Sei sulla pagina 1di 6

ASIA-PACIFIC JOURNAL OF CHEMICAL ENGINEERING

Asia-Pac. J. Chem. Eng. 2008; 3: 2429


Published online in Wiley InterScience
(www.interscience.wiley.com) DOI:10.1002/apj.113

Research Article

Cooling of spherical particles in a vertical airlift conveyor


C. E. Davies* and R. Y. G. Davies
Institute of Technology and Engineering, Massey University, Private Bag 11 222, Palmerston North, New Zealand

Received 26 July 2007; Revised 27 November 2007; Accepted 27 November 2007

ABSTRACT: The airlift conveyor modelled here is of a type used in industry to move a particulate feedstock from
ground level to an elevated position for gravity feeding to storage or further processing options. It consists of a feed leg
and a vertical riser which terminates in a de-entrainment device where the particles are removed from the conveying
gas. Typically the conveying gas enters the riser from a vertical or horizontal feed pipe, and blows the particles off the
face of an aerated bed in the horizontal leg of the feeder. The flow rate of the solids is controlled by aeration gas which
partially fluidises the material in the feed leg. In this article, we consider the cooling of spherical particles admitted to
a conveyor, for a variety of processing conditions. The riser in the model conveyor was 8 m high, the initial particle
temperature was 40 C and the temperature of the conveying air was 20 C in all calculations. The conveying velocities
used were, 12, 15 and 18 m s1 , and the particle sizes were 0.5, 1.0 and 1.5 mm. The particle flow rate was 1 tonne/h.
The calculation methodology was to find analytically the slip velocity profile of a single particle over the length of
the riser, and then to estimate heat transfer rates numerically using a Microsoft Office Excel spreadsheet. Temperature
profiles show that for all conditions examined, over 7080% of heat transfer takes place in the first half of the riser.
2008 Curtin University of Technology and John Wiley & Sons, Ltd.
KEYWORDS: heat transfer; particle; pneumatic conveying; spreadsheet model; airlift conveyor

INTRODUCTION
The work outlined here was undertaken following an
anecdotal report that simple airlift conveyors were being
used as coolers in several manufacturing applications
and had been running satisfactorily for many years. No
quantitative information was available at that time on
operating or material parameters and this investigation
was carried out to obtain information on the extent
of cooling that can be expected for typical process
conditions.

Physical system
An air-lift conveyor previously described by Davies
and Graham[1] is shown schematically in Fig. 1 to
illustrate a typical physical configuration. The solids
being conveyed move downwards in a feed leg and are
picked up by the conveying air from the free surface
of the material in the horizontal arm at the base of the
feed leg. Initially, they have no velocity component in
the vertical direction and are accelerated from rest by
the conveying air. Typically, flows are dilute and solids
*Correspondence to: C. E. Davies, Institute of Technology and
Engineering, Massey University, Private Bag 11 222, Palmerston
North, New Zealand. E-mail: C.Davies@massey.ac.nz
2008 Curtin University of Technology and John Wiley & Sons, Ltd.

to air mass ratios are in the range 12. The solids


move through the conveying tube, of specified length,
and are de-entrained by a knockout drum and cyclone.
The slip velocity, U , the difference between the
velocity of the conveying gas and the velocity of the
particle, decreases as the particle accelerates in the
conveying tube. For a specified conveying velocity,
the evolution of the slip velocity profile for a particle
depends on particle size, as small particles accelerate
rapidly and slip velocity is small over much of the
traverse up the conveyor. Conversely, large particles
accelerate more slowly and slip velocity is larger over
the whole of the time they are resident within the
conveying tube. It follows that residence times will
increase as particle diameter increases. The rate of heat
transfer from a particle moving in a fluid depends on
slip velocity and the residence time determines the time
available for heat transfer.
Velocities in a vertical conveyor of this type are
typically about 15 m s1 , conveyor diameters, 150
300 mm and lift heights 6 to 15 m. In the calculations
that follow, we have used three different velocities, 12,
15 and 18 m s1 , a single conveyor diameter of 150 mm
and a single lift height of 8 m. The inlet temperature of
the air was 20 C. As temperature changes were small,
we have neglected changes in the physical properties of
air, and in all calculations, have used the values at 20 C
interpolated from the values tabulated by Mayhew and

Asia-Pacific Journal of Chemical Engineering

COOLING OF SPHERICAL PARTICLES IN A VERTICAL AIRLIFT CONVEYOR

To Bag House

To storage or further processing


Powder Feed

Lift Air
Aeration Air

Figure 1. Schematic diagram of vertical conveyor.

Rogers[2] : air density, a , 1.207 kg m3 ; air viscosity,


a , 1.812 105 kg m1 s1 ; air thermal conductivity, a , 0.0257 W m1 C1 and air heat capacity, cpa ,
1004.6 J kg C1 . We have considered spherical particles only, and used three particle diameters, 0.5, 1 and
1.5 mm. The heat capacity of the particles was 2400 J
kg1 C1 and the density, 1250 kg m3 . The flow rate
of the particles was 1 tonne/h (0.278 kg s1 ) and the
initial particle temperature was 40 C.

HEAT TRANSFER
Our method for estimating heat transfer rates involves
two sequential calculations. First, the equation resulting
from the force balance on a particle in an air stream is
solved analytically to provide an expression giving the
relationship between height in the conveyor and slip
velocity.
The second step is the determination of air temperature and particle temperature. This is done by stepwise
heat balances over the height of the conveyor, which is
easy to implement in a Microsoft Office Excel spreadsheet. In each step, convective heat loss from a particle
is equated to the heat gained by the conveying air. The
heat transfer coefficient required for estimation of the
convective term is calculated using an expression from
Whitaker[3] and cited by Incropera and De Witt.[4]
Implicit in this approach are the following assumptions:
The heat transfer coefficient for a particle is not
affected by the presence of other particles in the
system.
Effects of turbulence can be neglected; the particles
travel in a straight line as they move through the
conveyor.
2008 Curtin University of Technology and John Wiley & Sons, Ltd.

Wall flow effects can be neglected; there is no down


flow at the wall.
All heat transfer is between the particles and the
conveying air. There are no losses, and there is no
heat exchange between the particles and the wall.
The mean velocity can be used as the air velocity in
all calculations of slip velocity; effects of a velocity
profile in the conveyor are not considered.
The particles do not interact.
Changes in the transport properties of the air can be
neglected.
All particles have the same initial temperature.
All particles are initially at rest; y = 0, up = 0, where
y is the Cartesian coordinate for the vertical direction
and up is the particle velocity.
Particle motion in the air stream can be determined
by a simple force balance; see Eqn (1) below.
Radiation heat transfer can be neglected.

Particle motion in conveyor


The motion of a particle in an air stream flowing at a
velocity ua is determined by the forces acting on it; see
Fig. 2. Applying Newtons second law and balancing
gravity force, FW , buoyancy force, FB , and drag force,
FD , with the force due to particle acceleration, we have:
FD + FB FW = mp

dup
dt

(1)

Where mp is the mass of the particle. Substituting for


FB and FW ,
FD + (ma mp )g = mp

dup
up
dy

(2)

FD

FB

up
FW

ua

Figure 2. Particle in air stream.


Asia-Pac. J. Chem. Eng. 2008; 3: 2429
DOI: 10.1002/apj

25

26

C. E. DAVIES AND R. Y. G. DAVIES

Asia-Pacific Journal of Chemical Engineering

where ma is the mass of air displaced and g is


acceleration due to gravity. With rearrangement Eqn (2)
becomes:
mp up dup
(3)
dy =
FD + (ma mp )g
By definition,
1
x2
FD = CD a U 2
2
4

(4)

Where CD is drag coefficient and x is particle


diameter.
Various expressions are available in the literature
relating drag coefficient, CD , to Rep , the Reynolds
number based on particle diameter and slip veloc Ux
ity; Rep = a . In this investigation, 400 < Rep <
a
1800. This range spans the upper end of the intermediate region, where CD is given by Eqn (5) due to
Dallavale,[5] and the lower part of the range of Rep
where Newtons law is obeyed and CD has the constant
value of 0.44; see Rhodes.[6] The notional boundary of
the intermediate region and the Newtons law region is
at Rep = 500, and although some Rep in this work are
greater than 500, Eqn (5) is asymptotic to the Newtons
law value. We have therefore used Eqn (5) for CD in
all calculations.
CD =

24
24a
+ 0.44
+ 0.44 =
Rep
x a U

(5)

Substituting for FD and CD in Eqn (3) and rearranging


in terms of U ,
dy =

mp (U ua )dU
3a xU + 0.055a x 2 U 2 + (ma mp )g

(6)

Integration of Eqn (6) gives:




 AU 2 + BU + C 
1


y=
ln  2


2A
Aua + Bua + C 


B


+ ua  
2 4AC 
2AU
+
B

B
2A


ln 




2
2
B 4AC
2AU + B + B 4AC 



 2Au + B B 2 4AC 
a


+ ln 

 (7)
 2Au + B + B 2 4AC 
a

3 x
0.055a x
, B = ma , and
where A =
mp
p
(ma mp )g
C =
mp
U cannot be obtained as an explicit function of
height, but this is a minor inconvenience in a spreadsheet computation.
2008 Curtin University of Technology and John Wiley & Sons, Ltd.

Heat transfer coefficient


The heat transfer coefficient, h, for a single particle is
estimated using Eqns (8) and (9) as in Whitaker[3] cited
by Incropera and De Witt.[4] Equation (9) is valid over
the range 0.71 < Pr < 380, 3.5 < Rep < 7.6 104 ; Pr
c
is Prandtl number, pa a . We have already noted above
a
that we neglect the changes in the physical properties
of air with temperature and use the values at the initial
temperature of 20 C.
h=

a Nup
x

(8)
2

Nup = 2 + (0.4Rep0.5 + 0.06Rep3 )




a 0.25
0.4
(Pr
)
sa

(9)

sa is air viscosity at thesurface


 of the particle; in all
a 0.25
= 1
calculations we have set
sa

Heat balance and computation


The trajectory of slip velocity, U , over the height of the
conveyor was found analytically as outlined above. We
have taken a Lagrangian approach, and have considered
heat transfer in successive height elements, yi , of the
conveyor as a particle moves through it. U was varied in
4050 decrements of 0.13 to 0.3 m s1 with smaller
decrements used as yi approached the height of the
conveyor, 8 m.
The residence time, pi , in a height element, i , is:
pi =

yi
upi

upi = ua Ui

(10)
(11)

In each height element hi is found from Eqns (8) and


(9), using the prevailing value of Ui .
Using the lumped capacitance method,[4] the temperature, pi reached by a particle during the time, pi , in
a height element is:

pi = ai + e

pi

(pi 1 ai )

(12)

where, ai is the temperature of the air in the element,


pi 1 is the temperature of the particle entering from the
previous element, i.e. the temperature of the particle at
the beginning of the computational interval and =
p cpp x
6h . This assumes thermal equilibrium within the
solid and is valid for Bi  1 , where Bi is Biot number,
hx
6 ; p is the thermal conductivity of the particle.
p

Asia-Pac. J. Chem. Eng. 2008; 3: 2429


DOI: 10.1002/apj

Asia-Pacific Journal of Chemical Engineering

COOLING OF SPHERICAL PARTICLES IN A VERTICAL AIRLIFT CONVEYOR

The heat balance linking particle temperature and the


temperature of the air is:

35

(13)

where Mp is the mass flow rate of the solids, cpp is the


specific heat capacity of the solids, and Ma is the mass
flow rate of the air.

Temperature C

Mp cpp (pi pi 1 ) = Ma cpa (ai +1 ai )

40

30
particle 18 m/s
air 18 m/s
particle 15 m/s
air 15 m/s
particle 12 m/s
air 12 m/s

25
20

RESULTS
15
0

4
Height (m)

Figure 4. Temperature profiles for 1 mm particles, filled


symbols; air, open symbols. This figure is available in colour
online at www.apjChemEng.com.

Ey/E8 (-)

The primary outputs of the heat transfer calculation


were the extent of cooling of the particles achieved
within the height of the conveyor tube, and the temperature of the conveying air. We have also determined
E
the fractional energy transfer, defined as Ey where Ey
8
is the cumulative energy transfer rate from the particles
as they travel from the entrance of the conveying tube
to a height y; E8 is the cumulative energy transfer rate
from the particles over the full 8 m of the tube.
Figure 3 shows the slip velocity profiles for the 1 mm
particles for the 3 air velocities used. The temperature
profiles for 1 mm particles and also the conveying air
are shown in Fig. 4.
The fractional energy transfer for 1 mm particles
is shown in Fig. 5. Similar trends to those seen in
Figs 35 were observed in the results obtained for the
0.5 and 1.5 mm particles.
Table 1 shows particle temperatures at the exit of an
8 m conveyor. For the particle sizes considered here,
0.5, 1 and 1.5 mm, the temperature reached by a particle
is only weakly dependent on particle size, and likewise,
does not change much as air velocity is varied from 12
to 18 m s1 .
It is apparent from Fig. 5, that approximately 70
80% of the total heat transfer from the particles to the

1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

18 m/s
15 m/s
12 m/s

4
Height (m)

Figure 5. Fractional energy transfer rate for one 8 m stage.


Particle diameter is 1 mm. This figure is available in colour
online at www.apjChemEng.com.

Table 1. Particle temperature ( C) at the exit of 8 m


conveyor.

20
18
18 m/s
15 m/s
12 m/s

Slip velocity (m/s)

16
14

ua
(m/s)
12
15
18

12
10
8

Particle diameter
0.5 mm

1 mm

1.5 mm

34.4
33.5
32.9

34.6
34.2
34.2

34.9
35
35.2

6
4
2
0
0

Height (m)

Figure 3. Slip velocity profiles for 1 mm particles. This figure

is available in colour online at www.apjChemEng.com.


2008 Curtin University of Technology and John Wiley & Sons, Ltd.

air takes place in the first 34 m of the tube, depending


on the air velocity. Figures 6 and 7 show the effect
of a second stage on the cooling of the particles; the
particles are collected after a (arbitrary) height of 4 m,
and transported for another 4 m with air initially at
20 C.
Asia-Pac. J. Chem. Eng. 2008; 3: 2429
DOI: 10.1002/apj

27

C. E. DAVIES AND R. Y. G. DAVIES

Asia-Pacific Journal of Chemical Engineering

Particle temperature C

39
37
35
33
31
one 8-m column

29

1st stage of two 4-m columns


27

2nd stage of two 4-m columns

25
0

4
5
Height (m)

Figure 6. Particle temperature for two 4 m stages. Air


velocity is 15 ms1 ; particle diameter is 1 mm. This figure is
available in colour online at www.apjChemEng.com.

one 8-m column


1st stage of two 4-m columns
2nd stage of 4-m columns

1.6
1.4
1.2
1
Eh/E8

28

0.8
0.6
0.4
0.2
0
0

4
Height (m)

Figure 7. Fractional energy transfer rate for two 4 m stages.


Air velocity is 15 ms1 ; particle diameter is 1 mm. This figure
is available in colour online at www.apjChemEng.com.

transfer rates due to linear increase in h with increasing


a , to a maximum of 5%, must be less than 5%, so the
use of constant transport properties does not introduce
significant error to the calculated values of particle and
air temperature. However, the validity of the lumped
capacitance model used for calculating particle temperature is contingent on Bi  1 and this requirement may
not be fulfilled where slip velocity, and hence h, is high,
or for large particles.
The temperature reached by the particles over the
(arbitrary) 8 m conveyor height used here is relatively insensitive to air velocity, so the performance
of coolers designed for low air velocities will not be
much compromised, although there could be opportunity for savings on capital and running costs. This is
of course subject to the proviso that the choice of air
velocity is adequate for proper hydrodynamic performance.
The trends shown in Fig. 5, when considered with
those in Figs 3 and 4, show that 7080% of the energy
exchange between the solids and gas takes place while
the solids are accelerating in the first 4 m of the
conveyor. The addition of a second stage can lower
the temperature reached by the particles significantly.
In a system consisting of two 4 m stages, with the
conveying air entering at 20 C in each stage, and the
partially cooled solids fed from the exit of the first stage
directly into the second stage, the energy transferred is
45% greater than in a single 8 m stage. An alternative
strategy, not modelled here, that might have some
practical benefit, would be to recycle both solids and the
conveying gas, while using a high conveying velocity.
For 18 m s1 , the second stage air inlet temperature
would be 30 C for 1 mm spheres, compared to 20 C
in the first stage, but there would be increased residence
time in the acceleration region and the particles would
be expected to cool several degrees further from the
34 C shown in Table 1.

DISCUSSION
CONCLUSIONS
In all calculations here, we have neglected changes in
the physical properties of the conveying air with temperature, and also used a constant conveying velocity.
Effectively this assumes a constant mass flow rate for
the air, and neglects changes in air density and viscosity
in drag calculations and air conductivity and viscosity in
estimates of heat transfer rate. The maximum changes
in air temperature observed were 15 C, and for this
increase, there is a decrease of 5% in density, and an
increase in viscosity and conductivity of 5%; there is
virtually no change in Pr 0.4 over this small temperature range. It is clear from consideration of the effect
of changes in velocity on heat transfer that relatively
large changes in Rep have only a small effect on final
particle temperature, so a change up to 5% in air viscosity will be of little consequence. Changes in heat
2008 Curtin University of Technology and John Wiley & Sons, Ltd.

The cooling of a dilute flow of spherical particles in


a vertical airlift conveyor has been modelled using
a two-step calculation method. In the first step, the
slip velocity profile over the conveyor was determined
analytically. In the second step, heat transfer from the
particles to the lift air was found numerically using
a Microsoft Office Excel spreadsheet. Approximately
7080% of the heat transfer took place over the first
half of the conveyor, and particle temperatures at the
exit of an 8 m conveyor, for the conditions considered
here, were only weakly dependent on air velocity. A
two-stage system comprising two 4 m conveyors, where
the particles leaving the first stage were recycled to
the second stage, was also considered. For the case
where the initial air temperature was the same in both
Asia-Pac. J. Chem. Eng. 2008; 3: 2429
DOI: 10.1002/apj

Asia-Pacific Journal of Chemical Engineering

COOLING OF SPHERICAL PARTICLES IN A VERTICAL AIRLIFT CONVEYOR

conveyors, the total energy transfer from the particles to


the conveying air was approximately 45% greater than
in a single 8 m stage.

NOTATION
A
B
Bi
C
CD
cp
E8
Ey
FB
FD
FW
g
h
m
M
Nu
Pr
Re
t
u
U
x
y

parameter, see Eqn (7) (m1 )


parameter, see Eqn (7) (s1 )
Biot number ()
parameter, see Eqn (7) (m s2 )
drag coefficient ()
specific heat capacity (J kg1 C1 )
heat transfer rate, 8m (W)
heat transfer rate, y (W)
buoyancy force (N)
drag force (N)
gravity force (N)
gravity acceleration (m s2 )
heat transfer coefficient (W m2 C1 )
mass (kg)
mass flow rate (kg s1 )
Nusselt number ()
Prandtl number ()
Reynolds number ()
time (s)
velocity (m s1 )
slip velocity (m s1 )
particle diameter (m)
Cartesian coordinate (m)

2008 Curtin University of Technology and John Wiley & Sons, Ltd.

Greek letters

thermal time constant (s)


temperature ( C)
thermal conductivity (W m1 C1 )
viscosity (kg m1 s1 )
mathematical constant ()
density (kg m3 )
residence time (s)

Subscripts
a
i
p
s

air
label in stepwise calculation
particle
surface

REFERENCES
[1] C.E. Davies, K.H. Graham. Pressure drop reduction by wall
baffles in vertical pneumatic conveying tubes, Chemeca 88,
Sydney, Australia, 2831 August, 1988.
[2] Y.R. Mayhew, G.F.C. Rogers. Thermodynamic and Transport
Properties of Fluids, Basil Blackwell, Oxford, 1968.
[3] S. Whitaker. AIChE J., 1972; 18(2), 361371.
[4] F.P. Incropera, D. De Witt. Fundamentals of Heat and Mass
Transfer, John Wiley and Sons, New York, 2002.
[5] J.M. Dallavale. Micromeritics, 2nd edn, Pitman, London, 1948.
[6] M. Rhodes. Introduction to Particle Technology, John Wiley and
Sons, England, 1998.

Asia-Pac. J. Chem. Eng. 2008; 3: 2429


DOI: 10.1002/apj

29

Potrebbero piacerti anche