Você está na página 1de 11

IIITIEIIIIIIIII011M10UIIIIIIL~

mlnERIIt
PRO(ENSlH6
ELSEVIER Int. J. Miner. Process. 51 (1997) 139-149

Role of hydrodynamic cavitation in fine particle


flotation
ZA. Zhou 1, Zhenghe Xu 2, J.A. Finch *, H. Hu, S.R. Rao
Department of Mining and Metallurgical Engineering, McGill University, Montreal, Que. H3A 2A7, Canada
Accepted 14 April 1997

Abstract

High ,energy dissipation, high feed velocity and relatively high feed pressures have been used
in some recently developed flotation cells to enhance fine particle flotation. One contributing
factor to the high flotation kinetics demonstrated in these new cells may be in-situ bubble
formation by hydrodynamic cavitation on hydrophobic particles. This could induce particle
aggregation through gas nucleus bridging with the consequently larger particles having a higher
collection rate. Both experimental measurements and numerical analysis indicated bubble forma-
tion whenever the water vapour pressure was reached. The presence of surfactants and dissolved
gas was found to preserve the bubbles generated by cavitation, thus producing more small
bubbles. Flotation of fine silica and zinc sulphide precipitates in a flotation device incorporating
hydrody~amic cavitation showed a significantly increased flotation kinetics. 1997 Elsevier
Science B.V.

Keywords: hydrodynamic cavitation; in situ gas nucleation; fine particle flotation; cavitation reactor

1. Introduction

Flotation is a versatile, surface property-based physical separation process which has


been widely used in recovering minerals and increasingly in a variety of environmen-
tally driven applications. In conventional dispersed air flotation, flotation efficiency

* Corresponding author. Tel.: + 1 (514) 398-4352; fax: + 1 (514) 398-4492; e-mail: jim@minmet.lan.mc-
gill.ca
1 Present address: Department of Chemical and Materials Engineering, University of Alberta, Edmonton,
Alta., T6G 2G6, Canada.
z Present address: Department of Chemical and Materials Engineering, University of Alberta, Edmonton,
Alta., T6G 2G6, Canada.

0301-7516/97/$17.00 1997 Elsevier Science B.V. All rights reserved.


PH S03bl-75 1 6 ( 9 7 ) 0 0 0 2 6 - 4
140 ZA. Zhou et al. l i n t . J. Miner. Process. 51 (1997) 139-149

tends to decrease significantly for particles finer than 10-20 /~m, mainly due to low
mass (and hence low inertia) which results in unfavourable hydrodynamic conditions tbr
the particle-bubble collision. In addition, fine particles do not have sufficient kinetic
energy to rupture the intervening liquid film between a particle and a bubble, thus
contributing to a low attachment efficiency. A key to enhancing fine particle flotation is
to develop an innovative approach for particle-bubble collision and attachment.
At least two mechanisms have been considered for particle-bubble attachment in
flotation: gas precipitation (as in dissolved air flotation and vacuum flotation), and
impaction/sliding (as in conventional mechanical flotation cells and flotation columns)
(Klassen and Mokrousov, 1963). Only the latter has been investigated in much detail
over the last 50 years, due to the relative simplicity of theoretical treatment and
experimental verification. Conventional flotation columns present an ideal situation for
particle-bubble attachment by the sliding mechanism. In mechanical flotation cells,
sliding coupled with some degree of gas nucleation is believed to contribute to
attachment (Klassen and Mokrousov, 1963). In such cells as Jameson cells (Jameson,
1988), pneumatic cells (Bahr, 1985), and those using static mixers (Adel et al., 1991;
Jordan and Susko, 1992; Xu et al., 1996), the reported high flotation rate may be due to
cavitation, or gas nucleation, coupled with strong particle-bubble impaction.
Direct formation of bubbles on particles by gas nucleation and cavitation is instanta-
neous as it eliminates the collision stage. This suggests that with appropriate engineering
gas nucleation/cavitation can be exploited to improve the collection of fine particles.
Fast flotation kinetics has been claimed for some new flotation devices. Some common
features of Jameson cells (Jameson, 1988), pneumatic cells (Bahr, 1985), contact cell
(Amelunxen, 1993), microcell (Yoon, 1993), air-sparged hydrocyclone (Miller et al.,
1988), etc., are: high energy dissipation, high feed stream flow velocities (6-12 m / s )
and relatively high feed pressures (2-3 atm). Under these conditions, bubble formation
by hydrodynamic cavitation is favoured. To understand the mechanisms of fast particle
collection rates in these novel flotation devices, a systematic investigation on the role of
hydrodynamic cavitation in fine particle flotation has been undertaken, and a newly
designed flotation reactor has been tested. A significant improvement in flotation
kinetics was observed when the flotation was assisted by the hydrodynamic cavitation.

2. Bubble formation by hydrodynamic cavitation

2.1. Principles

Hydrodynamic cavitation is defined as the formation of bubbles (either gas- or


vapour-filled cavities) in a flowing fluid due to the rupture of a liquid-liquid or a
liquid-solid interface caused by reduction of local pressure to a critical value (e.g.,
liquid vapour pressure) (Ross, 1976). It is evident that the formation of bubbles by
hydrodynamic cavitation must start at 'weak points' in a system. A reduction of local
pressure in a flowing liquid can be achieved, for example, by increasing flow velocity
within a restricted section, according to Bernoulli's principle (Fig. 1):
1 2 I 2
P1 + 2PI VI = P2 q- 2Pl V2 = C (constant) (1)
Z.A. Zhou et a l . / lnt. J. Miner. Process. 51 (1997) 139-149 141

Multimeter

1 (~P~ [ / ~Photodiode

D ~ v2,P2~ d

Lightsource/
I?1
Fig. 1. Schematicdiagramof the cavitationtube used in our experimentalhydrodynamiccavitationstudy.

in which V is the liquid (water) flow velocity at a point where the pressure is P, and Pl
is the density of liquid. By increasing V2 sufficiently, P2 will drop to (or below) liquid
vapour pressure, and cavity bubbles will form.
Favouring this mechanism in flotation is that solid particles and pre-existing gas
nuclei provide 'weak points' which facilitate hydrodynamic cavitation. The actual
energy required for cavitation may, therefore, be much less than that predicted by
theory. Dean (1944) reported that due to expansion of pre-existing nuclei, bubble
formation by cavitation can occur at less than one-tenth of the energy required when
nuclei are absent. Nicol et al. (1986) investigated the effect of pre-existing gas nuclei on
heterogeneous nucleation of microbubbles on hydrophobic particle surfaces in an
acoustic field, and observed an improvement in fine particle flotation. In addition, the
weaker adhesion between hydrophobic particles and water than the cohesion between
water molecules may provide another source of weak point for bubble formation by
cavitation. Although a recent investigation by Ryan and Hemmingsen (1993), using
particles with a diameter of 1 to 2.5 /xm, demonstrated that smooth surfaces, whether
hydrophilic or hydrophobic, did not facilitate gas bubble formation by dissolved gas in
water under static conditions, different responses under dynamic conditions were
reported in mineral flotation systems where the mineral particles are often irregular in
shape with a rough surface (Anfruns and Kitchner, 1977; Zhou et al., 1994).

2.2. E x p e r i m e n t a l o b s e r v a t i o n s

We have demonstrated bubble formation by hydrodynamic cavitation using so-called


cavitation tubes (Zhou et al., 1995a,b). In this set-up, water experienced a significant
velocity increase when passing through a nozzle. Bubble formation was detected by a
light transmittance method. The results (Fig. 2) showed that at a fluid flow velocity
above ca. 16 m / s , bubbles formed in tap water, as indicated by a significant decrease in
transmit~Led light. In the presence of surfactant molecules, the on-set velocity for
hydrodynamic cavitation reduced only marginally; however, the amount of bubbles
142 Z.A. Zhou et al. lint. J. Miner. Process. 51 (1997) 139-149

100 = ~ = ~'- :z~'LL"-r~..mSZ'L'--.~


-~. ~... m~..~
.~ 95

No Chemicals "\ "',;_


v 90 \ ~,. -

e-
DDA (1.7"10-5 M) ~ "", ~.
.~-
E
85 DDA (1.0"10-4 M) \ "[~.
-

DDA (1.0"10-3 M) ~ ~',,. []


~ 80 ? N
r: \
._~
..J

75 \
x~
(7
70 I i I i I ~ I i I
12 14 16 18 20
Liquid Flow Velocity Through Nozzle, (m/s)
Fig. 2. Effect of surfactant addition on bubble formationby hydrodynamiccavitation (d = 1.3 mm, no air
addition, pH = 5).

generated increased significantly (light transmittance was lowered). This finding indi-
cates that changes in surface tension or additional weak points introduced by surfactant
molecules have little effect on the activation energy needed to initiate cavitation. This
implies that hydrodynamic cavitation in real systems appears to be related to the
expansion of pre-existing gas nuclei: when such nuclei experience reduced pressures in
the cavitation tube, they expand. (The existence of gas nuclei in water has been
confirmed from coagulation, sedimentation and filtration tests of fine coal and silica
(Zhou et al., 1996a,b).) The surfactant adsorbs on the expanded gas nuclei, preventing
both their collapse and coalescence, thus generating more stable and smaller bubbles.
To test this explanation, a small amount of air (less than 3% of the liquid) was added
to the water. The effect (Fig. 3) was to reduce the velocity for onset of bubble formation,
and to produce more bubbles. The significant reduction of liquid flow velocities for
bubble initiation by cavitation (down to 5 - 7 m / s from ca. 16 m / s ) has been
demonstrated with different gas content and cavitation tube geometries (Zhou et al.,
1995a,b). The effect of increased air content on enhancing cavitation was also reported
by Holl (1960).

2.3. Numerical analysi s

To further understand the hydrodynamic behaviour of liquid flowing through a


cavitation tube, a numerical procedure has been developed (Hu et al., 1997). A wall
function with wall friction was incorporated into the turbulent energy and turbulent
dissipation ( k - e ) model. By using a novel boundary fitted coordinate approach for grid
generation, the stream function, vorticity, velocity and pressure distributions in the
Z.A. Zhou et al./ lnt. J. Miner. Process. 51 (1997) 139-149 143

100

~e
~',..
,.. ~ x ",,

~...-~
..... . ~

"'.,. ', ""~. ~


_'~' 95 noair ,(. ,,.. .. .."% \ ~, ""........
%%
ai [] ,~ .
o
f- air (10 mllmln)
[] ,. ,, ~'... \
~ 90 air(2Oml/min) (~.\v XX? ,,,_~
el air (30 ml/min) ". e, "~,,,~
t"

g~ 85 ~, ,,
'% \

80 , I , I , I , I , I ,"" ~ *
8 10 12 14 16 18 20 22
Liquid Flow Through Nozzle (m/s)
Fig. 3. EfFect of air addition on bubble formation by hydrodynamic cavitation (d = 1.3 ram, no chemicals).

cavitation tube were obtained. Simulation results were found to be in good agreement
with literature data (Hu et al., 1997). The pressure distributions calculated for the
cavitation tube compared with experimentally measured values as a function of fluid

velocity v
6.7 m/s

8.5 m/s
0 ~.t .... ..... . - - ~
-, .,. , , . . ~ / 10.6 m/s
-/

E 12.7 m/s

...."""--9 / 15.9 mls


i '**'"eo /
Q_ water vapor pressure
.s ro
I ~0 ........................................................................... e

-'15
0.05 0.1 0.15 0.2
Distance. Z (m)
Fig. 4. Comparison of predicted and measured pressure changes along the cavitation tube (no chemicals, no air
addition).
144 ZA. Zhou et al. lint. J. Miner. Process. 51 (1997) 139-149

Pressure
gauge
Q t i
Cavitation
tube ......
Airt--~ Manl~meter
pro)
1% silica (-5
10ppmDF250
pH=7.5- 7.8
1.25"10-4 MDAH
TailingsAGITAIRcell
Fig. 5. Schematicdiagramof set-up for continuous laboratoryflotation tests.

flow velocity are shown in Fig. 4. Again, excellent agreement was observed. At the flow
velocity of 16 m / s where bubble formation occurred, the pressure within the nozzle
reached - 10 m water, i.e., approached water vapour pressure. The simulation results
provide a scientific basis to examine hydrodynamic cavitation.

3. Fine particle flotation assisted by cavitation

To demonstrate the role of hydrodynamic cavitation in fine particle flotation, a


laboratory-scale flotation reactor employing hydrodynamic cavitation was designed (Fig.
5). Continuous flotation was conducted by pumping the slurry (1% w / w for - 5 /zm
silica and 0.2% w / w for zinc sulphide (ZnS) precipitates) through a cavitation tube into
a laboratory mechanical flotation cell (3 1 Agitair cell for silica and 5 1 Noranda
mini-cell (Xu et al., 1992) for ZnS precipitates). The cavitation tube was made from
glass with inner nozzle diameters of 1.30, 1.40 and 1.65 mm (Zhou et al., 1995a),
respectively. Chemicals used included 1.25 X 10 - 4 M dodecylamine (DDA) as collec-
tor, and 10 ppm Dowfroth 250 as frother (to produce more stable froth and to have an
enhanced bubble formation by cavitation). The concentrate and tails were collected
simultaneously for a given period of time and weighed to obtain the feed flow velocity
through the nozzle, and the retention time (~-) of slurry in the flotation cell. The
recovery, R c, was calculated using:

&- - - x lOO(%) (2)


Wc+Wt
where W~ and Wt are the dry weight of concentrate and tailings, respectively. Assuming
ZA. Zhou et al. l i n t . J. Miner Process. 51 (1997) 139-149 145

55
feed stream
no air
!50 []
feed stream
,45 air (0.15 Umin)

~
o
40

o
'~ 35

30

25 normal conditions

20 , I ~ I ~ I
10 15 20 25 30
Slurry flow velocity through nozzle (m/s)
Fig. 6. Effect of slurry flow velocity on - 5 /xm silica recovery. Conditions: 1% w/w silica; 1.25 10 -4 M
DDA; 10 ppm DF250; 3-1 Agitair cell; air rate in the cell 2 1/min; impeller speed: 850 rpm; r = 1.5 min;
nozzle size d = 1.3 and 1.65 mm. (Normal conditions indicate no cavitation tube nor added air in the feed
stream.)

perfect mixing and first-order kinetics, the fotation rate constant (k) was obtained by
solving the following equation for a given recovery and retention time:
k~-
R, (3)
1 +kr
In some of experiments, a small amount of air (less than 10% of the air used in the
mechanical cell) was added on the suction side of the pump to test the effect of added
air. Baseline flotation was conducted by pumping the slurry directly to the mechanical
cell. (It should be noted that flotation chemistry was not optimized in flotation tests to
investigate the role of hydrodynamic cavitation in fine particle flotation.) The results
obtained with fine silica and ZnS precipitates are shown in Figs. 6 and 7, respectively.

3.1. Silica

The results (Fig. 6) showed that a substantial increase in fine silica recovery for a
given flotation period when using the cavitation tube (without added air) at a flow
velocity greater than 17 m / s , corresponding to the velocity of hydrodynamic cavitation
(Figs. 2 and 3). (Note: to keep the same retention time in the mechanical cell, the slurry
velocity was adjusted by changing the nozzle size of the cavitation tube.) This finding
clearly demonstrates the role of hydrodynamic cavitation in increasing fine particle
flotation.
A much more significant increase in silica recovery (from 30 to 53%) was obtained at
a lower flow velocity (ca. 15 m / s ) when a small amount of air was introduced into the
feed line. This observation is consistent with our finding (Fig. 3) that cavitation occurs
146 Z.A. Zhou et al. lint. J. Miner. Process. 51 (1997) 139-149

60
k (min-1) I 2 - with~vitation tube ..
~ 50
>
1 0.135 . , , . . ' ..... "

8 40 i2 0.225
ne i

~ 30
u b e
"~ 20
Q..
mlO
C
N
o
0 1 2 3 4 5 6
Retention Time, (min)
Fig. 7. Comparison of ZnS precipitate flotation with and without cavitation tube in the feed stream (no air in
the feed stream, nozzle size d = 1.4 mm).

at a lower fluid flow velocity when the air is added. The most important finding from
these results is the presence of an optimal velocity (18 m / s in the case with added air)
of slurry passing through the nozzle to maximize the recovery. It seems that if the liquid
flow velocities within the nozzle become too high, the high shear force developed may
disrupt the bubble/particle aggregates, causing the observed reduction in recovery.
To further illustrate the role of cavitation, a comparison was made of flotation rate
constants (Table 1). A 40% increase in rate constant was obtained using a cavitation
tube (1.3 mm nozzle diameter) even though the overall aeration was less (2.15 1/min
compared to 3 1/min without the tube). This increase in flotation rate constant again
suggests that small bubbles generated by cavitation in the feed stream played a role in
enhancing flotation kinetics.

3.2. ZnS precipitates

The precipitates were prepared by mixing ZnSO4 with Na2S in tap water to give a
ZnS precipitate concentration of 2 g/1. The slurry was left overnight to minimize the
effects of the initial oxidation of freshly prepared sulphide precipitates. The pH of the
suspension was adjusted to 8.6 by NaOH, and the suspension was conditioned with

Table 1
Effect of different operating variables on fine silica ( - 5 ~ m ) flotation
Conditions Air ( I / m i n ) Rate constant k ( m i n - 1) Velocity in nozzle ( m / s )
cell feed
Normal a 3 - 0.40 -
Cavitation tube (nozzle d = 1.3 mm) 2 0.15 0.56 18.4

a Normal conditions means no cavitation tube nor added air in the feed stream.
Z.A. Zhou et al. ~Int. J. Miner. Process. 51 (1997) 139-149 147

DDA and frother for 10 min. Fig. 7 shows that the flotation rate increased significantly
when using a cavitation tube (even without the added air in the feed stream), and ca.
65% increase in flotation rate constant (see the insert) was obtained. These results again
illustrate the role of hydrodynamic cavitation.

4. Discussion

The present work has clearly demonstrated the potential benefit of increased recovery
by incoilaorating a cavitation tube in an otherwise conventional mechanical flotation cell.
The following two factors may have contributed to the increased flotation rate of fine
particle,~;. For one, the tiny bubbles formed in the feed stream, or more preferably in situ

a) Without tiny bubbles With tiny bubbles

@
000 #
b) One stage attachment Two-stage attachment

Q Q
+ 4.

@ &

flotation-sized bubble

particle (or aggregate)

0 tiny bubble
Fig. 8. Schematic illustration of possible roles of hydrodynamic cavitation in fine particle flotation: (a)
enhanced coagulation by bubble bridging; (b) one stage vs two stage attachment. (One stage attachment:direct
contact of a particle with a flotation-sized bubble; two-stage attachment: tiny bubbles attach to a mineral
particle, followed by the particle-bubble aggregate attaching to a flotation-sized bubble by coalescence of the
tiny bubble with the flotation size bubble.)
148 ZA. Zhou et aL / Int. J. Miner. Process. 51 (1997) 139-149

on hydrophobic particles, by hydrodynamic cavitation, may cause flocculation by a


bubble-bridging mechanism, resulting in an increase in the apparent particle size (Fig.
8a), thus increasing the collision probability with the bubbles in the flotation cell. In the
second, particles frosted with tiny bubbles may present a surface favourable to attach-
ment to flotation-sized bubbles, following the two-stage attachment mechanism (Fig. 8b)
proposed by Dziensiewicz and Pryor (1950), and Klassen and Mokrousov (1963).
An important consideration in applying hydrodynamic cavitation to fine particle
flotation is to minimize shear that otherwise may tear bubble-particle aggregates apart.
This appears to have occurred here (Fig. 6) and may be a designing factor in other new
devices (Tremblay et al., 1993). Clearly, both the total number of tiny bubbles produced
by cavitation and the stability of the particle-bubble aggregates formed need to be
considered in designing a cavitation tube. The presence of a maximum in flotation
response to flow velocity through a cavitation tube suggests that the main contribution
of fast slurry feed velocity to the increased flotation rate is not due to shear effects
dispersing the air and producing more small bubbles, but rather it is to generate bubbles
in situ on hydrophobic surfaces. If the role was only to generate fine bubbles, the
recovery should have increased with the flow velocity continuously until it levelled off.

5. Conclusions

(1) Experimental results and theoretical analysis showed that the local pressure of a
fluid stream can be reduced below its vapour pressure by increasing the flow velocity,
which induced bubble formation by hydrodynamic cavitation.
(2) The presence of surfactants and dissolved gas enhanced bubble formation by
hydrodynamic cavitation.
(3) Incorporating a cavitation tube in the feed line to a conventional flotation cell
increased flotation kinetics of fine silica and ZnS precipitates significantly, demonstrat-
ing the positive role of hydrodynamic cavitation in fine particle flotation.
(4) An optimum in the feed velocity was detected above which recovery decreased,
probably due to high shear disrupting bubble/particle aggregates.

Acknowledgements

The authors acknowledge the financial support from Natural Sciences and Engineer-
ing Research Council of Canada under Strategic Grant NSERC-STR0149414. The
in-kind support of Cominco, Inco and Noranda as well as the constructive comments
from the referees are also acknowledged.

References

Adel, G.T., Mankosa, M.J., Luttrell, G.H., Yoon, R.H., 1991. Full-scale testing of microbubble column
flotation. In: Agar, G.E., Huls, B.J., Hyma, D.B. (Eds.), Column '91. CIM, Montreal, pp. 263-274.
Z.A. Zhou et al. / lnt, J. Miner. Process. 51 (1997) 139-149 149

Amelunxen, R.L., 1993. The contact cell: A future generation of flotation machines. Eng. Mining J., April,
36-37.
Anfruns, JF., Kitchner, J.A., 1977. Rate of capture of small particles in flotation. Trans. Inst. Min. Metall. 86,
C9-C15.
Bahr, A., 1985. Application and sizing of a new pneumatic flotation cell. In: Proc. 15th Int. Mineral
Proces:dng Congr., Cannes, pp. 314-326.
Dean, R.B., 1944. The formation of bubbles. J. Appl. Sci., 15: 446-451.
Dziensiewicz, J., Pryor, E.J., 1950. An investigation into the action of air in froth flotation. Trans. IMM.,
Londola, 59: 455-491.
Holl, J.W., 1960. An effect of air content on the occurrence of cavitation. J. Basic Eng., 82: 941-946.
Hu, H., Zhou, Z.A., Xu, Z., Finch, J.A., 1997. Numerical and experimental study of a cavitation tube. Trans.
Inst. Min. metall., submitted.
Jameson, G.J., 1988. A new concept in flotation column design. In: Sastry, K.V.S. (Ed.), Column Flotation
'88. SlvIE Annual Meeting, Inc., Littleton, Co., pp. 281-286.
Jordan, C.E., Susko, F.J., 1992. Rapid flotation using a modified bubble-injected hydrocyclone and a
shallow-depth froth separator for improved flotation kinetics. Miner. Eng., 5: 1239-1257.
Klassen, V.I., Mokrousov, V.A., 1963. An Introduction to the Theory of Flotation. Butterworths, London, 493
pP.
Miller, J.D., Ye, Y., Pacquet, E., Baker, M.W., Gopalakrishnan, S., 1988. Design and operating valuables in
flotation separation with the air-sparged hydrocyclone. In: Forssberg, K.S.E. (Ed.), Proc. 16th Int. Mineral
Processing Congress. Elsevier, Amsterdam, pp. 499-510.
Nicol, S.I~., Engel, M.D., Teh, K.C,, 1986. Fine particle flotation in an acoustic field. Int. J. Miner. Process.,
17: 143-150.
Ross, D., 1976. Cavitation. Pergamon Press, New York, pp. 203-251.
Ryan, W.L., Hemmingsen, E.A., 1993. Bubble formation in water at smooth hydrophobic surfaces. J. Colloid
Interface Sci., 157: 312-317.
Tremblay, R., Roberts, K., Burrows, M., 1993. Jameson cell testing at the Kidd Creek Concentrator. In: Proc.
25th Annu. Mtg. Canadian Mineral Processors, Ottawa, Pap. No. 23.
Xu, M., Zhang, Q., Rao, S.R., Finch, J.A., 1992. An in-plant test of sphalerite flotation without Cu activation.
In: Proc. 24th Annu. Mtg. Canadian Mineral Processors, Ottawa, Pap. No. 14.
Xu, M., Quinn, P., Stratton-Crawly, R., 1996. A feed-line aerated flotation column, Part 1. Batch and
continuous testwork. Miner. Eng., 9: 499-508.
Yoon, R.H., 1993. Microbubble flotation. Miner. Eng., 6: 619-630.
Zhou, Z.A., Xu, Z., Finch, J.A., 1994. On the role of cavitation in particle collection during flotation--a
critical review. Miner. Eng., 7: 1073-1084.
Zhou, Z.A., Xu, Z., Finch, J.A., 1995a. Generation of small bubbles by hydrodynamic cavitation. Submitted to
Int. J. Miner. Process.
Zhou, Z.A., Xu, Z., Finch, J.A., 1995b. Fundamental study of cavitation in flotation. In: Proc. 19th Int. Miner.
Proce:~s. Congr., Vol. 3, San Francisco, pp. 93-97.
Zhou, Z.A., Xu, Z., Finch, J.A., 1996a. Effect of gas nuclei on hydrophobic coagulation. J. Colloid Interface
Sci., ]79: 311-314.
Zhou, Z.A., Xu, Z., Finch, J.A., Liu, Q., 1996b. Effect of gas nuclei on the filtration of fine particles with
diffenmt surface properties. Colloids Surf., 113: 67-77.

Você também pode gostar