Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

International Journal o[ Mineral Processing, 21 {1987) 241-260 241

Elsevier Science Publishers B.V., Amsterdam - - Printed in The Netherlands

Particle Size D e p e n d e n c e in Flotation D e r i v e d


from a F u n d a m e n t a l Model of the Capture P r o c e s s

G.S. DOBBY and J.A. FINCH


Department of Metallurgy and Materials Science, University of Toronto, Toronto, Ont.
MSS 1A4 (Canada)
Department of Mining and Metallurgical Engineering, McGiU University, Montreal, Que.
H3A 2A 7 (Canada)
(Received October 26, 1985; accepted after revision March 3, 1987)

ABSTRACT

Dobby, G.S. and Finch, J.A., 1987. Particle size dependence in flotation derived from a funda-
mental model of the capture process. Int. J. Miner. Process., 21: 241-260.

A flotation model is described wherein particle collection is considered to occur by parti-


cle-bubble collision followed by the particle sliding over the bubble during which attachment may
occur. Collision is quantified by a collision efficiency Ec. The existing model for Ec is extended
by including particle inertia. Attachment is quantified by an attachment efficiency EA calculated
as the fraction of particles which reside on the bubble for a time greater than the induction time.
These calculations use recent data on the liquid velocity distribution over the upper half of rigid
spheres. Collection efficiency EK is given by EK = Ec'EA. Detachment is not considered.
The model is examined and shown to agree with much experimental data. For example the
maximum observed in recovery vs particle size (dp) can be explained as Ec increases with increas-
ing dp, while EA decreases. The decrease in selectivity at fine sizes is in accord with the model
result that for very small particle sizes EA is large even for long induction times. A decrease in
bubble size does increase EK but does not improve selectivity. A high particle density ( > 5 g cm 3)
is shown to cause a sharp peak in recovery vs size and a decreased maximum size of flotability.
Finally, the effects of bubble concentration are explored.

INTRODUCTION

A fundamental model of particle collection by direct encounter with bubbles,


the coursing bubble hypothesis, is developed in this paper. This mechanism is
most closely realised in flotation columns. The prospect of developing a fun-
damental model to predict flotation rate constants for flotation columns pro-
vided the impetus of this work. In mechanically stirred flotation cells, collection
can occur by gas nucleation on particles in the low pressure region behind the
impellers, and the impellers impart turbulence which may lead to bub-
ble-particle detachment. These possibilities are not considered here. Never-

0301-7516/87/$03.50 © 1987 Elsevier Science Publishers B.V.


242

theless direct particle-bubble encounter must also play a principal role in


mechanical cells, thus the model will be applicable to these cells as well.

MODEL OF PARTICLE COLLECTION BY ENCOUNTER WITH BUBBLES

Overview

The proposed model is similar in concept to that of Sutherland (1948). Par-


ticle collection is considered to occur by particle collision followed by the par-
ticle sliding over the bubble surface.
Collisions are quantified by a collision efficiency Ec defined as the rate at
which particles collide with the bubble divided by the rate at which particles
flow across the projected area of the bubble (Weber, 1981 ).
Attachment efficiency EA is defined as the fraction of all colliding particles
that reside on the bubble for a time greater than the induction time ti, i.e. have
a sliding or contact time > ti.
The model does not consider detachment; consequently the collection effi-
ciency E~ is given by:

EK = E c 'EA (1)

( It will be shown that detachment of the particle from the bubble need not be
invoked to explain poor recovery of intermediate to large diameter particles).
The principal limitation in the model of Sutherland is the assumption of
potential flow. This work treats flotation sized bubbles (Reb up to 400) and
considers the recent findings on the fluid velocity distribution over the surface
of rigid spheres ( Seeley et al., 1975 ).
The assumption of bubbles acting as rigid spheres is reasonable in the sur-
factant solutions used in flotation (Clift et al., 1978). This assumption also
means that bubble rise velocity ub can be calculated by the method of Concha
and Almendra (1979). Particle settling velocity Up is also computed in this
manner (particles are assumed spherical and shape corrections are not
employed).
Collection efficiency is directly proportional to the flotation rate constant,
k (Jameson et al., 1977 ). Consequently it lies at the heart of flotation model-
ling. The observations regarding EK can, therefore, be extended to k and
recovery.
The model is developed initially for the single bubble-single particle system.
The model consequences are then examined and are shown to fit many obser-
vations in flotation. Later, an attempt is made to model the more practical case
of a bubble swarm.
243

i
i

2 tI

; x

Fig. I. Illustration of a particle approaching a gas bubble, x, y and r coordinates are dimensionless;
at the bubble surface r2=x2+y2= i.

Collision model

Fig. I illustrates a particle approaching a gas bubble. The equations of mo-


tion may be written:
dv* * * (2)
S k -d~z = ux - v x

S ~ t ~ = ( u* + u* ) - v * (3)

The Stokes number (Sk) is related to the ratio of inertial to drag forces.
Most work has focused on conditions where Sk < < 1 and thereby allowing this
inertial term to be ignored.

Low particle inertia, Sk < O.1


In the collision model of Weber and Paddock (1983), which assumes S k = 0,
total collision efficiency is the sum of gravitational and interceptional collision:
Ec = Ecg + Eci (4)
244

Interceptional collision alone I Ec~ = 0) occurs for neutrally buoyant particles


which follow the fluid streamlines exactly. Gravitational collision alone
I Eci= 0) is hypothetical; it would occur for particles having a finite settling
velocity but zero dimension.
E(:~ is given by (Weber and Paddock, 1983 ):
1.5 (dp y!r ('3/16)R% ]
Ec~ - l + u * \ d ~ ] k 1 + 1 + 0 . 2 4 9 R e h °5(~ (5a)

for 0 < Reb < 300. Ec, is given by ( Reay and Ratcliff, 1973 ) :

u* dp 2 2
E c ~ - i + u . ( 1 + = - ) ' sin' 0~ (5b)
db

where 0c is the angle (from the front stagnation point of the bubble) where
the fluid streamlines come closest to the bubble (see also the Notation). 0~
has been fitted to the results of Woo (1971) :
0c = 7 8 . 1 - 7 . 3 7 log Reb (20<Reb <400) (6a)
0¢=98.0-12.491og(10Reu) (l<Reb<20) (6b)
0~. = 9 0 . 0 - 2.5 log (100 Reu) (0.1<Reb < 1) (6c)

Intermediate particle inertia


Collision efficiencies for Sk > 0.1 are calculated in this work by determining
particle trajectories using a numerical solution to eqs. 2 and 3 and finding the
grazing trajectory by trial-and-error.
The information needed to complete the numerical solution are values for
Ux*and u~; the liquid velocity profile for the top half of the bubble. For Stokes
flow and potential flow these velocities can be determined from an analytical
solution of liquid flow around a solid sphere. Flint and Howarth (1971) used
the analytical solutions to numerically integrate eqs. 2 and 3 and thus derive
Ecg as a function of Sk for both Stokes flow and potential flow (they did not
consider Eci ). What is needed, however, is the velocity distribution for spheres
having Reynolds number in the range 20-300 (typical of flotation sized bub-
bles ). A good illustration of velocity distributions close to the surface of a rigid
sphere are the experimental results of Seeley et al., (1975) shown in Fig. 2.
Fig. 2a shows the liquid tangential velocity uS vs distance from the surface of
the sphere, r* - 1, at 0 = 45 °. The radial velocity is shown in Fig. 2b. The the~
oretical values for potential and Stokes flow are also indicated; neither is a
reasonable approximation. Note the steep tangential velocity gradient close to
the surface with Uo going to zero at the surface. The tangential velocity gradient
at the surface is the surface vorticity ~s. The slope of uS vs ( r* - 1 ) is approx-
imately ~s from ( r* - 1 ) = 0 to ~ 0.05. Therefore, close to the bubble, let:
245

0.07,

1.0 (b)
flow -o.o31 " 2 ~ ' ~ " - " - - - - ~ t ent io I flow

-0.131

-0.22l
o 0.6

z~ -0.32
-~
,= 0.~ zx • w

"6 -0.z.2[
0.2
-0.52

0.1 0.2 0.3 0.,¢


Radial distance, r*-1
o'i 0'.2 o'.3 oI~ o'.s
Radial distance, r*-1
Fig. 2. Liquid velocity on the surface of a solid sphere versus distance from surface ( r * - 1) at
0 = 45 ° (Seeley et al., 1975). Velocity and distance are dimensionless. (a) Tangential velocity,
(b) radial velocity. Experimental data: • R e = 290, A R e = 760.

uS = ~ s ( r * - l ) (7)
An estimate of u* close to the bubble is obtained, knowing that for axisym-
metric flow around a sphere:

u* = - [ ( r * - 1)2 sin0] -16P'


,~o (8)
where ~] is the dimensionless stream function. Near the surface of the sphere
~] is given by ( Weber, 1981 ):
1 .
= ~ ( r - 1 ) 2 ~ s sin0 (9)

Since ~s is known, ~ can be calculated and 6 ~ / 6 0 can be derived at any value


of r and 0 close to the bubble surface.
The surface vorticity ~s has been estimated by correlating with 0 and Reb
using data from Woo (1971) and Masliyah (1970). The correlation is of the
form:
~ = a + b O + c O 2 +dO 3 (10)
where a, b, c and d are in turn third-order polynominal regression functions of
R% (see Appendix 1 ).
A less fundamental approach is used to estimate liquid velocities further
from the bubble. If u~, denotes potential flow velocity (either tangential or
radial) and u~ denotes Stokes flow velocity, then let the true velocity (away
from the bubble surface) URe be:
246

Fie b
,:, 20 ~J :) 05
70
• tO0
<~ 3 0 0
60
Potential Flow . //
Ec "~"'L / /
50
/
(%]
40 J
j" 0
30 /

20

10 Stokes Flow. ~ ~

I I I I ] I I
05 1 2 .4 .6 .8 t 0

Stokes N u m b e r , Sk

Fig. 3. Trajectory calculations, Ec versus Sk and Reb.u* = 0.05

URe= Xup + (1 - X ) u~
where 0 < X < 1. The value of X is adjusted to give a calculated collision effi-
ciency at low Sk ( < 0.1 ) equal to that predicted by Weber's model.
Thus, to calculate collision efficiency at intermediate values of Sk, eqs. 2 and
3 were solved with a Runge-Kutta numerical routine employing the equations
described in this section. Ec is then given by (x*)2 where x* is the starting
value, i.e. far from the bubble, of the trajectory calculation.
An example of calculated collision efficiencies as a function of Stokes num-
ber (Sk < 1 ) is given in Fig. 3. Note the collision efficiencies fall approximately
midway between the Stokes flow and potential flow collision efficiencies for
Sk > 0.4.
Collision efficiency was correlated with Sk by multiple regression using 40
sets of conditions representing realistic combinations of Sk, Reb and u*. The
result is:

Ec = Eco (1.627 Re°.°6Sk°.S4u *-°.1~ ) (11)

where Eco is the collision efficiency obtained as Sk--*O, i.e. Weber's result
( Sk = 0.05 was used). The range of application is:

( Reb°.°6Sk°.S4u*-°.16 ) > 0.614

20 < Reb < 300; Sk < 0.8; u* < 0.25.


To give some perspective, the system consisting of an 80/~m particle with
247

pp=5 g cm -3 and a 0.10 cm bubble with Reb=100 gives S k = 0 . 3 6 and


Ec/Eco = 1.74.

A t t a c h m e n t model

The underlying assumption is that after collision the particle slides over the
bubble and attachment occurs when the intervening liquid (disjoining) film
thins and ruptures. (Particle bounce need not be considered for particles less
than 100/xm (Dobby, 1984 ). ) The time required for the film to rupture is the
induction time, ti. Consequently, particles with a sliding time greater than t i
are considered to have attached.
The calculation of the fraction of particles which attach, i.e. EA, requires a
knowledge of: (a) the distribution on the bubble surface of particle collision
angles, (b) the angle at which fluid streamlines start to carry the particle ra-
dially away from the bubble, i.e. the maximum angle of contact Or,, and (c)
the particle sliding velocity.
(a) Distribution of collision angles. The distribution is quantified by no, the
fraction of all colliding particles that collide between the front stagnation point
and angle 0n. This is calculated using the trajectory model. Some results are
shown in Fig. 4. As an approximation,
sin 2On
no -- sin20c (12)

is employed, which is also shown in Fig. 4.


(b) M a x i m u m angle of contact, Or,. To complete the calculation of sliding time,
0m must be known. This is done by determining the angle at which the radial
component of the particle settling velocity upr (directed toward the bubble
surface) is equal to the radial component of the liquid velocity ur (directed
away from the bubble surface) calculated at ( r* - 1 ) = dJdb, using eqs. 8 and
9. At 0 > 0~ the particle no longer contacts the bubble, unless attachment has
already occurred.
Although not used in the present model, a correlation between 0m, pp and 0c
is given by:
0m = 9 + 8.1p, + 0c (0.9-- 0.09p,) (13)
(C) Particle sliding velocity. It is evident from Fig. 2a that there is a significant
tangential velocity gradient on the upper surface of a bubble. Fig. 5 shows the
Reb = 290 data on Fig. 2a along with the calculated surface vorticity (eq. 10).
The velocity data is approximated by two functions:
for ( r * - l ) <0.06
uS = 0.7~s (r* - 1 ) (14a)
and for ( r* - 1 ) > 0.06
248

i:l
('6

' 2 j

08
i
,/
~

//

°2tS '~/J/// '


OE From
trajectory
no O~ calculations
/ c
04 sin20n
02 sin2Oc
V

08 f/~/ '
06 / a
04

' /@/2 '


()8
0E / / / e

20 40 60
On °
Fig. 4. Distribution of collision angle, pp=7.5 g cm ~, Reb=100. (a) dp=20/zm, Sk=0.033,
u* = 0.013; (b) dp= 30/tin, Sk = 0.075, u* = 0.026; (c) dp= 40/tin, Sk = O.133, u* = 0.042; (d) dp= 50
#m, Sk = 0.208, u* = 0.064; (e) dp= 60 #m, Sk = 0.300, u* = 0.084.

u $ = 0 . 7 ~s(0.06) (14b)

T h e factor 0.7 fits the velocity gradient at R e b = 2 9 0 and 0 = 4 5 ° and the as-
s u m p t i o n is made t h a t this factor also applies at other Reb and 0. (In the
trajectory simulations, rigorous calculation of u% close to the bubble was n o t
so important, c o n s e q u e n t l y the 0.7 factor was n o t included in eq. 7.) For
dp/db < 0.03 ( i.e. ( r* - 1 ) < 0.03 ), eq. 14a alone is adequate. Therefore particle
tangential velocity is:

dp
Upo = 0.7~s Ub ~ + Up sinO (15a)

For dp/db > 0.03 (i.e. ( r* - 1 ) > 0.03 ), the particle velocity is calculated by
dividing the particle into two zones, one t h a t sees a velocity gradient described
by eq. 14a and the other t h a t sees a c o n s t a n t velocity described be eq. 14b.
T h e n Up6 is given by:
249

1.0 / ( s : 15.7

0.8

u; 0.6

0.1.

0.2

o.os .'1 .2' .'3 .1,I


Dislance from Sphere Surface, r~-I
Fig. 5. Tangential velocity gradient at surface of sphere at Re = 290 and 0 = 45 °. From eq. AI-1,
~s= 15.7 at these conditions.

upe =0.7~Ub
(d. - dp
0.03rib)o "06 + o.o3] + sin0 (15b)

Particle sliding time


The particle sliding time t~ can now be calculated, and is given by:

ts =
(o j ) n re (dp +db)/(t~po) (16)

where 0 is in degrees and ~vo is the average particle tangential velocity deter-
mined from eq. 15a or 15b using average values, ~s and s~n0.
Fig. 6 shows a test of the sliding time calculations. Schulze and Gottschalk
(1981) measured the time that 160/~m galena particles remained in contact
with a 0.3 cm bubble after collision at various angles (On). Eq. 16 provides a
reasonable fit; the potential flow assumption does not (the apparent fit of
potential flow in the original paper of Schulze and Gottschalk resulted from
an error in Sutherland's model; this is discussed elsewhere (Dobby and Finch,
1985) .)

A t t a c h m e n t efficiency E A
A particle attaches to a bubble when it resides on the bubble surface for a
time ts equal to or greater t h a n the induction time ti. Let 0;, be the angle 0n ( in
eq. 16) when ts = t i. After rearrangement this gives:
250

6O

"~ Equation (15) V O l l • •

d *

~
Z(.9 40 i i ~ k ~ : •ill••

floten "

0 10 20 30 40 50 60 70 80 90
ANGLE QF CONTACf.0n(degrees)
Fig. 6. Particle sliding times. Measurement by Schulze and Gottschalk (1981); predictions from
eq. 16 and potential flow assumption for 0m = 90 °.

0~=0m 360@eti (17)


n (du + d p )
Consequently, a t t a c h m e n t efficiency is given by:
sin2 0~
E . - sin2 0c (18)

T h e a t t a c h m e n t model is now complete. Fig. 7 shows a plot of EA vs ti for a


series of particle sizes at d b = 0.1 cm ( R% = 100). At a c o n s t a n t i n d u c t i o n time
a t t a c h m e n t efficiency increases with decreasing particle size. For example, if

(%) 6C
lOum

40 20 um

10 20 30 40 50 60
Induction Time (ms)
Fig. 7. Attachment efficiency versus induction time and particle size. Conditions: db=0,1 era,
ub= 10 cm s- 1,#=0.01 poise, pp = 4.0 g cm -3, Cg=0.
251

EK
OOl,
(%)

003
E
01 o

g
u 0.02

#
0.0

I I 1 I l I I
' 2'o ' 4b ' 6b ' 5 10 20 aO 60 80

ParticLe Size (.urn) Particle Size (urn)

Fig. 8. Collection efficiency versus particle size a n d induction time, from flotation model. Condi-
tions: d b = 0.1 cm, Ub= 10 cm S- ~, # = 0.01 poise, pp = 4.0 g c m - 3, ~g = 0.

Fig.9. Rate c o n s t a n t versus particle size for galena flotation in a 2.5 cm diameter, 2 m long labo-
ratory flotation column (Dobby, 1984). Gas bubbles generated using a porous steel sparger, vg = 0.33
cm s -1, db=0.10 cm.

ti = 20 ms then EA ( 80 #m) = 1.5% but EA ( < 10#m) --~100%. This strong par-
ticle size effect is not unexpected, given the relationship between particle size
and sliding velocity.

MODEL RESULTS

A computer program was written to solve Ec, EA and E K for the range
O.1 < Reb < 400. In the following sections the model is used to examine the effect
of flotation variables.

Particle size

Fig. 8 illustrates the relationship between EK and dp for given ti. (For con-
ditions see figure caption.) The figure is constructed assuming particles of
different size have similar induction time. The first feature of note is that EK
passes through a maximum. The maximum is explained by the opposing effect
of particle size upon collision and attachment; as dp increases Ec increases but
EA decreases. The peak is in accord with recovery-size data reported for many
mineral systems ( Trahar, 1981 ), including galena flotation data illustrated in
252

Fig. 9 (Dobby, 1984). Galena size-by-size rate constant data shown in Fig. 9
was obtained in a laboratory flotation column, where physical entrainment was
virtually eliminated and no mechanical agitation was employed.
The peak in size-by-size data has been explained in the past by bub-
ble-particle detachment (Woodburn et al., 1971 ) or by an increase in ti with
d, (Jowett, 1980). The present model shows that invoking such explanations
is not necessary. (A forthcoming article will report experimentaal results on
measurement of the dynamic tj-dp relationship. )
A second observation is that the peak shifts to smaller dp as t~ increases, i.e.
as particle hydrophobicity decreases. This is in agreement with observation;
for example Anthony et al. (1975) observed the recovery peak for sphalerite
as increasing, from about 15 pm to 80 ~m, with increasing copper sulphate
additions (implying reducing induction time i.
Finally, Fig. 8 shows that collection efficiency for very small particles
(dp_<5-10/~m) is quite insensitive to induction time. This is because small
particles have a low velocity over the bubble and consequently a high ts; thus
EA is insensitive to moderate changes in ti. This observation has important
implications for selectivity. Consider the separation of two minerals, one
strongly floating with ti--15 ms, the second weakly floating with t~= 40 ms.
The relative collection efficiency with dp is shown in Table I. The separation
ranges from excellent at the coarse and intermediate sizes (dp> 20 zm) to
virtually non-existant for dr, < 5 pm. The decrease in selectivity with decrease
in particle size is well known. In mechanical cells the problem is compounded
by entrainment but this model suggests the origin may be fundamental.

Bubble size

The results in Fig. 8 were for db=0.10 cm. The effect of bubble diameter is
explored in Fig. 10 It is assumed ti is independent of bubble diameter ( Klassen
and Mokrousov {1963) and Leja and Poling (1960) have reported results that
imply a decrease in ti with decreasing db.) Decreasing bubble size clearly in-
creases collection efficiency. This results from increases in both Ec ( see eqs. 5
and 6 ) and EA. The increase in EA is because the fractional decrease in particle
sliding velocity on a smaller bubble ( ub decreases with decreasing db) exceeds
the fractional decrease in sliding distance. Clear experimental evidence of this
bubble size effect has been provided by Anfruns and Kitchener (1977).
The increase in E K with decreasing db is attractive. However, smaller bub-
bles do not improve selectivity. Table I shows that db= 0.05 cm gives similar
selectivity to dh ----0.10 cm.

Particle density
Fig. 11 illustrates the pronounced effect particle density has on the EK vs dp
relationship. With increasing particle density, EK increases, the curve becomes
253

3.0

0.07 cm

1.0

E K (%)
0.10 cm

0.5

~ cm

I I I I I I I
20 40 60

Particle Size (~m)


Fig. 10. Bubble diameter effect on EK, from flotation model, ti=20 ms, #=0.01 poise, pp=4.0 g
cm -3, and ~Og=0. ub=12.7 cm s -1 at db=0.13 cm, 10.0 cm s -1 at db=0.10 cm, 7.0 cm s -1 at
db=0.07 cm, 4.8 cm s -1 at db=0.05 cm.

progressively more peaked and the maximum floatable size of particle de-
creases (especially for #p > 5 g cm -3 ).
The increase in E K is related to the increase in Stokes number with a con-
sequent increase in Ec (Fig. 3 ). The sharper peak and reduced maximum size

TABLE1

Relative collection efficiency of weakly to strongly floating mineral at two bubble sizes

dp [EK (ti=40 ms) 1


(pm) EK(ti=15 ms)
d b
0.05 cm 0.10 cm

30 0.13 0.09
20 0.28 0.19
10 0.74 0.69
5 0.97 0.97
254

10
\
\
EK
(%) 06

0~,

02

Particle Size(urn)
Fig. 11. Particle density effect on EK, from flotation model, ti = 20 ms, db = 0.1 cm, uh = 10 cm s- ~,
/~=0.01 poise, ¢~=0.

for flotation occurs because the increase in pp increases the particle settling
velocity up (eqs. 15a and b). As a result the sliding velocity increases, t8 de-
creases and EA decreases. Changes in t8 have most impact on EA at coarse sizes
(Fig. 7).
The effect of particle density on the shape of E K VS dp appears to be con-
firmed in practice. Trahar (1981) reports a sharp peak at around 20-40 ttm for
cassiterite flotation (pp= 7.0 g cm-3). At the other end of the density scale,
coal flotation (pp = 1.3 g cm-3) is practised up to 800/tm (Aplan, 1980).

Liquid viscosity

Viscosity,/z, affects both bubble rise velocity and particle setting velocity,
but more important it likely affects induction time. It is assumed here that
t~ocZ, following the work of Jowett (1980). Fig. 12 was then constructed for
initial conditions of ti--~20 m s at ]~=0.01 poise. (Also dbaC ~o.25 (Jameson,
1983 ) was assumed but this assumption made little impact on EK compared to
assuming ti3cl/). It is evident that a decrease in viscosity increases collection
efficiency. This observation may be relevant to milling operations where a
summer to winter change in water temperature occurs.

Collection in a bubble swarm

To this point the model is a single bubble model. It is intuitive that there
would be interaction between bubbles in a bubble swarm. LeClair and Ham-
255

EK(%)0"5 ~ [ - ~

0.1

0.05

I
20 4'0 610
I I I

ParticleSize(~rn)
Fig. 12. Viscosity effect on EK, from flotation model, pp=4.0 g cm -3, Cg=0; at #=0.007 poise,
ti= 14 ms, db=0.091 cm, Ub= 11.9 cm s - l ; at ~=0.010 poise, ti=20 ms, db=0.1 cm, Ub= 11.8 cm
s - 1; at/~ = 0.015 poise, ti = 30 ms, db= 0.111 cm, Ub= 11.5 cm s - 1.

ielec (1968) have calculated surface vorticity as a function of Reynolds number


and volume fraction of spheres (or in this case gas holdup, Og). The data of
LeClair (1970) were fitted by a polynomial regression of the form:
~so = ~ + n O g (19)
where ~so is the surface vorticity of a gas bubble at gas holdup 0g and n is a
function of Reb and 0 (see Appendix 2 ).
In this manner the effect of bubble concentration on surface vorticity and,
therefore, on attachment efficiency is accounted for. The assumptions are made
that collision efficiency of a bubble is not affected by a bubble swarm, except
by the lowering of the bubble rise velocity (calculated using the Richardson
and Zaki (1954) relationship) and that bubble diameter is constant. Under
these conditions, Fig. 13 was constructed.
An increase in gas holdup increases collection efficiency. The principal rea-
son is the decrease in Ub. One way to increase Cg is to increase gas rate, best
quantified by superficial gas velocity Vg. However, this will also cause db to
increase. Flotation rate constant is related to EK by:

k~ vgEK (20)
db
256

/. . . . " , 10%
0.5

EK(%) ', \ \\
', 5%
\ ',

0.1

0.05

' 2'0 ' 4'0 60

Particle Size (#m)


Fig. 13. Gas holdup effect o n EK, from flotation model, db = 0.1 cm, p = 0.01 poise, pp = 4.0 g c m ~;
at 0 = 0 . 0 5 , u~= 10.1 c m s '; at 0~=0.10, Ub=8.6 c m s 1

Consequently the flotation rate constant may pass through a maximum with
Vg as the effect of increasing the number of bubbles (through Vg) is progres-
sively offset by an increasing db and (consequently) decreasing EK. A maxi-
m u m in flotation rate constant with gas rate has been reported (Laplante et
al., 1983; Dobby and Finch, 1986).

CONCLUSIONS

A fundamental model of flotation has been developed. Particle inertia in


particle-bubble collision is accounted for, thus extending the range of the ex-
isting collision model. Recently derived solutions to the flow equations around
spheres are employed to develop a model of particle attachment. When the
collection is tested two broad observations are made which agree with much
experimental data.
(1) A peak in collection efficiency with particle size occurs when induction
time is constant with particle size. This is the result of collision efficiency
increasing with dp while attachment efficiency decreases. The shape of the
curve and the location of the peak depends on induction time, bubble diameter
and particle density.
( 2 ) Collection efficiencies become similar for particles less than 5-10/lm;
257

this is because of the long contact time of the smaller particles. Thus even at
relatively long induction times (40-100 ms ) EA is very high for small particles.
This means separability decreases with decreasing particle size. In the absence
of entrainment, smaller bubbles do not improve separability.
Specific applications of the collection model, including the effect of solids
concentration, will be addressed in a forthcoming paper.

ACKNOWLEDGEMENTS

The authors are grateful to the Centre de Recherche Noranda for providing
a scholarship for three years. Funding for flotation research was provided by
Energy, Mines and Resources Canada, the Natural Sciences and Engineering
Research Council (strategic grants program) and the Minist~re d'l~ducation
de Quebec (FCAC program). This funding is gratefully acknowledged.

APPENDIX 1 -- SURFACE VORTICITY ~

T h e surface vorticity data of Woo (1971) h a s been correlated to bubble Reynolds n u m b e r Reb
(given as Re in equations) a n d angle 0 m e a s u r e d from t h e front s t a g n a t i o n point. Woo reported
~8 for t h e following Reynolds n u m b e r s : 0.2,0.5,0.75,1,2,3,5,10,20,30,40,100,200,300 a n d 400; a n d
at every 12 ° for Re < 40, a n d every 3 ° for Re > 100. T h e following correlations are for 0 < 0 < 84 °:

~ =a+bO+cO 2+d~ 3 (AI-1)

where, for 20 < R e < 400:

a=-0.01082-7.273.10-4Re+ 1.735.10 -6 Re 2 - 2 . 0 4 6 . 1 0 -9 Re 3 (A1-2)

b = - 0 . 0 7 4 5 + 3 . 0 1 3 . 1 0 -3 R e - 7 . 4 0 2 . 1 0 -6 Ree + 8 . 9 3 1 . 1 0 -9 Re 3 (A1-3)

c = - 4 . 2 7 6 " 1 0 -4 - 1 . 9 7 7 " 1 0 - ° R e + 5 . 1 9 4 . 1 0 - s Re e - 6 . 5 2 0 . 1 0 - " Re 3 (AI-4)

d - - - 1.103.10 -6 _ 1.032.10 - s R e + 1.397' 10 - ,o R e 2 _ 1.334.10 - 13 Re 3 (A1-5)

and, for 0.2 < Re < 20:

a=-1.217"10-3-1.745.10 -3 R e + 5 . 1 4 3 . 1 0 -~ R e 2 - 1 . 1 6 5 . 1 0 -6 Re 3 (AI-6)

b = 0 . 0 2 8 5 9 + 9 . 2 2 9 ' 1 0 -3 R e - 3 . 8 5 ' 1 0 -4 Re e +9.190- 10 -6 Re 3 (A1-7)

c = - 4 . 0 6 0 ' 1 0 -5 - 5 . 8 5 7 . 1 0 -5 R e + 1.620.10 -6 Re e - 2 . 9 9 2 . 1 0 - s Re 3 (A1-8)

d = - 9 . 6 1 0 ' 1 0 -7 - 2 . 5 4 ' 1 0 -7 R e + 1.74'10 - s R e 2 - 5 . 1 0 . 1 0 - ' 0 R e 3 (AI-9)


In each of t h e regressions, eqs. A1-2-A1-9, correlation coefficient > 0.99.

A P P E N D I X 2 - - S U R F A C E V O R T I C I T Y A N D GAS H O L D U P

T h e effect of gas holdup u p o n surface vorticity h a s been obtained from t h e data of LeClair
258

(1970). LeClair reported ~ at R e = 0.1,1,10,100 and 500. The following correlation was obtained
using LeClair's data at R e = 0.1,1,10,100 and 500;0 = 12,24,36,48,60,72 and 84 ~; and 0~ = 0, 0.091,
0.165 and 0.259.
Let:
~ = ~ +n¢~e (A2-1)
where:
n=a' +b'O+c'O 2 + d ' O :~ (A2-2)
~o is surface vorticity at gas holdup O~, and where:
a'= -0.0199+3.30' -:~ Re-4.780'10-5 R e 2 +7.939"10-s Re,~ (A2-3)

b' =0.6579-7.91"10 -:~ Re+ 1.269"10 -4 R e 2 -2.107'10 -7 R e :~ (A2-4)

c' = -6.165'10 -4 +3.499'10 -5 R e - 7.024'10 -7 R e 2 +1.160'10 -9 R e :~ (A2-5)

d'=-2.339"10 -~ +5.305'10 -s R e - 1.177'10 -~ R e 2 +2.156'10 -12 R e :t (A2-6)


For each of eqs. A2-3-A2-6 correlation coefficient > 0.99.

REFERENCES

Anfruns, J.F. and Kitchener, J.A., 1977. IMM., pp. C9-15.


Anthony, R.M., Kelsall, D.F. and Trahar, W.J., 1975. Proc. Aust. Inst. Min. Metall., 254: 47-58.
Aplan, F.F., 1976. In: M.C. Fuerstenau (Editor), Flotation - A.M. Gaudin Memorial Volume.
AIME, New York, N.Y., Ch. 45.
Clift, R., Grace, J.R. and Weber, M.E., 1978. Bubbles, Drops and Particles. Academic Press, N.Y.,
Chap. 5 and 7.
Concha, F. and Almendra, E.R., 1979. Int. J. Miner. Process., 5: 349-367.
Dobby, G.S., 1984. Ph.D. Thesis, McGill University, Montreal, Que.
Dobby, G.S. and Finch, J.A., 1986. C.I.M. Bulletin, 79 (889): 89-96.
Dobby, G.S. and Finch, J.A., 1986. J. Colloid Interface Sci., 109 (2) : 493-498.
Flint, L.R. and Howarth, W.J., 1971. Chem. Eng. Sci., 26: 1155-1168.
Jameson, G.J., 1983. In: K.J. Ives (Editor), The Scientific Basis of Flotation. NATO.
Jameson, G.J., Nam, S. and Moo Young, M., 1977. Min. Sci. Eng., 9(3): 103-118.
Jowett, A., 1980. In: P. Somasundaran (Editor), Fine Particles Processing. Proc. Int. Symp., Las
Vegas, Nev. AIME, New York, N.Y., Chap. 37.
Klassen, V.I. and Mokrousov, V.A., 1963. An Introduction to the Theory of Flotation. Butter-
worths, London, Chap. 5.
Laplante, A.R., Toguri, J.M. and Smith, H.W., 1983. Int. J. Miner. Process., 11: 203-219.
LeClair, B.P., 1970. Ph.D. Thesis, McMaster University, Hamilton, Ont.
LeClair, B.P. and Hamielec, A.E., 1968. Ind. Eng. Chem. Fundam., 7 (4): 542-549.
Leja, J. and Poling, G.W., 1960.5th Int. Miner. Process. Congr. London, IMM, pp. 325-341.
Masliyah, J.H., 1970. Ph.D. Thesis, Univ. of British Columbia, B.C.
Reay, D. and Ratcliff, G.A., 1973. Can. J. Chem. Eng., 51 (4): 178-185.
Richardson, J.F. and Zaki, W.N., 1954. Trans. Inst. Chem. Eng., 32: 35-52.
Schulze, H.J. and Gottschalk, G., 1981. J. Laskowski (Editor), Proc. 13th Int. Miner. Process.
Congr., Warsaw. Elsevier, N.Y., pp. 63-84.
259

Seeley, L.E., Hummel, R.L. and Smith, J.W., 1975. J. Fluid Mech., 68(3 ): 591-608.
Sutherland, K.L., 1948. J. Phys. Colloid Sci., 52" 394-425.
Trahar, W.J., 1981. Int. J. Miner. Process., 8: 289-327.
Weber, M.E., 1981. J. Sep. Process. Tech., 2 (1) : 29-33.
Weber, M.E. and Paddock, D., 1983. J. Colloid Interface Sci., 94 (2): 328-335.
Woo, S.W., 1971. Ph.D. Thesis, McMaster University, Hamilton, Ont.
Woodburn, E.T., King, R.P. and Colborn, R.P., 1971. Metall. Trans., (2): 3163-3174.

NOTATION

db bubble diameter
dp particle diameter
E~ attachment efficiency
Ec collision efficiency
Ecg collision efficiency by gravitation
Eci collision efficiency by interception
Eco Ec calculated assuming S k = 0 when S k > 0.05
Eco Ec at gas holdup 0g
E~ collection efficiency = Ec'EA
k flotation rate constant
n constant in eq. 19
no fraction of colliding particles that collide between 0 = 0 and 0 = 0~
r* dimensionless radial co-ordinate from bubble centre
r* -- 1 dimensionless radial co-ordinate from bubble surface
Re Reynolds number
Reb Reynolds number of bubble~ Ubpldb/p
Sk 1 pp dp
Stokes number =~ ~ ( ~ ) R %
t time
ti induction time
t~ sliding time (time of particle-bubble contact)
U velocity, any phase
Ub bubble rise velocity relative to liquid
Up particle settling velocity relative to liquid
Upr radial particle velocity = ur + Up cos0
UpO tangential particle velocity = uo + Up sin0
Ur radial component of liquid velocity around a sphere
UxUy components of liquid velocity in x and y directions
Uo tangential component of liquid velocity around a sphere
U* = uplub
u~ U* or uS at potential flow
u~ u* or uS at Stokes flow
u~e u* or uS for bubble with R%
u~ .-~ Ur/ Ub
u~ ~- U,x/Ub
U*y Uy/U b
uS Uo/l.~b
Vx particle velocity in x direction
260

U/ particle velocity in y direction


v~
v~ ~- L~y/t~ b
superficial gas velocity
0 angular co-ordinate measured from front stagnation point of bubble
angle of closest approach for fluid streamlines
~m maximum angle of contact between sliding particle and bubble, i.e. the angle at which
the particle begins to move radially away from the bubble surface ( if attachment has
not yet occurred)
~n collision angle
O. that yields t~= t,
jl liquid viscosity
surface vorticity, dimensionless
~ at gas holdup ¢~
Pl density of liquid
Pp density of particle
gas holdup, fractional
VJ stream function, dimensionless

You might also like