Sei sulla pagina 1di 12

Powder Technology 67 (1991) 163-174

163
The behaviour of L-valves with granular powders
D. Geldart and P. J ones
School of Powder Technology, Departmen of Chemical Engineering, Universiw of Bradford, Bradford, West Yorkshire
BD7 I DP (UK)
(Received July 19, 1990; in revised form February 4, 1991)
Abstract
An extensive experimental program was carried out on three sands, all group B powders, in L-valves
of diameter 40, 70 and 100 mm. The radius of the elbow, the number of aeration points and the
inclination of the horizontal section influence the solids flow rate. In a system in which both the feed
hopper and the discharging solids are at atmospheric pressure, the maximum solids fluxes achievable
can be calculated from hopper discharge correlations and are in the range 600- 1200 kg/(m*.s). The
minimum aeration gas requirement is determined by the incipient fluidization velocity of the powder.
Correlations are given from which the aeration gas flow rate and the pressure drop across the L-valve
can be calculated for any solids flux, and a design procedure is given for systems using group B solids.
Introduction
L-valves belong to a class of devices for controlling
the flow of granular materials with injected gas alone.
These non-mechanical devices also include J- and
V-valves, and because there are no moving parts in
contact with the solids, they are therefore particularly
suitable for use in environments which are hostile
to materials of construction, such as high temper-
atures, abrasive particles and corrosive gases. They
are cheap to construct, do not seize up, and can be
easily replaced.
In this study, carried out under cold conditions,
the objectives were to study the influence of pipe
diameter and type of solids on the operability of
the L-valve system, to understand the parameters
which influence its behaviour and to devise a strategy
for design and scale-up.
Equipment
The basic solids circulation system is shown in
Fig. 1. It consists of a mechanical conveyor, a large
hopper, an L-valve and an inclined chute. Both the
top of the feed hopper and the discharge end of
the L-valve are vented to atmosphere. The tubular
chute, of diameter 150 mm, is inclined downwards
to give free flow of solids into the elevator feed
hopper and has a segment cut out of its upper surface
half way down its length. Rotation of the chute
AERO-MECHANICAL
ELEVATOR
AIR-SLIDE
Fig. 1. Solids circulating system.
through 180 on its axis allows the solids to be
diverted into a container for a time, so that mass
flow rates of the solids can be measured. Some early
experiments were done using an air slide instead of
the chute.
0032-5910/91/$3.50 0 1991 - Elsevier Sequoia, Lausanne
164
The arrangement of the various L-valves is shown
in Fig. 2. The aeration point (AP) is in every case
50 or 100 mm above the centre line of the horizontal
section. The pressure tappings contain a small wad
of cotton wool to prevent ingress of particles and
are connected to water manometers by rubber tubing.
In most cases, there are four tappings around the
circumference at each of the lower levels. With one
exception, all the L-valves are made of Perspex to
permit visual observation of the flow. The pipes are
40, 70 and 100 mm internal diameter, have 555mm
long horizontal sections and 3..52- or 3.81-m long
vertical standpipes. In addition, two 40-mm internal
diameter steel pipes were used, one being fitted with
a flexible elbow to allow the horizontal section to
be inclined downwards at various angles. A limited
number of tests was carried out in which a 280-mm
long horizontal section of 20-mm bore Perspex tube
replaced the same length of 40-mm bore pipe, thus
providing a constriction in the discharge end of the
L-valve.
The feed hopper consists of a 600-mm diameter
cylinder with a lower conical section having a half-
angle of 30. For the powders used, this acted as a
mass flow hopper. Smaller conical sections could be
40mm
40mm
c
65s
St eel L-Vdv. wt t h
bolted to the bottom flange of the main cone to
match up with the different sizes of downcomer.
This general arrangement has several advantages
over the more usual method involving return of solids
to the feed hopper by pneumatic conveying.
- Solids mass flow rates can be measured directly,
eliminating the need to use calibrations or mea-
surement of particle velocities at the wall, both of
which can be unreliable.
- The pressure balance around the loop depends
only on the L-valve operation and not on pressure
drops in the return loop.
- Attrition of the solids is negligible.
Solids used
The full experimental programme involved four
sands, vermiculite, and a fine coal char. However,
the behaviour of the group A powders (one of the
sands, the vermiculite and the char) wasvery different
from that of the group B solids and will be discussed
in a separate paper. The properties of the three
sands discussed in this paper are summarized in
Table 1.
The mean sieve sizes are calculated from
-Q
%
s
49
%
-a 2
Q i
-AP o-
p,
---- -
-
70 and 1OOmm
Perspex L-valves
::
h
1
43
/
0
L-Val va w i t h Fi x ad 9O El bow Fl ex l bl . El bow Dbn* nml ons I n mm
Fig. 2. Dimensions and positions of pressure tappings in 40-, 70- and loo-mm diameter valves.
16.5
TABLE 1. Properties of the powders used
Property Sands
Sl s3 s4
Mean sieve size ci, (pm) 280 500 790
Surface-volume diameter 2, (em) 243 435 650
Particle density pr (kg/m3) 2645 2604 2661
Bulk densities:
Loosely packed paLp (kg/m3) 1562 1.527 1534
Tapped IJ~T (kg/m3)
1668 1685 1687
Voidages:
l LLP
0.409 0.413 0.423
Q
0.369 0.353 0.366
Minimum fluidization
velocity U,, (cm/s) 6.4 20.4 40
1
dp= -
P
and the mean surface-volume diameter 8, was ob-
tained by measuring the pressure drop across a fixed
bed of the powder and back-calculating d,, using
the Ergun equation. The minimum fluidization ve-
locities U,r and the tapped and loosely packed bed
voidages were also separately measured.
Experimental procedure
After filling the system with the chosen solids, the
aeration gas flow rate was set at a value which gave
a high solids circulation rate, and the rig was run
for a few minutes. It was then shut down and restarted
at a low air flow so that solids just began to trickle
from the discharge. Invariably, after a short time,
the solids flow stopped and the air had to be increased
until a condition of minimum continuous discharge
prevailed. The solids flow rate was measured five
times and an average taken, the air flow rate noted,
and the pressures up the standpipe recorded. The
aeration rate was increased by stages until a maximum
was reached. No hysteresis was observed on de-
creasing the air flow. In the 40- and 70-mm diameter
pipes, the maximum mass flow was heralded by the
disappearance of the moving bed in the upper part
of the standpipe, and its replacement with streaming
solids flow. In the lOO-mm diameter valves, the
maximum was limited by the capacity of the me-
chanical conveyor, which was about 3 kg/s. Another
constraint on the solids flow rate was that with the
coarsest sand, in the 70- and lOO-mm diameter pipes,
sonic velocity was reached in a single aeration nozzle,
thus limiting the air flow. Because it was not prac-
ticable to increase the diameter of the nozzle, we
used three nozzles at the same level.
Results
General observations
On the whole, our visual observations agree with
those made in the pioneering work of Knowlton and
Hirsan [l] and are illustrated in Fig. 3. At very low
flow rates (Fig. 3(a)), there is a narrow, fast-moving
stream of solids at the top of the horizontal pipe
and virtually no movement detectable in the stand-
pipe. The width of the channel increases as the size
of the sand in the rig is increased. However, there
is significant segregation by size in the horizontal
section, with the smallest particles in the size dis-
tribution predominating at the top of the pipe.
At medium solids flows, dunes or ripples move
across the top of the pipe at a frequency of about
4 per second and cause fluctuations both in the
solids discharge rate and also in the pressures mea-
sured just above the aeration position.
At high solids rates, the dune frequency falls to
about 1 per second. When the maximum solids flow
is reached, accompanied by streaming and flow fluc-
tuations, any further increase in gas injection causes
a sharp reduction in solids flow. Knowlton and Hirsan
[l] attributed this to the fluidization which occurred
in their downcomer when the pressure gradient in
la)
LOW
AERATI ON
RATE
l b)
MEDI UM
AERATI ON
RATE
I CI
HI GH
AERATI ON
RATE
Fig. 3. Mode of solids discharge from L-valves.
166
it reached bMF*g. These large pressure gradients
could occur in their rig because the L-valves dis-
charged into a vertical pneumatic conveying line
across which the pressure drop increased with the
solids discharge rate.
In our rig, the discharge end of the L-valves was
always at atmospheric pressure and we believe that,
under these circumstances, it is the size of the hopper
outlet which limits the solids discharge rate. If the
L-valve delivers a greater solids flow to the discharge
end than is fed to it by the hopper outlet, the
standpipe starts to empty and a bed level appears
at the top of the standpipe and moves downwards
with streaming solids above it. The resistance to gas
flow above the aeration point is reduced and some
of the aeration gas passes up the standpipe, leaving
less available to transport the solids through the
horizontal section and causing the discharge rate to
fall momentarily. The level, and the resistance to
flow above the aeration point, then increase, and
more injection gas passes downwards. Thus, the
system starts to operate like an automatic L-valve
[2], but with flow instability. Further increase in
aeration gas flow then starts to choke off the flow
from the hopper outlet (injecting air just below a
hopper outlet has been used successfully as a method
of controlling solids egress [3]).
To check the validity of this hypothesis, we cal-
culated the maximum flow rates out of a hopper
using the Carleton [4] equation.
g=
4V. sin p + 15p1cL2/3VS4B
D ,a;~
where V, =4MJ(peLprrD2).
The results are shown in Table 2A together with
the maximum fluxes achieved in the various L-valves
using the three sands. In addition, a separate test
was carried out in which the 40-mm diameter down-
comer was removed and the flow rate of sand Sl
measured as it flowed freely out of the hopper. It
can be seen that where comparisons are possible,
the maximum L-valve throughput is between 84 and
97% of the maximum flows predicted by the Carleton
equation.
In Table 2B, the highest solids fluxes quoted in
Knowlton and Hirsans paper for four different valve
sizes are compared with values calculated from
Carletons equation. Their equipment layout diagram
shows a ball valve of unspecified diameter below the
hopper outlet, and this may have reduced the max-
imum solids flux attainable; with the exception of
the 6-in valve, the maximum solids fluxes attained
were between 47 and 80% of the predicted values.
There are, therefore, reasonable grounds for be-
lieving that when the L-valve discharges at atmos-
pheric pressure the maximum solids flux capable of
passing through it is determined by the size of the
hopper outlet, or of any other restricting orifice,
such as a ball or slide valve, below it.
If, however, the L-valve has to feed against an
appreciable back pressure, such as a fluidized bed
or a pneumatic conveying line, a lower maximum
operating flux will be reached when the pressure
gradient in the downcomer approaches that of the
fluidized solids.
Effect of L-valve geometry
Elbow construction
In all fluid-particle operations, visual observation
is very helpful in bringing about an understanding
of the hydrodynamics, and it is for this reason that
most of our L-valves are made of Perspex. The
method of construction used, two tubes cut at 45
and cemented together, resulted in sharp 90 elbows.
In commercial applications, equipment has to be
made of steel, or some other robust material, and
the elbow may not be sharp, especially if the pipes
are lined with insulating bricks or cement. We there-
fore replaced the entire 40-mm diameter Perspex
system by 40-mm diameter smooth-walled steel tubing
and used a standard small-radius cast iron 90 elbow
to connect the standpipe to the horizontal section.
Because we also wanted to investigate the effect of
using a sloping horizontal section, we subsequently
replaced the cast elbow by a piece of wire-reinforced
flexible plastic pipe which, in effect, gave a large
radius elbow.
As shown in Fig. 4 for the 90 elbows (curves 1,
2, 3) at a given aeration rate, the solids discharge
rate increases as the elbow geometry changes from
sharp to gradual. Conversely, slightly less gas is
required to transport the same solids rate when the
bend is gradual. The gradual bend also improved
sensitivity of control near the minimum discharge
rate condition. It was observed that at high solids
discharge rates, considerable electrostatic charging
occurred when using the Perspex L-valves. However,
it seems unlikely that this was the cause of the
difference between the sharp (Perspex) bend and
the more gradual (steel) bends, since the biggest
differences in solids flow occurred when the elec-
trostatics were least evident.
I nclined elbows
Sloping the horizontal section downwards (curves
4, 5 and 6 in Fig. 4) increases the solids flow rate
at any given aeration rate and gives greater flow
stability and control at low discharge rates. However,
both with the 280-pm (Fig. 4) and the SOO-pm sand
(Fig. S), at an angle of 106, flow becomes uncon-
167
TABLE 2. Comparison of maximum solids flow rate from Gvalves with maximum flow rate from a hopper calculated
using Carletons equation
A This work
Diameter of L-valve (mm) 40 70 100
Mean sieve size of sand (pm) 280 500 790 280 500 790 280 500 790
Max. solids flux from 637 637 - 706 792 > 792 > 381
L-valve (measured) =3 kg/s
kg/(m* . s)
Maximum flow out Predicted (eqn. (2)) 652 660 676 839 86.5 899 980 1026 1069
of hopper Measured
780 _ _ _ _ _ _ - -
kg/(m.s)
Limited by capacity of conveyor.
B Knowllon and Hirsan [l]
Diameter of L-valve (in)
Mean sieve size of powder (pm)
Largest solids flux from
Lvalve quoted in paper
kg/(m* . s)
Maximum flow out of
hopper predicted by
Carleton equation
kg/(m* .s>
1.5 2 3 6 3
260-pm sand 188~pm siderite
354 324 553 196 875
595 684 833 1071 1100
250

250 500 0 500 1000
Aeration rate (cms/sl
Aeration rate (cm 3/s1
Fig. 4. Discharge rates of 280-pm sand from 40-mm L-
valves with various elbow configurations.
Fig. 5. Discharge rates of 500~pm sand from 40-mm steel
L-valve with various elbow configurations and one aeration
point.
trollable at about two-thirds of the maximum solids
flow rate and no further increase is possible. The
angles of repose of unaerated sands are shown on
Fig. 6, but in practice the angles of partly-aerated
Elbow angle
0 goo
e g4o
sands will be lower. This, and visual observation,
leads us to believe that the inability of the 106 valve
to deliver the maximum solids rate is caused by
extensive bypassing of gas along the top of the tube.
168
500~m5ANO [----I ,4.26
GLE OF REWSE
DI AGRAM : A B C 0
ELBOWANGLE : 90' 94' 98' 106'
Fig. 6. Fraction of L-valve discharge section occupied by solids at various elbow angles.
Construction of horizontal section
In a limited number of tests using sand Sl, the
final 280 mm of the horizontal section of the 40-
mm diameter Perspex L-valve was replaced by an
equal length of 20-mm diameter tube positioned
concentrically. The effect is shown in Fig. 7. The
ability to control finely the solids flow rate was much
improved, particularly at low aeration rates. Thus,
a steady minimum discharge of 6 g/s was possible
with the constriction in place compared with 35
g/s in the basic valve. However, the maximum dis-
charge rate achievable was much reduced. In contrast,
Knowlton and Hirsan [l] reduced the diameter of
l Wlthout constriction
q With constriction
to 20 mm I . D.
1
\
p \
I
I L
/
Aeration rate Icm3/d
Fig. 7. Discharge rates of 280-pm sand from 40-mm Perspex
Lvalve showing effect of constriction on horizontal section.
their entire horizontal section and found that it had
little effect upon the discharge rate for a given
aeration rate.
Number of aeration points
Most of the experiments were done using only
one aeration point positioned on the outside of the
elbow. However, when using coarse sand S3 in the
larger valves, high air flow rates were required, and
sonic velocity was approached in the injection nozzle.
Since the supply pressure was limited to 5 bar, this
restricted the air flow rate to less than about 1.5
l/s. With the thin-walled Perspex standpipes, it was
not practicable to use larger nozzles, so we resorted
to splitting the air supply between three injection
nozzles with that on the inside of the elbow not
being used. Some of the results are shown in Fig.
8. Because of the sonic velocity limitation described
above, as expected, higher maximum air and solids
rates were achieved with three nozzles than with
Fig. 8. Discharge rates of 280-, 500- and 790-pm sand
from 70-mm L-valve showing effect of number of aeration
points.
169
only one. What is surprising though is that at any
given aeration rate the solids rates were somewhat
lower when using three aeration points. It is con-
jectured that the air from the two nozzles at 90 to
the single nozzle was bypassing the inside of the
elbow. Thus, although all the injected air is used in
dense phase transport in the horizontal section, much
less is available to assist solids flow at the inside of
the elbow.
Effect of pipe diameter and particle size
Knowlton and Hirsan [l] noted that more aeration
gas is required to achieve a given solid flow rate as
particle size and pipe diameter increase. This is
confirmed by our data shown in Figs. 8 and 9,
respectively. As Knowlton [2] pointed out, the total
gas flow through the valve, Qt, is larger than that
injected because in almost all cases a gas flow, Q,,,
is entrained downwards in the void spaces between
the particles so that
Qt = QDC + Qeti
(3)
Knowlton and Hirsan [l] showed how Qnc may be
calculated from the pressure gradient in the down-
comer, and we shall address that question later in
the paper. Until now, however, there has been no
way of correlating Qext with solids flow rate.
The data for the 280-pm sand are replotted in
Fig. 10 as solids mass flux G, versus the superficial
velocity V,, of the aeration gas. The form of the
4000
3000
z
2
0
E
I
; 2000
f
.E
=
*
P
3
1000
0
00 0
Aeration rate (cm3/s1
Fig. 9. Discharge rate of 280-wrn sand VS. aeration rate
for 40-, 70- and lOO-mm Perspex L-valves with one aeration
point.
1000 -
800 -
70mm 9
Uert km/s)
Fig. 10. Solids discharge flux for 280-pm sand vs. aeration
rate expressed as superficial gas velocity in the L-valve.
curves in Fig. 10 suggested that the data might lie
on one curve if GJ D were plotted instead of G,
and, indeed, the use of GJ D for the correlation of
both dilute and dense phase pneumatic conveying
data has a sound theoretical basis [5]. When data
for the other sizes of sands were plotted in this way,
it was noticed that the minimum superficial aeration
velocities were similar to the minimum fluidization
velocities for the respective sands, and this gave the
idea of plotting GJ D against U,,lU,,,,. The results
are shown on Fig. 11 for all the sharp 90 elbows.
The least-squares fitted equation is
GS
- =3354+ -2965
D
(4)
mf
The results of Knowlton and Hirsan [l] were not
included in calculating the correlation, but they are
shown on Fig. 12 compared with eqn. (4). Apart
from the data for their finest sand, the agreement
is quite good. It follows from earlier comments that
eqn. (4) gives minimum values for G,/D since having
elbows rounded at the inside, or horizontal sections
which slope downwards, give higher discharge rates
for a given aeration rate.
Pressure profiles and gas J ?OW in the downcomer
The maximum pressure in the rig occurs at the
aeration point, and for all three sands it decreases
linearly with increasing vertical height (Fig. 13).
Because the profiles are linear, the pressure gradients
can be calculated and cross-plotted against aeration
170
V?llW Symbol
0
0 1 2 3 4 5 6
hxt /b F
Fig. 11. G,lD vs. U,,llJ,, for sharp 90 elbows - this
work.
is
E
r
y
5000
e
d
0
4
Aeration
0
0 500 1000
Pressure fmmWGl
Fig. 13. Pressure profiles with 280-pm sand in 40-mm
diameter Perspex L-valve (one aeration point).
.260 jlnl sand
. 188 urn slderlte
.
. /
H Equation 4
.
/
.
.
/_, , , ,
0 1 2 3 4 5 6 7
Uext h4F
Fig. 12. G,D vs. Ue,lUMF Knowlton and Hirsan [l] data
compared with eqn. (4).
rate, thus permitting a convenient summary of most
of the data on two graphs (Figs. 14, 15). It should
be noted that even at the maximum aeration rates,
which, on the 40- and 70-mm diameter valves cor-
respond to the maximum solids rates permitted by
the hopper outlet, the maximum pressure gradients
are generally below 2 000 N/m2 per metre. This
should be compared with the values which would
be achieved if the downcomer were to be fluidized,
and which can be calculated from
= /%MFg = f%LPg
\ J-DC jrnt
s
c 1500
t
i;
E
P
8
.r 1000
2
5
2
m
?
3
0
t 500
0
0 500 1000
Aeration rate km3/sI
Fig. 14. Pressure gradients VS. aeration rates for 280~pm
sand in Perspex L-valves of various diameters.
Assuming &MF is equal to the loosely packed densities
of 1562 and 1527 kg/m3 for the 280- and 500-pm
sands, respectively, the fluidized bed pressure gra-
dients would be 15 323 and 14 980 N/m2 per metre,
171
where
v,= G,
PBLP
3000
1
7
.+
E
40 mm Isteel. 900)
\ D
t
70 mm
f 2000
8
5
G
0
.E
f
100 mm
E
4 1000
2
a
f
v
IO1 , , )
0 1000 2000 3000
Aeration rate Icm3/s)
Fig. 15. Pressure gradients vs. aeration rates for SO&pm
sand in Gvalves of various diameters.
respectively, factors of about 7 higher than the highest
values measured. Clearly then, the solids are far
from being in a fluidized condition in the downcomer
even at the highest solids rate.
At first sight, the form of the pressure profiles
seems to imply that there is a flow of gas up the
downcomer. However, as shown by Knowlton [2], a
decrease in pressure from bottom to top of the
downcomer will also be obtained if the velocity of
the gas relative to the particles (often called the slip
velocity) is upwards. In the general case, the relative
velocity between gas and particles is
v,=
.!h
ELP
(6)
(7)
and lJ, is the superficial velocity of gas in the standpipe
relative to the wall.
Consider case (a) in Fig. 16, in which the downward
direction is positive; because there is a net flow of
gas upwards relative to the wall (- V), to an observer
travelling with the particles the gas velocity relative
to the particles
Iv,1 = K - (- V,)
(8)
=v,+vs
In case (b), the gas is travelling downwards relative
to the wall (in the same direction as the particles)
at a velocity V,, but because it is moving more slowly
than the particles, its velocity relative to an observer
+
! I-IV, Down is the positive direction
_i
---
I
j l+)Vs
t
--I-
t
i
IV,1
WV.
-_ --
(b)
Fig. 16. Relative gas velocities in a downcomer. (a) Gas
travels upwards relative to wall; (b) gas travels downwards
relative to wall.
on the particles is again upwards and is given by
lKl=K-v, (9)
In both the cases shown, because the gas is moving
upwards relative to the particles, the pressure de-
creases from bottom to top, and may be calculated
from the Ergun equation. It should be noted that
I V,1 is an interstitial velocity, whereas the Ergun
equation is normally written in terms of the superficial
velocity U,, where V, =U,/E. Written in terms of V,,
and assuming that the voidage in the downcomer is
equal to the loosely packed voidage eLp, we have
If we know the pressure gradient AP&LDC, I V,1 can
be calculated from eqn. (10). If AP decreases from
bottom to top, then the gas is moving upwards relative
to the particles. However, the absolute direction and
magnitude of the gas (relative to the wall) must be
found by calculating V, from eqn. (6) and inserting
it in eqn. (ll), which is obtained by rearranging eqn.
(5), i.e.,
vg=K-lKl
(11)
If V, is negative, then some of the aeration gas
passes upwards from the aeration position whilst the
rest moves downwards through the valve. If V, is
positive, then all the aeration gas, Q_.., passes down-
wards together with a flow Qno entrained by the
solids, where
QDC = Vd ELP
(12)
Absolute interstitial gas velocities V, have been cal-
culated for the 280-pm and 790-pm sands using
experimental data and are shown on Fig. 17. For
the 280-pm sand the procedure was as follows:
- From Fig. 14, find the aeration rate Q,, cor-
responding to a chosen pressure gradient AP,,,I L,,.
172
5
B
10 -
1
15 -
Fig. 17. Gas and particle velocities in 70-mm diameter
downcomer for 280~pm and 790~pm sands.
- From Fig. 8, find the solids mass flow rate
corresponding to this value of Qea and express it
as a mass flux G,.
- Use eqn. (6) to calculate V,, which is always
positive.
- Insert APDclLDc into eqn. (10) and calculate IV,].
- Calculate VP from eqn. (11). If V, is positive, the
net flow of gas above the aeration point in the
downcomer is in the same direction as the solids,
i.e., downwards and appears on Fig. 17 below the
x-axis. Conversely, if V, is negative, then, relative to
the wall, gas is travelling upwards.
As Fig. 17 shows, there are critical fluxes G,r and
Gsz above which gas is carried downwards from the
hopper. The influence of particle size is quite marked:
for the smaller sand, G,r =20 kg/(m.s) and for the
larger GsZ.= 165 kg/(m.s). At a flux of 340 kg/(m*.s),
the gas entrained by the 280-pm sand constitutes
46% of the external aeration air, whilst in the 790-
pm sand it is less than 3%. In order to confirm that
the gas flow directions were indeed as calculated in
Fig. 17, it was decided to inject CO2 as a tracer gas
into the 280-pm sand (i) at the top of the downcomer
and (ii) mixed with the aeration gas at the bottom.
The experimental set up and details of the tests are
described in detail elsewhere [6]. Briefly, a cylinder
of pure CO2 was used to provide an injection mixture
of 4 000 ppm CO2 in compressed air. Initially, the
CO2 was injected through pressure tapping 7 (see
Fig, 2) situated flush with the wall, and gas samples
were taken at various heights and peripheral positions
around the tube. However, it was found that at high
solids flows the CO2 did not disperse radially, since
none was detected at the opposite wall, even at the
bottom of the downcomer. Subsequently, when in-
jecting CO2 at the top, an injection tube was inserted
to a position beyond the pipe centre line, and good
radial dispersion was achieved. When CO2 was in-
jected at the top of the downcomer and samples
taken at position 1, CO2 concentrations above back-
ground were undetectable up to 13 kg/(m**s), but
were -3 500 ppm at 26 kg/(m*.s) and higher. Con-
versely, when injecting CO2 at the aeration point
and sampling at position 4, CO2 concentrations were
-4 000 ppm when the solids fluxes were less than
13 kg/(m**s) or higher. Thus, the experiments were
broadly in agreement with the predicted critical flux
value of 20 kg/(m*.s).
Correlation of pressure drop
The data for pressure drops AP, between the
aeration point and the solids discharge were com-
pared with the Wen and Simons [7] correlation.
However, because their predicted values were much
too low, we correlated our data against valve diameter,
mean sieve size and mass flux of the solids. The
densities of the sands were so similar that inclusion
of 4 in the correlation could not be justified. The
best fit was obtained from
216G . 1
2 (N/m3) = s Y
~0. 63 d 0. 15
P
(13)
where LN is the length of the horizontal section of
the valve measured from the injection point in the
downcomer. However, LH was not varied, so all our
data refer to a length of 0.55 m.
Knowlton and Hirsans [l] data are of limited use
for comparative purposes, because they are expressed
in terms of pressure drop per unit length of down-
comer, and the system included a vertical lift line
followed by a bend and a horizontal pipe section
in addition to the horizontal L-valve section. Thus,
the pressure gradient in the downcomer in their
system would change not only in response to changes
in A? across the L-valve elbow and horizontal section,
but also because of changes in AP across the lift
line and upper bend. Nevertheless, the qualitative
dependencies of eqn. (13) are in agreement with
their work, which found that the effects on APu of
mass flux and particle size were small and propor-
tional to (G&J, where m is small.
Design procedure
We are now in a position to set out a stepwise
procedure for the design of an L-valve for any
particular configuration, mass flow rate and group
B solid.
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
Identify the mean sieve size and particle density
of the solids to be controlled. Check that they
belong to Geldarts Group B 181.
Specify the maximum mass flow rate to be
controlled by the L-valve, M, ,,,=, and the hor-
izontal length of the valve, LH.
Calculate from eqn. (2) the size of the outlet
D from the storage hopper situated at the top
of the downcomer using a value of M, = 2M, ,,,-.
Calculate the minimum controllable solids mass
flow rate from eqn. (4) using U&J,,= 1. If this
is too high, consider sloping the horizontal section
of the value downwards to 5. Alternatively,
recalculate D in step 3 using MS=lSM, ,,.,=.
Calculate Umr for the solids at the prevailing
conditions of temperature and pressure in the
L-valve using, for example, the Wen-Yu equation
(see Geldart [9]).
Using D calculated in steps 3 and 4, calculate
the maximum operating mass flux G, ,,,=, and
from eqn. (4) calculate the maximum value of
u .
TlZ maximum volumetric flow of aeration gas
is
Q,, = U,, rrD=/4.
From eqn. (13), calculate APH.
Taking into account the system as a whole,
calculate the absolute pressure at the end of
the horizontal section of the L-valve, PL.
Calculate or specify the absolute pressure at the
top of the downcomer, Pr The pressure drop
which needs to be generated across the down-
comer is then
PL+APH-PT=APDc
(14)
If there is restricted height available such that
LDc is virtually specified, calculate the pressure
gradient and check that
@DC
- <pBLPiit
LDC
(15)
which is the pressure gradient which could cause
the solids in the downcomer to be fluidized. If
AP&LDc is too large, then, according to eqn.
(14), PL must be decreased or PT increased,
otherwise the design is not viable. If AP&LD,
approaches pBLpg, gas bubbles may form locally
at the aeration point and choke the flow.
If LDc is not a major constraint, choose various
values of LDc which satisfy the criterion in eqn.
(15), and from eqn. (lo), calculate IV,l.
Using these values, and V, from eqn. (6), in
eqn. (ll), calculate the magnitude and sign of
V, for the range of solids fluxes to be used in
13.
173
normal operation. If V, is negative, some of the
aeration gas will pass up the downcomer. This
may be desirable to strip out the gas associated
with the downward moving solids, or it may be
unacceptable. Calculate QD, from eqn. (12) and
(taking into account its sign) insert it into eqn.
(3) together with Q., from step 6, to find the
total gas Qt leaving the L-valve with the solids.
Depending on the process requirements, select
the optimum downcomer length.
Conclusions
Three sands (all belonging to Geldarts group B)
have been used in three sizes of L-valve.
(1) If the pressures at the top of the downcomer
and the end of the horizontal section are similar,
the maximum solids fluxes attainable can be cal-
culated from a hopper discharge rate equation, such
as that proposed by Carleton. Depending on the
system these can be in the range 600-l 200 kg/(m=*s).
(2) If the L-valve has to discharge against a
significant back-pressure, the maximum flux attain-
able can be significantly lower and is reached when
the pressure gradient in the downcomer equals that
in a fluidized bed of the particles.
(3) The radius of the elbow, the number of aeration
points and the slope of the horizontal section in-
fluence the solids flow rate.
(4) The minimum aeration gas flow rate required
to produce the minimum stable solids flow is given
by the product of the cross-sectional area of the
pipe and the minimum fluidization velocity of the
powder.
(5) Correlations are given from which the aeration
gas flow rate and the pressure drop across the L-
valve can be calculated for any solids flux.
(6) A stepwise procedure is given for the design
of an L-valve for any group B solid.
List of symbols
A
JP
d,
cross-sectional area of downcomer pipe, m*
mean sieve size of powder, m
average size of particles between two ad-
jacent sieves, m
2%
mean surface/volume size of powder, m
D diameter of downcomer, m
g
acceleration due to gravity, 9.81 m/s*
G,
mass flux of solids through L-valve, kg/(m2. s)
L
DC
length of downcomer, m
LH
length of horizontal section of L-valve, m
M,
mass flow of solids through L-valve, kg/s
174
Q DC
Q exf
Qt
u
cxt
volumetric flow rate of gas in downcomer,
m3/s
volumetric flow rate of injected aeration gas,
m3/s
total flow rate of gas leaving L-valve, m3/s
superficial velocity of injected aeration gas
based on diameter of horizontal section,
m/s
minimum fluidization velocity of powder
transported, m/s
superficial velocity of gas in downcomer,
m/s
absolute velocity of gas in downcomer,
m/s
numerical value of relative velocity between
gas and particles in downcomer, m/s
absolute velocity of particles in downcomer,
m/s
weight fraction of particles having a size dpi
Greek symbols
P
half-angle of hopper cone section
@DC
pressure drop between aeration point and
top of downcomer, N/m2
AH pressure drop between aeration point and
end of horizontal section of L-valve, N/m2
ELP
EL
P
PBMF
PSLP
4
voidage of loosely packed bed of powder,
-
gas viscosity, N - s/m*
density of gas surrounding particles, kg/m3
bulk density of powder at minimum flui-
dization conditions, kg/m3
bulk density of powder in loosely packed
condition, kg/m3
particle density (including all open and
closed pores), kg/m3
References
1 T. M. Knowlton and I. Hirsan, Hydrocarbon Proc., 57
(1978) 149.
2 T. M. Knowlton, in D. Geldart (ed.), Gas Fluidizution
Technology, Wiley, Chichester, 1986, p. 406.
3 Y. Yuasa and H. Kuno, Powder Technol., 6 (1972) 97.
4 A. J. Carleton, Powder Techno/., 6 (1972) 91.
5 D. Geldart and S. J. Ling, Powder TechnoL, 62 (1990)
241.
6 P. Jones, M. Phil. Dissertation, Univ. Bradford (1988).
7 C. Y. Wen and H. P. Simons, AIChE J., 5 (1959) 263.
8 D. Geldart, Powder Technol., 7 (1973) 285.
9 D. Geldart, in D. Geldart (ed.), Gas Fluidimtion Tech-
nology, Wiley, Chichester, 1986, p. 24.

Potrebbero piacerti anche