Vous êtes sur la page 1sur 8

Liquid Transport in a Coalescing Froth

Peter M. Ireland* and Graeme J. Jameson


Centre for Multiphase Processes, University of Newcastle, NSW 2308 Australia

A previous study (Ireland and Jameson, J. Colloid Interface Sci., 314, 207-213 (2007)) demonstrated that a “drift-flux” model could describe
liquid transport in a stable rising froth with added “wash water.” In the present study, a drift-flux model was used to describe a rising coalescing
froth. This model incorporated the effect of liquid released into the froth by coalescence. Vertical profiles of liquid fraction and bubble size were
obtained in a laboratory cell; a novel technique was used for measuring bubbles deep within the cell. These data were consistent with the
predictions of the drift-flux model.

Une étude antérieure (Ireland et Jameson, J. Colloid Interface Sci., 314, 207-213 (2007)) démontre qu’un modèle à dérive de flux pourrait
décrire le transport liquide dans une mousse en ascension stable avec ajout d’ «eau de lavage». Dans la présente étude, on utilise un modèle
à dérive de flux pour décrire une mousse coalescente en ascension. Ce modèle incorpore l’effet du relâchement de liquide dans la mousse
par coalescence. Les profils verticaux de fraction liquide et de taille de bulles ont été obtenus dans une cellule de laboratoire; une nouvelle
technique a été utilisée pour mesurer les bulles en profondeur dans la cellule. Ces données sont consistantes avec les prédictions du modèle
à dérive de flux.

Keywords: foam, flotation, drift-flux, fluidization, coalescence

INTRODUCTION “froth”(rather than “foam”) is used for the upper layer in the

I
n a froth column, gas is released at a constant rate into a flotation cell, whether or not particles are present.
surfactant solution in a cylindrical vessel; the resulting froth At steady state, assuming a very stable froth with no film
rises until it overflows. Two layers are observed—the froth breakage or bubble coalescence, the volumetric liquid fraction of
and the underlying bubbly liquid, as in a flotation cell, where air the froth reaches a value uniform with height in the column. The
bubbles are introduced to a suspension of particles in water. superficial liquid velocity, or volumetric flow rate of liquid per
Hydrophobic particles attach to the bubbles and are carried unit cross-sectional area of the column, is also constant with
upwards to the surface of the liquid where they form a froth height. The transport of liquid in a froth column without coales-
layer that then flows over the lip of the vessel to form the cence has previously been analyzed successfully using a drift-
flotation product. Hydrophilic particles remain in the liquid and flux model; key advances in this regard were made by Pal and
flow out of the bottom of the cell as tailings. Flotation is often Masliyah (1989a, b). In this paper, we demonstrate that this type
used to separate finely ground ore containing small quantities of of model can also be used to describe the transport of liquid and
a valuable mineral from other fine particles consisting primarily the liquid content of a coalescing froth, which better reflects the
of a gangue such as quartz. The liquid that is entrained into the froth in a flotation cell than an idealized stable froth. A new
froth from the pulp contains some of this hydrophilic gangue, representation of the drift-flux relation is introduced, allowing
and this can contaminate the flotation product. To reduce the the changes in liquid fraction and bubble size with depth in a
contamination, clean wash water is commonly applied to the coalescing froth layer at steady state to be expressed by a single
froth layer, to flush the entrained particles out of the froth,
resulting in a higher-grade product. In order to optimize the
grade of the product, it is clearly important to understand how * Author to whom correspondence may be addressed.
liquid is transported in the froth layer. In this paper, the term E-mail address: peter.ireland@newcastle.edu.au

654 The Canadian Journal of Chemical Engineering Volume 85, october 2007
curve. A number of such “coalescence curves” for different drainage term as in Equation (3) and representative parameter
liquid superficial velocities are then verified by experimental values (including a constant bubble radius). As can be seen, a
measurements in a laboratory cell. It should be noted that drift- whole family of these “drift-flux curves” exists for different
flux theory is only capable of relating the liquid fraction and values of the superficial gas velocity. It should be noted that a
bubble size to each other at each point in a coalescing froth. In decrease in the bubble radius will generally result in a decrease
order to model the variation of each quantity individually with in Vt; thus, a decrease in bubble radius can have a similar effect
depth in the froth, one would need to understand the actual to an increase in Jg.
coalescence process. Without coalescence or addition of extra liquid to the froth, the
The conditions for co-existence of the liquid and froth layers liquid superficial velocity at steady state must be the same in all
in a froth column without wash water were studied by Langberg regions of the column. Langberg and Jameson (1992) argued that
and Jameson (1992). These authors adapted an analytic structure for moderate values of the gas velocity (e.g. curve (a), Figure 1)
developed by Jones et al. (1980) and Jones and Leung (1985) to the liquid fractions of the liquid and froth layers will be solutions
describe multi-phase flows in standpipes. The flow is assumed to of Equation (2) corresponding to the given value of the liquid
be one-dimensional, and in any particular froth or liquid layer, velocity (Figure 2). Ireland and Jameson (2007a) and Stevenson
the bubble size is assumed constant. The basis of all of this work (2006b) argued that the froth phase without added liquid must
is in the so-called ‘drift-flux’ theory of Wallis (1969) for liquid sit at the peak of the curve, (eLmax, JLmax), and this is indeed
transport in dispersed two-phase systems, as elaborated by observed experimentally (Ireland and Jameson, 2007a). The
Rietema (1982). The superficial gas velocity, Jg, the superficial other intercept, with the ascending arm of the curve, must
liquid velocity, JL and the liquid volume fraction, eL, are related therefore correspond to the underlying bubbly liquid layer. It is
by a so-called “drift-flux equation”: important to note that only the descending arm of the drift-flux
Jg JL curves (between the maximum and minimum) has been verified
− = U s (eL ) (1) experimentally. At large values of the gas velocity or small bubble
1 − eL eL
sizes, as in curve (e), Figure 1, the drift-flux equation predicts
The “slip velocity,” US, is the difference between the linear that the liquid velocity will increase monotonically with liquid
velocities of the gas and liquid phases. Langberg and Jameson fraction. Under these circumstances, the liquid-froth interface
(1992) devised a very convenient and conceptually lucid form of will vanish, as there will only be one intercept for a given flow
the drift-flux equation with JL as the dependent variable: rate. The entire system is filled with a single bubbly liquid layer,
a state known as flooding. In practice, flooding is often accompa-
 e   e 
J L = J g  L  − e LU s ( e L ) ≡ J g  L  − J Ld (2) nied by instability, turbulence and even explosive behaviour, as
 1 − eL   1 − eL  in a champagne bottle that is shaken and then opened. The onset
of this behaviour is not well-understood at present.
The first term represents the superficial velocity of liquid rising
Ireland and Jameson (2007a) extended the above analysis to
in the column in the absence of drainage. The drainage term J Ld
include the addition of liquid at a point inside the froth. It was
can be given by one of an assortment of drainage models; J dL
asserted that the difference between the superficial liquid
depends on the bubble size, the viscosity and other properties of
velocity above and below the injection point must be equal to
the interstitial liquid, and the interactions of the liquid with the
the superficial injected or ‘wash’ liquid velocity JLW. The froth
surfactant films separating the liquid and gas bubbles. A variety
above the injection point is assumed to be unchanged from its
of drainage terms have been proposed, either rigorously derived
(e.g. Leonard and Lemlich, 1965; Langberg and Jameson, 1992,
adapted from Ergun, 1952; Neethling et al., 2002) or deduced
from dimensional analysis (Stevenson, 2006a). Langberg and
Jameson (1992) originally used a drainage term originating from
the Richardson-Zaki equation for flows through fluidized media
(Richardson, 1971):

J Ld = Vt enL (3)

Here, Vt is the terminal rise velocity of a bubble in an infinite


liquid and n is an exponent which may depend upon the partic-
ular surfactant solution used. Although the Richardson-Zaki
equation is empirical rather than physically rigorous, it has a
number of advantages that promote its use in the present
context. It is clear that for liquid fractions approaching unity (i.e.,
a liquid with a few bubbles) the relative velocity between the
phases will be the terminal bubble rise velocity, and thus the
relative superficial velocity between the phases (i.e., J Ld) will
approach Vt. The drainage model derived from the Richardson-
Zaki equation automatically satisfies this requirement. It has
only one adjustable parameter, rather than the two required by Figure 1. Curves of dimensionless superficial liquid velocity vs.
several of the alternatives mentioned above, and is thus a degree liquid fraction obtained using a drift flux equation of the form
less arbitrary when fitted to experimental data. It is also highly JL/Jg = eL/ (1– eL)–(Vt /Jg)enL with a drainage term derived from the
mathematically tractable. The liquid superficial velocity Richardson-Zaki equation, where in this case n = 3 (an arbitrarily chosen
calculated from Equation (2) is plotted in Figure 1, with a value). The gas superficial velocity Jg has values (a) Vt/9; (b) Vt/7; (c) Vt/5;
(d)Vt/3; (e) Vt.

Volume 85, october 2007 The Canadian Journal of Chemical Engineering 655
state without wash liquid. Thus, the superficial liquid velocity yet have a combined two-phase flow/coalescence model for
below the injection point will adjust to accommodate a superfi- rising froth in a column, and thus cannot predict the precise
cial liquid velocity of JL = JLmax – JLW. The liquid fraction will variation of eL and rb with column height. However, we can
likewise adjust in a manner that can be predicted using the drift- verify that eL and rb vary as predicted by Equation (4).
flux equation, as illustrated in Figure 2; as JLW increases from In the experiments discussed in the next section, the superfi-
zero, the state of the froth beneath the injection point moves cial gas velocity was held constant, while the froth depth, and
down the drift flux curve from the peak, becoming wetter in thus the liquid overflow rate and superficial velocity, were
the process. altered in a controlled manner; eL and rb were then measured at
A coalescing froth is more complex than a non-coalescing various heights in the froth. It is therefore convenient to
froth in several respects. In a rising froth, the bubble size introduce another graphical representation of the drift-flux
increases with height in the froth, altering the drainage rate as relation. Consider the drift-flux equation with a Richardson-
described by Equation (3). In addition, coalescence releases Zaki-type drainage term:
interstitial liquid, which then flows back through the underlying
 e 
froth, altering the liquid superficial velocity (and thus liquid J L = J g  L  − Vt ( rb ) enL (5)
fraction) in that region. In a froth column, continuity dictates  1 − eL 
that at steady state, both the gas and liquid superficial velocity The terminal bubble velocity as a function of the bubble size,
must remain constant with height. Thus, we can write the drift- Vt(rb), is taken from the experimental bubble rise data of
flux equation in the form: Motarjemi and Jameson (1978). Rearranging this, we have
 e     e 
J g  L  − J d ( e L , rb ) = J L = constant (4)  
 1 − e L
rb = Vt−1   J g  L  − J L  e L −n  (6)
   1 − e L   
The system is illustrated by the drift-flux curves in Figure 3.
where Vt–1 is the inverse of the function Vt, i.e., Vt−1 Vt ( rb ) = rb.
As the froth at a given height coalesces, releasing some intersti-
We can also find a comparable form that describes the states
tial liquid into the froth below, the liquid fraction of the underly-
corresponding to the maximum and minimum in the drift-flux
ing layers must adjust itself such that Equation (4) is satisfied
curves; at these points,
there as well. Both the change in bubble size with height and the
flow-back of interstitial liquid due to coalescence can thus be dJ L 1
= Jg − Vt ( rb ) nenL−1 = 0 (7)
accommodated by the drift flux equation via adjustments in the de L (1 − e L )2
liquid fraction. This implies a specific relationship between eL
Rearranging Equation (7), we obtain a relationship between
and rb for the prevailing superficial gas and liquid velocities. In
bubble size and liquid fraction for these maxima and minima:
this model, only the froth at the very top of the column is not
subject to liquid back-flow from above. This means that eL and  1 e1−n 
JL will sit at the peak in the relevant drift-flux curve only at the rb = Vt−1  J g L  (8)
 n (1 − e L )2 
top of the froth; this uppermost state will then mediate the flow  
of liquid to the rest of the froth. Since liquid fraction is a critical Figure 4 shows the bubble radius plotted against the liquid
determinant of coalescence behaviour, the adjustment to the fraction for a number of values of the dimensionless liquid
liquid fraction caused by coalescence-induced back-flow will velocity and representative parameter values, using Equation
likely influence the coalescence rate, which will in turn affect (6). Also plotted on the same axes is the curve corresponding to
the liquid fraction in underlying layers, and so on. We do not as Equation (8), i.e., the maxima and minima of the drift-flux

Figure 3. A typical family of drift flux curves for different bubble sizes
and constant gas velocity, showing the effect of coalescence. The liquid
Figure 2. Typical curve of superficial liquid velocity vs. liquid fraction, superficial velocity remains constant; as one rises through the froth
showing the effect injecting wash water into the froth layer, the bubbles become larger and the froth drier.

656 The Canadian Journal of Chemical Engineering Volume 85, october 2007
curves (which we dub the “extremal curve”). If JL were displayed Experimental
as the vertical coordinate of a surface plot, with el and rb as the
The primary experimental apparatus is shown schematically in
horizontal coordinates, the curves in Figure 4 would be contours
Figure 5. It consisted of a modified Jameson cell with a cylindri-
of equal height on that surface. The drift-flux curves in Figure 3
cal profile and a diameter of 300 mm. Air and water were fed
would be slices through the surface parallel to the el axis.
together under pressure into an 80 mm diameter stainless steel
Variation of the superficial gas velocity would produce a whole
downcomer pipe, producing a bubbly mixture with bubbles of
family of these surfaces. Each “contour” of constant JL corresponds
initial diameter ~0.4 mm. The total cell cross-section area,
to the set of solutions (eL, rb) that together represent a coalescing
allowing for the space occupied by the downcomer, was 661 cm2.
froth in a column. We therefore refer to these contours as
The cell was capped by a launder to collect the overflow. The gas
“coalescence curves.” For a non-coalescing froth with a constant
and liquid flows were measured using rotameters. The surfactant
bubble size, this set of solutions reduces to the points of intersec-
solution was 30 ppm MIBC (methyl isobutyl carbinol; 4-methyl-
tion of the coalescence curve with the horizontal line correspond-
2-pentanol) by volume in tap-water at 22-25ºC. A level control-
ing to the constant bubble size. The top of the froth, whether
ler consisting of an inverted U-shaped tube of adjustable length
coalescing or not, will correspond to that point of intersection of
was used to determine the position of the interface between the
the coalescence curve with the extremal curve with the lowest
liquid and froth layers, and thus to control the liquid overflow
liquid fraction. As one moves deeper into the froth, the froth
rate from the cell. In all cases, the total gas flow rate was 80 L/
state (eL, rb) moves down the coalescence curve to the right; the
min (corresponding to 2.017 cm/s), and the total liquid flow rate
bubbles becomes smaller and the froth wetter, until the initial
was 40 L/min. This allowed a maximum liquid superficial
(minimum) bubble size rbmin is reached. The underlying liquid
velocity (for the case where there was no flow from the level
layer then corresponds to the right-hand intercept of rb = rbmin
controller) of 1 cm/s.
with the coalescence curve. We designate the minimum in the
Profiles of the liquid fraction in the column were obtained
coalescence curve the “drift-flux limit”. Its significance, particu-
using a simple technique attributable to Cutting et al. (1981). A
larly if the minimum bubble size is less than the drift-flux limit,
1 mm internal diameter copper tube was inserted into the froth
is discussed later in this paper.
to the desired depth of tip. A small flow of air was passed into
Coalescence curves at larger values of JL do not intersect the
the froth from the tube using an aquarium pump. As individual
extremal curve at all; this corresponds to a flooded column.
bubbles formed and were released, the maximum pressure
Along these non-intersecting curves, the liquid fraction increases
monotonically with the bubble size, a trend that is the opposite
of that familiar in stable coalescing froths. Indeed, it implies that
in a flooded column with coalescing bubbles (where the gradient
in bubble size with height must always be positive), the liquid
fraction will increase with height. While not necessarily impossi-
ble, this would be a highly unstable situation, and buoyancy-
driven downward flows of the denser overlying material would
very likely occur. It is possible that the turbulent behaviour
observed in flooded froth columns is related to this kind of
instability, although there is no direct evidence for such a link.

Figure 4. A typical family of coalescence curves for different superficial


liquid velocities and constant gas velocity. The extremal curve (dashed) is
also shown; the descending arm of the extremal curve represents the top Figure 5. Schematic diagram of the modified Jameson cell used in
of the froth, whereas the ascending arm represents the drift-flux limit. this study

Volume 85, october 2007 The Canadian Journal of Chemical Engineering 657
(assumed equal to the ambient pressure at the tip opening) was flow rate of each phase was kept approximately equal to the
measured using a micromanometer (Furness Controls FCO14), estimated total in-flow of each (froth plus mixing liquid). Any
and repeated at 5 mm intervals up the column. The liquid net flow of material into or out of the sampling device, other
fraction was given by: than that in the ambient rising froth, could thus be minimized.
1 dP Inside the viewing chamber, the bubbly liquid flowed through a
eL = − (9) narrow (1 mm wide) perspex-walled space, which brought all of
rliquid g dy
the bubbles into approximately the same plane. The measured
where rliquid is the density of the liquid phase and P is the bubble distributions contained bubbles whose diameter was
pressure. larger than 1 mm, and these were omitted from the sample. The
Bubble diameters were measured at a specific point in the cell statistical consequences of omitting these larger bubbles are
using a novel technique devised specifically for this study. This discussed later in this paper. The bubbles were then photographed
technique allowed further coalescence to be suppressed while through the wall of the viewing chamber along with an appropri-
the bubbles were brought out of the cell for measurement. The ate scale object, and measured using the OPTIMAS software
sampling device could simply be dipped into the cell, eliminat- (MediaCybernetics, Silver Spring, MD 20910, U.S.A.). 350–500
ing the necessity for drilling access holes in the side. Figure 6 bubbles were measured for each data point.
shows the sampling device schematically. The froth (or bubbly The bubble sampling technique described above was effective
liquid) rose through the entry tube into the mixing chamber, in all but the top 50 mm of the cell below the overflow lip, where
where it was combined with liquid at approximately the same too large a proportion of the bubbles were larger than the viewer
total volumetric flow rate. The total flow rate of added liquid was gap width of 1 mm. However, in this region, it was often possible
less than 1% of the total liquid flow rate in the cell. The resulting to clamp a small rubber block to the outside of the cell just
mixture was a bubbly liquid in which the bubbles were separated below the overflow lip. This ensured that a narrow strip of the
by enough liquid to halt coalescence. This was then sucked up a outside wall of the column, inside the launder, was free from
tube through a viewing chamber into a separation chamber, overflowing froth. The bubble distribution could then be
where drainage and coalescence occurred rapidly. It was photographed through the cell wall. The width of the clear strip
intended that the froth would rise more or less undisturbed into of wall was much smaller than the radius of curvature of the
the mixing chamber at the local ambient rate. The gas and liquid column, and could be considered locally flat. For analysis
were drawn from the chamber through separate rotameters; the purposes, the photographed region was divided into 10 mm
zones, with median depths of 15, 25, 35, and 45 mm. 200-300
bubbles were generally available for measurement in each zone.
This technique had the advantage that it offered the entire
bubble size distribution for measurement, with no upper limit.
However, due to the construction of the launder, this was
impractical elsewhere in the cell.

RESULTS AND DISCUSSION


Pressure profiles within the cell are shown for various values of
the liquid superficial velocity in Figure 7. To obtain the liquid

Figure 7. Plots of measured pressure in the cell vs. depth below the
overflow lip, for three different superficial liquid velocities. The lines
Figure 6. Schematic diagram of the bubble sampling device used in represent the composite polynomial fit curves used to determine the
this study liquid fraction.

658 The Canadian Journal of Chemical Engineering Volume 85, october 2007
fraction as a function of depth in the cell, each pressure profile through the cell wall, on the assumption that this was representa-
was approximated by a smooth curve constructed from sections tive of the normalized cumulative volume of bubbles in the bulk
of second-order polynomials, as shown. The least-squares R2 froth. Again, a least-squares fit with the upper-limit distribution
value of this combined approximation was greater than 0.99 in gave R2 > 0.99 in all cases. For comparison, we also performed
every case. This allowed the liquid fraction to be calculated as a an upper-limit fit to the cumulative volume distribution of the
continuous function of depth in the cell via Equation (9). As visible bubbles, as if it were a volume-sampled distribution
mentioned earlier, bubbles larger than the width of the interior of obtained using the sampler tube. Profiles of Sauter mean bubble
the bubble viewer (1 mm) were sometimes observed. Although size vs. depth in the froth are shown in Figure 8. The bubble sizes
these bubbles were excluded from the sample of measured obtained using the upper-limit distribution fit to the visible area
bubble diameters, they often represented a substantial proportion of the bubbles (rather than their volume) photographed through
of the total bubble volume, and could not simply be ignored. To
avoid underestimating the (Sauter) mean bubble diameter, the
entire size distribution needed to be inferred from the measured
bubble diameters. Grau and Heiskanen (2005) had previously
found that the size distribution in a number of types of mechan-
ical flotation cell was well-described by a so-called ‘upper limit
distribution’, also referred to as a three-parameter log-normal
distribution, given by Mugele and Evans (1951). The distribution
of bubble volume, according to this model, can be expressed as:
2
δ dbmax      
dV  ad
= exp −δ2 In  max b   (10)
ddb (
π db dbmax − db  )  d
  b − db 
  
where a and δ are dimensionless constants and dbmax is the
eponymous upper limit of the distribution. Integration of
Equation (10) (Goering and Smith, 1978) gives the cumulative
volume distribution (note that this is dimensionless):

∫−∞ exp ( −u ) du = 0.5{sign ( z ) erf ( δ z ) + 1}


1 δz
V= (11)
π
where erf(x) is the error function, commonly available as an
intrinsic function in spreadsheet software, etc. The argument z
is defined as:
 ad 
z = 1n  max b  (12)
d − db 
 b
The Sauter mean bubble diameter for this distribution is given
by:
dbmax
db =
(
1 + a exp 1 / 4δ2 ) (13)

(Mugele and Evans, 1951). Since the measured distribution of


bubble sizes did not go up to dbmax, the distribution in Equation
(10) had also to be multiplied by a normalization constant to
achieve a fit to the experimental distribution. The upper limit
distribution provided a least squares fit with R2 > 0.99 in all
cases, and thus was a good description of the measurable part of
the bubble size distributions.
A similar fit was performed to the data from the bubble size
distributions obtained by photographing the bubbles through the
cell wall (these were assumed to be ‘complete’ distributions, in
the sense that there was no upper limit to the measurable bubble
size). It was important to recognize that the number distribution
of bubbles of different sizes in a given volume sample of froth is
not the same as the number distribution of bubbles visible
through a wall. Indeed, for a very large number of bubbles, we
Figure 8. Plots of Sauter mean bubble radius vs. log depth below the
expect the distribution of volume of a randomly-packed bubble
overflow lip. Note the difference in the degree of continuity between
ensemble within a three-dimensional region to approach the the mean bubble radii obtained using the sample tube and those
distribution of cross-sectional bubble area in any plane through obtained by through-the-wall photography, for (a) and (b). As
that region, when normalized by the total bubble volume in the expected, when the mean for through-the-wall photography is
region or area on the plane, respectively. With this in mind, we determined by an upper-limit distribution fit to the cumulative area
attempted a fit with the upper-limit distribution to the normalized of the visible bubbles (a), the degree of continuity is higher than
cumulative cross-sectional area of the bubbles photographed when the fit is to the cumulative volume of the visible bubbles (b).

Volume 85, october 2007 The Canadian Journal of Chemical Engineering 659
the cell wall (Figure 8a) showed a high degree of continuity with maintain a steady state, one would expect a wet layer of froth
the bubble size data obtained using the sample tube. (Note that to start forming above a dry layer—a highly unstable
the Sauter mean bubble diameter is not dependent on the normal- situation, likely to promote buoyancy-driven flows. Indeed,
ization constant but merely on the un-normalized shape of the the drift-flux limit may be an important determinant of the
distribution.) On the other hand, the bubble sizes obtained by onset of buoyancy-driven flows in froths. It is not possible at
fitting to the cumulative volume of the visible bubbles (Figure 8b) this stage to determine what relationship might exist between
provided a lower degree of continuity with the other data. Thus, the liquid fraction and bubble size as a result of non-uniform
we consider our assumption, that the normalized bulk volume drainage mechanisms such as buoyant flows.
distribution is well-represented by the normalized area distribu- • Our prediction that the top of the froth would correspond to
tion of the visible bubbles, to be validated. the peak in the coalescence curve was not verified directly.
The overall validity of our model was assessed by rearranging However, the trend in the measured data approaching the top
Equation (5) into the form: of the froth seems consistent with this prediction.
• These data are also inconclusive as regards the co-existence
  eL  
 − J L  / Vt ( rb ) = e L of the bubbly liquid and coalescing froth layers. We predicted
n
Jg  (14)
  1 − e L  that if the minimum bubble size were larger than the drift
flux limit, the froth state would move down the descending
Figure 9 shows the left-hand side of (5) plotted against eL for the
arm of the coalescence curve until it reached the intersection
collected data. These are in good agreement with a power law
with rb = rbmin. The bubbly liquid layer would then
distribution with n = 2.339. Figure 10 shows the measured
correspond to the intersection of the ascending arm of the
values of the Sauter mean bubble radius plotted against the
curve with rb = rbmin. No “jump” from the descending to the
measured liquid fraction, along with coalescence curves derived
ascending arm of the curve is clearly discernable in the data.
using the above value of n, and the corresponding extremal
However, as explained above, the minimum bubble size may
curve. Again, there is good agreement between each of the
well have been close to or less than the drift-flux limit for
coalescence curves and the corresponding experimental data.
some or all of the superficial liquid velocities. Since we are
Some aspects of the behaviour of this coalescing froth remain
unable to determine the shape of the coalescence curves in
unresolved. The outstanding issues are as follows:
the presence of non-uniform flows, speculation about the
• The significance of the “drift flux limit” is not yet established.
behaviour at the bubbly liquid/froth interface would be
Inspecting Figure 10, it is difficult to discern whether the
somewhat premature.
minimum bubble size in any of the experiments was smaller
• In an actual flotation process, the presence of hydrophobic
than the minimum in the coalescence curves. There is no
particles on the bubble surface would likely stabilize the froth
particular reason for the minimum bubble size to be larger
and reduce the speed of coalescence (Pugh, 2005). This in
than this limit. We can therefore only speculate as to what
itself would not affect the form of the coalescence curves, as
will happen should the bubble size be smaller. It may be the
they are independent of the actual coalescence mechanism.
case that the drift-flux limit represents the smallest bubble
However, slower coalescence would tend to produce a greater
size for which the given liquid and gas superficial velocities
liquid superficial velocity in the cell for the same gas
can co-exist in a uniform flow state. If this is the case, other
superficial velocity and initial bubble size. One would
non-uniform flow states may take over and “short-circuit” the
therefore expect the system to move to a different coalescence
uniform drift-flux-type flow. Buoyancy-driven flows, which
curve. The presence of particles on the liquid-gas interface
are a good example of non-uniform flow mechanisms, are
frequently observed in froths when wash liquid is added,
producing a large slip velocity between the two phases
(Ireland and Jameson, 2007b; Vera et al., 2000). If liquid were
unable to drain quickly enough between the bubbles to

Figure 10. Plots of measured bubble radii vs. measured liquid fraction,
for different values of superficial liquid velocity. The corresponding
Figure 9. Experimental data from the present study, plotted in the form coalescence curves, generated using the value of n from Figure 9,
Equation (14). The closest power law fit is shown. are shown.

660 The Canadian Journal of Chemical Engineering Volume 85, october 2007
may also alter the slip conditions between the liquid and gas REFERENCES
phases, which would be reflected in a different value of the
Cutting, G. W., D. Watson, A. Whitehead and S. P. Barber, “Froth
index n. This would alter the form of the coalescence curves,
Structure in Continuous Flotation Cells: Relation to the
although it is not clear how.
Prediction of Plant Performance from Laboratory Data using
Process Models,” Int. J. Miner. Process. 7, 347–369 (1981).
CONCLUSIONS Ergun, S., “Fluid Flow through Packed Columns,” Chem. Eng.
Process. 48, 89 (1952).
In this paper, we set out to demonstrate that a drift-flux model
Goering, C. and D. Smith, “Equations for Droplet Size
describing the liquid flow through a fluidized bed of bubbles
Distributions in Sprays,” Transactions of the ASAE 21, 209
could be used to describe the transport and degree of retention of
(1978).
liquid in a coalescing froth. A new graphical representation of the
Grau, R. A. and K. Heiskanen, “Bubble Size Distribution in
drift-flux relationship was presented, in which the concurrent
Laboratory Scale Flotation Cells,” Minerals Eng. 18, 1164
changes in bubble size and liquid fraction with depth in a coalesc-
(2005).
ing froth layer at steady state could be expressed by means of a
Ireland, P. M. and G. J. Jameson, “Liquid Transport in a Multi-
single curve. A set of experimental measurements of bubble size
Layer Froth,” J. Colloid Interface Sci. 314, 207-213 (2007a).
and liquid fraction variations in a coalescing froth was presented,
Ireland, P. M. and G. J. Jameson, “Behaviour of Wash Water
and was found to be consistent with the corresponding coales-
Introduced into a Froth,” Int. J. Miner. Process. in press,
cence curves generated using the drift-flux relations.
corrected proof (2007b).
While drift-flux theory allows the bubble size and liquid
Jones, P. J. and L. S. Leung, “Downflow of Solids through Pipes
fraction to be related to each other in each part of a coalescing
and Valves,” in “Fluidisation,” J. F. Davidson, R. Clift and
froth, it does not explain how each of these quantities individu-
D. Harrison, Eds., 2nd ed., Academic Press, London (1985).
ally varies with depth in the froth. In order to do that, one would
Jones, P. J., C. S. Teo and L. S. Leung, “The Stability of Vertical
require some knowledge of the actual coalescence process itself.
Gas-Solid Downflow in Bottom-Restrained Standpipes,” in
In addition, the relationship between the bubble size and liquid
“Fluidization,” J. R. Grace and J. M. Matsen, Eds., Plenum,
fraction beyond the drift-flux limit, and possibly in the presence
New York (1980), pp. 469–476.
of buoyancy-driven flows, is not well-understood. Work contin-
Langberg, D. E. and G. J. Jameson, “The Coexistence of the
ues to resolve these outstanding issues.
Froth and Liquid Phases in a Flotation Column,” Chem.
Eng. Sci. 47, 4345 (1992).
ACKNOWLEDGEMENT Leonard, R. and R. Lemlich, “A Study of Interstitial Liquid
We are grateful to the Australian Research Council for its support Flow in Foam,” AIChE J. 11, 18–25 (1965).
of the Centre for Multiphase Processes, under the Special Motarjemi, M. and G. J. Jameson, “Mass Transfer from Very
Research Centre scheme. Small Bubbles—The Optimum Bubble Size for Aeration,”
Chem. Eng. Sci. 33, 1415–1423 (1978).
Mugele, R. and H. Evans, “Droplet Size Distribution in Sprays,”
NOMENCLATURE Ind. Eng. Chem. 43, 1317–1324 (1951).
a parameter in upper-limit distribution (dimensionless) Neethling, S. J., H. T. Lee and J. J. Cilliers, “A Foam Drainage
db bubble diameter (mm) Equation Generalized for all Liquid Contents,” J. Phys.:
dbmax maximum bubble diameter (mm) Condensed Matter 14, 331 (2002).
g gravitational acceleration (m s-2) Pal, R. and J. Masliyah, “Flow Characterization of a Flotation
Jg gas superficial velocity (cm s-1) Column,” Can. J. Chem. Eng. 67, 916 (1989).
JL liquid superficial velocity (cm s-1) Pal, R. and J. Masliyah, “Flow in Froth Zone of a Flotation
J max Column,” Can. Metallurgical Quarterly 29, 97 (1989).
L liquid superficial velocity at peak in drift-flux curve
(cm s-1) Pugh, R. J., “Experimental Techniques for Studying the
JLd drainage superficial velocity (cm s-1) Structure of Foams and Froths,” Adv. Colloid Interface Sci.
JLW wash liquid superficial velocity (cm s-1) 114–115, 239–251 (2005).
n index in Richardson-Zaki equation (dimensionless) Richardson, J. F., “Incipient Fluidization and Particulate
P pressure (Pa) Systems,” in “Fluidization,” J. F. Davidson and D. Harrison,
rb bubble radius (mm) Eds., Academic Press, London (1971).
rmin Rietema, K., “Science and Technology of Dispersed Two-Phase
b minimum bubble radius (mm)
u dummy integration variable (dimensionless) Systems—I and II,” Chem. Eng. Sci. 37, 1125–l 150 (1982).
US slip velocity (cm s-1) Stevenson, P., “Dimensional Analysis of Foam Drainage,”
V cumulative fractional bubble volume (dimensionless) Chem. Eng. Sci. 61, 4503–4510 (2006a).
Vt terminal bubble rise velocity (cm s-1) Stevenson, P., “The Wetness of a Rising Foam,” Ind. Eng.
y vertical coordinate (m) Chem. Res. 45, 803 (2006b).
z argument in upper-limit distribution (dimensionless) Vera, M. U., A. Saint-Jalmes and D. J. Durian, “Instabilities in
a Liquid-Fluidized Bed of Gas Bubbles,” Phys. Rev. Lett. 84,
Greek Symbols 3001 (2000).
eL liquid volume fraction (dimensionless) Wallis, G., “One-Dimensional Two-Phase Flow,” McGraw-Hill,
eLmax liquid fraction at peak in drift-flux curve New York, Ch. 4 (1969).
(dimensionless­)
rliquid density of liquid phase (kg m-3)
Manuscript received March 2, 2007; revised manuscript received
δ parameter in upper-limit distribution (dimensionless) May 31, 2007; accepted for publication June 11, 2007.

Volume 85, october 2007 The Canadian Journal of Chemical Engineering 661

Vous aimerez peut-être aussi