Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 31

Flow through Packed Beds and Fluidized Beds

Chemical engineering operations commonly involve the use of packed and fluidized beds. These
are devices in which a large surface area for contact between a liquid and a gas (absorption,
distillation) or a solid and a gas or liquid (adsorption, catalysis) is obtained for achieving rapid
mass and heat transfer, and particularly in the case of fluidized beds, catalytic chemical reactions.
First, let us consider flow through a packed bed.

Packed Beds:
A typical packed bed is a cylindrical column that is filled with a suitable packing
material. You can learn about different types of packing materials from Perrys
Handbook. The liquid is distributed as uniformly as possible at the top of the
column and flows downward, wetting the packing material. A gas is admitted at
the bottom, and flows upward, contacting the liquid in a counter-current fashion.
An example of a packed bed is an absorber. Here, the gas contains some carrier
species that is insoluble in the liquid (such as air) and a soluble species such as
carbon dioxide or ammonia. The soluble species is absorbed in the liquid, and
the lean gas leaves the column at the top. The liquid rich in the soluble species
is taken out at the bottom.
From a fluid mechanical perspective, the most important issue is that of the
pressure drop required for the liquid or the gas to flow through the column at a
specified flow rate. To calculate this quantity we rely on a friction factor
correlation attributed to Ergun. Other fluid mechanical issues involve the proper
distribution of the liquid across the cross-section, and developing models of the
velocity profile in the liquid film around a piece of packing material so that
heat/mass transfer calculations can be made. Design of packing materials to
achieve uniform distribution of the fluid across the cross-section throughout the
column is an important subject as well. Here, we only focus on the pressure
drop issue.
The Ergun equation that is commonly employed is given below.
fp = 150/Rep + 1.75

Here,
Here, the friction factor pf for the packed bed, and the Reynolds number pRe,
are defined as follows.

f p=

p D p 1
L V 2s 3

( )

Re =
p

Dp V s
( 1 )

The various symbols appearing in the above equations are defined as follows.
p: Pressure Drop

L: Length of the Bed


DP: Equivalent spherical diameter of the particle defined by D P = 6

Volume of the particle


Surface area of the particle

: Density of the fluid


: Dynamic viscosity of the fluid

Q
A , where Q is the volumetric flow rate of the fluid

Vs: Superficial velocity =

and A is the cross-sectional area of the bed)


:Void fraction of the bed ( is the ratio of the void volume to the total volume
of the bed)
Sometimes, we may use the concept of the interstitial velocity Vi, which is
related to the superficial velocity by Vi = V s/
The interstitial velocity is the average velocity that prevails in the pores of the
column.
Two simpler results, each obtained by ignoring one or the other term in the
Ergun equation also are in use. One is the Kozeny-Carman equation, used for
flow under very viscous conditions.
fp = 150/Rep when Rep

The other is the Burke-Plummer equation, used when viscous effects are not
as important as inertia.
fp = 1.75,

Rep

1,000

It is suggested that the student simply use the Ergun equation.

Fluidisation & mapping of regimes:


When a bed of solids is kept suspended by fluid up-flow, the bed can behave in
various waysi) smoothly fluidised; ii) Bubbling; iii) Slugging; iv) Spouting
etc.

Before fluidisation of a particle bed in a column tower, the particles will form a
fixed bed and fluidisation sets on when the velocity of the fluid ,either liquid or
gas attains a minimum level and then with the increase of velocity , the process
of fluidisation takes different form as mentioned above. Simultaneously with
different degrees of increase of fluid velocities, the fluid pressure drop gradually
increases with increasing velocities and at certain minimum value of the
velocity , the bed starts expanding and pressure drops starts decreasing and at
the onset fluidisation, the pressure drop becomes constant.

Fixed bed of particles:


Characterization of particles of bed:
The size of spherical particle can be measured without ambiguity; however,
question arises with non-spherical particles. There are various ways of defining
particle sizes. We can adopt deff that is useful for both flow and pressure drop.
For particle size larger than or > 1 mm: This can be measured by means of
caliper, micrometer, mass determination, if density is known. From these
measurements, we can first calculate the equivalent spherical diameter, defined
as follows:
dsph = ( Diameter of spheres having the same volume as the particles)
Various measures of non-sphericities are available, and are summarized in
tables as under

Sphericity:
Sphericity is defined as the ratio of surface of a sphere having the same volume
as the particle to the surface of that particle.

s =

Surface of sphere
( Surface
of particle )

of same volume

With this definition, s=1 for spheres & 0 < s<1 for all other particles
The example of relationship between sphericity, shape factor and porosity is
presented in Table below.

Table 1. Sphericity, shape factor and porosity for granulated


minerals

Description

Sphericity,

Shape factor, S

Porosity,

Spherical

1.0

6.0

0.38

Raunded

0.98

6.1

0.38

With grooves

0.94

6.4

0.39

Pointed

0.81

7.4

0.40

Angular

0.78

7.7

0.43

Crushed

0.70

8.5

0.48

Now,
we

represent a bed of non-spherical p[articles by a bed of spherical particles of


diameter deff such that the two beds have the same total surface area and same
fractional voidage m. This representation should ensure almost the frictional
resistance to flow in these two beds. Then by geometry, we can show that
deff = s dsph
The specific surface of particles in either bed is then found to be
2

'

d / s
SURFACE OF THE APARTICLE
6
= sph3
=
[ m1 ]
VOLUME OF THE PARTICLETHE BED
s d sph
d sph/ 6

and for the whole bed

SURFACE OF ALL THE APARTICLE


=
TOTAL VOLUME OF THE PARTICLESTHE BED

6 ( 1 m )
s d sph

[ m1 ]
Where, s is measured directly, estimated from table above or evaluated by
other procedures.
For irregular particles with no seemingly longer or shorter dimension, d eff = s
dsph= s dp; for very small particle (< 40 m , we cannot use screen analysis.
So we then rely on

Scanning of magnified photographs of particles.

Sedimentation of particles in a known fluid; the terminal velocity of these


particles will give the diameter of the equivalent sphere.

For particles with regular shapes the sphericity may be determined on the base
of that definition. For particles with irregular shapes the sphericity is determined
with approximate methods. One of them is the measurement of three
characteristic dimensions, which may be obtained with optical method, or
density and volumetric, or gas-expansion methods. According to Dharamajah
and Cleasby, as well as Zenz and Othmer, with these dimensions being known,
the sphericity may be calculated from the following formula

s =

r
s. t

Where: r is the smallest dimension of the particle, whereas s and t are the
remaining dimensions.
Intuitively, one can imagine that a relationship exists between sphericity and
porosity, as well as that porosity varies with/during bed compacting. Razumow,
as well as Cooke and Rowe, indicated that, for uniform spherical particles,
porosity depends only on the number of points of contact of individual spheres
and not on grain diameter. Thus, theoretically, the porosity of such bed may
range from 0.259 for strongly compacted bed to 0.476 for a loose bed in
which each sphere has only 6 contact points with the neighboring spheres. As in
practice a bed is never built of uniform and perfect spheres, the porosity may
deviate significantly from these values. The relationship between sphericity and
porosity and the degree of compacting (densification) is presented in Fig. 1.

Porosity:
As the total volume of solid particles forming a fixed bed is smaller than the
volume of the bed, the ratio of free spaces between these particles to bed
volume is called porosity, which may be determined by different methods. Most
frequently, it is calculated as:

where: s specific density of the skeleton, p apparent density calculated as


the ratio of grain mass to grain volume, as if they had no pores.

Fixed beds- one size particles:


According to Brown et al., the fraction void m in a packed bed is related to
particle sphericity; in addition, for vessels of small diameter, the wall effect
becomes important and influences the bed voidage.
The frictional pressure drop, always positive, through fixed beds of length L
containing a single size of isotropic solids of screen size d p has been correlated
by Ergun by equation,

Pfr
Lm

150

( 1 m )
2m

1 m g u 20
+1.75 3
2
m s d p
( s d p )
u0

The measured pressure drop is


pmeasured p fr +

g Lm
gc

all

is positive, where + sign indicates up-flow

of fluid. The last term may be appreciable for flowing liquids, but it can be safely
ignored for flowing gases unless one is dealing with deep beds at high pressure.
Thus , in most cases with gases, we may write,

p= pfr = p measured ,

all

is positive. For randomly packed granular materials, this expression has

been found to represent data within

25% accuracy. However, it may not be

expected to extend this to non-randomly packed beds or to beds of abnormal


void content such as raschig ring, or to highly porous beds( e.g., fibrous beds
where

m =0.6 0.98 . At this high porosity, pressure drop can be much greater

than that expected by Ergun equation. Other procedures for finding pressure
drop are given by Perry, based on Chilton & Colburn, by Carman & by Brown,
based on the work of Brownell & Katz. Browns work account for all ranges of s
&

m .

Fixed beds Solids with size distribution:


Before considering the behaviour of beds containing solids of different sizes, we
must be able to describe usefully the size distribution of a batch of solid
particles. For this, define the size distribution function P and p as follows. Let P
be the volume fraction of the particles smaller than d p, and let p d (dp) be the
volume fraction of particles between dp & dp + d (dp).

dp1

P 1= P d ( d p )
0

&

Minimum Fluidisation Velocity:

Fluidisation without carryover of particles:


Consider a bed of particles resting on a distributor designed for uniform up-flow
of gas for instance, a porous sintered metal plate. As discussed earlier, the
onset of fluidisation occurs when

( Drag force by upward moving gas ) = ( Weight of particles )


Or,

Pressure Crosssectional
fraction Specific weight
Volume of
=
drop
are
consisting
of
Bed
across Bed
of the Tube
of solids
Solids

)(

Or, with

pb A t

)(

)(

)(

always positive,

=W=

A t Lmf ( 1 mf ) ( s g )

g
gc

By rearranging, we find for minimum fluidizing condition that

pb
g
=( 1 mf )( s g )
L mf
gc
At the onset of fluidisation, the voidage is a little larger than that in a packed
bed, actually corresponding to the loosest sate of a packed bed of hardly any
weight. Then, we may estimate

mf from random packing data.

The superficial velocity at the minimum fluidisation condition, u mf, is obtained by


combining the last eqn & Ergun eqn.
Actually , in case of isotropic shaped solids, this gives a quadratic in u mf ;

d p g ( s g ) g
1.75
2 150 ( 1 mf ) d p umf g
=
( d p u mf g / ) +
3
3
2

mf s
mf s
2

Or,
1.75
2 150 ( 1 mf )
( p , mf ) +
( p , mf )= Ar
3
mf s
3mf 2s

Where, Archimedes Number is defined as Ar =

d 3p g ( s g ) g
2

Some authors call this number as Galileo Number, Ga. In the special case, for
very small particles, the eqn simplifies to

umf =

d 2p ( s g ) g 3mf 2s
150
( 1 mf ) , when

p , mf <20

For very large particles,


2
mf

d p ( s g ) g 3mf s
1.75 g

p , mf >1000

Effect of pressure & Temperature:


The effect of pressure has been studied by many investigators and the
observations may be summarized as under:

mf

increases slightly (1-4%) with a rise in operating pressure.

umf

decreases with a rise in operating pressure. However, this decrease is

negligible for beds of fine particles (d p <100


to 40%) for larger particles (dp

360

but becomes significant ( up


). These experimental findings

are consistent with predictions discussed above.

If

u mb
umf

for coarse alumina (dp

450

m ) increases upto 30% for rise in

operating pressure. This suggests that an increase in operating pressure


widens the range of particulate fluidisation in gas-solid system.
The effect of temperature has also been studied by numerous researchers and
also the combined effect of high temperature and high pressure has been
studied which may be summarized below.

mf

increases with temperature for some particles (upto 8% for temp. up to


0

500 C), but seems to be unaffected by temperature for coarse particles.

umf

can be reasonably predicted by the eqn discussed above.

Problems:
1. Calculate the minimum fluidisation velocity

umf

for the bed of sharp sand

particles used by some Plant Engineer and reported as in the figure below.
Data:

Bed : mf =0.55

Ambient Air:

g =0.0012 g /cc ,

= 0.00018g/(cm s)

Solids: Sharp irregular sand, not longish or flattish

d p

= 160

s=0.67 &

s =2.6 g /cc

Solution: Because the particles are small, we use eqn

to find umf

d p ( s g ) g 3mf 2s
150
( 1 mf ) , when

umf =

p , mf <20

0.0162 ( 2.60.0012 ) 980 x 0.553 0.672


150 x ( 0.00018 ) ( 10.55 )

. Thus,

= 4.01 cm/s

Now we have to check whether the above eqn is applicable by calculating


particle Reynolds number Rep,mf =

d p umf g

( 0.016 )( 4.01 ) (0.0012)


0.00018

= 0.43

which <20 ; hence valid. Then conclude that

umf = 4.01 cm/s


Q. Explain why the pressure drop at or during fluidisation remains constant?
The

- versus-

u0

diagram is particularly useful as a rough indication of

the quality of fluidisation, especially visual observations are not possible.

During the operation of fluidisation process, the bed is not visible from outside
always unless arranged for with feasible operation e.g. glass window sort of
arrangement. However, in general such visibility of fluidisation of a bed of solid
particles with either a liquid or a gas, normally pressure drop is measured in a
series of increasing velocity before and after fluidisation, notionally. By plotting
the results, we get graphical nature of the pressure drop vs. superficial velocity
as above.
However, when the particles which are not too small and uniformly sized say
160

sand, is typical of uniformly sized particles that are not too small. For

the relatively low flow rates in a fixed bed, the pressure drop is approximately
proportional to gas velocity, as indicated in the figure and usually reaching a

pmax , slightly higher than the static pressure of the bed. With a

maximum

further increase in gas velocity, the fixed bed unlocks, in other words voidage
increases from m to mf, resulting in a decrease in pressure drop to the static
pressure of the bed. With the gas velocities beyond fluidisation, the bed
expands and gas bubbles are seen to be present, resulting in non-homogeneity.
Despite this gas velocity increase, the pressure drop remains constant. To
explain the fact, it may be described as that the dense gas-solid phase is well
aerated and can deform easily without appreciable resistance. In its
hydrodynamic behaviour, we can take the gas-solid dense phase as a liquid. If a
gas is introduced at the bottom of a tank containing a liquid of low viscosity, we
find that pressure required for injection is roughly the static pressure of the
liquid and is independent of the flow rate of gas. The constancy in pressure drop
in the two situations, bubbling liquid and the bubbling fluidized bed, are
somewhat analogous.
When the gas velocity is further decreased, , the fluidized particles settle down
to form a loose fixed bed of voidage mf . with gas flow eventually turned off, a
gentle tapping or vibration will reduce the voidage to its stable initial value of
m. Usually umf is taken as the intersection of the

pvsu0

line in the above

figure and the corresponding horizontal line to W/A t . Large, fairly regular
fluctuations suggest that slugging is occurring on the other hand, an observed
pressure drop lower than W/At indicates a partly fluidized bed.

Terminal Velocity of particles:


When particle of size dp falls through a fluid, its terminal free fall velocity can be
estimated from fluid mechanics by the expression

ut =

4 dp( ) g
s

3 gc CD

1
2

, where CD is an experimentally determined drag coefficient.

Following useful approximation, the direct evaluation of terminal velocity of


particles can be expressed as considering

d p as the dimensionless particle

size u* is the dimensionless gas velocity such as


( CD

2p )1/3

d =d p

g ( s )
g

1
3

= Ar1/3 =

And u* = u

18 2.3351.744 s
+
0.5
d p2
(d p )

g
( s g ) g

1
3

p
=

AR

1
3

=(

4 p
)
3 CD

1/3

, 0.5 <s < 1

For spherical particles, the above expression reduces to

18 0.591
+
2
0.5
d p ( d p )

, where s = 1

To avoid carryover of particles from a fluidized bed, keep the gas velocity
between ut and umf. In calculating, use the mean diameter d p for the size
distribution actually present in the bed, whereas for u t, use the smallest size
present in the bed in appreciable quantities.
Example 3: Calculate ut sharp irregular sand particles used by an engineer for
the fluidisation operation in the plant and the report of his observation is as
follows:
Data:
Air:

g = 1.2 x 10-3 g/cm3, Viscosity of the gas,

Density of the gas,

1.8 x 10-4 g/cm. s


Sand:

d p

s = 0.67,

= 160 m ,

First, calculate,

d =d p

g ( s )
g

1
3

= 2.60 g/cm3

= 0.0016

0.0012 ( 2.60.0012 ) 980


0.00018

1
3

7.28

And

from

eqn

18 2.3351.744 x 0.67
+
7.282
( 7.28 )0.5

18 2.3351.744 s
+
0.5
d p2
(d p )

= 1.2954 Finally from eqn

ut

u* = u

2g
( s g ) g

1
3

or,

0.00018 ( 2.60.0012 ) 980


0.00122

= u

ut
1
3

( s g ) g
2g

1
3

= 1.2954

= 88 cm/s

Example 4: Predict the mode of fluidisation for particles of density

s =1.5

g/cm3 at superficial gas velocity of u0 = 40 and 80 cm/s


a) dp = 60
b) dp = 450

m ,

g = 1.5 x10-3 ;

g = 1 x10-3 ;

m ,

Solution: For small particles of bed, d =d p

0.006

(1.5 x 103 )( 1.51.5 x 103) (980)

& u0* = u0

( 2.4 x 104 )

g
( s g ) g

1
3

= u0

2 x 10-4 g/cm. s
2 .5 x 10-4 g/cm. s

g ( s )
g

1
3

1
3

= 2.28

( 1.5 x 103 )
2 x 104 ( 1.5 1.5 x 104 ) 980

1
3

= 0.00785 and

1.577, for 40 & 80 cm/s.

For u0 = 40 cm/s, the u0* = 0.00785 indicates that onset of turbulent fluidisation
in an ordinary bubbling bed.
& at u0*=80 cm/s; fast fluidisation occurs which requires a solid circulating
system.

Fluidization:
Fluidization (or fluidisation) is a process similar to liquefaction whereby a
granular material is converted from a static solid-like state to a dynamic fluid-

like state. This process occurs when a fluid (liquid or gas) is passed up through
the granular material.
When a gas flow is introduced through the bottom of a bed of solid particles, it
will move upwards through the bed via the empty spaces between the particles.
At low gas velocities, aerodynamic drag on each particle is also low, and thus
the bed remains in a fixed state. Increasing the velocity, the aerodynamic drag
forces will begin to counteract the gravitational forces, causing the bed to
expand in volume as the particles move away from each other. Further
increasing the velocity, it will reach a critical value at which the upward drag
forces will exactly equal the downward gravitational forces, causing the
particles to become suspended within the fluid. At this critical value, the bed is
said to be fluidized and will exhibit fluidic behavior. By further increasing gas
velocity, the bulk density of the bed will continue to decrease, and its
fluidization becomes more violent, until the particles no longer form a bed and
are conveyed upwards by the gas flow.
When fluidized, a bed of solid particles will behave as a fluid, like a liquid or gas.
Like water in a bucket: the bed will conform to the volume of the chamber, its
surface remaining perpendicular to gravity; objects with a lower density than
the bed density will float on its surface, bobbing up and down if pushed
downwards, while objects with a higher density sink to the bottom of the bed.
The fluidic behavior allows the particles to be transported like a fluid, channeled
through pipes, not requiring mechanical transport (e.g. conveyor belt).
A simplified every-day-life example of a gas-solid fluidized bed would be a hotair popcorn popper. The popcorn kernels, all being fairly uniform in size and
shape, are suspended in the hot air rising from the bottom chamber. Because of
the intense mixing of the particles, akin to that of a boiling liquid, this allows for
a uniform temperature of the kernels throughout the chamber, minimizing the
amount of burnt popcorn. After popping, the now larger popcorn particles
encounter increased aerodynamic drag which pushes them out of the chamber
and into a bowl. The process is also key in the formation of a sand volcano and
fluid escape structures in sediments and sedimentary rocks.
A fluidized bed is formed when a quantity of a solid particulate substance
(usually present in a holding vessel) is placed under appropriate conditions to
cause a solid/fluid mixture to behave as a fluid. This is usually achieved by the
introduction of pressurized fluid through the particulate medium. This results in
the medium then having many properties and characteristics of normal fluids,
such as the ability to free-flow under gravity, or to be pumped using fluid type
technologies.

Application:
The resulting phenomenon is called fluidization. Fluidized beds are used for
several purposes, such as fluidized bed reactors (types of chemical reactors),
fluid catalytic cracking, fluidized bed combustion, heat or mass transfer or
interface modification, such as applying a coating onto solid items. This
technique is also becoming more common in aquaculture for the production of
shellfish in integrated multitrophic aquaculture systems.

Fluidized beds are used as a technical process which has the ability to promote
high levels of contact between gases and solids. In a fluidized bed a
characteristic set of basic properties can be utilised, indispensable to modern
process and chemical engineering, these properties include:

Extremely high surface area contact between fluid and solid per unit bed
volume

High relative velocities between the fluid and the dispersed solid phase.

High levels of intermixing of the particulate phase.

Frequent particle-particle and particle-wall collisions.

Taking an example from the food processing industry: fluidized beds are used to
accelerate freezing in some IQF tunnel freezers (IQF means Individually Quick
Frozen, or freezing unpackaged separate pieces). These fluidized bed tunnels
are typically used on small food products like peas, shrimp or sliced vegetables,
and may use cryogenic or vapor-compression refrigeration. The fluid used in

fluidized beds may also contain a fluid of catalytic type; that's why it is also
used to catalyse the chemical reaction and also to improve the rate of reaction.
Fluidized beds are also used for efficient bulk drying of materials. Fluidized bed
technology in dryers increases efficiency by allowing for the entire surface of
the subject of the drying to be suspended and therefore exposed to the air. This
process can also be combined with heating or cooling, as the application, if
necessary.

Fluidized bed types


Bed types can be coarsely classified by their flow behavior, including: .

Stationary or bubbling bed is the classical approach where the gas at low
velocities is used and fluidization of the solids is relatively stationary, with
some fine particles being entrained.

Circulating fluidized beds (CFB), where gases are at a higher velocity


sufficient to suspend the particle bed, due to a larger kinetic energy of the
fluid. As such the surface of the bed is less smooth and larger particles can
be entrained from the bed than for stationary beds. Entrained particles are
recirculated via an external loop back into the reactor bed. Depending on the
process, the particles may be classified by a cyclone separator and separated
from or returned to the bed, based upon particle cut size.

Vibratory fluidized beds are similar to stationary beds, but add a mechanical
vibration to further excite the particles for increased entrainment.

Transport or flash reactor (FR). At velocities higher than CFB, particles


approach the velocity of the gas. Slip velocity between gas and solid is
significantly reduced at the cost of less homogeneous heat distribution.

Annular fluidized bed (AFB). A large nozzle at the center of a bubble bed
introduces gas as high velocity achieving the rapid mixing zone above the
surrounding bed comparable to that found in the external loop of a CFB.

Fluidized Beds
A fluidized bed is a packed bed through which fluid flows at such a high
velocity that the bed is loosened and the particle-fluid mixture behaves as
though it is a fluid. Thus, when a bed of particles is fluidized, the entire bed
can be transported like a fluid, if desired. Both gas and liquid flows can be
used to fluidize a bed of particles. The most common reason for fluidizing a
bed is to obtain vigorous agitation of the solids in contact with the fluid,
leading to excellent contact of the solid and the fluid and the solid and the
wall. This means that nearly uniform temperatures can be maintained even
in highly exothermic reaction situations where the particles are used to
catalyze a reaction in the species contained in the fluid. In fact, fluidized beds
were used in catalytic cracking in the petroleum industry in the past. The
catalyst is suspended in the fluid by fluidizing a bed of catalytic particles so
that intimate contact can be achieved between the particles and the fluid.

Nowadays, you will find fluidized beds used in catalyst regeneration, solidgas reactors, combustion of coal, roasting of ores, drying, and gas adsorption
operations.

First, we consider the behaviour of a bed of particles when the upward


superficial fluid velocity is gradually increased from zero past the point of
fluidization, and back down to zero. Reference is made to the figure on page
3.
At first, when there is no flow, the pressure drop zero, and the bed has a certain
height. As we proceed along the right arrow in the direction of increasing
superficial velocity, tracing the path ABCD, at first, the pressure drop gradually
increases while the bed height remains fixed. This is a region where the Ergun
equation for a packed bed can be used to relate the pressure drop to the
velocity. When the point B is reached, the bed starts expanding in height while
the pressure drop levels off and no longer increases as the superficial velocity is
increased. This is when the upward force exerted by the fluid on the particles is
sufficient to balance the net weight of the bed and the particles begin to
separate from each other and float in the fluid. As the velocity is increased
further, the bed continues to expand in height, but the pressure drop stays
constant. It is possible to reach large superficial velocities without having the
particles carried out with the fluid at the exit. This is because the settling
velocities of the particles are typically much larger than the largest superficial
velocities used.
Now, if we trace our path backward, gradually decreasing the superficial

velocity, in the direction of the reverse arrows in the figure, we find that the
behaviour of the bed follows the curves DCE. At first, the pressure drop stays
fixed while the bed settles back down, and then begins to decrease when the
point C is reached. The bed height no longer decreases while the pressure drop
follows the curve CEO. A bed of particles, left alone for a sufficient length of
time, becomes consolidated, but it is loosened when it is fluidized. After
fluidization, it settles back into a more loosely packed state; this is why the
constant bed height on the return loop is larger than the bed height in the initial
state. If we now repeat the experiment by increasing the superficial velocity from
zero, well follow the set of curves ECD in both directions. Because of this reason,
we define the velocity at the point C in the figure as the minimum fluidization
velocity V. We can calculate it by balancing the net weight of the bed against
the upward force exerted on the bed, namely the pressure drop across the bed
p multiplied by the cross-sectional area of the bed . In doing this balance, we
ignore the small frictional force exerted on the wall of the column by the flowing
fluid.

Upward force on the bed =p A


If the height of the bed at this point is L and the void fraction is , we can
write
Volume of particles = (1- ) AL
If the acceleration due to gravity is , the net gravitational force on the
particles (net weight) is g
Net Weight of the particles =

( 1 ) ( p p pf ) ALg

Balancing the two yields


p =

( 1 ) ( p p pf )

By using an expression relating p to the superficial velocity, which is the


fluidization velocity at this point, we can obtain a result for the latter.
Typically, for a bed of small particles (Dp 0.1 mm), the flow conditions at
this stage are such that the Reynolds number is relatively small (Re10), so
that we can use the Kozeny-Carman Equation, applicable to the viscous flow
regime, for establishing the point of onset of fluidization. This yields
( p f ) g D 2p 3
Vf=
150 ( 1 )
When the superficial velocity V s is equal to Vf, we refer to the state of the
bed as one of incipient fluidization. The void fraction at this state

depends upon the material, shape, and size of the particles. For nearly
spherical particles, McCabe, Smith, and Harriott (2001) suggest that lies in
the range 0.40 0.45, increasing a bit with particle size.
For large particles (mm), inertial effects are important, and the full Ergun
equation must be used to determine V. When in doubt, use the Ergun
equation instead of a simplified version of it.
Dp f 1
Now, we consider the condition we must impose on the superficial velocity so
that particles are not carried out with the fluid at the exit. This would occur if
the superficial velocity is equal to the settling velocity of the particles.
Restricting attention to small particles so that Stokes law can be used to
calculate their settling velocity, we can write

V settling =

( p f ) g D 2p
18

If we now use the result for the minimum fluidization velocity for the case of
small particles, given above, we see that the ratio

V settling
Vf

25 ( 1 )
3
3

For lying in the range 0.40 0.45, this yields a ratio ranging from78 -50.
McCabe et al. suggest that it is common to operate fluidized beds at velocities as
high as 30

V f , and values as large as 100 V f

are used on occasion.

Recognizing that not all particles are of the same size and that D p is only an
average size, we see that fine particles are likely to be carried out with the
exiting fluid in such a situation. They can be recovered by filters or cyclone
separators and returned, in order to obtain the benefits of operating a bed at
such large superficial velocities.
Fluidization can be broadly classified into particulate fluidization or bubbling
fluidization. Particulate fluidization occurs in liquids. As the velocity of the liquid
is increased past the minimum fluidization velocity, the bed expands uniformly,
and uniform conditions prevail in the liquid solid mixture. In contrast, bubbling
fluidization occurs in gas-fluidized beds. Here, when the bed is fluidized, large
pockets of gas, free of particles, are seen to rise through the bed. Where there
are particles, the bed void fraction is approximately at the value that prevails at
the point of incipient fluidization. The bubbles grow until they fill the crosssection, and then successive bubbles move up the column, a condition known as
slugging.
The above classification should not be interpreted rigidly. Sometimes, very dense
particles in a liquid can show bubbling and gases at high pressure when
flowing through beds of fine particles, can give rise to particulate fluidization.
Usually, this occurs at lower velocities, and at higher velocities, the bed shows
bubbling.

Basic model
When the packed bed has a fluid passed over it, the pressure drop of the fluid is

approximately proportional to the fluid's superficial velocity. Superficial


velocity (or superficial flow velocity), in engineering of multiphase flows and
flows in porous media, is a hypothetical (artificial) fluid velocity calculated as if
the given phase or fluid were the only one flowing or present in a given cross
sectional area. Other phases, particles, the skeleton of the porous medium, etc.
present in the channel are disregarded. In order to transition from a packed bed
to a fluidized condition, the gas velocity is continually raised. For a free-standing
bed there will exist a point, known as the minimum or incipient fluidisation
point, whereby the bed's mass is suspended directly by the flow of the fluid
stream. The corresponding fluid velocity, known as the "minimum fluidization
velocity",
. Beyond the minimum fluidization velocity (
), the bed
material will be suspended by the gas-stream and further increases in the
velocity will have a reduced effect on the pressure, owing to sufficient
percolation of the gas flow. Thus the pressure drop for
is relatively
constant.
At the base of the vessel the apparent pressure drop multiplied by the crosssection area of the bed can be equated to the force of the weight of the solid
particles (less the buoyancy of the solid in the fluid).

where:
is the bed pressure drop
is the bed height
is the bed voidage, i.e. the fraction of the bed volume that is occupied by the
voids (the fluid spaces between the particles)
is the apparent density of bed particles
is the density of the fluidizing fluid
is the gravity acceleration
is the total mass of solids in the bed
is the cross-sectional area of the bed

Geldart Groupings
In 1973, Professor D. Geldart proposed the grouping of powders in to four socalled "Geldart Groups". The groups are defined by their locations on a diagram
of solid-fluid density difference and particle size. Design methods for fluidized
beds can be tailored based upon the particle's Geldart grouping:

Group A For this group the particle size is between 20 and 100 m, and the
particle density is typically less than 1.4g/cm 3. Prior to the initiation of a
bubbling bed phase, beds from these particles will expand by a factor of 2 to 3
at incipient fluidization, due to a decreased bulk density. Most powder-catalyzed
beds utilize this group.
Group B The particle size lies between 40 and 500 m and the particle density
between 1.4-4g/cm3. Bubbling typically forms directly at incipient fluidization.
Group C This group contains extremely fine and consequently the most
cohesive particles. With a size of 20 to 30 m, these particles fluidize under
very difficult to achieve conditions, and may require the application of an
external force, such as mechanical agitation.
Group D The particles in this region are above 600 m and typically have high
particle densities. Fluidization of this group requires very high fluid energies and
is typically associated with high levels of abrasion. Drying grains and peas,
roasting coffee beans, gasifying coals, and some roasting metal ores are such
solids, and they are usually processed in shallow beds or in the spouting mode.

Distributor
Typically, pressurized gas or liquid enters the fluidized bed vessel through
numerous holes via a plate known as a distributor plate, located at the bottom
of the fluidized bed. The fluid flows upward through the bed, causing the solid
particles to be suspended. If the inlet fluid is disabled, the bed may settle, pack
onto the plate or trickle down through the plate. Many industrial beds use a
sparger distributor instead of a distributor plate. The fluid is then distributed
through a series of perforated tubes.
Superficial velocity can be expressed as:

where:

us - superficial velocity of a given phase, m/s

Q - volume flow rate of the phase, m3/s

A - cross sectional area, m2

Using the concept of porosity, the dependence between the advection velocity
of fluid and the superficial velocity can be expressed as (for one-dimensional
flow):

where:

is porosity, dimensionless

v is the advection velocity, m/s.

The local physical velocity can still be different than the advection velocity
because the vector of the local fluid flow does not have to be parallel to that of
average flow. Also, there may be local constriction in the flow channel.

Derivation of Drag Equation:


The drag equation may be derived to within a multiplicative constant by the
method of dimensional analysis. If a moving fluid meets an object, it exerts a
force on the object. Suppose that the variables involved under some
conditions are the:

speed u,

fluid density ,

viscosity of the fluid,

size of the body, expressed in terms of its frontal area A, and

drag force FD.

Using the algorithm of the Buckingham theorem, these five variables can be
reduced to two dimensionless parameters:

drag coefficient CD and

Reynolds number Re.

Alternatively, the dimensionless parameters via direct manipulation of the


underlying differential equations may also be followed.
That this is so becomes apparent when the drag force FD is expressed as part of
a function of the other variables in the problem:

This rather odd form of expression is used because it does not assume a one-toone relationship. Here, fa is some (as-yet-unknown) function that takes five
arguments. Now the right-hand side is zero in any system of units; so it should
be possible to express the relationship described by fa in terms of only
dimensionless groups.
There are many ways of combining the five arguments of fa to form
dimensionless groups, but the Buckingham theorem states that there will be
two such groups. The most appropriate are the Reynolds number, given by

and the drag coefficient, given by

Thus the function of five variables may be replaced by another function of only
two variables:

Where, fb is some function of two arguments. The original law is then reduced
to a law involving only these two numbers.
Because the only unknown in the above equation is the drag force FD, it is
possible to express it as

or
and with
Thus the force is simply A u2 times some (as-yet-unknown) function fc of the
Reynolds number Re a considerably simpler system than the original fiveargument function given above.
Dimensional analysis thus makes a very complex problem (trying to determine
the behavior of a function of five variables) a much simpler one: the
determination of the drag as a function of only one variable, the Reynolds
number.
The analysis also gives other information for free, so to speak. The analysis
shows that, other things being equal, the drag force will be proportional to the

density of the fluid. This kind of information often proves to be extremely


valuable, especially in the early stages of a research project.
To empirically determine the Reynolds number dependence, instead of
experimenting on huge bodies with fast-flowing fluids (such as real-size
airplanes in wind-tunnels), one may just as well experiment on small models
with more viscous and higher velocity fluids, because these two systems are
similar.
In fluid dynamics, the drag equation is a formula used to calculate the force of
drag experienced by an object due to movement through a fully enclosing fluid.
The formula is accurate only under certain conditions: the objects must have a
blunt form factor and the fluid must have a large enough Reynolds number to
produce turbulence behind the object. The equation is

is the drag force, which is by definition the force component in the direction
of the flow velocity,
is the mass density of the fluid
is the velocity of the object relative to the fluid,
is the reference area, and
is the drag coefficient a dimensionless coefficient related to the object's
geometry and taking into account both skin friction and form drag.
The equation is attributed to Lord Rayleigh, who originally used L2 in place of A
(with L being some linear dimension).
The reference area A is typically defined as the area of the orthographic
projection of the object on a plane perpendicular to the direction of motion. For
non-hollow objects with simple shape, such as a sphere, this is exactly the same
as a cross sectional area. For other objects (for instance, a rolling tube or the
body of a cyclist), A may be significantly larger than the area of any cross
section along any plane perpendicular to the direction of motion. Airfoils use the
square of the chord length as the reference area; since airfoil chords are usually
defined with a length of 1, the reference area is also 1. Aircraft use the wing
area (or rotor-blade area) as the reference area, which makes for an easy
comparison to lift. Airships and bodies of revolution use the volumetric
coefficient of drag, in which the reference area is the square of the cube root of
the airship's volume. Sometimes different reference areas are given for the
same object in which case a drag coefficient corresponding to each of these
different areas must be given.
For sharp-cornered bluff bodies, like square cylinders and plates held transverse
to the flow direction, this equation is applicable with the drag coefficient as a
constant value when the Reynolds number is greater than 1000. For smooth
bodies, like a circular cylinder, the drag coefficient may vary significantly until

Reynolds numbers up to 107 (ten million).

Drag of Blunt Bodies and Streamlined Bodies


A body moving through a fluid experiences a drag force, which is usually divided
into two components: frictional drag and pressure drag. Frictional drag
comes from friction between the fluid and the surfaces over which it is flowing.
This friction is associated with the development of boundary layers, and it
scales with Reynolds number as we have seen above. Pressure drag comes from
the eddying motions that are set up in the fluid by the passage of the body. This
drag is associated with the formation of a wake, which can be readily seen
behind a passing boat, and it is usually less sensitive to Reynolds number than
the frictional drag. Formally, both types of drag are due to viscosity (if the body
was moving through an an inviscid fluid there would be no drag at all), but the
distinction is useful because the two types of drag are due to different flow
phenomena. Frictional drag is important for attached flows (that is, there is no
separation), and it is related to the surface area exposed to the flow. Pressure
drag is important for separated flows, and it is related to the cross-sectional
area of the body.
We can see the role played by friction drag (sometimes called viscous drag) and
pressure drag (sometimes called form drag or profile drag) by considering an
airfoil at different angles of attack. At small angles of attack, the boundary
layers on the top and bottom surface experience only mild pressure gradients,
and they remain attached along almost the entire chord length. The wake is
very small, and the drag is dominated by the viscous friction inside the
boundary layers. However, as the angle of attack increases, the pressure
gradients on the airfoil increase in magnitude. In particular, the adverse
pressure gradient on the top rear portion of the airfoil may become sufficiently
strong to produce a separated flow. This separation will increase the size of the
wake, and the pressure losses in the wake due to eddy formation Therefore the
pressure drag increases. At a higher angle of attack, a large fraction of the flow
over the top surface of the airfoil may be separated, and the airfoil is said to be
stalled. At this stage, the pressure drag is much greater than the viscous drag .

When the drag is dominated by viscous drag, we say the body is streamlined,
and when it is dominated by pressure drag, we say the body is bluff. Whether
the flow is viscous-drag dominated or pressure-drag dominated depends
entirely on the shape of the body. A streamlined body looks like a fish, or an
airfoil at small angles of attack, whereas a bluff body looks like a brick, a
cylinder, or an airfoil at large angles of attack. For streamlined bodies, frictional
drag is the dominant source of air resistance. For a bluff body, the dominant
source of drag is pressure drag. For a given frontal area and velocity, a
streamlined body will always have a lower resistance than a bluff body. For
example, the drag of a cylinder of diameter $D$ can be ten times larger than a
streamlined shape with the same thickness (see figure 1).

You might also like