Вы находитесь на странице: 1из 11

Int. J. Miner. Process.

78 (2005) 11 – 21
www.elsevier.com/locate/ijminpro

Nanobubbles, hydrophobic effect, heterocoagulation


and hydrodynamics in flotation
Heinrich Schubert
TU Bergakademie Freiberg, D-09596 Freiberg, Germany
Received 14 March 2005; received in revised form 15 July 2005; accepted 16 July 2005
Available online 2 September 2005

Abstract

It has been shown in recent years that the long-range attractive interaction forces between hydrophobic surfaces in aqueous
systems are caused by the capillary forces of gas bridges which form at the coalescence of nanobubbles adhering on the
surfaces. The coalescence of nanobubbles on selectively hydrophobized particles with coarser bubbles initiates the jump into the
three-phase contact at the attachment events in flotation. Therefore, one should no longer speak of hydrophobic forces in these
events but of hydrophobic effects because they are caused by capillary forces. However, in the selective hydrophobization of
particles by long-chain collectors (surfactants), there is another situation. Here, the association of nonpolar groups in the
adsorption layers plays an important role in the energy balance of adsorption. This association is caused by btrulyQ hydrophobic
interactions, which support the spotty distribution of the adsorbed collector ions on the particle surfaces and promote the
formation of nanobubbles. This paper is intended to show to what extent the results obtained by basic research on the
nanobubble formation as well as the force–distance dependence of collision events can be applied to flotation processes. This
particularly requires the consideration of the highly turbulent flow conditions in the impeller stream of the flotation machines, in
which the attachments almost exclusively occur. Various phenomena which occur in these machines and affect the reagent
regime and the hydrodynamics point to the fact that nanobubbles can form, exist and even grow into microbubbles in that
region. Therefore, so-called combined attachment events should predominate as already imagined several decades ago. The
highly turbulent pressure fluctuations in the impeller stream in addition to the dispersion of the bubbles effect also their
oscillations in size and shape, so that adsorption equilibria in the interfaces liquid/gas cannot be supposed. However, the
pressure fluctuations present the possibility to overcome potential barriers in the wetting films at attachment events.
D 2005 Elsevier B.V. All rights reserved.

Keywords: flotation; nanobubbles; hydrophobic effect; collector association; hydrodynamics; particle–bubble attachment

1. Introduction

The interpretation of the physical nature of the


interactions between hydrophobic particles in aqueous
E-mail address: schuberh@mvtat.tu-freiberg.de. systems has been increasingly taken up recently. The
0301-7516/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.minpro.2005.07.002
12 H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21

difficulties of their explanation resulted from their While for the first force measurements crossed
relatively long-range character (Ninham, 1999). The cylinders were used as contacting solids (Israelachvili,
recent results of several research teams have provided 1991; Israelachvili and Pashley, 1982; Pashley and
an explanation of the phenomena such that it is appro- Israelachvili, 1981), for the AFM studies mainly the
priate to speak of a hydrophobic effect instead of a solid plate/solid sphere (Ishida et al., 2000, 2002;
hydrophobic force. These facts also have conse- Nguyen et al., 2003; Sakamoto et al., 2002; Tyrrell
quences for the interpretation of the heterocoagulation and Attard, 2001, 2002) and solid sphere/bubble (Pre-
between hydrophobic particles and bubbles inclusive uss and Butt, 1998, 1999) systems were studied. As
of the hydrodynamic conditions, in that they occur in materials silicates were predominantly used, initially
flotation processes, too. mica, later on, above all, SiO2, glass, Si-wafer with
oxidized surface (Christenson and Claesson, 2001;
Ishida et al., 2000, 2002; Nguyen et al., 2003; Preuss
2. Hydrophobic effect and Butt, 1998, 1999; Sakamoto et al., 2002; Tyrrell
and Attard, 2001, 2002; Yang et al., 2003), but also
When two particles in an aqueous suspension plastics (Christenson and Claesson, 2001; Nguyen et
approached closely, the resulting interaction force al., 2003) and occasionally other materials. The sili-
always includes the van-der-Waals force and an elec- cate materials were made hydrophobic either by che-
trostatic component. The theoretical basis for this is mical preparation (methylation or silanation) (Attard,
given by the DLVO-theory (Derjagin–Landau–Ver- 2003; Christenson and Claesson, 2001; Ishida et al.,
wey–Overbeek theory). If hydrophilic particles come 2000, 2002; Nguyen et al., 2003; Snoswell et al.,
very close to each other, these interactions are aug- 2003; Tyrrell and Attard, 2001, 2002; Yang et al.,
mented by a steric hindrance which is caused by the 2003) or by adsorption of cationic surfactants, parti-
structurization of their hydration layers. In the case of cularly quaternary alkylammonium salts (Preuss and
hydrophobic and hydrophobized particles relatively Butt, 1998, 1999; Sakamoto et al., 2002). The majo-
strong and long-range attraction forces can be noticed. rity of the investigations were carried out with
The first direct force measurements between two relatively strong hydrophobicities (macroscopic
hydrophobic surfaces were reported in the early eigh- advancing contact angles N808). Furthermore, besides
ties of last century (Israelachvili, 1991; Israelachvili the chemical heterogeneity of the solid surfaces their
and Pashley, 1982; Pashley and Israelachvili, 1981). physical heterogeneity (above all roughness) was
But only the development and application of the found to be an essential parameter and, therefore,
atomic force microscopy (AFM) combined with varied (Nguyen et al., 2003; Snoswell et al., 2003;
respective periphery made it recently possible to mea- Yang et al., 2003).
sure the force–distance dependence between interact- By means of the tapping mode of the atomic force
ing surfaces as well as to analyze the surface states. In microscopy the existence of nanobubbles on hydro-
this way the physical character of the long-range phobic surfaces can be indirectly determined by scan-
attractive interactions could be elucidated on the one ning the position of the liquid/gas interface, and as a
hand, and the knowledge about the course of bubble result a visual representation of the surface coverage
attachments on hydrophobic solid surfaces could be with nanobubbles is generated (Hansma et al., 1994;
broadened on the other hand (Attard, 2003; Ishida Johnson, 1999). Further indications of their existence
et al., 2000, 2002; Nguyen et al., 2003; Preuss and can be derived from the force–distance dependence.
Butt, 1998, 1999; Tyrrell and Attard, 2001, 2002; The figures of the surface coverage known up to now
Sakamoto et al., 2002; Snoswell et al., 2003; Steitz can be subdivided into two groups, namely those
et al., 2003; Yang et al., 2003; Stöckelhuber et al., where single bubbles, which have approximately the
2004). The most important result of these investiga- shape of spherical segments, exist at more or less large
tions is the finding that in aqueous systems hydropho- distances (Sakamoto et al., 2002; Steitz et al., 2003;
bic solid surfaces are in general coated with Yang et al., 2003), and others where the nanobubbles
nanoscopic bubbles which can coalesce and form gas form a close-packed network (Attard, 2003; Tyrrell
bridges at the approach of surfaces. and Attard, 2001, 2002). These differences may
H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21 13

reflect the respective hydrophobicity state as well as


the heterogeneity of the surfaces and/or the composi-
tion of the aqueous phase.
The existence of nanobubbles on hydrophobic
solid surfaces in aqueous systems can be considered
omnipresent. Even under very careful experimental
conditions it is hardly possible to preclude them com-
pletely (Sakamoto et al., 2002).
On close approach of interacting partners coated
with nanobubbles, the coalescence of colliding nano-
bubbles occurs as shown in Fig. 1 schematically. Each
step in the force–distance curve indicates such a coa-
lescence event and demonstrates a sharp rise of the
attractive capillary force caused by forming or grow-
ing gas bridges. Fig. 2 represents a force–distance
dependence having several coalescence events from
one of the first measured examples (Attard, 2003). Fig. 2. Force–distance dependence between hydrophobic surfaces
If, as in flotation, one of the colliding partner is a coated with nanobubbles (Attard, 2003; Parker et al., 1994).
coarser air bubble, this can result in heterocoagulation,
i.e. the attachment of the hydrophobic particle onto trolled by the DLVO-regime, if necessary supple-
the bubble. Up to now only a few experimental results mented by steric hindrance, and a hydrophobic
are available about the respective force–distance attraction. After the three-phase contact a relatively
dependences (Ralston et al., 2001). Because of the strong capillary force is effective. In these measure-
deformability of a bubble, the correct distance mea- ments normally the first sharp rise caused the jump
surement represents a severe problem. For the force– into the three-phase contact. But in some cases, there
distance measurement between a spherical particle of have also been noticed smaller rises which did not
few micrometers in diameter and an air bubble of yet lead to the three-phase contact (Preuss and Butt,
few hundreds of micrometers in diameter, Preuss and 1998). Important for the attachment events are also
Butt (1998, 1999) developed an experimental set-up the findings by Preuss and Butt that potential barriers
which is similar to AFM and has measuring instru- in the water film can be overcome by external forces
ments adjusted to the specific demands. They have similarly to flocculation. Therefore, in flotation the
interpreted their results on the force–distance depen- flow forces can play an indispensable role for colli-
dence in the following way. Before the three-phase sion events resulting in attachment.
contact and apart from external forces, the approach Recently Stöckelhuber et al. have presented model
of the hydrophobic particle to the bubble is con- developments which can explain the destabilizing

Fig. 1. Formation and rupture of gas bridges between solid surfaces coated with nanobubbles.
14 H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21

effect of nanobubbles at the collision of hydrophobic bles. The latter angles are much larger and, therefore,
particles with air bubbles (Stöckelhuber, 2003; Stöck- the nanobubbles show flatter shapes of their approxi-
elhuber et al., 2004; Slachachov et al., 2005). When mately spherical segments (Yang et al., 2003).
nanobubbles exist in the contact region (Fig. 3), this ! Nanobubbles form only in aqueous systems. In
situation controls the stability and rupture, respectively, H2O/C2H5OH-systems the disposition to their for-
of the water film and because of that the jump into the mation decreases with increasing alcoholic content,
three-phase contact and consequently the attachment. and disappears completely at high alcohol concen-
The water film which has now to be considered is no trations (Christenson and Claesson, 2001; Nguyen
more the wetting film between the hydrophobic particle et al., 2003).
and the air bubble but the bfoam lamellaQ between the
air bubble and the nanobubble (Fig. 3). Its stability is A question which is not yet solved sufficiently
controlled by the DLVO-regime exclusively within the concerns the stability (better: metastability) of the
lamella, capillary forces (note the high capillary pres- nanobubbles. All observations made under the respec-
sure within the nanobubble!), and by strong flow forces tive experimental conditions give evidence of their
in flotation process. Therefore, in general the prerequi- existence over at least several tens of minutes without
sites for heterocoagulation should be met. discernible alterations (Attard, 2003; Yang et al.,
From the findings about the nanobubble formation 2003). This is particularly remarkable because the
published up to now, the following conclusions can be capillary pressure p k of nanobubbles should be high.
derived: For a spherical nanobubble of radius R = 50 nm in
water of surface tension of c lg = 0.072 N/m, for
! The formation of nanobubbles on solid surfaces instance, on the basis of the Laplace equation a capil-
requires a minimum hydrophobicity as well as lary pressure of p k = 2c lg / R c 29 bar results. Even if
roughness and increases with further increase of one considers that a nanobubble which shows a dia-
the hydrophobicity and roughness (Christenson meter of 2R = 100 nm in the AFM image, has a mean
and Claesson, 2001; Sakamoto et al., 2002; Yang curvature radius R m H 50 nm because of the large
et al., 2003). microscopic contact angle on the one hand, and that
! Physical and chemical heterogeneity further the the surface tension c lg can be b0.072 N/m because of
formation of nanobubbles. The heterogeneous adsorption in the liquid/gas interface and/or supersa-
nanobubble nucleation takes place in roughness turation of the liquid phase on the other hand, a con-
cavities. The roughness state is, therefore, a deci- siderable capillary pressure would still remain. The
sive parameter (Brennen, 1995; Nguyen et al., following possible explanations which can be only
2003; Yang et al., 2003; Ralston et al., 2001). briefly discussed here were given by (Attard, 2003):
! The macroscopic contact angle of a hydrophobic
surface differs from the microscopic contact angles, ! The nanobubbles are not in equilibrium with the
i.e. the angles determined by AFM at the nanobub- gas–supersaturated liquid phase.
! The difference between macroscopic and micro-
scopic contact angle is the consequence of the
fact that the line tension j in the three-phase con-
tact cannot be neglected for nanobubbles. There-
fore, the Young equation must be modified by an
additional term, and it follows for the microscopic
contact angle # (Yang et al., 2003):
j
cos # ¼ cos #Y  ð1Þ
clg r

#Y contact angle according to Young equation


Fig. 3. Water film between a hydrophobic solid surface coated with 1/r local curvature of the bubble base on solid
a nanobubble and an air bubble. surface.
H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21 15

! There are also doubts if the Laplace equation is


generally applicable to nanobubbles. This is ex-
plained by the effect of the solid surface on their
volume and the gas/solid interface.

The question about the stability of the forming gas


bridges with regard to their capillary pressure does
not arise because their two radii of curvature have
different signs.

3. Hydrophobization of solid surfaces by surfactant


adsorption
Fig. 4. Interactions at surfactant adsorption.
The selective hydrophobization of hydrophilic
solid surfaces plays a process-determining role in
flotation processes. The reagents used for this pur- or physisorption), those between the polar groups in
pose have a polar–nonpolar molecule structure and the adsorption layer as well as the dehydration
belong, apart from some reagents having short chains effects connected with adsorption. The nonpolar
(e.g. xanthates), to the ionic surfactants, which show group contributes by its association with neighbour-
in aqueous solutions the ability to form micelles and ing nonpolar groups within the adsorption layer to
other associations. the energy balance (Gaudin and Fuerstenau, 1955).
The polar surfactant groups promote dissolution in For the interaction of association per surfactant ion
water as well as cause direct bonding to solid surfaces (in the following termed association energy), it can
by physisorption (electrostatic adsorption) and in be written on the basis of the monolayer model of
some applications also by chemisorption. adsorption (Schubert and Baldauf, 1967; Schubert,
The nonpolar groups try to bescapeQ from sur- 1972, 1996):
rounding water molecules. If a nonpolar group pene- WA ¼ uðm  1ÞskT ð3Þ
trates an aqueous solution, hydrogen bonds must be
broken. But the interaction forces between nonpolar
groups and surrounding H2O molecules are much u change of the free enthalpy per –CH2– group
smaller than those of the hydrogen bridges between at complete association
water molecules. The consequence of this situation m equivalent or effective chain length
is described as hydrophobic bonding and results s association grade (0 b s b 1)
from the fact that the nonpolar groups associate in k* pre-association coefficient (1 z k* N 0).
order to reduce their contact area with water mole-
cules as far as possible (hydrophobic solvation) In solutions of low ionic strength, one can put
(Brezesinski and Mögel, 1993; Israelachvili, 1991). u =  1.39 kT, i.e. the amount which is equivalent to
At the surfactant adsorption on solid surfaces from a complete transfer of a –CH2– group from sur-
aqueous solution, the interaction W P which is caused rounding water molecules into surrounding saturated
by the polar group (or groups) can be differentiated nonpolar groups (Mukerjee, 1967). But this amount
from the interaction WA which the nonpolar group could only be realized if it would lead to the forma-
contributes (Fig. 4). The interaction of adsorption, tion of a complete surfactant layer having an ideal
W y , is given by: arrangement, and on the additional condition that
before adsorption no pre-associations (dimerization,
Wy ¼ WP þ WA ð2Þ
micellation, self-association) have taken place. For
W P includes the interactions of the polar group with straight-chain surfactants the actual chain length
the solid surface or its electric double layer (chemi- agrees with the effective chain length. Chain branch-
16 H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21

ing, double bonds, non-terminal polar groups and the tion of the surface coverage. It can be seen that su is
like impede the association. Under the condition that already considerable at very low surface coverages
the adsorption density C y and W P are constant, the (b 3%) and increases with the surface coverage up
association energy su can be determined from the to a limit. In solutions of low ionic strength this
adsorption isotherms of homologous series as fol- limit amounts to about  1 kT for straight-chain
lows (Schubert and Baldauf, 1967; Schubert, 1972, surfactants with one terminal polar group (curves
1996): 1, 2, 3 and 6 in Fig. 6). This means that on the
condition of u =  1.39 kT, the association grade s
Blnc su
¼ : ð4Þ in complete monolayers does not exceed about
Bm kT 0.7. By comparison of the curves 1 and 2 in
In Fig. 5 two examples of such series are Fig. 6, one can note that on adsorption of n-alkane
represented. Fig. 6 shows the respective associa- carboxylic ions on carborundum as result of the
tion energies su, calculated by Eq. (4), as a func- hydrolysis at pH 4.5 coadsorption of alkane car-

Fig. 5. Adsorption isotherms of homologous surfactant series: (a) n-alkyl ammonium chlorides on quartz (200–250 Am) at 25 8C and pH 6; (b)
n-alkane carboxylates on corundum (160–200 Am) at 60 8C and pH 4.5; (c k critical micelle concentration); (Schubert and Baldauf, 1967;
Schubert, 1972, 1996).
H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21 17

Fig. 6. Association energy su of surfactant layers as function of surface coverage (Schubert, 1996): (1) quartz (200–250 Am), n-alkyl
ammonium chlorides, 25 8C and pH 6; (2) corundum (160–200 Am), n-alkane carboxylates, 60 8C and pH 4.5; (3) quartz (200–250 Am), n-alkyl
ammonium chlorides, 10 g/l KCl, 25 8C; (4) quartz (200–250 Am), n-alkyl ammonium chlorides, 100 g/l KCl, 25 8C; (5) sylvite (200–250 Am),
n-alkyl ammonium chlorides, saturated KCl-solution, 25 8C; (6) cassiterite (40–63 Am), Na-n-alkane carboxylate, pH 6, 20 8C; (7) cassiterite
(40–63 Am), 1,1-n-alkane carboxylate, pH 6, 20 8C.

boxylic molecules has occurred. Such a coadsorp- bert and Bischofberger, 1998), the following spe-
tion improves the association in a surfactant layer. cial features must be emphasized:
Furthermore, a comparison of the curves 1, 3, 4 and One has to start from the fact that the selective
5 in Fig. 6 suggests the conclusion that with hydrophobization prevailingly occurs only in the flo-
increasing ionic strength, and beyond that at con- tation cells. Because of the energetical heterogeneity
stant conditions, both u and s grow because of the of the particle surfaces, the surfactant (collector) ions
salting-out effect. In the case of 1,1-alkane dicar- are adsorbed in a spotty distribution. This has been
boxylates the limit of su reaches only about 0.6 su. directly shown by microautoradiographic methods
This must be the effect of an association hindrance (Boudriot et al., 1977; Plaksin, 1960; Schlesier,
caused by the second polar group, but also points to 1965; Schubert, 1996). With long-chain surfactants
specific characteristics of the adsorption layer for- – above all straight-chain ones – the association of
mation (Schubert, 1996). the nonpolar groups furthers the spotty distribution
already at very low surface coverages and in that way
the formation of nanobubbles, too. The necessary
4. Heterocoagulation and hydrodynamics in the surface coverage for flotation, however, depends on
flotation process several parameters (solid surface properties, surfactant
type, character and intensity of the adsorption mecha-
Now the question arises, as to what extent the nism, size and structure of the nonpolar group, parti-
findings discussed above can be applied to indus- cle size and others). Based on available results, one
trial flotation processes with the hydrodynamic can assume that it is mainly between 10% and 50%
conditions in mind? For flotation machines having (Schubert, 1996). This means that the macroscopic
impeller–stator systems which supply the energy contact angles should be smaller than found in the
required for the realization of the microprocesses AFM studies. This has an unfavourable effect on the
into the aerated suspension (Schubert, 1999; Schu- formation of nanobubbles, whereas the essential
18 H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21

increase of the physical heterogeneity (irregular, On the condition that ē = 1.5 W/kg, as an example one
rough particle shapes!) as well as chemical heteroge- can assume that the energy dissipation rate e R in the
neity (inclusive of the surfactant association) should impeller stream at least amounts to 30ē c 45 W/kg.
further the nanobubble formation substantially. High stresses are the consequence in air dispersion,
According to the results of our own investigations, because for the bubble diameter d B in the inertial sub-
the microprocesses of air dispersion and particle–bub- range of microturbulence it is valid d B ~ 1 / e R0.4 as
ble attachment almost exclusively occur in the impeller well as high collision rates Z between bubbles and
streams of the flotation machines. These streams make particles for which one can correspondingly apply
up only a small percentage of the total volume of a Z ~ e R4/9(Schubert, 1996, 2003).
flotation cell (Schubert, 1999; Schubert and Bischof- It has been demonstrated by high-speed photo-
berger, 1998). This fact is in accordance with physical graphs in a specially designed flotation cell that the
considerations which suggest that the microprocesses kinetics of flotation processes are determined primar-
occur where the energy required for them dissipates. ily by the conditions in the impeller stream (see Figs.
Visually, one can get an idea of the position and 7 to 9 in Schubert and Bischofberger (1998)). Because
extension of the zone of high dissipation rate from of the highly turbulent flow conditions in the impeller
its small-bubbly (bmilkyQ) state as seen in Fig. 7, stream, it was impossible to identify attachment
where air bubbles were used as tracer (only a small events there. Notwithstanding the careful analysis of
air concentration was used in water without surfactant about 1400 m film, attachment events could also not
addition) (Schubert, 1999; Schubert and Bischofber- be detected outside of the impeller stream but only
ger, 1998). Furthermore, from stirring operations as already completed ones.
well as own studies it is known that the energy dis- The specific volume flow rate of air q L = V̇ L / V S
sipation rate e R in the impeller stream is a multiple of (V̇ Lvolume flow rate of air; V S suspension volume)
the mean dissipation rate ē = P / m ( P power input into amounts to about 0.5 to 1.5 m3/(m3 min) and the air
the mass m of the multi-phase system). Values of (e R / flow number c L = V̇ L / (nD 23) to about 0.07 to 0.3 (D 2
e )max c 5 to 200 are quoted for stirring equipment, and impeller diameter; n rotational speed of impeller)
¯ higher ones are valid for impeller–stator systems
the (Schubert, 1996). From these, the prerequisites for
(Schubert, 2003). In industrial flotation processes the supersaturation of the liquid phase with air can be
mean specific power input P / V S (V S aerated volume derived. Air dispersion proceeds in the regions at the
of the suspension) is mainly in the range 1 to 2 kW/m3. downstream side (wake) of the impeller elements

Fig. 7. Position of the impeller stream of a double-finger impeller (Schubert, 1999).


H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21 19

(blades, bars, fingers etc.) at peripheral speeds of the bles and coarser bubbles, i.e. so-called combined
impeller which are generally between 6 and 9 m/s. attachment events. The occurrence of such events
Under these conditions the liquid flow around an was already imagined by Taggart (1951) in the mid-
impeller element breaks away at the edges of the ele- 20th century and has been proved by Klassen (1960)
ment and a cavity forms, which is air-loaded, extended and Klassen and Mokrousov (1959) theoretically and,
in the flow direction and whose surface pulsates (Fig. later, experimentally. Eventually the pressure stresses
8). The latter is caused by the large eddies, which detach in the inertial sub-range according to Eq. (5) contri-
themselves on the flow around the element, periodically bute to overcome potential barriers in the water films
tear off parts from the cavity and transport them down- which have been found by Preuss and Butt in some of
stream, in the course of these events they are further their force–distance measurements.
dispersed. In the wake of such cavities one can observe
bubbles which are deformed as shreds, extended in the
flow direction and progressively further dispersed. 5. Conclusions
Furthermore, the turbulent pressure fluctuations in
the impeller stream have to be considered. Their All respective phenomena in flotation machines
RMS-value (root-mean-square value) in the inertial which concern the reagent regime and the hydrody-
sub-range can be written (Schubert, 2003): namics support the assumptions that nanobubbles can
qffiffiffiffiffiffiffiffiffiffiffiffi form on hydrophobic particle surfaces, exist there and
P P
DpV2 cð1 to 2Þq1 DuV2 c2q1 ðeR DrÞ2=3 ð5Þ even grow into microbubbles. Therefore, so-called
qP ffiffiffiffiffiffiffiffiffiffiffiffi combined attachment events should predominate in
DpV2 =RMS-value of turbulent velocity fluctua- the impeller stream.
tions which appear between two points at a dis- On account of the highly turbulent flow conditions
tance Dr in the microturbulence field (Dr = K); in the impeller stream and the consequent high fre-
K=macroscale of turbulence. quency velocity and pressure fluctuations, bubbles are
Although the frequencies of the pressure fluctua- subjected to continuous alterations of their size and
tions in the impeller streams are smaller than those in shape as well as oscillations. The consequence is that
the ultrasound range, it can be deduced from test adsorption equilibria in the interfaces liquid/gas can-
results obtained with the latter in air-saturated water not be supposed. This particularly affects the transfer-
that they further the heterogeneous nucleation as well ability of model calculations on the stability of
as the growth of already existing nanobubbles into wetting films during particle collisions. However,
microbubbles (Brennen, 1995). the pressure fluctuations in the impeller stream open
One can proceed from the fact that in the wakes the chance to overcome existing potential barriers.
downstream of the impeller elements both the air It would be welcome if future further research
dispersion and the events of particle–bubble attach- work on particle–bubble attachment would approach
ment occur and that they more or less run parallel. In the real situation in flotation processes more closely,
this case appears that the attachment events occur including, above all, consideration of the hydrody-
between particles coated with nano- and/or microbub- namic conditions.

Fig. 8. Air dispersion in the wake of an impeller element (Charčenko, 1986; Grainger-Allen, 1970; Schubert and Bischofberger, 1996).
20 H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21

Symbols Grainger-Allen, T.J.N., 1970. Bubble generation in froth flotation


c Concentration machines. Trans. Inst. Min. Metall. C 79, C15 – C22.
Hansma, P.K., et al., 1994. Tapping mode atomic force microscopy
cL Air flow number (c L = V̇L / (nD 23)) in liquids. Appl. Phys. Lett. 64, 1738 – 1740.
D2 Impeller diameter Ishida, N., Sakamoto, M., Miyahara, M., et al., 2000. Attraction
dB Bubble size between hydrophobic surfaces with and without gas phase.
k Boltzmann constant Langmuir 16, 5681 – 5687.
Ishida, N., Sakamoto, M., Miyahara, M., et al., 2002. Optical
m Mass
observation of gas bridging between hydrophobic surfaces in
n Rotational speed of impeller water. Colloid Interface Sci. 253, 112 – 116.
P Power input Israelachvili, J.N., 1991. Intermolecular and Surface Forces, 2nd ed.
p Pressure Academic Press.
pk Capillary pressure Israelachvili, J.N., Pashley, R.M., 1982. The hydrophobic interac-
qffiffiffiffiffiffiffiffiffiffiffiffi Root-mean-square value of turbulent pres- tion is long range, decaying exponentially with distance. Nature
P
300, 341 – 342.
DpV2 sure fluctuation
Johnson, Ch.A., 1999. Tapping Modek AFM imaging in fluids
qL Specific volume flow rate of air ( q L = V̇L / V S) for the study of colloidal particle adsorption. Digit. Instrum.
r Radius, generally Rev. 7/99.
R Radius of a principal curvature of an inter- Klassen, V.I., 1960. Theoretical basis of flotation by gas precipita-
tion. Proc. Int. Mineral Process. Congr., London 1960, Instn.
face
Min. Metall., London, pp. 253 – 268.
T Temperature Klassen, V.I., Mokrousov, V.A., 1959. Vvedenie v Teoriju Flotacija.
u Fluid velocity Gosgortechizdat, Moskva.
qffiffiffiffiffiffiffiffiffiffiffiffi Root-mean-square value of turbulent velo- Mukerjee, P., 1967. The nature of the association equilibria and
P
DuV2 city fluctuation hydrophobic bonding in aqueous solutions of association col-
V Volume, generally loids. Adv. Colloid Interface Sci. 1 (3), 241 – 275.
Nguyen, A.V., Nalaskowski, J., Miller, J.D., et al., 2003. Attraction
VS Suspension volume between hydrophobic surfaces studied by atomic force micro-
V̇L Volume flow rate of air scopy. Int. J. Miner. Process 72, 215 – 225.
W Energy Ninham, B.W., 1999. On progress in forces since the DLVO theory.
e Energy dissipation rate Adv. Colloid Interface Sci. 83, 1 – 17.
eR Energy dissipation rate in impeller stream Parker, J.L., Claesson, P.M., Attard, P., 1994. Bubbles, cavities and
the long-ranged attraction between hydrophobic surfaces.
# Contact angle J. Phys. Chem. 98 (34), 8468 – 8480.
c lg Surface tension of liquid Pashley, R.M., Israelachvili, J.N., 1981. A comparison of surface
ql Liquid density forces and interfacial properties of mica in purified surfactant
solution. Colloids Surf. 2, 169 – 187.
Plaksin, I.N., 1960. Study of superficial layers of flotation reagents on
minerals and the influence of the structure of minerals on their
References interaction with reagents. Proc. Int. Mineral Process. Congr.,
London 1960, Instn. Min. Metall., London, pp. 253 – 268.
Attard, P., 2003. Nanobubbles and the hydrophobic attraction. Adv. Preuss, M., Butt, H.-J., 1998. Direct measurement of particle–
Colloid Interface Sci. 104, 75 – 91. bubble interactions in aqueous electrolyte: dependence on sur-
Boudriot, H., Matthé, P., Schneider, H.A., 1977. Freib. Forsch.-H. factant. Langmuir 14 (12), 3164 – 3174.
A564, 25 – 30. Preuss, M., Butt, H.-J., 1999. Direct measurement of forces between
Brennen, E., 1995. Cavitation and Bubble Dynamics. Oxford Uni- particles and bubbles. Int. J. Miner. Process 56, 99 – 115.
versity Press, New York. Ralston, J., Fornasiero, D., Mishchuk, N., 2001. The hydrophobic
Brezesinski, G., Mögel, H.-J., 1993. Grenzflächen und Kolloide. force in flotation — a critique. Colloids Surf. A 192, 39 – 51.
Spektrum Akademischer Verlag, Heidelberg. Sakamoto, M., Kanda, Y., Miyahara, M., et al., 2002. Origin of
Charčenko, Ju.V., 1986. Osobennosti kavitacionnych javljenie vo long-range attractive force between surfaces hydrophobized by
flotacionnych mašinach. Obogašč. Rud 31 (3), 16 – 20. surfactant adsorption. Langmuir 18 (15), 5713 – 5719.
Christenson, H.K., Claesson, P.M., 2001. Direct measurements of Schlesier, G., 1965. Über die Adsorption radioaktiv markierter
the force between hydrophobic surfaces in water. Adv. Colloid langkettiger Flotationsamine an Alkalisalzen. Freib. Forsch.-H.
Interface Sci. 91, 391 – 436. A335, 63 – 79.
Gaudin, A.M., Fuerstenau, D.W., 1955. Quartz flotation with cationic Schubert, H., 1972. Die Rolle der Assoziation der unpolaren
collectors. Trans. Am. Inst. Min. Metall. Pet. Eng. Inc. 202, Gruppen bei der Sammleradsorption. Freib. Forsch.hefte,-H.
958 – 962. A514.
H. Schubert / Int. J. Miner. Process. 78 (2005) 11–21 21

Schubert, H., 1996. Aufbereitung fester Stoffe, Bd. II: Sortier- Snoswell, D.R.E., Duan, J., Fornasiero, D., et al., 2003. Colloid
prozesse, 4.Aufl. Deutscher Verlag für Grundstoffindustrie, stability and the influence of dissolved gas. J. Phys. Chem. B
Stuttgart. 107, 2986 – 2994.
Schubert, H., 1999. On the turbulence-controlled microprocesses in Steitz, R., Gutberlet, Th., Hauss, Th., et al., 2003. Nanobubbles and
flotation machines. Int. J. Miner. Process 56, 257 – 276. their precursor layer at the interface of water against a hydro-
Schubert, H. (Ed.), 2003. Handbuch der Mechanischen Verfahrens- phobic substrate. Langmuir 19, 2409 – 2418.
technik, 2 Bände. Wiley-VCH. Stöckelhuber, K.W., 2003. Stability and rupture of aqueous wetting
Schubert, H., Baldauf, H., 1967. Die Rolle der Assoziation der films. Eur. Phys. J. E 12, 431 – 435.
unpolaren Gruppen bei der Adsorption ionogener Tenside an Stöckelhuber, K.W., Radoev, B., Wenger, A., et al., 2004. Rupture of
Oxidoberflächen. Tenside 4 (6), 172 – 176. wetting films caused by nanobubbles. Langmuir 20, 164 – 168.
Schubert, H., Bischofberger, C., 1996. Zu den Mikroprozessen Bla- Taggart, A.F., 1951. Elements of Ore Dressing. John Wiley & Sons,
senzerteilen und Korn-Blase-Haftung in mechanischen Flotation- Inc., New York.
sapparaten sowie Konsequenzen für die Mahstabsübertragung der Tyrrell, J.W.G., Attard, P., 2001. Images of nanobubbles on hydro-
Makroprozesse. Aufbereit.-Tech. 37 (5), 193 – 202. phobic surfaces and their interactions. Phys. Rev. Lett. 87 (17),
Schubert, H., Bischofberger, C., 1998. On the microprocesses air 176104-1-4.
dispersion and particle–bubble attachment in flotation machines Tyrrell, J.W.G., Attard, P., 2002. Atomic force microscope images
as well as consequences for the scale-up of the microprocesses. of nanobubbles on a hydrophobic surface and corresponding
Int. J. Miner. Process 52, 245 – 259. force-separation data. Langmuir 18, 160 – 167.
Slachachov, R., Radoev, B., Stöckelhuber, K.W., 2005. Equilibrium Yang, J., Duan, J., Fornasiero, D., et al., 2003. Very small bubble
profile and rupture of wetting film on heterogeneous substrates. formation at the solid–water interface. J. Phys. Chem. B 107,
Colloid Surf., A Physicochem. Eng. Asp. 261, 135 – 140. 6139 – 6147.

Вам также может понравиться