J.C. (1959) Trans. Inst. Chern. Engrs 37, 108.
Brown, R.L. and Richards, J.e. (1960) Trans. Inst. Chern. Engrs. 38, 243.
Brown, R.L. and Richards, J.C. (1965) Rheo!. Acta 4, 153.
Buchele, M.V. and Wynn, P. (1980) Chemical Engineering Tripos. Part 2. Research
Project Report, University of Cambridge.
Bulsara, P.U., Zenz, F.A. and Eckert, R.S. (1964) Ind. Engng Chern. Proc. Des.
Dev. 3,348.
Carleton, A.J. (1972) Powder Techno!. 6, 91.
Crewdson, B.J., Ormond, A.L. and Nedderman, R.M. (1977) Powder Techno!. 16,
197.
Davidson, J.F. (1981) Private communication; see also Nedderman and Laohakul
(1981).
Davidson, J.F. and Nedderman, R.M. (1973) Trans. Inst. Chern. Engrs. 51, 29.
Demming, W.E. and Mehring, A.L. (1929) Ind. Engng Chern. 29,661.
Dosekun, R. (1981) PhD Thesis. Department of Chemical Engineering, University of
Cambridge.
Drescher, A. (1991) Analytical Methods in Bin-Load Analysis. Elsevier Science,
Amsterdam.
Fowler, R.T. and Glastonbury, J.G. (1959) Chern. Engng Sci. 10,150.
Franklin, F.C. and Johanson, L.N. (1955) Chern. Engng Sci. 4, 119.
Hagen, E. (1852) Ber. Preuss. Akad. d. Wiss. 35.
Harmens, A. (1963) Chern. Engng Sci. 18,297.
Harrison, A. and Mushin, S.A. (1979) Chemical Engineering Tripos. Part 2. Research
Project Report, University of Cambridge.
Hinchley, J.W. (1926) Chernical Engineering. Encyclopaedia Britannica.
Holland, J., Miller, J.E.P. Schofield, e. and Shook, C.A. (1969) Trans. Inst. Chern.
Engrs. 47, 154.
Hosseini-Ashrafi, M.E. and Tiiziin, U. (1992) Chern. Engng Sci. 48, 158.
Janssen, H.A. (1895) Z. Ver. dt. Ing. 39, 1045.
Jenike, A.W. (1961) Gravity Flow of Bulk Solids. Utah Engineering Experimental
Station, University of Utah. Bulletin 108.
Jenike, A.W. (1964) Storage and Flow of Solids. Utah Engineering Experimental
Station, University of Utah. Bulletin 123.
Jenike, A.W. (1967) Quantitative design of mass flow bins. Powder Techno!. 1,237.
Jenike, A.W. (1991) Powder Technol. 50,229.
Jenike, A.W., Johanson, J.R. and Carson, J.W. (1973) Bin loads - part 3. Mass flow
bins. J. Engng Ind. Trans. ASME B 95,6.
Johanson, J.R. (1965) Trans. Soc. Min. Engrs AIME 232, 69.
Johanson, J.R. (1968) Powder Technol. 1,328.
Kay, J.M. and Nedderman, R.M. (1974) An Introduction to Fluid Mechanics and Heat
Transfer, Cambridge University Press, Cambridge.
Ketchum, M.S. (1929) The Design of Walls, Bins and Grain Elevators, McGraw-Hill,
New York.
Kotchanova, 1.1. (1970) Powder Techno!. 4,32.
Laird, B.W. and Roberts, P.M. (1979) Chemical Engineering Tripos. Part 2. Research
Project Report, University of Cambridge.
366 Storage and discharge of particulate bulk solids
Langmaid, R.N. and Rose, H.E. (1966) J. Inst. Fuel 30, 157.
Laohakul, C. (1978) PhD Thesis. Department of Chemical Engineering, University of
Cambridge.
Lee, S., Cowin, S.C. and Templeton, I.S. (1974) Trans. Soc. Rheology 18, 247.
Leung, L.S., Jones, P.J. and Knowlton, T.M. (1978) Powder Technol. 19,7.
Litwiniszyn, J. (1971) Symposium Franco-Polonais, Problemes des Rheologie,
Warsaw.
McCabe, RP. (1974) Geotechnique 1, 45.
McDougall, I.R and Evans, A.C. (1966) Trans. Inst. Chern. Engrs 44, 15.
McDougall, I.R. and Knowles, G.H. (1969) Trans. Inst. Chern. Engrs 47, 73.
Mullins, W.W. (1974) Powder Technol. 9, 29.
Mullins, W.W. (1979) Powder Techno!. 23,115.
Nedderman, RM. (1992) Statics and Kinematics of Granular Materials, Cambridge
University Press, Cambridge.
Nedderman, R.M. and Laohakul, C. (1981) Powder Technol. 25, 91.
Nedderman, R.M. and Tiiziin, U. (1979) Powder Techno!. 22,243.
Nedderman, R.M., Tiiziin, U. and Thorpe, RB. (1983) Powder Technol. 35,69.
Newton, R.H., Dunham, G.S. and Simpson, T.P. (1945) Trans. Am. Inst. Chern.
Engrs 41,215.
Nguyen, T.V., Brennen, C. and Sabersky, R. (1979) Mechanics Applied to the
Transport of Bulk Materials. ASME, AMD-31.
Nielsen, J. (1983a) International Conference on Bulk Materials Storage, Handling and
Transport, Newcastle, Australia.
Nielsen, J. (1983b) Proceedings of the 15th Annual Symposium of the Fine Particle
Society, Hawaii, USA.
Rausch, J.M. (1948) PhD Thesis, Princeton University.
Resnick, W. (1972) Trans. Inst. Chern. Engrs 50, 289.
Resnick, W., Heled, Y., Klein, A. and Palm, E. (1966) Ind. Engng Chern. Funds 5,
392.
Roberts, A.W. (1994) First International Particle Technology Forum, Denver, USA,
Part 3,64.
Roscoe, K.H. (1970) The influence of strains in soil mechanics. Tenth Rankine
Lecture. Geotechnique 20, 122.
Rose, H.F. and Tanaka, T. (1959) The Engineer (London) 208, October 23.
Savage, S.B. (1965) Br. J. Appl. Phys. 16, 1885.
Savage, S.B. (1967) I. J. Mech. Sci. 9, 651.
Savage, S.B. and Sayed, M. (1979) Mechanics Applied to the Transport of Bulk
Materials, ASME, AMD-31.
Savage, S.B. and Sayed, M. (1981) Z. Angew. Math. Phys. 32, 125.
Shaxby, J.H. and Evans, J.C. (1923) Trans. Faraday Soc. 19,60.
Sokolovskii, V.V. (1965) Statistics of Granular Media, Pergamon Press, Oxford.
Spink, C.D. (1976) PhD Thesis, Department of Chemical Engineering, University of
Cambridge.
Sullivan, W.N. (1972) PhD Thesis, California Institute of Technology.
Takahashi, H. and Yanai, H. (1973) Powder Technol7, 205.
Takahashi, H. and Yanai, H. (1974) Kagaku Kogaku 38, 74.
Thiemer, O.F. (1969) Trans. ASME J. Engng 1ndust. Series B, 91, 460.
Thorpe, R.B. (1984) PhD Thesis, Department of Chemical Engineering, University of
Cambridge.
Tiiziin, U. (1979) PhD Thesis, University of Cambridge.
Tiiziin, U. and Nedderman, R.M. (1982) Powder Techno!. 31,27.
Tiiziin, U. and Nedderman, RM. (1985a) Chern. Engng Sci. 40, 325.
Tiiziin, U. and Nedderman, R.M. (1985b) Chern. Engng Sci. 40,337.
References 367
Tiiziin, U., Houslby, G.T., Nedderman, R.M. and Savage, S.B. (1982) Chern. Engng
Sci. 37, 1691.
Walker, S.M. (1966) Chern. Engng Sci. 21, 975.
Walters, J.K. (1973a) Chern. Engng Sci. 28, 13.
Walters, J.K. (1973b) Chern. Engng. Sci. 28, 779.
Weighhard, K. (1952) Ingen.-Arch. 20, 109.
Williams, J.C. (1977) Chern. Engng Sci. 32,247.
Yuasa, Y. and Kuno, H. (1972) Powder Techno/. 6,97.
Zenz, F.A. (1975) Fluidisation Technol. 2,239.
Index
Lambert-Beer law, see Light scattering count median diameter (CMD) 16-17,
Laplace equation 110 49-52
Light scattering diameter of the sphere of average mass 9
extinction 41-2 diameter of the sphere of average surface 9
Franhofer theory 36-41 geometric mean 7
Lambert-Beer law 41 geometric standard deviation (GSD) 16-18,
Mie theory 35-8 49-52
particle shape effects 39 mass mean diameter 10-12
refractive index 37-42 median 7
Rayleigh theory 36 mode 7
see also Particle size measurement moment averages 9-10
Liquid bridges Sauter mean 10, 279
capillary forces between spheres 110-13 surface mean diameter 10-12, 16, 279
capillary state 109 Particle size distribution
condensation and evaporation 116-20 conversion between weighted distributions
cone and plate contact 113-16 12-14, 51-2
dynamic forces 116-17 cumulative 5-9
effect of surface roughness 112-14, 119 discrete and continuous 4-5
funicular state 109 frequency distribution function 4-7
pendular state 109 frequency histogram 3-5
toroidal approximation 110 in general 3
Log-normal distribution, see Particle size log-normal distribution 16-8, 49-52
distribution log-probability axes 16-19
Log-probability axes, see Particle size mass distribution 9-15
distribution number distribution 3
surface distribution 9
Mean free path 1,77 weighted distributions 9-14
Microscopy, see Particle size measurement Particle size measurement
Mohl-Coulomb failure criterion 164 'aerodynamic particle sizer' 47
Mohr's circle of stresses 163-4 coincidence errors 39
Momentum balance, for particle-fluid 'Coulter counter' 42-4
mixtures in pipe flow 91-4 'electrozone' method 42-4
extinction meters 34, 41-2
Newton's law (for drag), see Drag on a particle Fraunhofer scattering instruments 33
Normal distribution 49 inertial impactor 45-6
methods compared 32
Particle, definition of 1 microscopy 31
Particle density 45 settling methods 45
Particle diameter sieving 31
aerodynamic diameter 21 single particle optical counters 33
equivalent diameters in general 3, 18 see also Light scattering
equivalent-projected-area circle diameter Pipe flow of particlelfluid mixtures 87-96
20 see also Pneumatic conveying; Hydraulic
equivalent-surface sphere diameter 20 conveying
equivalent-volume sphere diameter 18 Pneumatic conveying
relationship between Stokes' and choking 222
equivalent-volume diameters 69-70 comparison with hydraulic conveying
Stokes' diameter 20-1 201-5
Particle shape entry loss due to acceleration 223-5
dynamic shape factor 26 flow in horizontal pipes 223-5
fractal dimension 25-6 flow in vertical pipes 225
Heywood's approach 21-4 flow regimes and transitions 221-2
Wadell sphericity 21 saltation 221-2
Particle size, see Particle size, averages; Pyknometer, see Particle density
Particle size distribution; Particle size
measurement Resolution
Particle size, averages of optical microscope
arithmetic mean 7 of size measuring devices 3
372 Index
Restitution, coefficient, of, see Impact and Stresses in bins and hoppers
rebound active and passive states 165-7,350,354--61
Reynolds' lubrication equation 116 effect of a surcharge 355--6
Reynolds' number Janssen-Walker analysis 356--61
for a cyclone 268 method of characteristics 349
at minimum fluidization 228 method of differential slices (Janssen's
for a particle 55, 263 analysis) 350--6
Richardson-Zaki correlation 85--6, 234 Silo collapse 301-2, 362
stress and velocity coupling 349
Sampling switch stress (Walkers' analysis) 361-2
extractive 27-8 Walkers' distribution factor 358
in general 26--30 Strouhal number 75--6
in situ 27-8
isokinetic 27-30 Terminal velocity
Sauter mean, see Particle size, averages effect of temperature and pressure 63-5
Segregation 27 glide/tumble motion 75
Settling for non-spherical particles in
collection mechanism in filtration 280-3 creeping flow 65-72
of particle assemblies 85--6 for non-spherical particles at intermediate
Shape, see Particle shape Reynolds' numbers 72-5
Shear cell, see Friction, in bulk solids for non-spherical particles in the Newton's
Sieving, see Particle size measurement law range 75-6
Silo, see Bulk storage vessels, design; Stresses sideslip 76
in silos and hoppers for spheres 59-65
Size, see Particle size, averages; Particle size see also Drag on a particle
distribution; Particle size measurement
Skewness, of a distribution 7 Unsteady motion of single particles, see Drag
Slip correction factor, see Cunningham slip on a particle
correction factor
SolI Van der Waals forces
Spouting, of particles 226, 229-30, 235--6, 239 combining relations 104--7
Stability, of a suspension 1 Derjaguin's approximation 107
Stokes diameter, see Particle diameter geometrical effects 105
Stokes-Einstein equation 79 Hamaker constant 103
Stokes' law, see Drag on a particle Lifshitz theory 104
Stokes number pair potential 99-101
in acceleration and curvilinear motion 82-5 practical calculations 108-9
in agglomeration 253-7 Von Karman vortex street 58
in filtration 281 Virtual mass, see' Added mass'
in impaction 45--6
in isokinetic sampling 29 Wadell sphericity, see Particle shape
Stopping distance, for a particle 81
see also Stokes number Yield locus, see Friction in bulk solids