Sie sind auf Seite 1von 10

Powder Technology 161 (2006) 135 – 144

www.elsevier.com/locate/powtec

Parametric study of fine particle fluidization under mechanical vibration


Chunbao Xu 1, Jesse Zhu *
Particle Technology Research Centre, Department of Chemical and Biochemical Engineering The University of Western Ontario,
London, Ontario, Canada, N6A 5B9

Received 28 July 2004; received in revised form 20 September 2005; accepted 13 October 2005
Available online 28 November 2005

Abstract

Investigations into the effects of vibration on fluidization of fine particles (4.8 – 216 Am average in size) show that the fluidization quality of
fine particles can be enhanced under mechanical vibration, leading to larger bed pressure drops at low superficial gas velocities and lower values
of u mf. The effectiveness of vibration on improving fluidization is strongly dependent on the properties (Geldart particle type, size-distribution and
shape) of the primary particles used and the vibration parameters (frequency, amplitude and angle) applied. The possible roles of mechanical
vibration in fine particle fluidization have been studied with respect to bed voidage, pressure drop, agglomeration, and tensile strength of particle
bed. Vibration is found to significantly reduce both the average size and the segregation of agglomerates in the bed, thus improving the fluidization
quality of cohesive particles. Also, vibration can dramatically reduce the tensile strength of the particle bed. Obviously, vibration is an effective
means to overcome the interparticle forces of fine powders in fluidization and enhance their fluidization quality.
D 2005 Elsevier B.V. All rights reserved.

Keywords: Fluidization; Fluidization quality; Fine particles; Mechanical vibration; Tensile strength; Particle agglomeration; Particle shape

1. Introduction efficiency and the efficiency of heat and mass transfer can be
significantly improved. Using a self-developed vibro-fluidized
Fluidization is a widely used process for powder handling in bed, Mori et al. [4] found that particles smaller than 24 Am that
industry because of the favourable gas –solid and solid – solid are not fluidizable under non-vibrated conditions could be
contacting efficiencies. However, fine particles smaller than completely fluidized under vibration (50 Hz/0.5 mm) at a gas
about 35 Am in size are very cohesive and are generally velocity less than 4.5 cm/s, giving appreciable bed expansion.
believed to be not suitable for fluidization due to the strong According to Dutta and Dullea [9], when applying vibration to
interparticle forces [1,2]. Such fine particles are classified as cohesive powders, improved fluidization quality was observed
Geldart group C (cohesive) particles [3]. Group C particles tend by increased bed pressure drop and bed expansion and
to cling to each other as a consequence of the interparticle decreased elutriation. As a result, vibro-fluidized beds are
forces and therefore form channels and agglomerates in now used in industries for handling particulate materials such
fluidization, which often lead to poor fluidization or even as fluidization, mixing, granulation, drying, coating, etc.
complete de-fluidization. [10,11].
Mechanical vibration has proved to be an effective means Despite many previous studies on the effect of mechanical
for improving the fluidization of cohesive particles [4 –8]. vibration, a comprehensive and parametric investigation on the
Vibration prevents the solids from channelling and seriously fluidization behaviour of fine particles under vibration is still
agglomerating by supplying the bed with the energy required to not available in the literature. Therefore, in the present work,
overcome the interparticle forces. With the application of the the effects of vibration on the fluidization of a wide range of
vibration, the fluidization quality, the gas – solid contact fine particles including Al2O3, TiO2, CaCO3 and glass beads,
4.8– 216 Am in average size, are studied with respect to bed
pressure drop, bed expansion ratio, agglomeration of the fine
* Corresponding author. Tel.: +1 519 661 3807.
E-mail address: jzhu@uwo.ca (J. Zhu).
particles, and tensile strength of bed materials. Influencing
1
Present address: Department of Chemical Engineering, Lakehead Univer- factors controlling the performance of vibration, including
sity, Thunder Bay, Ontario, Canada, P7B 5E1. properties of primary particles and vibration parameters, and
0032-5910/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2005.10.002
136 C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144

the possible roles of vibration in fine particle fluidization are fluidization state of fine particles reduces greatly as the static
extensively discussed. bed height (L 0) increases, due to the dissipation of the vibration
energy along the bed height. Consequently, many previous
2. Experimental studies on the vibration-assisted fine particle fluidization have
employed a small value of the static bed height [8,13]. In the
The experimental setup is illustrated in Fig. 1, consisting of present study, the static bed height is fixed to about 10 cm for
a fluidized bed column made of Plexiglas (100 mm i.d. and all experiments, with L 0 / D å 1.0, so as to avoid the problem
1000 mm tall), a vibration generation system, a data acquisition related to dissipation. In addition, for every experimental run,
system with LabVIEW, and a gas flow control system. the particle bed is loosened with a flow of air at 2.0 cm/s for
Compressed air stripped of trace humidity through a fixed 10– 20 min prior to the fluidization experiments. During the
bed of silica gel is used as the fluidizing gas, with the flow rate bed-loosening process as well as the subsequent fluidization
controlled by a series of rotameters (Omega Engineering Inc.) process, the entrained fines are returned to the bed, to prevent
and a digital mass flow controller (Fathom Technologies). All the lowering of the static bed height due to possible particle
the flowmeters are carefully calibrated with a Wet Test Meter carryover. The entrained fines during the experiments are
(GCA/Precision Scientific). A polyethylene porous plate (15 – collected by a bag filter, made of densely woven cloth, placed
45 Am pore size, 30% pore volume and 6.4 mm thick) serves as externally right above the free board in a cylindrical column,
the gas distributor. It has been measured that for all types of and returned manually to the bed by gently tapping the column
particles studied, the grid pressure drop is larger than 30% of upon the completion of the bed-loosening process or in about
the bed pressure drop at a superficial gas velocity at 2 cm/s, every 5 min during the fluidization. Then, with decreasing gas
ensuring a uniformly distributed fluidizing gas in the bed. velocity (u g), pressure drops across the whole bed (DP) and the
Pressure drops across the whole bed are obtained with a bed height (L) are measured simultaneously. When vibration is
differential pressure transducer. Mechanical vibration is gener- applied to the bed, the differential pressure signals measured
ated by a pair of vibrators mounted opposite one another on the across the particle bed fluctuate due to the influence of the
two sides of a steel-made base and driven by an ABB inverter vibration. This fluctuation strongly depends on the vibration
with a vibration frequency ( f) varying from 0 to 50 Hz. By parameters (frequency, amplitude and direction). To reduce the
changing the unbalanced weights of the vibrator, various negative effects of the signal fluctuation and ensure enough
amplitudes (A) between 0 and 3 mm (depending on the quantities of data for acquiring accurate average bed pressure
frequency used) can be obtained. The amplitude of the drop, the pressure signals are sampled by a data acquisition
vibration base is measured using a quartz shear ICP\ system with LabVIEW at a frequency of 1000, much higher
accelerometer (Dalimar Instruments Inc.) with the sensitivity than the vibration frequency applied to the particle bed, and
of 10.09 mV/m/s2 at 100 Hz. The vibration angle or direction over long enough sampling period (60 –120 s).
can also be varied within 0 – 90- through adjusting the To assess the fluidization quality, the normalized bed
mounting angles of the vibro-motors on the vibration base, pressure drop is adopted. It is defined as the ratio of the
where the angle of 0- corresponds to horizontal vibration and measured pressure drops across the whole bed (DP) to the
90- the vertical vibration. normal pressure caused by particle weight (m sg / S), where m s
Earlier work such as Chen et al. [12], has demonstrated that denotes the weight of solids in the bed and S the cross-sectional
the effectiveness of mechanical vibration for improving the area of the fluidization column. When the entire bed is
fluidized, the normalized pressure drop, DP / (m sg / S), should
Gas Outlet attain unity and remain stable thereafter even if the gas velocity
to Bag Filter Compressed Air
further increases, provided that the bed is completely fluidized.
Also worth mentioning is that in this study, the average bed
Silica Gel voidage (( b) rather than the bed expansion ratio (convention-
Column ally used by other researchers) is used in this study to
Fluidized Bed characterize the bed expansion in fluidization for the following
Pressure
Transducer
Column two reasons: (1) ( b is calculated from the bed height data, thus
Rotameter
directly related to the bed expansion ratio, and (2) more
Data Acqusition importantly, it is more advantageous to use ( b instead of the
System bed expansion ratio for the vibration-assisted fluidization,
because often no bed expansion but consolidation is observed
for a particle bed under vibration particularly when the
Mass-flow
superficial gas velocity is below the minimum fluidization
Controller velocity [8,13].
Piezoelectric The particles used in this study are Al2O3 (4.8 Am), TiO2
Accelerometer Vibrator (5.2 Am), CaCO3 (5.5 Am) and glass beads (6.1, 10, 39, 65 and
Inverter 216 Am average in size), pertaining to Geldart groups C, A and
B particles. Table 1 gives the key physical properties of these
Fig. 1. Schematic diagram of the experimental apparatus. powders and Fig. 2 shows the size distribution. To minimize
C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144 137

Table 1
Particles used in the experiments
Powders d (1)
p (Am) Density (kg/m3) H R(3) ( – ) AOR(4) (-) Geldart group(5) Morphology(6)
qp q bt(4) q ba(4)
Al2O3 4.8 3850 1610 790 2.04 52.6 C Irregular
TiO2 5.2 3880 760 450 1.69 45.9 C Irregular
CaCO3 5.5 2700 1160 520 2.23 50.0 C Irregular
GB-S(2) 6.1 2500 1160 600 1.93 53.6 C Spherical
GB-S(2) 10 2500 1550 1000 1.55 49.1 C Spherical
GB-S(2) 39 2500 1510 1360 1.11 33.5 A Spherical
GB-I(2) 39 2500 1660 990 1.68 44.7 A or C Irregular
GB-S(2) 65 2500 1510 1370 1.10 30.4 A Spherical
GB-S(2) 216 2500 1620 1590 1.02 20.4 B Spherical
1
The volume-weighted mean diameter by laser diffraction (Malvern Mastersizer 2000).
2
Soda lime silica glass beads (GB): -S (spherical shape), -I (irregular shape).
3
Hausner ratio = q bt / q ba.
4
Analyzed by a Hosokawa Powder Tester.
5
Geldart [3].
6
Determined by SEM.

the influence of moisture, all types of particles were subject to essentially no fluidization. When mechanical vibration is
drying overnight at 80 -C under vacuum before being loaded applied, however, the fluidization state of all these particles
into the bed for the fluidization tests. To investigate the can be significantly improved. When the gas velocity increases
influence of the particle shape, two glass bead samples of same to a certain value, depending on the powder type and the
mean diameter at 39 Am but different shapes (spherical and vibration conditions, the plug and channel phenomena disap-
irregular) are also examined in this study. The irregular shaped pear, the bed pressure drop increases and smooth bed
glass bead sample (39 Am) is prepared by jet milling of the expansion starts. For fine particles of TiO2 (5.2 Am), Al2O3
glass beads (65 Am) followed by sieving to the desired size. (4.8 Am) and CaCO3 (5.5 Am), the entire fluidization of the bed
Fig. 3 shows the SEM photos of these two samples of glass is still not obtainable even under vibration. In this case, the
beads of different shapes.

3. Results and discussion

3.1. Fluidization behaviour of fine particles under mechanical


vibration

For the Group C particles of Al2O3 (4.8 Am), TiO2 (5.2 Am),
CaCO3 (5.5 Am) and glass beads (6.1 and 10 Am) under no
vibration, the whole bed usually tends to lift as a plug with a
relatively high pressure drop when the air flow is initially
turned on and increased slowly. Channelling and agglomerat-
ing take place when the gas velocity further increases, with
Cumulative volume percentage (%)

Al2O3 (4.8µm)
100 TiO2 (5.2µm)
CaCO3 (5.5µm)
GB-S (6.1µm)
80
GB-S (10µm)
GB-S (39µm)
60 GB-I (39µm)
GB-S (65µm)
GB-S (216µm)
40

20

0
0.01 0.1 1 10 100 1000
Particle size (µm)
Fig. 3. SEM photos of glass beads (39 Am) of different shapes: (a) spherical and
Fig. 2. Size distribution of the powders used. (b) irregular.
138 C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144

observation is that only a partial fluidization can be achieved at noted that for all the cases shown in Fig. 4a, except for the
the top-bed, while stagnant agglomerates begin to form at the 6.1 Am glass beads under vibration, the u mf obtained, should
bottom of the bed. On the other hand, an entire fluidization is be more appropriately called the apparent minimum fluidiza-
attained for the glass beads (6.1 and 10 Am) under vibration, tion velocity, since the fluidization of an entire bed is not
and a minimum fluidization velocity can be determined from achieved in the testing range of u g. Meanwhile, the vibration
the plots of bed pressure drop against superficial gas velocity, causes consolidation of the particle bed at a lower gas
as will be discussed later. velocity, resulting in much smaller values of ( b than those
Fig. 4 shows plots of the normalized pressure drop and the without vibration, as shown in Fig. 4b. Similar observations
average bed voidage against the superficial gas velocity of the consolidation effect of mechanical vibration have been
during fluidization of the typical group C particles of TiO2 reported in some previous studies [4,13,15,16]. It is shown in
(5.2 Am), Al2O3 (4.8 Am) and glass beads (6.1 Am) with and Fig. 4b that for TiO2 (5.2 Am), the consolidation phenomenon
without mechanical vibration (50 Hz/0.3 mm). For these three appears essentially in the whole range of the gas velocities
group C powders under no vibration, a nearly linear (0¨5 cm/s) tested. As for the other two powders, namely
relationship between the pressure drop and the superficial Al2O3 (4.8 Am) and glass beads (6.1 Am), there exists a
gas velocity is clearly observed in the log – log coordinates, transition point of superficial gas velocity at about 0.4 cm/s,
which is typical for a fixed bed. However, by applying the beyond which the average bed voidages under vibration will
vibration (50 Hz/0.3 mm), higher pressure-drops are obtained exceed those without vibration. It should be noted that all of
for all these powders, suggesting an improved gas – solid these three powders belong to Geldart group C powders and
contacting efficiency. In particular, for glass beads (6.1 Am), there is no clear trend in the Hausner ratio, AOR, moisture
the normalized bed pressure drop attains approximately unity level or particle shape that would suggest the TiO2 should be
when the gas velocity reaches around 0.1 cm/s, as shown in different in bed expansion behaviors under vibrated condi-
Fig. 4a, indicating a thorough fluidization. The incipient or tions from the other powders. A possible reason is that the
minimum fluidization velocity (u mf) is determined by the TiO2 powder (5.2 Am) used in this study has a much wider
conventional means [14], as illustrated in Fig. 4a. It should be size-distribution compared with those of the other two
powders, as shown in Fig. 2. This assertion may be further
proven by comparing the bed expansion behaviors between
(a)
the 39 Am smooth glass beads of a narrow size-distribution
umf
and the 39 Am irregular-shaped glass beads of a wider size-
Normalized pressure drop (-)

1
distribution, as will be shown later in Fig. 7.
The effects of mechanical vibration on the fluidization
behaviour of glass beads (10, 65 and 216 Am), pertaining to
typical Geldart groups C, A and B, are shown in Fig. 5. The
0.1 pressure drops across the bed are significantly increased for all
types of particles under vibration, leading to a much smaller
u mf than that without vibration. The u mf for 10 Am glass beads
becomes the lowest after vibration is applied, in comparison to
being the highest among the three powders without vibration.
0.01 The vibration appears to have ‘‘rectified’’ the trend to that u mf
0.01 0.1 1 10
Superficial gas velocity (cm/s)
decreases with particle size. In addition, similar to that formerly
shown in Fig. 4b for Al2O3 (4.8 Am) and glass beads (6.1 Am),
(b) there is a transition gas velocity at about 0.1 cm/s for the 10Am
0.90
glass beads, beyond which the average bed voidage under
0.85 vibration exceeds that without vibration. For the other two
Average bed voidage (-)

powders of larger sizes, i.e., glass beads (65 and 216 Am),
0.80 however, the particle beds are consolidated almost in the whole
range of gas velocities tested and the consolidation is more
0.75 remarkable at lower gas velocities, as illustrated in Fig. 5b.

0.70 TiO2 (5.2µm)


3.2. Influencing factors
Al2O3 (4.8µm)
0.65 GB-S (6.1µm)
Open symbols: No vibration The factors that may influence the performance of mechan-
Solid symbols: 50Hz/0.3mm ical vibration on fine particle fluidization are the properties of
0.60
0.01 0.1 1 10 the primary particles (e.g., Geldart type, size-distribution and
Superficial gas velocity (cm/s) shape) and the vibration parameters (e.g., frequency, amplitude
Fig. 4. Effects of mechanical vibration on hydrodynamic behaviours of typical and angle), among which the effects of particle shape and size-
Geldart group C particles of TiO2 (5.2 Am), Al2O3 (4.8 Am) and glass beads distribution on the fluidization behaviour of fine particles under
(6.1 Am): (a) the normalized bed pressure drop and (b) the average bed voidage. vibration are studied for the first time.
C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144 139

(a) the results of u mf for glass beads fluidized with and without
umf vibration (50 Hz/0.3 mm) are plotted against the mean particle
Normalized pressure drop (-)

1
size. It is shown that the vibration leads to a decrease in u mf for
all types of the glass bead samples. Such effect is particularly
remarkable for the group C particles. However, the effective-
ness of vibration decreases with the particle size. It thus
0.1 suggests that the effect of vibration on fluidization of fine
particles strongly depends on the Geldart type of particles.
In addition to Geldart type, the influence of particle shape
on the performance of vibration has also been investigated by
using two glass bead samples of the same mean diameter at 39
0.01 Am but different shapes, spherical and irregular. Fig. 7 shows
0.01 0.1 1 10 the comparison of the average bed voidage for the two samples
Superficial gas velocity (cm/s) fluidized with and without vibration. Although these two
(b) powders are in the same category (group A) according to
0.90
Geldart classification [3], the irregular shaped sample has a
0.85 GB-S (10µm)
much higher bed voidage in fluidization and shows a larger bed
GB-S (65µm)
Average bed voidage (-)

0.80
GB-S (216µm)
expansion ratio. The difference may be ascribed to the
0.75 Open symbols: No vibration differences in surface properties and hence the change in
0.70 Solid symbols: 50Hz/0.3mm
interparticle forces. In addition to the van der Waals force being
0.65 considered as the dominant interparticle force for most fine
0.60 particles [1], mechanical forces due to interlocking of re-entrant
0.55 surfaces can be significant for the particles of irregular shapes
0.50 [2]. As given in Table 1, the Hausner ratio (q bt / q ba = 1.68) and
0.45 the angle of repose (AOR = 44.7-) for the irregular glass beads
0.40 are much higher than those for the spherical one, suggesting an
0.35
0.01 0.1 1 10
increased interparticle force due to the irregular shape. In term
Superficial gas velocity (cm/s) of the Hausner ratio and the angle of repose, the 39 Am
irregular shaped glass beads are more likely classified as the
Fig. 5. Effects of mechanical vibration on hydrodynamic behaviours of glass group C powder according to the new criteria of Geldart et al.
beads (10, 65 and 216 Am): (a) the normalized bed pressure drop and (b) the
average bed voidage.
[17] and Zhao et al. [18]. Although it is obviously not true to
conclude that a group C powder will exhibit higher bed
From the discussion in the previous section regarding Figs. expansions than a group A powder, the theory of Jaraiz et al.
4 and 5, the performance of vibration varies significantly with [5] do suggest that for group A particles, an increased
the types of particles. The influence of the Geldart type on fine interparticle force will result in an enhancement in tensile
particle fluidization can be further revealed from Fig. 6, where
0.70
Glass Beads (39 µm)
Glass Beads (6.1-216 µm)
10.00 0.65
Average bed voidage (-)

0.60
umf (cm/s)

0.55

1.00
0.50

0.45 Spherical
Irregular

No vibration 0.40 Open symbols: No vibration


0.10
50Hz/0.3mm Solid symbols: 50Hz/0.3mm

0.05 0.35
5 10 100 300 0.0 1.0 2.0 3.0 4.0 5.0
dp (µm) Superficial gas velocity (cm/s)

Fig. 6. u mf as a function of particle size for glass beads with and without Fig. 7. Dependency of bed voidage on particle shape for glass beads (39 Am)
vibration. fluidized with and without vibration.
140 C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144

0.3 4.0
strength of the particle bed and lead to a larger bed expansion
Glass Beads (10 µm)
ratio in fluidization. On the other hand, the increased
Glass Beads (39 µm)
interparticle forces appear also to cause difficulty in fluidiza-
tion of the particles and result in a higher u mf as shown in Fig. 3.0
8, where it is shown that vibration (50 Hz/0.3 mm) leads to a
0.2
remarkable decrease of about 50% in u mf for the spherical

umf (cm/s)
sample, but only a slight reduction in u mf for the irregular

Λ (-)
shaped one. In addition to different shapes, another difference 2.0
between these two types of glass beads is the size-distribution.
As revealed in Fig. 2, the smooth glass beads have a narrow
0.1
size-distribution while the irregular-shape sample has a much
wider size-distribution. Such a difference in particle size 1.0

distribution, as well as particle shape, may both contribute to


their different behaviors in bed expansion.
Other influencing factors are the vibration parameters
0.0 0.0
including frequency, amplitude and angle. Because of the 10 20 30 40 50
nature of the mechanical vibration system, it is difficult to vary Frequency (Hz)
only the frequency while maintaining the same amplitude, Fig. 9. Dependencies of u mf and K on vibration frequency for glass beads (10
although it is obviously advantageous to keep the other and 39 Am).
parameters such as the vibration amplitude unchanged when
examining the influence of vibration frequency. Given fixed quency for both types of particles, whereas the decreasing
conditions of the vibrator, e.g., the unbalanced weight, the tendency is levelled off at about 30 Hz. At f > 35 Hz, however,
amplitude of vibration is strongly dependent on the frequency u mf increases slightly with frequency, suggesting that 35 Hz
used. With a fixed unbalanced weight of the vibrator, for could be the optimum frequency to achieve the best perfor-
instance, the vibration amplitudes are measured at 0.62, 0.34, mance of the vibration for the current vibrated fluidized bed
0.29, 0.27 and 0.30 mm at the frequency of 10, 20, 30, 40 and system. On the other hand, a monotonous increasing trend is
50 Hz, respectively. By taking both frequency and amplitude shown for the vibration strength with frequency. K increases
into account, a vibration strength (K) is adopted in this work to relatively slowly with f prior to 35 Hz while more sharply after
characterize the intensity of mechanical vibration. K is a this point, also suggesting the frequency of 35 Hz being a
dimensionless parameter defined as the ratio of acceleration of critical point in the relationship between K and f. The
vibration to that of gravity, i.e., experimental results on the influence of vibration amplitude
on u mf are depicted in Fig. 10, where the dependencies of u mf
K ¼ Að 2pf Þ2 =g: ð1Þ and K on vibration amplitude are plotted at a fixed frequency
A same definition of the vibration strength can be found in (20 Hz) for glass beads (10 and 39 Am). Clearly, for both
other works of Noda et al. [8] and Wang et al. [15]. powders, u mf decreases with both amplitude and the vibration
Fig. 9 shows dependencies of u mf and K on vibration intensity. It should be noted that although the frequency of 35
frequency ( f) for glass beads (10 and 39 Am). At lower
frequencies ( f  35 Hz), u mf decreases with increasing fre- 0.4 1.0
Frequency at 20 Hz
Glass Beads (39 µm)
1.0 Glass beads (10 µm)
Glass beads (39 µm) 0.8
0.3

0.8
0.6
umf (cm/s)

Λ (-)
umf (cm/s)

0.2
0.6
No vibration 0.4
50Hz/0.3mm
0.4
0.1
0.2

0.2

0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.5
0.0
Irregular shape Spherical shape Amplitude (mm)

Fig. 8. Dependency of u mf on particle shape for glass beads (39 Am) fluidized Fig. 10. Dependencies of u mf and K on vibration amplitude for glass beads (10
with and without vibration. and 39 Am).
C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144 141

Hz is shown to be a critical point for u mf And K in Fig. 9, the 1000


results illustrated in Fig. 10 are for a fixed frequency at 20 Hz. ms g/S σ σ σ σ
The reason why the frequency was fixed at a value other than
the critical value (35 Hz) in examining the effects of vibration ε 0 =0.605

Bed pressure drop (Pa)


amplitude is to avoid the possible predominance of the 500 ε 0 =0.647
frequency over the amplitude.
In addition to the influences of vibration frequency and ε 0 =0.678

amplitude, the effect of vibration angle (or direction) on


fluidization is also investigated with glass beads (39 Am) under ε 0 =0.704

frequencies of 10 –50 Hz, and the results are shown in Fig. 11.
As expected, the effect of vibration on the fluidization of the
powder changes greatly with the vibration angle. For most
frequencies tested, the best performance (leading to lowest u mf Glass Beads (6.1 µm)
values) is observed at the vibration angle of 0- (horizontal
100
vibration), worse at 45- and the worst at 90- (vertical 0.05 0.1 0.4
vibration). One possible reason is that the horizontal vibration Superficial gas velocity (mm/s)
could disrupt the channels formed or reduce the tensile strength
Fig. 12. Pressure drops across the bed for tensile strength tests for glass beads
of the particle bed more efficiently than the vertical vibration.
(6.1 Am).
In addition, the influence of frequency shown in Fig. 9 can also
be observed in Fig. 11: u mf decreases with the frequency at
lower frequencies and levels off at about 35 Hz. Like in many study, the following method is applied to determine the tensile
other mechanical vibration systems, the horizontal vibration strength of a powder bed with and without vibration. The bed
generated by the present vibrator-system may inevitably of fine particles is first consolidated to obtain a compacted bed
contain a minor component of vertical vibration that may also of different states of compaction by vibration for different
contribute to the improvement of fluidization quality. Due to lengths of time before being subject to the tensile strength tests.
the limited length of the paper, however, this effect is not Then, the fluidizing gas is introduced to the bed at a flow rate
discussed. increased slowly with a very small step at 1 – 5 ml/min, with the
bed pressure drop recorded simultaneously, until the eventual
3.3. Tensile strength of a particle bed rapture of the compacted bed. The tensile strength can then be
determined from the plot of the bed pressure drop in a similar
Tensile strength is a fundamental property of materials. In a manner to that reported by Watson et al. [19], as shown in Fig.
particle bed, the value of tensile strength provides a terminal 12 for 6.1 Am glass beads. The tensile strength is the extra
point of the yield loci of failure for samples in given states of pressure drop over and above the normal pressure drop
compaction. It presents the inherent interparticle attractive corresponding to the particle weight, required to rupture the
forces developed when a mass of particles is compacted. In this bed. This ‘‘extra force’’ is needed to overcome the interparticle
force, so that it can be another measure of the interparticle
0.25
force. The tensile strength thus obtained for all the group C
Glass Beads (39 µm) particles without vibration and for TiO2 (5.2 Am) particles with
and without vibration are shown in Fig. 13. In all cases, the
tensile strength reduces dramatically in a linear relationship
Vabration angle:
0.20 with the initial bed voidage. As the initial bed voidage
90° (Vertical)
increases, the bed compaction becomes less and so does the
0° (Horizontal)
45°
interparticle force. This obviously leads to much lower tensile
umf (cm/s)

strength. More importantly, it is shown that vibration can


0.15 dramatically lower the tensile strength of the particle bed.

3.4. Roles of vibration in the fluidization of fine particles

0.10 Despite the vital importance, a comprehensive understand-


ing on the roles of vibration in the fluidization of fine particles
has not yet been reported in the literature. Therefore, a general
discussion is presented here for the possible roles of vibration
0.05 with respect to bed voidage, pressure drop, agglomeration and
10 20 30 40 50
tensile strength of a particle bed.
Frequency (Hz) The bed voidage (or the bed expansion ratio) and the
Fig. 11. Effect of vibration angle on u mf for glass beads (39 Am) fluidized under pressure drop across the bed are the most commonly used
various frequencies. parameters to characterize the fluidization behaviours of
142 C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144

350
particles. Results from the previous section (Figs. 4 and 5) have
indicated a decrease in bed voidage and an increase in bed
300
TiO2 (5.2 µm)
pressure drop at a low gas velocity for all the particles with
vibration. Obviously, consolidation effect of the vibration may
account for the decrease in bed voidage. The cause for the
Tensile strength, σ (Pa)

250
increase in bed pressure drop may lie in two aspects. First, from
Al2O3 (4.8 µm)
200
the well-known Ergun equation [20],
CaCO3 (5.5 µm)  
DP 150ð1  eb Þ2 lg lg 1:75ð1  eb Þ qg u2g
150 ¼  þ
Glass Beads (6.1 µm) L drag eb3 /s d̄ p
2
e3b /s d¯ p

100
ð2Þ

a decrease in bed voidage (( b) due to vibration may lead to an


50
increase in bed pressure drop. Second, it has been observed that
TiO2 (5.2 µm)
(10Hz/0.3mm)
vibration can efficiently eliminate the formation of channels,
0
0.5 0.6 0.7 0.8 0.9
which also leads to more uniform fluidization, and therefore
Initial bed voidage, ε0 (-)
also an increase in pressure drop.
To examine the self-agglomeration of fine particles during
Fig. 13. Tensile strengths for group C particles and the effect of vibration. fluidization, a method to sample the agglomerates out of the

Fig. 14. Agglomerate formation in fluidization of the 5.5 Am CaCO3 powder at 9.5 cm/s for 3 h with and without vibration (45 Hz/0.45 mm): (a) the top-bed
agglomerates, (b) the middle-bed agglomerates and (c) the bottom-bed agglomerates.
C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144 143

bed without disruption is crucial. This is very challenging due (2) The effectiveness of vibration is strongly dependent on
to the fragile structure of the agglomerates. In this work, a the particle properties (Geldart particle type, size-
novel ‘‘on-line sampling’’ method has been developed to distribution and shape). The effectiveness is found to
prevent the agglomerates from being disrupted while main- be more significant for group C powders or smaller
taining their size and shape during sampling. The on-line particles than for groups A and B powders or larger
sampling method has been introduced in Xu et al. [16] and particles. In comparison with the spherical particles, the
described in more detail in another paper by Xu and Zhu [21]. irregular particles exhibit higher expansion ratios or
In brief, the agglomerates are sampled in situ, i.e., without larger bed voidages, but are more difficulty in becoming
stopping the fluidizing gas or making any other changes to the fluidized (leading to a higher u mf) probably due to the
fluidizing conditions (e.g. vibration) by using a sampling ladle enhanced interparticle forces.
from the top of the column. To sample the agglomerates in the (3) Vibration parameters also show strong influence on fine
upper region of the bed, the ladle is used to pick up the particle fluidization. At low vibration intensities, u mf
agglomerates directly. To sample the agglomerates in the decreases as the vibration intensity increases (by increas-
middle and bottom regions, an intensity-controllable vacuum ing either frequency or amplitude) but levels off when the
is used to carefully remove all the particulate materials above intensity reaches a critical value. For the first time, the
the sampling plane before a sample is taken. Fig. 14 shows the effectiveness of mechanical vibration is demonstrated to
microscopic images of agglomerates sampled from both the strongly depend on the vibration angle or direction, with
top and the bottom of the bed of the CaCO3 particles (5.5 Am) the best performance being observed at 0- (horizontal
after being ‘‘fluidized’’ in air at 9.5 cm/s for 3 h with and vibration) and the worst at 90- (vertical vibration).
without mechanical vibration (45 Hz/0.45 mm). Under no (4) Vibration can dramatically lower the tensile strength of
vibration, the maximum size of the top-bed agglomerates is the particle bed, suggesting that vibration can overcome
around 500 Am and the bottom-bed agglomerates around 6000 the interparticle forces of fine powders and improve their
Am, suggesting a very severe segregation of agglomerate size. fluidization quality.
With vibration, however, the maximum agglomerate sizes (5) The possible roles of mechanical vibration in fluidization
from the bottom-bed are reduced to about 500 Am and become of fine particles are examined with respect to bed
similar to that from the top bed at around 400 Am. Obviously, voidage, pressure drop, agglomeration and tensile
vibration can substantially reduce both the average size and strength of the particle bed. The decrease in bed voidage
the segregation of agglomerates in the bed, both very at a lower gas velocity (caused by the consolidation
important for improving the fluidization quality of cohesive effect of the vibration) and the elimination of the
particles. channels by vibration may account for the increase in
Since tensile strength is a fundamental property of a particle bed pressure drops. Vibration can also significantly
bed, representing the inherent interparticle attractive forces reduce both the average size and the segregation of
developed when a mass of particles is compacted, the role of agglomerates in the bed, both vitally important for
vibration may also be evaluated with respect to the tensile improving the fluidization quality of cohesive particles.
strength of a particle bed. As shown formerly in Fig. 13, for the
particle bed of TiO2 (5.2 Am), the tensile strength is Nomenclature
dramatically reduced when vibration is applied. It therefore A Amplitude of vibration (m)
suggests that vibration can overcome the interparticle forces of AOR Angle of repose (-)
fine particles and improves their fluidization quality. D Inner diameter of the fluidized bed column (m)
Therefore, the roles of mechanical vibration in fluidization dp Mean particle size (m)
of fine particles may be summarized as follows: the consoli- f Frequency of vibration (Hz)
dation effect of vibration causes lower bed voidages, the g Gravity acceleration (9.81 m/s2)
disruption of the channels and agglomeration leads to more L0 Static (unexpanded) bed height (m)
uniform fluidization and thus higher bed pressure drops as well L Bed height (m)
as a lower u mf, and the vibration also reduces interparticle ms Weight of solids in the fluidized bed (kg)
forces and thus the tensile strength. In one words, the external DP Pressure drop across the fluidized bed (Pa)
energy introduced by vibration helps the fluidization of RH Hausner ratio (= q bt / q ba) (– )
cohesive particles by disrupting channels, breaking the S Cross sectional area of the fluidized bed column (m2)
agglomerates and reducing the tensile strength of the particle ug Superficial gas velocity (m/s)
bed. u mf Minimum fluidization velocity (m/s)

4. Conclusions
Greek letters
(1) For all the powders examined, the fluidization quality (b Average bed voidage (– )
can be enhanced by mechanical vibration, leading to (0 Initial average bed voidage (– )
higher bed pressure drops at low superficial gas K Vibration strength ( – )
velocities and lower values of u mf. q ba Aerated bulk density of the particles (kg/m3)
144 C. Xu, J. Zhu / Powder Technology 161 (2006) 135 – 144

q bt Tapped bulk density of the particles (kg/m3) [9] A. Dutta, L.V. Dullea, A comparative evaluation of negatively and
qg Gas density (kg/m3) positively charged submicron particles as flow conditioners for a cohesive
powder, AIChE Symp. Ser. 86 (276) (1990) 26 – 40.
qp Particle density (kg/m3) [10] A.S. Mujumdar, Aerodynamics, heat transfer and drying in vibrated fluid
r Tensile strength of the particulate bed (Pa) beds, Am. J. Heat Mass Trans. 7 (1983) 99 – 110.
lg Gas viscosity (kg/m/s) [11] J.R. Wank, S.M. George, A.W. Weimer, Vibro-fluidization of fine boron
/s Particle sphericity ( –) nitride powder at low pressure, Powder Technol. 121 (2001) 195 – 204.
[12] J. Chen, Z. Wang, Y. Jin, Z. Yu, Vibro-fluidization of groups A, B and D
particles, Proceedings of the 6th National Conference on Fluidization,
Acknowledgements Wuhan, China, 1993, pp. 111 – 116.
[13] Y. Mawatari, T. Koide, Y. Tatemono, S. Uchida, K. Noda, Effect of
The authors are grateful to the Ontario Research and particle diameter on fluidization under vibration, Powder Technol. 123
Development Challenge Fund for supporting this study. (2002) 69 – 74.
[14] M. Kwauk, Fluidization-Idealized and Bubbleless with Applications,
Science Press and Ellis Horwood Ltd., Beijing and New York, 1992.
References [15] Y. Wang, T.-J. Wang, Y. Yang, Y. Jin, Resonance characteristics of a
vibrated fluidized bed with a high bed hold-up, Powder Technol. 127
[1] H. Krupp, G. Sperling, Theory of adhesion of small particles, J. Appl. (2002) 196 – 202.
Phys. 37 (1966) 4176 – 4180. [16] C. Xu, Y. Cheng, J. Zhu, Fine particle fluidization—effects of mechanical
[2] J.N. Staniforth, Ordered mixing or spontaneous granulation? Powder and acoustic vibrations, in: U. Arena, R. Chirone, M. Miccio, P. Salatino
Technol. 45 (1985) 73 – 77. (Eds.), Fluidization XI, Engineering Conferences International, New York,
[3] D. Geldart, Types of gas fluidization, Powder Technol. 7 (1973) 285 – 297. 2004, pp. 627 – 634.
[4] S. Mori, A. Yamamoto, S. Iwata, T. Haruta, I. Yamada, Vibro-fluidization [17] D. Geldart, N. Harnby, A.C. Wong, Fluidization of cohesive powders,
of group-C particles and its industrial applications, AIChE Symp. Ser. 86 Powder Technol. 37 (1984) 25 – 37.
(276) (1990) 88 – 94. [18] J. Zhao, H. Liu, H. Xu, Research on fluidization characteristics of group C
[5] E. Jaraiz, S. Kiruma, O. Levenspiel, Vibrating beds of fine particles: powders, Ind. Chem. 12 (1991) 249 – 255.
estimation of interparticle forces from expansion and pressure drop [19] P.K. Watson, J.M. Valverde, A. Castellanos, The tensile strength and free
experiments, Powder Technol. 72 (1992) 23 – 30. volume of cohesive powders compressed by gas flow, Powder Technol.
[6] R. Yamazaki, Y. Kanagawa, G. Jimbo, J. Chem. Eng. Jpn. 7 (1974) 115 (2001) 45 – 50.
373 – 378. [20] S. Ergun, Fluids flow through packed column, Chem. Eng. Prog. 48 (2)
[7] A.W. Siebert, D. Highgate, M. Newborough, Heat transfer characteristics (1952) 88 – 94.
of mechanically-stimulated particle beds, Appl. Therm. Eng. 19 (1999) [21] C. Xu, J. Zhu, Experimental and theoretical study on the agglomeration
37 – 49. arising from fluidization of cohesive particles—effects of mechanical
[8] K. Noda, Y. Mawatari, S. Uchida, Flow patterns of fine particles in a vibration, Chem. Eng. Sci. 60 (2005) 6529 – 6541.
vibrated fluidized bed under atmospheric or reduced pressure, Powder
Technol. 99 (1998) 11 – 14.

Das könnte Ihnen auch gefallen