You are on page 1of 18

Advances in Colloid and Interface Science 114–115 (2005) 9 – 26

www.elsevier.com/locate/cis

Zeta potentials in the flotation of oxide and silicate minerals


D.W. Fuerstenaua,T, Pradipb
a
Department of Materials Science and Engineering, University of Califonia, Berkeley, CA 94720, United States
b
Tata Research Development and Design Centre 54B, Hadapsar Industrial Estate Pune 411 013, India
Available online 23 March 2005

Abstract

Adsorption of collectors and modifying reagents in the flotation of oxide and silicate minerals is controlled by the electrical double layer
at the mineral–water interface. In systems where the collector is physically adsorbed, flotation with anionic or cationic collectors depends on
the mineral surface being charged oppositely. Adjusting the pH of the system can enhance or prevent the flotation of a mineral. Thus, the
point of zero charge (PZC) of the mineral is the most important property of a mineral in such systems. The length of the hydrocarbon chain of
the collector is important because of chain–chain association enhances the adsorption once the surfactant ions aggregate to form hemimicelles
at the surface. Strongly chemisorbing collectors are able to induce flotation even when collector and the mineral surface are charged similarly,
but raising the pH sufficiently above the PZC can repel chemisorbing collectors from the mineral surface. Zeta potentials can be used to
delineate interfacial phenomena in these various systems.
D 2005 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2. The electrical double layer and electrokinetic potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1. Electrical double layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2. Electrokinetic potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3. Effect of electrolytes on the zeta potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3. Application of electrokinetics in flotation systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1. Determination of PZC’s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2. Relationship of flotation response to the PZC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3. Role of the hydrocarbon chain in surfactant adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4. Effect of pH and surface potential on oxide flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5. Flotation depression with inorganic ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.6. Flotation activation with inorganic ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.7. Collector uptake by physisorption or chemisorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.8. Chemisorption and physisorption in oxide flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.9. The adsorption of dodecylammonium acetate on talc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

T Corresponding author. Fax: +1 510 643 5792.


E-mail address: dwfuerst@berkeley.edu (D.W. Fuerstenau).

0001-8686/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.cis.2004.08.006
10 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

1. Introduction

Although useful for determining the efficacy of reagents


for making flotation separations, conducting flotation experi-
ments under carefully controlled physical–chemical condi-
tions simply cannot be carried out with standard laboratory
test machines. So one of the earliest research tools used for
fundamental investigation of surface chemical phenomena in
flotation was the determination of factors that affect contact
angles, a technique used with great success by Sutherland and
Wark [1]. Since most flotation separations are based on the
selective adsorption of collectors (organic surfactants) at
mineral/water interfaces, direct measurement of the amount
of the reagent adsorbed permits a physical–chemical analysis
of adsorption phenomena in relation to wettability. Gaudin
[2] and his co-workers determined adsorption isotherms of a
number of collectors on a variety of minerals, leading the way
in this important approach to flotation chemistry. A signifi-
cant step forward in delineating the physical chemistry of
flotation systems was bringing the concepts of the electrical
double layer and electrokinetic potentials into widespread use
in the interpretation of flotation phenomena [3–5]. Real
progress in correlating flotation response with the physical
chemistry of the system was not made until the modified Fig. 1. Correlation among adsorption density, contact angles, flotation
Hallimond tube was designed and a controlled experimental response and zeta potentials of quartz as a function of pH with 4105M
procedure devised [1,6]. Earlier, vacuum flotation experi- dodecylammonium acetate solutions (after Ref. [8]).
mental methods delineated the limits of flotation, that is float/
no-float conditions, but these could provide no information by electrical double layer interactions. Flocculation–
about anything in between [7]. Fig. 1 shows the first dispersion phenomena are of crucial importance not only
correlation of the flotation of quartz with surface chemical to flotation but also to such other unit operations as
and electrical double layer phenomena when using dodecy- comminution, gravity separation, thickening, filtration
lammonium acetate (DAA) as collector [8]. This figure and tailings management.
illustrates how measurements that reflect conditions at the ! The sign and magnitude of the surface charge controls
solid/liquid interface (adsorption density and zeta potential) the adsorption of physically adsorbing flotation agents.
can be correlated directly with surface phenomena that reflect ! A high surface charge can inhibit the chemisorption of
complex conditions at solid/liquid/gas interfaces (contact chemically adsorbing collectors.
angle and flotation behavior). That the zeta potential becomes ! Slime coatings are controlled by electrical double layer
more negative with increasing pH in the acidic region is due and hydrodynamic phenomena.
to the effect of pH on the surface potential of quartz, as will be ! The double layer on air bubbles has a significant effect
discussed subsequently. on the flotation of naturally floating mineral systems.
Significant progress has been made during the past several ! Flotation kinetics relates directly to the effect of double
decades in delineating the physical chemistry of flotation layers on the rate of film thinning during bubble–particle
systems through utilization of a number of experimental contact
techniques, such as those used in obtaining the results ! Fine particle/air bubble/oil droplet interactions are
presented in Fig. 1. This review, in particular is concerned dominated by surface charge effects.
with electrical double layer phenomena in mineral separation
systems and primarily with the application of electrokinetic Electrical double layer parameters can be evaluated by a
studies for delineation of the relevant double layer effects in number of different experimental methods, depending on
the flotation of oxide and silicate minerals. the properties of the material being investigated. These
The electrical double layer at interfaces plays a broad role methods include the measurement of surface charge by
in many mineral processing operations and is of primary titration or other techniques, direct determination of the
importance in flotation. Some of the surface-physico-chemical capacitance of the double layer, electrocapillarity effects,
effects encountered in these systems include the following: and measurement of electrokinetic potentials which result
from relative tangential motion when at least one of the
! The flocculation and dispersion (and as a consequence phases is a liquid. Electrokinetic potentials, though difficult
also the rheology) of mineral suspensions is controlled to interpret precisely, are relatively simple to measure and
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 11

do yield valuable information on adsorption effects and on figure also shows the drop in potential across the double
potentials between interacting phases. Therefore, these layer. The closest distance of approach of hydrated counter
methods have become both popular and useful in the study ions to the surface, d, is called the Stern plane. The total
of basic phenomena in mineral processing. The purpose of double layer potential, or surface potential, is w 0 and that at
this paper is to review briefly the nature of electrokinetic the Stern plane is w d . From the Stern plane out into the bulk
potentials and the use of electrokinetic information in of the solution, the potential drops exponentially to zero.
understanding flotation behavior. For the sake of brevity, The electrokinetic or zeta potential is usually taken as an
this paper emphasizes the fundamental aspects of electro- approximation of w d . In the case of ions that interact with
kinetic phenomena and their interpretation in relation to the surface sites, either chemically or by some other strong
behavior of flotation separation systems only rather than specific adsorption forces, the adsorbed ions may lie closer
details of experimental measurement techniques. The read- to the surface at a plane b. These two planes are now
ers are also referred to several excellent review papers and known as the outer (d) and inner (b) Stern planes,
recent monographs on the topic [9–17]. respectively, but in earlier literature they were referred to
as the outer and inner Helmholtz planes. As a first
approximation, one often simply refers to adsorption in
2. The electrical double layer and electrokinetic potentials the Stern plane and assumes for simplicity that d and b
coincide. In the case of large organic ions, plane b could lie
2.1. Electrical double layer outside plane d.
Several different parameters that quantify the electrical
The immersion of a solid into an aqueous solution double layer are useful in interpreting flotation behavior,
produces a region of electrical inhomogeneity at the solid– particularly the selective adsorption of collectors. This
solution interface. An excess (+ or ) charge apparently includes such factors as the magnitude of the surface
fixed at the solid surface is exactly balanced by a diffuse charge, the point of zero charge of the mineral, interfacial
region of equal but opposite charge (termed counter ions) potentials, thickness of the electrical double layer, specific
and is called the electrical double layer. adsorption of collectors, and ion exchange phenomena.
Since adsorption of collectors at mineral–water inter- The surface charge in systems of importance to mineral
faces is controlled in many cases by the electrical double processing may arise from a number of different sources.
layer, we must be concerned with factors responsible for the For example, the surface charge on an oil droplet may result
surface charge on the solid and with the behavior of ions from the adsorption of long-chained surfactant ions at the
that adsorb as counter ions to maintain electroneutrality. oil/water interface. In the case of layer-silicate minerals
Fig. 2 is a schematic representation of a simple model of the (such as clays and micas), because of substitution of A13+
electrical double layer at a mineral/water interface, showing for Si4+ in the silica tetrahedra and Mg2+ for A13+ in the
the charge on the solid surface and the diffuse layer of octahedral layer of the crystal lattice, the surfaces of these
counter ions extending out into the aqueous phase. This crystal faces carry a constant negative charge that is
independent of solution conditions. In the case of salt-type
minerals such as barite (barium sulfate), fluorite (calcium
fluoride), argentite (silver sulfide), iodyrite (silver iodide),
etc., the surface charge arises from the preference of one of
the lattice ions for the solid relative to the aqueous phase.
Equilibrium is attained when the electrochemical potential
of these ions is constant throughout the system. Those
particular ions which are free to pass between both phases
and therefore establish the electrical double layer are called
potential-determining ions.
For oxide minerals, hydrogen and hydroxyl ions have
long been considered to be potential-determining [18], but
today there still remains a difference in opinion as to how
pH controls the surface charge on oxides. Since oxide
minerals form hydroxylated surfaces when in contact with
water vapor, a hydroxylated surface should also be expected
when the solid is in equilibrium with an aqueous solution.
Adsorption/dissociation of H+ from the surface hydroxyls
can account for the surface charge on the oxide.
Fig. 2. Schematic representation of the electrical double layer and potential MOHðsurf Þ ¼ MO þ
ðsurf Þ þ Haq ;
drop through the double layer at a mineral/water interface. The planes b and þ þ
MOHðsurf Þ þ Haq ¼ MOH2ðsurf Þ ð1Þ
d represent the inner and outer Stern planes.
12 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

If M +z and A z are the potential-determining ions of solution is increased, the thickness of the double layer
valence z for a given mineral, then the surface charge (r 0) is decreases. For 1-1 valent electrolytes, 1/j is 100 nm in 105
simply given by M, 10 nm in 103 M, and 1 nm in 10l M solutions, for
example.
r0 ¼ zF ðCM þz  CAz Þ ð2Þ If there is only adsorption in the outer Stern plane, then
the surface charge, r 0, is oppositely equal to the diffuse
where F is the Faraday constant and the adsorption density layer charge, r d. The charge in the diffuse double layer r d
(C M +z C A z ) mole per cm2 can be measured by titration of a given by the Gouy–Chapman relation [10] as modified by
suspension of the mineral in water [20]. Stern for a symmetrical electrolyte is:
With regard to flotation behavior, the single most
important parameter that describes the electrical double rd ¼  r0 ¼ ½ 8ee0 RT C 1=2 sinh ð zFwd =2RT Þ ð7Þ
layer of a mineral in water is the point of zero surface charge
(PZC). The PZC is expressed as the condition in the aqueous where w d is the potential in the outer Stern plane. Further, if
solution at which r 0 is zero and this is determined by a w d does not change appreciably, this equation shows that
particular value of the activity of the potential-determining the adsorption density of counter ions should vary as the
(PD) ion (a +M +z (pzc)) for the PD cation). Assuming that square root of the concentration of added electrolyte, as
potential differences due to dipoles, etc., remain constant, the deBruyn [23] found for the adsorption of dodecylammo-
surface potential is considered to be zero at the PZC. The nium acetate on quartz at low concentrations without ionic
value of the surface potential at any activity of potential- strength control.
determining electrolyte, a M mz is given by If counter ions are adsorbed only by electrostatic
attraction, they are called indifferent electrolytes. The
RT aþ þz Boltzmann factor gives the probability of finding an ion i
w0 ¼ ln þ M ð3Þ
zF aM þZ ð pzcÞ of charge z i e at a particular point with potential w, so that
the distribution of ions in a potential field is given by
where R is the gas constant and T the absolute temperature.
In the case of oxides, the relation for the change of w 0 with  
zi Fw
the activity of potential-determining electrolyte (pH) is more Ci ¼ Ci0 exp  ð8Þ
complicated because the activity of species at the surface may RT
not be unity as in the case of the normal reversible electrode.
Because of this [12,17,10–22], the change in w 0 with pH can where C i is the concentration of i at the point where the
be less than the Nernstian slope of 59 mV. However, some potential is w, C i0 is the concentration of i at the point
studies show that oxide electrodes do follow Nernstian where the potential is zero (out in the bulk solution), and z i
behavior. As a first approximation, the potential of an oxide is the valence of i (including sign).
surface can be considered to vary with pH as As presented here, concentrations are given in terms of
moles and not numbers of ions, so Eq. (8) is written in terms
RT h i h i
w0 ¼ 2:3 pH pzc  pH ¼ 0:059 pH pzc  pH volts: of the Faraday constant and not the Boltzmann constant.
F The concentration of indifferent ions in the outer Stern plane
ð4Þ (d) would be that given by Eq. (8) at the potential w d . On
The importance of the PZC is that the sign of the surface the other hand, some ions exhibit surface activity in addition
charge has a major effect on the adsorption of all other ions to electrostatic attraction and adsorb strongly in the Stern
(and particularly those ions charged oppositely to the plane because of such phenomena as covalent bond
surface) because those ions function as the counter ions to formation, hydrophobic bonding, hydrogen bonding, sol-
maintain electroneutrality. In contrast to the situation in vation effects, etc. The term specific adsorption is used for
which the potential-determining ions are special for each such situations. Because of their surface activity, the charge
system, any ions present in solution can function as the of such surface-active counter ions adsorbed in the inner
counter ions. As has been well established [10], the counter Stern plane (b) can exceed the surface charge. In such cases,
ions occur in a diffuse layer that extends from the surface
out into the bulk solution. For a symmetrical electrolyte r0 þ rb þ rd ¼ 0: ð9Þ
consisting of ions of valence z (z +=z =z) at a concentration
C, the bthicknessQ of the diffuse double layer, is 1/j where j For some purposes, the inner and outer Stern planes can
is given by be considered simply to coincide, since much of the
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi interpretation of collector adsorption mechanisms will be
2z2 F 2 C done in terms of electrokinetic potentials, as will be
j¼ ð5Þ discussed shortly. For adsorption densities with less than
ee0 RT
about 30% coverage, Grahame [24] derived a Boltzmann-
where e is the dielectric constant of the liquid and e 0 is the type relation, now generally called the Stern–Grahame
permittivity of a vacuum. As the ionic strength of the equation, which will be expressed here in terms of
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 13

adsorption free energies and adsorption densities, rather than with the radius of the two kinds of ions having to be taken
in terms of charge and potentials as done by Grahame: into account if they differ in size. If the competing ion a is
  an ion that adsorbs only electrostatically (such as Na+ on
Cd ¼ 2rCexp  DG0ads =RT ð10Þ quartz), and b is an ion that has specific affinity to the
surface (such as a flotation collector), then the exponential
where C d is the adsorption density in mol/cm2, r is the
term would simply be exp[(DG 0spec)b /RT]. In ion exchange,
effective radius of the adsorbed ion, here C is the bulk
divalent ions dominate over monovalent ions because of the
concentration in mol/cm3, and DG 0ads the standard free
valence effect in the Boltzmann relation (Eq. (12)).
energy of adsorption. If ions are adsorbed only because of
electrostatic interactions, then the standard free energy of
2.2. Electrokinetic potentials
adsorption is given by

DG0ads ¼ DG0elec ¼ zFwd : ð11Þ Electrokinetic phenomena, which involve the interrela-
tion between mechanical and electrical effects at a moving
When an ion exhibits surface activity, such as a flotation interface, have found widespread use in colloid and surface
collector, then the standard free energy of adsorption has chemistry. The electrical potential involved is that at the
additional terms slipping plane, that is the plane at which the relative motion
takes place. This potential is commonly referred to as the
DG0ads ¼ zFwd þ DG0spec ð12Þ zeta potential, f. The problem is what potential does f
represent in the double layer. Across the diffuse layer, the
where DG 0ads represents the specific interaction terms. There
potential decays with distance according to the equation:
can be various contributions to the specific adsorption free
energy: wðxÞ ¼ wd expð  jxÞ ð16Þ
DG0spec ¼ DG0chem þ DG0h þ DG0hpb þ DG0solv
where x represents the distance from the outer Stern plane to
þ DG0hpb T þ N ð13Þ the slipping plane. Thus, the closer the slipping plane is to
the Stern plane, the closer will w (x) [the zeta potential in this
where the individual terms represent changes in the standard case] approximate w d . In applying electrokinetic results, the
free energy due to chemical bonding, hydrogen bonding, assumption that w (x) iw y is very useful. This approximation
hydrophobic bonding, and solvation effects, respectively. seems permissible because the potential difference between
The term DG 0T
hpb represents the specific adsorption phenom- the Stern plane and the slipping plane is small compared to
ena through surfactant chain interaction with a hydrophobic the total potential difference across the double layer. In the
solid, such as talc or graphite. Depending on the mecha- case of micelles, Stigter [25] showed that the slipping plane
nisms involved in the interaction of the collector with the might occur within 0.1 nm of the polar heads of the ions
mineral surface, the contributions to the change in adsorp- constituting the micelle. Lyklema and Overbeek [26] have
tion free energy can be essentially zero or have a finite suggested a gradual transition between the mobile and
value. immobile parts of the double layer in such systems.
The structure of the diffuse part of the double layer However, it is of interest to note that the electrokinetic
depends upon the composition of the bulk solution and any potential has been found to be independent of Reynolds
change in the composition of the bulk solution can cause ion number, over a wide range from laminar to turbulent flow
exchange within the double layer [10]. If two different kinds conditions [27]. Smith [12,28,29] has presented an interest-
of counter ions, a and b are present, then ing approach to confirm the validity of the w x iw y
assumption in actual colloidal systems, It turns out that
Ca =Cb ¼ K ðCa =Cb Þ ð14Þ the assumption that the slipping plane and the Stern plane
coincide is not only valid but extremely useful in dealing
where C a and C b are the adsorption densities in the double
with the electrical double layer theory of adsorption.
layer, C a and C b are their respective bulk concentrations,
There are four well-known types of electrokinetic
and K is a constant. For indifferent counter ions such as K+
phenomena, from which zeta potentials can be determined
and Na+, K has a value close to unity [10]. However, ions
[10,14]. Electro-osmosis is the process whereby the
having a strong affinity for the surface will tend to
application of an electric field causes flow of liquid.
concentrate in the double layer over simple indifferent
Electrophoresis is the process by which the application of
electrolytes. Where appreciable adsorption occurs in the
a field causes suspended particles to move through a liquid.
Stern layer, the competition for sites would include specific
Streaming potential is the process in which the application
adsorption phenomena:
of a pressure gradient causes a liquid to flow through a bed
"   0  #
DG0ads b DGads a of particles or a capillary, generating an electric current
Ca ra Ca
¼ exp  ð15Þ which leads to an electric field. Sedimentation potential (the
Cb rb Cb RT RT
Dorn effect) is the process by which particles are moved
14 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

under the influence of gravity or centrifugal force, thereby compositions on the surface and in the inner and outer Stern
generating an electric field. Both electrophoresis and planes can give the same zeta potential, depending on
streaming potential methods have been widely used for surface potential, ionic strength, and specific adsorption
investigation of mineral/aqueous solution systems of interest [18]. Thus, to understand the significance of the zeta
to flotation. Sedimentation potentials have been used for the potential at some single condition, determination of the
study of air bubbles. These methods have complications. change in zeta potentials as solution conditions are varied is
When determining streaming potentials with beds of mineral generally required. From these changes, modes of adsorp-
particles, the overlapping of double layers, surface con- tion of various kinds of ions can be ascertained if one makes
ductance and varying pore diameters can cause the the useful assumption that the slipping plane and the Stern
calculated zeta potentials to be too low at low solution plane coincide.
concentrations. Electrophoresis techniques have been Table 1 summarizes how the different types of electro-
widely used for flotation chemistry studies, but the mineral lytes can affect various double layer parameters. How
must be finely ground for these measurements and the various types of electrolytes may affect the zeta potential
suspensions must have low solids content. Evaluation of the can be illustrated with potential–distance curves of the
zeta potential from electrophoretic mobilities can be electrical double layer. Fig. 3a, b, c illustrate schematically
complicated, depending on particle diameter and double the effects of adding potential-determining (PD) electrolytes
layer thickness [14]. In general, the size of particles used in (salts containing either the PD cation or anion), indifferent
flotation chemistry studies is sufficiently large that the electrolytes and surface-active electrolytes respectively. Fig.
simple Smoluchowski equation can be used to convert 3a and c show that the addition of two kinds of electrolytes
electrophoretic mobilities (l E) to zeta potentials (f) and is can reverse the sign of f, one being the potential-
given here since some of the results in this review have been determining ions and the other specifically adsorbed ions.
presented in terms of mobilities: The solution conditions where f is zero is termed the
ee0 f isoelectric point (IEP) or the point of zeta potential reversal
lE ¼ ð17Þ (PZR). The addition of a PD ion can bring f to zero because
g
the surface is uncharged. This IEP is also the point of zero
where g is the viscosity of the solution. With aqueous charge, PZC. Fig. 3c shows that a specifically adsorbed
solutions at 25 C, f (volts)=13 l E. The readers are referred counter ion can also reverse the zeta potential. This is
to several excellent reviews for more details about electro- because contributions to the specific adsorption free energy
kinetics [10,14,17,32–35]. in addition to electrostatic interactions lead to increased
More recently for measurements with relatively concen- adsorption in the Stern plane such that r d Nr 0. Here the
trated suspensions, instruments based on the electro- IEP or PZR is not the PZC.
acoustophoretic effect are available [36–40]. Even though Usually the results of electrokinetic studies are presented
the theory of electro-acoustic effects in concentrated in terms of f-vs-log C curves. As can be ascertained from
colloidal suspensions is still not fully worked out, this Fig. 3 and Table 1, the addition of a positively charged PD
technique offers certain distinct advantages over conven- ion causes f to increase and then decrease because the
tional electrokinetic techniques since colloids containing double layer is compressed as the ionic strength is increased.
particles of sizes ranging from a few nanometers to several The addition of a PD anion leads to the reversal of f when
microns can be analyzed and samples having a wide range the surface potential is reversed. As indicated in Fig. 3b, the
of solid-to-liquid ratios can be conveniently used without lower the ionic strength, the more f approaches w 0. So when
dilution. Electro-acoustophoretic techniques thus provide a the zeta potential is determined as a function of pPD, zeta
means to directly measure the zeta potential (or at least potentials are reduced but still cross at the PZC. If
relative surface charge characteristics) of the actual pulps
during flotation and flocculation processes. However, for
Table 1
relating behavior to solution chemistry, the concentration of The effect of adding potential-determining ions, indifferent electrolytes and
electrolytes in solution must also be determined (because of surface-active electrolytes to a suspension of a material with a reversible
the very high surface area of solids in the system). electrochemical double layer on double layer parameters: surface potential
(w 0) double layer thickness (l/j), surface charge (r 0), and zeta potential (f)
2.3. Effect of electrolytes on the zeta potential Parameter Potential-determining Indifferent Surface-active
electrolytes electrolytes electrolytes
Electrokinetic measurements can provide very useful w0 Increases or Remains Remains constantT
information on adsorption phenomena since changes in the decreases constant
1/j Reduced Reduced Reduced
zeta potential reflect directly adsorption in the Stern plane. r0 Increases or Increases Increases
Since the zeta potential is the potential at the slipping plane decreases
when liquid is forced to move relative to the solid, only ions f Increases, decreases, Only tends Can reverse sign
in the diffuse layer outside the slipping plane are those that or reverses sign to zero
are involved in the electrokinetic process. All sorts of charge T Assuming any shift in PZC to be very small.
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 15

utilization of electrokinetic techniques for distinguishing


between physical and chemical adsorption of flotation
collectors and for delineating conditions under which
various kinds of specific adsorption phenomena might
occur. Often small changes in adsorption are reflected in
pronounced changes in electrokinetic potentials. In the
following paragraphs, we illustrate the utility of electro-
kinetic studies in flotation systems with the help of a few
selected examples.

3.1. Determination of PZC’s

As already stated, an isoelectric point determined by


electrokinetic measurements may or may not correspond to
the point of zero charge of the material. A careful set of
experiments must be carried out to ascertain if an IEP is a
PZC. The true PZC is the condition where the surface is
uncharged. Fig. 4 (top) presents the results of titrating
alumina with acid/base solutions to determine the adsorption
density (surface charge) of H+ and OH on alumina [14,41].
Only at pH 9.1 is the net adsorption of these ions nil, clearly
showing that only at pH 9.1 is the surface uncharged and,
thus, pH 9.1 is the PZC of alumina. Fig. 4 (bottom) shows
the electrophoretic mobility of the same alumina as a
Fig. 3. Schematic representation of the effect of the addition of various
function of pH in solutions of 0.001 M of several indifferent
types of electrolytes on the surface potential, the zeta potential and
thickness of the electrical double layer: (a) the addition of potential-
determining electrolytes, (b) indifferent electrolytes, and (c) specifically
adsorbed electrolytes.

electrokinetic measurements are made as a function of the


concentration of an indifferent electrolyte (such as NaCl
with quartz), f only decreases towards zero with increasing
electrolyte addition. However, effects due to overlapping
double layers reduce the apparent value of the electrokinetic
potential at low ionic strengths when streaming potential or
electro-osmotic methods are used with beds of particles
[14,30] so that f can appear to increase until roughly
millimolar concentrations are reached. If counter ions have a
constant specific adsorption potential, typically f-vs-log C
curves are straight lines that reverse the zeta potential
(similar to those observed upon the addition of the PD
electrolyte but the slope is generally lower). If the
specifically adsorbing electrolyte is a surfactant that forms
hemimicelles or is a multivalent cation that hydrolyzes, the
f-vs-log C curves follow those for indifferent electrolytes
and then deviate sharply when hemimicelles form or cations
hydrolyze. From simple double layer theory, it is possible to
model f-vs-log C curves [30,31].

3. Application of electrokinetics in flotation systems


Fig. 4. Determination of the point of zero charge (PZC) of alumina by acid–
Electrokinetic measurements can be used to delineate base titration of suspensions with KCl as indifferent electrolytes (upper
interfacial phenomena where electrical double layer effects plots) and measurement of the electrophoretic mobility of alumina with
are of relevance to flotation. Of particular importance is the three different indifferent electrolytes (after Refs. [18,41]).
16 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

electrolytes. All curves exhibit an IEP or PZR at pH 9.1, flotation collector, flotation separations depend on differ-
indicating that the IEP is the PZC. In using electrokinetics to ences in the PZCs of the minerals. Typical values of the PZC
determine the PZC, one must decide which electrolyte is for some of the oxide minerals are the following [9]: quartz
potential-determining and then choose another salt as the [SiO2] pH 2, cassiterite [SnO2] pH 4.5, zirconia [ZrO2] pH 4,
indifferent electrolyte. The curves given in Fig. 4 show that rutile [TiO2] pH 5.6, natural hematite [Fe2O3] pH 4.8–6.7,
the surface charge increases with increasing ionic strength, synthetic hematite pH 8.6, corundum [Al2O3] pH 9.1,
in accordance with Eq. (7). On the other hand, as salt is magnesia [MgO] pH 12. The PZCs of silicate minerals are
added to the system, the double layer is compressed (Fig. affected by crystal chemistry and selective leaching of metal
3b) and the zeta potential decreases in absolute magnitude. cations or silica from the surface. Typical values of the PZC
Fig. 4 also shows that when the surface is negatively of some silicate minerals are the following [42]: kyanite
charged, all of the salts have a similar effect on f, since the [Al2SiO5] pH 7.8, zircon [ZrSiO4] pH 5.8, olivine [(Mg,Fe)2
counter ions are K+ in each case, whereas the different SiO4] pH 4.1, almandine [Fe3Al2(SiO4)3] pH 5.8, beryl
electrolytes have a small effect on f below the PZC because [Be3Al2(Si6O18)] pH 3.4, spodumene [LiAl(SiO3] pH 2.6,
of differences in ionic size and polarizability of perchlorate, rhodonite [MnSiO3] pH 2.8, talc [Mg6(Si8O20)(OH)4] pH
nitrate and chloride anions. 3.6, muscovite [K2Al4(Al2Si6O20)(OH,F)4] pH 1, and
Conducting electrokinetic measurements as a function of orthoclase [K(AlSi3O8] pH 1.8.
PD ion activity at several ionic strengths yields curves that
all cross at a point and this IEP should be the PZC (as can be 3.2. Relationship of flotation response to the PZC
seen by the plots given in Fig. 5 for corundum). Electro-
kinetic measurements provide a very convenient and widely Over the period 1953–1956, Fuerstenau began to evolve
used tool for determining the PZC of a mineral. the flotation concept that flotation collectors which physi-
As will be discussed in the next section, for systems cally adsorb must function as counter ions in the electrical
involving the physisorption of ionic surfactants as the double layer, that is, that oxide mineral flotation with an
anionic collector is appreciable only at pHs below the PZC
and with a cation collector on at pHs above the PZC [43].
These concepts, which have been termed the electrostatic
model of flotation collection, are summarized in the
paragraphs that follow. As part of his doctoral thesis, Modi
[44] determined the PZC of corundum and conducted
modified Hallimond tube flotation experiments with a
variety of physisorbed collectors as a function of concen-
tration and pH. Fig. 5 presents the results of the very first
experiments on these concepts which were conducted as
part of Modi’s thesis [43,44], but were not published until
1960 [4]. The upper part of Fig. 5 presents the zeta potential
of corundum (synthetic sapphire) determined by streaming
potential measurements as a function of pH at various
additions of sodium chloride as indifferent electrolyte [45].
All the curves intersect and cross at zero zeta potential at
about pH 9, which is the PZC of this material. Flotation
experiments were conducted at a solution concentration of
4105 M with three different high-purity collectors:
dodecylammonium chloride, sodium dodecylsulfate, and
sodium dodecylsulfonate. The lower part of Fig. 5 shows
clearly that corundum responds to the anionic collectors at
pHs below the PZC where corundum is positively charged,
and to the cationic collector at higher pHs where corundum
is negatively charged. The pKa of dodecylamine is 10.4, and
at pH 10.4 the DAC in solution will be 50% aminium ions
and 50% amine molecules. Under these conditions, flotation
is maximal due to coadsorption of aminium ions and amine
Fig. 5. The dependence of the flotation of corundum (alumina) on surface molecules. As the pH is raised to about 12, flotation drops
charge. Plots of the zeta potential as a function of pH at different
sharply and ceases at pH 12.6. Under these conditions, there
concentrations of NaCl show the PZC of corundum. Lower curves are the
flotation recovery of corundum in 4105 M sodium dodecylsulfonate, are insufficient aminium ions to bind the collector to the
sodium dodecylsulfate, and dodecylammonium chloride (after Refs. surface. This upper pH limit for flotation with dodecylami-
[4,44,45]). nium collectors is virtually universal.
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 17

Subsequently, Iwasaki and Cooke and several co-workers


carried out similar detailed research on the flotation of a
wide range of iron ore minerals with high-purity reagents
and presented and interpreted their results in this same
manner [46,47]. The flotation response of oxide and silicate
minerals to these types of collectors is characteristically
similar to that presented in Fig. 5. Fig. 6 is a composite
showing the flotation response of four different oxides
whose PZCs range from pH 2–9 with dodecylammonium
chloride and sodium dodecylsulfate as collector (data from
[44,46,48]). Here the results are plotted relative to the PZC
of each mineral. Since the flotation time differed with each
investigation (15 s in the case of corundum and 5 min in the
case of goethite), the plots do not fall on top of each other,
but the four different oxide minerals clearly behave quite
similarly. The flotation of all the minerals with the 12-
carbon amine ceases above pH 12, but by presenting the
data relative to the PZC, this does not appear in the plots
except for corundum. Cases [49] conducted a thorough
investigation of the flotation behavior of a series of well-
characterized silicate minerals also using dodecylammo-
nium chloride and sodium dodecylsulfate as the collector.
Fig. 7 summarizes his results for the flotation response of
kyanite (disthene), zircon, garnet, beryl, augite and biotite
with these two collectors, with flotation recovery being
plotted relative to the PZC of each mineral. Presenting the
flotation results in this manner shows whether the collectors Fig. 7. The flotation response of a number of silicate minerals relative to the
interact only by physical forces or whether there are any pH of their PZC in the presence of 104 M dodecylammonium chloride and
additional forces interacting between the collector and the sodium dodecylsulfate as collectors [49]).
mineral surface (Eq. (13)). As can be seen from the plots
given in Fig. 7, only garnet appears to have any extra this suite seem to be governed quite well by electrostatic
affinity (although not strong) for both anionic and cationic interactions with the surface. Because the PZC of biotite,
collectors. The flotation behavior of the other minerals in and also augite, is very low, only limited flotation with
anionic collectors was observed with these minerals.

3.3. Role of the hydrocarbon chain in surfactant adsorption

The hydrocarbon chain of the surfactant contributes to


enhanced adsorption due to the chain–chain interactions in
the Stern plane, as first suggested in 1953 [39]. This can
quantitatively be represented as two terms in the free energy
of adsorption, namely DG 0elec and DG 0hpb. The adsorption of
surfactant ions at low concentration occurs as individual
counter ions, but once the adsorbed ions reach a certain
critical concentration at the solid–liquid interface they begin
to associate into two-dimensional patches of ions at the
surface in much the same way as they associate by self-
assembly into three-dimensional micelles in bulk solution.
The forces responsible for this association at the surface will
be the same as those operating in the bulk during micelle
formation, except that the Coulombic attraction for the
surface adsorption sites will aid the association. Because
the head groups are oriented towards the solid surface before
Fig. 6. The flotation response of a number of oxide minerals with 104 M
dodecylammonium chloride and 104 M sodium dodecylsulfate as the zeta potential reverses sign, the patches of associated ions
collectors as a function of pH relative to the pH of their PZC (after Refs. at the water–solid interface have been termed bhemimicellesQ
[44,46,48]). [3,18]. The bulk critical concentration at which the associ-
18 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

ation is observed is referred to as the hemimicelle (HMC) the formation of micelles is 1.0 to 1.1 RT (about 25
concentration. The HMC is not constant, but will depend on kJ) per mole of CH2 groups [51]. If the cohesive free energy
the potential-determining ion activity in solution. In the case per mole of CH2 group in the hydrocarbon chain is / and
of alkylammonium ions on quartz at neutral pH, the hemi- the number of carbon atoms in the alkyl chain is N, then for
micelle concentration is about 0.01 of the bulk critical micelle hemimicelle formation (if N is about 8 or greater)
concentration (the CMC), but a simple calculation shows that
the concentration of adsorbed ions within the Stern layer DG0hpb ¼ N /: ð18Þ
approximates the CMC [50,51].
To help visualize these concepts, a simple schematic If all other specific adsorption effects are negligible, then
drawing is given in Fig. 8 for an anionic surfactant or above the hemimicelle concentration the adsorption density
collector adsorbing on a positively charged mineral surface. of surfactant ions in the Stern layer is given by
   
Fig. 8a represents the case for low adsorption densities  DG0ads  zFwd  N /
where the surfactant ions adsorb individually, both in the Cd ¼ 2rCexp ¼ 2rCexp ð19Þ
RT RT
diffuse layer and in the Stern layer. Flotation would be
limited under these conditions. Fig. 8b illustrates the This equation shows that the amount of surfactant adsorbed
association of adsorbed surfactant ions into hemimicelles in the Stern plane should depend strongly on the hydro-
in the Stern layer. As drawn, the system is at its IEP or PZR, carbon chain length, once the hemimicelle concentration is
where the double layer can simply be considered to be a exceeded. This has been demonstrated for the adsorption of
bimolecular condenser (without a diffuse layer). Flotation alkylammonium ions on quartz [51] and alkyl sulfonate ions
should be optimal under these conditions since the on alumina [52].
surfactant head groups are oriented towards the solid and Somasundaran conducted a detailed systematic study of
the zeta potential is zero. Further surfactant adsorption the effect of the chain length of aminium collectors on the
causes the zeta potential to be reversed because the charge in flotation of quartz and in the course of that research also
the Stern layer exceeds the surface charge, and a triple layer measured the effect of chain length on the zeta potential
builds up with the head groups of some of the surfactant of quartz. Fig. 9 presents the results of measurements of
ions now being oriented towards the water. Flotation will the zeta potential of quartz at neutral pH as a function of
tend to decrease under these conditions. the concentration of alkylammonium acetates (without
Since the length of the hydrocarbon chain of a surfactant ionic strength control) ranging from 10 to 18 carbon
is important in adsorption processes, the number of carbon atoms [51]. As expected, the zeta potential of quartz with
atoms in the chain should directly relate to adsorption the various surfactants approximates that with ammonium
behavior. By conducting a series of streaming potential acetate until hemimicelle formation sets in. Earlier
measurements on quartz with alkylammonium acetates measurements showed that octylammonium acetate
having alkyl chains ranging from 8 to 18 methylene groups, behaves almost identically to ammonium acetate, indicat-
Fuerstenau [50] clearly showed that hemimicelle formation ing that the hydrocarbon chain must exceed, say, 8 carbon
indeed depends on hydrocarbon chain length. In these atoms to be sufficiently hydrophobic for hemimicelle
electrokinetic studies, octylammonium acetate behaved formation to occur [50]. Increasing the length of the
identically to ammonium acetate. Studies of micelle hydrocarbon chain causes a systematic decrease in the
formation have shown that the cohesive free energy for hemimicelle concentration.

Fig. 8. Schematic representation of the aqueous adsorption of an anionic surfactant (collector) on a positively charged mineral. (a) Adsorption as individual
counter ions at low concentrations. (b) Aggregation of adsorbed surfactant ions into hemimicelles at higher adsorption densities with the charged head groups
oriented towards the solid surface and at the PZR in this case. (c) At adsorption densities where the charge in the Stern plane exceeds that of the surface, a triple
layer forms with surfactant ions being adsorbed in reverse orientation due to electrostatic repulsion.
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 19

ions at w d =0 should be a linear function of the alkyl chain


length with a slope equal to //2.3RT. Fig. 8 (top) includes
such a plot for alkylammonium ions on quartz, and from this
plot / is calculated to be 1RT (about 0.6 kcal/mol of
CH2 groups) in agreement with values obtained from
solubility and micelle data [51].
For rapid flotation to occur, the collector must be adsorbed
in the Stern plane and in sufficient quantity for the mineral to
be hydrophobic. Rapid flotation takes place once hemi-
micelles form. The one system in which chain length and
flotation response has been investigated in detail is the
flotation of quartz with high-purity alkylammonium acetates.
Vacuum flotation experiments were conducted by deBruyn
[53] and also by Somasundaran [54]. Vacuum flotation
experiments have been conducted to determine incipient
flotation conditions with collector concentration and pH
being the two variables. Often the resulting plots have been
termed critical flotation curves. The term flotation edge has
been used to designate the conditions for incipient flotation.
The results obtained by Somasundaran and by deBruyn both
show that incipient flotation exhibits a chain-length effect,
probably assisted by transfer of surfactant adsorbed on the
bubble to the mineral surface. In Fig. 9 (top) the results of
Fig. 9. The zeta potential of quartz as a function of the chain length of Somasundaran’s measurements of the aminium concentration
alkylammonium acetates at neutral pH. The upper curves give a plot of the for incipient flotation of quartz (the flotation edge) at pH 6.5
cationic surfactant concentration at which the zeta potential is zero as a
is also plotted as a function of the number of carbon atoms in
function of the number of carbon atoms in the hydrocarbon chain, the
concentration at which Hallimond tube flotation increases sharply (termed the collector chain. These experiments again show a system-
the flotation parameter), and the flotation edge or critical flotation atic chain-length effect. With all of the alkylaminium
concentration obtained from vacuum flotation experiments (after Refs. collectors, deBruyn and Somasundaran found that flotation
[51,54,55]). ceases slightly above pH 12, where the collector has
hydrolyzed to amine molecules, and below about pH 2.5–
When the zeta potential is reduced to zero after 3.5 near the PZC of quartz.
association of the alkyl chains, it follows from Eq. (19) that Detailed Hallimond tube flotation experiments were
carried out for quartz with a full range of high-purity
lnC0 ¼ N ð/=RT Þ þ lnðCd Þ0  ln2r ð20Þ alkylammonium acetate collectors, from 18 down to 4 carbon
atoms [55]. These flotation results, which are presented in
where C=C 0 at w d =0. If the ratio of the adsorption density Fig. 10, show that the concentration of collector needed for
at w d =0 to the factor 2r is relatively independent of chain flotation decreases systematically as the chain length of the
length, then the logarithm of the concentration of surfactant aminium salt is increased. Flotation is even possible with

Fig. 10. Flotation response of quartz at pH 6–7 as a function of the concentration of alkylammonium acetates of different hydrocarbon chain lengths (after Ref.
[55]).
20 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

butylammonium acetate, but at a collector concentration of


about 1 M. Here again, the collector adsorbed on the bubble
must also be assisting adsorption since chain–chain inter-
action phenomena would be negligible. Of interest is that
with a strongly chemisorbing collector, such as ethyl
xanthate, sulfide minerals respond to flotation at a residual
collector concentration of about 1 ppm in solution. The quartz
flotation results were analyzed in a manner analogous to that
for the effect of aminium ion concentration on the PZR, using
an equation similar to Eq. (20), except in terms of the
hemimicelle concentration which was obtained by extrap-
olation of the flotation recovery curve and was termed the
flotation parameter [55]. In Fig. 9, this flotation parameter
(essentially the HMC) is plotted as a function of the number
of carbon atoms in the aminium collector, and a straight line
similar to that for the PZR is obtained, again giving a value
for / as 1RT.
Somasundaran and Fuerstenau [56] conducted the first
experiments on the adsorption of sodium dodecylsulfonate on
alumina with controlled ionic strength, using sodium chloride
as the supporting electrolyte. With ionic strength fixed, they
observed a distinct three-stage isotherm when the results are
plotted logarithmically (sometimes called an SF isotherm). At
low concentrations, surfactant ions are adsorbed individually
by exchange with chloride counter ions; but when the
hemimicelle concentration is reached, the slope of the Fig. 11. Effect of the concentration of sodium dodecylsulfate on the
isotherm increases markedly due to chain–chain interaction; flotation recovery of corundum at various pH values. The upper curves
and when the zeta potential is reversed, the slope of the present the zeta potential of corundum as a function of the concentration of
isotherm decreases again (due to reversal of f). Later work sodium dodecylsulfate. In addition, the effect of sodium chloride on the zeta
potential of corundum is also shown, to illustrate the difference between an
showed that these isotherms exhibit a fourth region as the
indifferent electrolyte and a surfactant forming hemimicelles (after Refs.
CMC is approached. This review will not discuss details of [4,57]).
adsorption isotherms, but the contact angle exhibited abrupt
changes at the boundaries of the three regions.
mineral being floated. At pH 9.3 there is limited flotation,
3.4. Effect of pH and surface potential on oxide flotation which was also observed in the curves given in Fig. 5. This is
because there are still a certain number of positive sites on the
Modi and Fuerstenau [4,57] investigated how pH controls surface at pHs above the PZC. When pH 11 is reached, the
the effect of sodium dodecylsulfate on the zeta potential and surface is so negatively charged that there is no flotation of
flotation response of corundum, and their results are corundum with dodecylsulfate as collector. Comparison of
presented in Fig. 11. For the alumina/dodecylsulfate system, the upper and lower parts of Fig. 11 shows that flotation
decreasing the pH from 6.5 to 4.0 increases the surface increases sharply as the hemimicelle concentration is reached
potential and surface charge, leading to increased adsorption and flotation is about 100% when the IEP or PZR is reached.
of counter ions. For a physisorbing surfactant, such as sodium
dodecylsulfate, the hemimicelle concentration is reduced 3.5. Flotation depression with inorganic ions
nearly 10 fold, as can be seen by the sharp change in the
slopes of the f-log C curves for SDS given in the upper part of With physisorbed collectors that float insoluble oxides
Fig. 11. It is at these points where the SDS plots deviate from and silicates, reducing and reversing the surface charge by
the plots for sodium chloride. At pH 11 where the corundum changing the pH depresses flotation, as exemplified by
is negatively charged, the counter ions are sodium ions and raising the pH when floating corundum with SDS (Fig. 10).
the dodecylsulfate ions do not take part in the adsorption Since physisorbed surfactants act as flotation collectors by
processes. Because of hemimicelle formation occurring at a adsorption as counter ions in the double layer, the addition
lower concentration as the pH is lowered, flotation of of any salt should compete with the surfactant ions as
corundum also occurs at a lower reagent concentration, as counter ions and at sufficiently high concentration depress
can be seen by the plots given in the lower part of Fig. 11. flotation. This might be important in the industrial flotation
Clearly, reduction in reagent consumption can be achieved if of nonmetallic minerals using seawater. To illustrate this
selectivity is maintained at higher surface potential for the phenomenon, Fig. 12 presents plots of the effect of sodium
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 21

potentials at different planes and knowledge of hydration


phenomena, these estimates starting with electrokinetic
results are interestingly close.

3.6. Flotation activation with inorganic ions

We have just seen the effects of sodium sulfate in


inhibiting the flotation of positively charged alumina with
an anionic collector. Fig. 13 illustrates how the same
inorganic salt can promote flotation. At pH 6, alumina does
not respond to flotation with dodecylammonium chloride as
collector since the organic cations do not adsorb on a
positively charge solid. On the other hand, by adding
Fig. 12. Depression of the flotation of corundum using sodium dodecyl
sufficient SO42 to the system, the net charge in the Stern
sulfate as collector at pH 6 by the addition of sodium chloride and sodium layer and the surface is reversed, and the solid can now
sulfate. Included in this figure is the effect of the two inorganic salts on the adsorb organic cations [4]. In flotation terminology, sulfate
zeta potential of corundum (after Refs. [4,45,57]). ions are said to function as an activator; they serve as a link
between the surface and similarly charged collector ions. In
sulfate and sodium chloride on the zeta potential of such systems the electrical double layer is in reality a triple
corundum and also their effect on corundum flotation with layer.
4105 M SDS at pH 6 [4,45,57]. Sulfate ions exhibit Opposite to the case of activation of corundum with
specific adsorption on corundum, as evidenced by the sulfate ions shown in Fig. 13, industrial activation of oxides
reversal in the zeta potential at 3104 M Na2SO4. The generally involves the linkage of an anionic collector to a
plots given in Fig. 12 show excellent correlation of the negatively charged mineral through the addition of strongly
effect of the two inorganic salts on the zeta potential of adsorbing metal cations. Since hydrolyzed metal cations are
corundum (in the absence of SDS) and in depressing extremely surface-active, they will function best as activa-
flotation. This figure shows that both Cl and SO42 inhibit tors at the incipient hydrolysis pH. Fig. 14 shows the results
the flotation of positively charged alumina but with the of such an investigation of quartz flotation carried out by
effect of SO42 being nearly 500 times that of Cl. M.C. Fuerstenau and Palmer [58] using 104 M sulfonate as
The greater effect of sulfate over chloride ions in collector and various metal cations at 104 M as activator.
depressing alumina flotation results from the valence effect As the pH is increased, activation occurs sharply when the
and the specific adsorption potential of SO42 on alumina. activating cation begins to hydrolyze in solution. Hydroxy-
This can be interpreted in accordance with Eq. (15). The lated metal ions are extremely surface active and readily
concentration of SDS used in these experiments was adsorb on all sorts of adsorbents. In Fig. 14, only the initial
4105 M, and hemimicelles had formed at the surface, flotation edge is shown for clarity in making the plots.
with the adsorption free energy being 5.6RT under these However, flotation ceases at about pH 3.8 with Fe(III), at
conditions [57]. From conditions leading to the PZR of pH. 4.5 with Al(III), and at pH 11.7 with Mg(II), which are
corundum with sodium sulfate, the specific adsorption free the pH values at which those metal hydroxides precipitate.
energy of divalent sulfate ions is about 6.9RT. The ionic Since the PZC of quartz occurs at about pH 2, quartz
radii are 0.18 nm for chloride ions and 0.29 nm for both
sulfate ions and the head group of dodecylsulfate ions. For
these calculations, we will consider conditions that reduced
flotation recovery to 50%. Fig. 12 shows that 5.6103 M
NaCl is required to do this when the collector concentration
is 4105 M, or 140 times as much. Assuming that half the
dodecylsulfate ions in the Stern plane are replaced by
chloride ions, Eq. (15) estimates that the required ratio of
chloride to dodecylsulfate is 170. To carry out the
calculation for depression with divalent sulfate ions, we
have to take into account that one divalent sulfate ion will
displace two dodecylsulfate ions in the Stern plane and
bring the zeta potential into the calculation. Doing so shows
that the concentration of divalent sulfate ions necessary to
reduce the recovery to 50% is one-fourth that of the SDS in
solution, or 0.9105 M compared to the 1.3105 M Fig. 13. The activation of positively charged corundum at pH 6 with sodium
observed. Given that a rigorous calculation would require sulfate for flotation with dodecylammonium chloride (after Ref. [4]).
22 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

At point A, the surface charge is zero and ideally there is no


double layer. Nearly all minerals encountered in mineral
processing are hydrophilic in the absence of a collector,
except for graphite, molybdenite and talc which are
naturally hydrophobic because of their crystal structure.
If a physisorbing anionic surfactant is added to the
system (such as alkyl sulfates on alumina) with the
concentration of indifferent salt being the same, a curve
similar to the dashed one in Fig. 15 is obtained. Starting at
high pPD, the curves coincide because there is no specific
adsorption of the anionic surfactant on the negatively
charged mineral. As the pPD is decreased, point A is
reached, and this still is the PZC of the solid. As the pPD is
decreased further and the solid becomes positively charged,
the negatively charged surfactant ions begin to adsorb.
When the pDP is lowered to the region around point D, the
Fig. 14. The activation of quartz with 104 M metal cations and 104 M adsorption density of the surfactant ions reaches the
sulfonate collector as a function of pH. The minimum flotation edges are hemimicelle concentration and chain–chain association
plotted for clarity, but maximal flotation occurs only where the data points leads to an increasing specific adsorption potential such
are shown (after Ref. [63]).
that at point C the zeta potential is reversed. Point C is an
IEP (or PZR) and not the PZC. Even though the surface
responds to flotation with a sulfonate or a soap only after itself is becoming more highly positively charged, the
extensive adsorption of the activating metal cation in the increased adsorption of negatively charged surfactant ions
Stern layer. In the case of complex silicates, autoactivation increases the specific adsorption potential because of
can occur by metal ions derived from the crystal lattice by increased chain–chain interaction. At point C the double
their dissolution and readsorption as a flotation activator. layer is a molecular condenser consisting of a surface charge
and a molecular layer of surfactant anions. In the case of a
3.7. Collector uptake by physisorption or chemisorption hydrophobic material, such as talc, the f-vs-pPD curve will
have the appearance of the curve with open circles in Fig.
If counter ions are adsorbed only through such forces as 15, where the reversal of f occurs at point BV, somewhat
electrostatic attraction and hydrophobic bonding (van der lower than point A. This is again an IEP that is not the PZC
Waals interaction between the hydrocarbon chains), the but reflects specific adsorption due to the hydrophobic chain
process is termed physical adsorption or physisorption. If of the surfactant displacing water molecules from the
the surfactant forms covalent bonds with metal atoms in the
surface, then the process is called chemisorption. As already
discussed, examples of physical adsorption in mineral/water
systems include alkylammonium ions on quartz and other
oxide minerals and alkyl sulfates and sulfonates on alumina.
Examples of chemisorption of surfactants are the, oleate/
hematite, hydroxamate/hematite, hydroxamate/MnO2 and
oleate/fluorite systems. Infrared spectroscopic studies have
helped confirm the existence of calcium oleate, ferric oleate,
ferric hydroxamate and Mn-hydroxamate species on the
surface, indicating chemisorption by covalent bond linkages
[59–62].
The nature of the plot of zeta potential as a function of
the potential-determining (PD) ion activity can be used to
distinguish whether the adsorption occurs by physical or
chemical interactions. Plotting f-vs-pPD for systems at
constant concentration of an indifferent electrolyte with and
without added surfactant yields visible evidence of various
adsorption mechanisms. Fig. 15 is a schematic illustration
Fig. 15. Schematic illustration of the effect of potential-determining ions
showing the effect of various interactions on the zeta
(pH for oxides) on the zeta potential in the presence of an indifferent
potential of a solid. The solid line in this figure is typical for electrolyte, a physisorbing anionic surfactant on a hydrophilic mineral, a
that of an indifferent electrolyte on a hydrophilic material, physisorbing collector on a hydrophobic mineral, and a chemisorbing
with the reversal of zeta potential at point A being the PZC. anionic collector.
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 23

hydrophobic surface. In this case, the curves come together and the double layer can be considered to be a molecular
0 0
again at point EV where the mineral surface is sufficiently condenser with DG ads =DG spec . Under these conditions, the
negatively charged to overcome the specific adsorption Stern–Grahame relation can be expressed as
potential due to hydrophobic interactions with the surface.
For surfactants that chemisorb at the surface of a mineral, w0 Km ¼ 2rC0 zFexp  DG0spec =RT ð21Þ
bonding to the surface can be very strong and, for an anionic
collector, only when the surface becomes sufficiently
where K m is the capacitance of the molecular condenser, r is
negatively charged by raising the pPD will adsorption
the effective radius of the ionic head of the adsorbed ion,
cease. This occurs at point EW where the two curves come
and C 0 is the concentration in mole per cm3 at the IEP. For a
together again. At this point chemical contribution to the
given activity of potential-determining ions, w o is obtained
free energy of adsorption is counter balanced by electro-
from Eq. (3) (or other form of the Nernst equation modified
static repulsion. Where the two curves join again generally
for oxides). If K m is assumed constant (for example, 15 AF/
marks the upper limit for flotation by a chemisorbing
cm2), then DG 0spec can be estimated. With this method,
collector. Also, again point BW is an IEP and not the PZC.
Fuerstenau and Modi [57] calculated DG 0spec to be 9RT at
Detailed experiments were conducted by Han et al. [63]
the PZR for dodecylsulfonate ions on corundum at pH 6.5
where the electrophoretic mobilities of synthetic hematite
(Fig. 9). Its value would be zero at the HMC and at the
we measured for various concentrations of sodium dodecyl
break in the zeta potential curve above the HMC, which
sulfate, potassium octylhydroxamate, and oleic acid in
appears to represent monolayer coverage in streaming
aqueous solutions containing 0.002 M NaCl as a function
potential experiments, extrapolation indicates that DG 0spec
of pH. In Fig. 16, their results for 2104 M hydroxamate
is 13RT at this point.
and dodecylsulfate are plotted together with results for
At Point E in Fig. 15, specific adsorption in the Stern
1104 M oleic acid. The oleate chain must enter into the
plane must be nil and the specific adsorption potential is just
adsorption process also because the curve remains in the
counter balanced by the electrostatic contribution, that is,
negative region (and even becomes slightly more negative)
DG 0elec=DG 0spec. In the case of chemisorption, this relation
as the pH is lowered.
permits estimation of DG 0chem.
At any IEP, the free energy of adsorption can be
Specific interactions during the adsorption of a surfactant
calculated or estimated. For the sake of simplicity, assume
on a mineral surface result in a shift of isoelectric point
that the inner and outer Stern planes coincide and that
(IEP). Pradip [64] has shown that this shift in IEP,
specific adsorption occurs in the plane d. The adsorption
DpH=(pPZRPZCp) is related through the following
density in the Stern layer is given by Eq. (10). At any IEP or
0 equations derived on the basis of electrical double layer theory:
PZR (such as point B in Fig. 13), f=w d =0, so that DG ads

DpH ¼ 1:040C0 exp  DG0spec =RT ð22Þ

where C0 is the equilibrium molar concentration of surfactant


at the IEP (or PZR) and DG 0spec is the specific free energy of
interaction in kcal/mol or RT units. The validity of this
relationship is illustrated in Fig. 17 for the chemisorption of
octylhydroxamate for two oxide minerals. The calculated
values are comparable to those computed from the corre-
sponding adsorption isotherms [64].

3.8. Chemisorption and physisorption in oxide flotation

The flotation response of manganese dioxide with


sodium dodecylsulfonate [62], potassium octylhydroxamate
[62] and sodium oleate [65], are given in Fig. 18 (top, center
and bottom). These plots illustrate the complicated nature of
physical and chemical adsorption in flotation processing.
The PZC of this manganese dioxide (high-purity g-MnO2)
occurs at pH 5.6 [62]. The results given in Fig. 18 (top)
indicate that the flotation of manganese dioxide with anionic
Fig. 16. The electrophoretic mobility of hematite as a function of pH in
sulfonate behaves as expected for a physisorbing collector.
solutions containing 2103 M NaCl, 2104 M sodium dodecylsulfate
and potassium octylhydroxamate, and 1104 M sodium oleate. All Flotation only occurs below about pH 6, where the surface
solutions with the three different surfactants also contained 2103 M carries a positive charge and hence adsorbs the anionic
NaCl as supporting electrolyte for ionic strength control (after Ref. [63]). sulfonate collector as a counter ion in the double layer. The
24 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

containing the same amount of indifferent electrolyte but


also a constant concentration of surfactant (such as oleate),
the shift in the zeta potential is a measure of the specific

Fig. 17. Shift in the isoelectric point (IEP) through specific adsorption of a
surfactant (hydroxamate) on hematite and pyrolusite and the calculated shift
expected from electrical double layer theory (after Ref. [64]).

results presented in Fig. 18 (center) show the flotation


response with potassium octylhydroxamate, a chelating
agent that strongly coordinates with manganese ions at the
mineral surface, as substantiated by infrared spectroscopy.
In this chemisorbing system, flotation takes place at pHs
where the solid surface is highly negatively charged. The
maximum in flotability occurs at about pH 9, similar to the
case for hematite, which is approximately the pKa of
hydroxamic acid. As the pH is increased above this value,
flotation decreases, ceasing by about pH 11, because of
strong repulsion of the hydroxamate anions by the
negatively charged surface. Hydroxamic acid molecules
must be the chemisorbing species.
The flotation of manganese dioxide with oleate collector
is particularly complicated, with two flotation peaks being
present in this system, as can be seen in Fig. 16 (bottom). The
peak at about pH 8 is typical for the chemisorption of oleate
on many minerals. Ananthapadmanabhan et al. [66] have
suggested that the acid–soap dimer, which is at a maximum
under these conditions, is extremely surface-active. Adsorp-
tion at these alkaline pHs occurs under conditions where the
solid is negatively charged, and strong chemical bonding
must be responsible for the observed mineral–collector
interaction. Infrared spectroscopy showed the presence of
ferric–carboxylate bonds when oleate is adsorbed on
hematite [59]. In the vicinity of pH 5.5, there is little
flotation of manganese dioxide. However, as the pH is
lowered further, a second flotation maximum appears. This
must now be due to physical adsorption that involves a
positively charged surface adsorbing oleate anions. Since the
pKa of oleate is about 4.7, the actual collecting species must
be oleate ions together with oleic acid molecules. At pHs less
than 3 or so, most of the oleate has been transformed to oleic
acid, which is not the reactive species, and flotation ceases
(similar to the action of amines at high pH).
Fig. 18. The effect of pH on the flotation of manganese dioxide (PZC=pH
Fuerstenau and Shibata [67] pointed out that if f-vs-pH 5.6) with various additions of three different collectors: sodium dodecylsul-
curves are measured at constant concentration of indifferent fonate, potassium octylhydroxamate and sodium oleate (after Refs.
electrolyte (such as sodium chloride) and also with solutions [62,65]).
D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26 25

adsorption free energy and that this correlates quite well nitrate as supporting electrolyte [67]. These results without
with flotation response under the same conditions. surfactant indicate that the condition for zero net surface
charge (the isoelectric point) of talc occurs at pH 2.5.
3.9. The adsorption of dodecylammonium acetate on talc Because a talc particle has a highly anisotropic surface, the
net charge on the particles must be controlled to a large
Because of the natural flotability of talc, specific extent by adsorption/dissociation reactions of H+ and OH
adsorption phenomena on talc differs from that on oxides ions with broken Si–Mg–O bonds at the crystal edges. As
and other silicate minerals. Talc belongs to the family of can be seen, f becomes only slightly more positive over all
silicate minerals that are comprised of three repeating pHs with aminium additions up to 105 M. It is over this
binfiniteQ sandwich-like layers: a layer of octahedrally same range that flotation increases to a plateau. It appears
coordinated magnesium–hydroxyl ions between two layers that at low amine concentrations, the most significant
of tetrahedrally coordinated silicon–oxygen ions. Since each adsorption of the organic cations takes place on the charged
three-layer sheet is electrically neutral, the crystal is held edges. At higher amine concentrations, there is a sharp
together by weak van der Waals forces acting across change in the adsorption mechanism, as can be seen from
adjoining sheets, and the crystals readily cleave along these the jump in the f-vs-pH curve for 5105 M surfactant. The
sheets, forming faces that are uncharged and nonpolar in pronounced shift of the curves towards more positive zeta
water. On the other hand, the edges of talc particles are potentials probably results from the hydrophobic surface
formed by breaking ionic/covalent bonds and therefore contribution to the adsorption free energy. Here adsorption
exhibit a high pH-dependent electrical charge in water and must take place by the hydrocarbon chains displacing water
are highly polar. Because the faces are hydrophobic and the molecules from the hydrophobic surface. Because the face
edges hydrophilic, the flotation behavior of talc can be of a soft talc crystal will have steps on it, these small edge
rather complex. effects must also contribute to adsorption on the faces. At
Some insight into the behavior of talc/aminium systems higher DAA concentrations at acidic pHs, the major
can be gained from the measurement of zeta potentials. The contribution to the adsorption free energy must result from
plots given in Fig. 19 present the f-potential of talc as a hydrophobic interactions of surfactant chains with the
function of pH with concentrations of dodecylammonium hydrophobic surface since the zeta potential is highly
0
acetate ranging from 0 to 0.001 M in 0.002 M potassium reversed (the DGhpb* term in Eq. (13)). The sharp increase
in f at pH 10 in millimolar surfactant solutions results from
coadsorption of aminium ions and amine molecules in
hemimicelles. Above pH 12, adsorption of aminium ions
cease and the zeta potential drops sharply to the same
negative value observed without collector.

4. Summary

The electrical double layer has a dominant role in


collector adsorption and flotation of oxide and silicate
minerals. Measurements of zeta potentials provides a
straight-forward way of ascertaining adsorption mechanisms
and conditions in such systems. Cationic collectors adsorb
when the surface charge on the mineral is negative, and
vice-versa for anionic collectors. Separations are based on
differences in the points of zero charge (PZCs) of minerals
in an ore. For physically adsorbing collectors, surfactant
ions can aggregate in the Stern plane, forming hemimicelles
through hydrocarbon chain association. For rapid flotation
rates, collector adsorption in the Stern plane must be
pronounced. Collector adsorption can be enhanced by
increasing the surface potential by adding potential-deter-
mining ions (through pH regulation in the case of oxides),
increasing the chain length of the collector, and adding
neutral organic molecules. Flotation can be depressed by
Fig. 19. The effect of pH on the zeta potential of talc in the presence of
various additions of dodecylammonium acetate with 2103 M potassium adding inorganic salts that compete with the collector.
nitrate as supporting electrolyte, illustrating specific adsorption through Minerals will respond to flotation with anionic chemisorb-
hydrophobic bonding with a hydrophobic mineral surface (after Ref. [67]). ing collectors even when the mineral surface is highly
26 D.W. Fuerstenau, Pradip / Advances in Colloid and Interface Science 114–115 (2005) 9–26

negatively charged, that is above the PZC. However, [29] Foxall T, Peterson GC, Rendall HM, Smith AL. J Chem Soc, Faraday
electrostatic effects are important even in chemisorbing Trans I 1979;75:993.
[30] Russel WB, Saville DA, Schowalter WR. Colloidal Dispersion. UK7
surfactant systems since strongly opposing electrostatic Cambridge University Press; 1989.
interaction can overcome chemisorption, thereby preventing [31] Nguyen AV, Schulze HJ. Colloidal Science of Flotation. NewYork7
collection. Whether flotation collectors are physisorbed or Marcel Dekker; 2004.
chemisorbed, they have distinct effects on the zeta potential [32] Delgado AV, editor. Interfacial Electrokinetics and Electrophoresis.
Marcel Dekker; 1998.
of the mineral, and these effects can be used to evaluate
[33] Oshima H, Furusawa K, editors. Electrical Phenomena at Interfaces:
specific adsorption free energies. Fundamentals, Measurements and Applications, 2nd ed. 1991.
[34] O’Brien RW. Fluid Mech 1988;190:71.
[35] Marlow BJ, Fairhurst D, Pendse HP. Langmuir 1988;4:611.
Acknowledgements [36] Malghan SG, editor. Electroacoustics for Characterization of Partic-
ulate and Suspensions. USA7 NIST Publication; 1993.
[37] O’Brien RW, Midmore BR, Lamb A, Hunter RJ. Faraday Discuss
Pradip wishes to acknowledge support from the Depart- Chem Soc 1990;90:301.
ment of Science and Technology, India, under an Indo–U.S. [38] Babchin AJ, Chow RS, Sawatzky RP. Adv Colloid Interface Sci
STI program during the preparation of this paper. 1989;30:111.
[39] Fuerstenau DW, ScD. Thesis Massachusetts Institute of Technology;
1953.
[40] Hunter RJ, Wright HJL. J Colloid Interface Sci 1971;37:564.
References [41] Yopps JA, Fuerstenau DW. J Colloid Sci 1964;19:61.
[42] Fuerstenau DW, Raghavan S. Proceedings, XII Int Mineral Processing
[1] Sutherland KL, Wark IW. Principles of Flotation. Melbourne7 Congress, vol. II. Sao Paulo7 Nacional Publicacoes & Pubicadado;
Australasian Institute of Mining and Metallurgy; 1955. 1980. p. 368.
[2] Gaudin AM. Flotation. New York7 McGraw Hill; 1957. [43] DW Fuerstenau, in Adaptation of New Research Techniques to
[3] Gaudin AM, Feurstenau DW. Trans AIME 1955;202:66. Mineral Engineering Problems, Mass. Institute of Technology, NYO
[4] Modi HJ, Fuerstenau DW. Trans AIME 1960;217:381. 7178, MITS 31: NYO 7179, MITS 32 1956.
[5] Aplan FF, Fuerstenau DW. In: Fuerstenau DW, editor. Froth [44] Modi HJ, ScD. Thesis, Massachusetts Insitute of Technology; 1956.
Flotation—50th Anniversary Volume. New York7 AIME; 1962. p. 170. [45] Modi HJ, Fuerstenau DW. J Phys Chem 1957;61:640.
[6] Fuerstenau DW, Metzger PH, Seele GD. Eng Min J 1957;158:93. [46] Iwasaki I, Cooke SRB, Colombo AF. Report of investigation RI 5593.
[7] Schuhmann R, Prakash B. Trans AIME 1950;187:591. U.S7 Bureau of Mines; 1960.
[8] Fuerstenau DW. Trans AIME 1957;208:1365. [47] Iwasaki I, Cooke SRB, Kim YS. Trans AIME 1962;223:113.
[9] Fuerstenau DW, Healy TW. In: Lemlich R, editor. Adsorptive Bubble [48] Nakatsuka K, Matsuoika I, Shimoiizaka J. Proceedings 9th Int.
Separation Techniques. New York7 Academic Press; 1972. p. 92. mineral processing congress. Prague7 Ustav pro Vyzkum Rud; 1970.
[10] Overbeek JThG. Electrokinetic phenomena. Kruyt HR, editor. Colloid p. 251.
Science, vol. I. Amsterdam7 Elsevier; 1952. [49] Cases JM. Trans AIME 1970;247:123.
[11] Van Oss CJ. Sep Purif Methods 1975;4(1):167. [50] Fuerstenau DW. J Phys Chem 1956;60:981.
[12] Smith AL. In: Parfitt GD, editor. Dispersion of Powders in Liquids. [51] Somasundaran P, Healy TW, Fuerstenau DW. J Phys Chem
London7 Applied Science; 1973. 1964;68:3562.
[13] Dukhin SS, Derjaguin BV. Electrokinetic phenomena. Mistetsky A, [52] Wakamatsu T, Fuerstenau DW. Adv Chem Ser 1968;79:161.
Zimmerman M, editors. Surface and Colloid Sci, vol. 7. New York7 [53] deBruyn PL. Trans AIME 1955;202:291.
John Wiley & Sons; 1974. [54] Somasundaran P, Fuerstenau DW. Trans AIME 1968;214:102.
[14] Hunter RJ. Zeta Potential in Colloid Science—Principles and [55] Fuerstenau DW, Healy TW, Somasundaran P. Trans AIME
Applications. NewYork7 Academic Press; 1981. 1964;229:321.
[15] Shaw DJ. Electrophoresis. New York7 Academic Press; 1969. [56] Somasundaran P, Fuerstenau DW. J Phys Chem 1966;70:90.
[16] Ney P. Zeta Potentiale and Flotierbarkeit von Mineralien. Appl [57] Fuerstenau DW, Modi HJ. J Electrochem Soc 1959;106:336.
Mineral vol. 6. Wien, (NY)7 Springer Verlag; 1973 214 pp. [58] Fuerstenau MC, Palmer BR. Fuertenau MC, editor. Flotation, AM
[17] Hunter RJ. Foundations of Colloids Science, vol. I & II. Oxford7 Gaudin Memorial Volume, vol. 1. New York7 AIME; 1976. p. 148.
Clarendon Press; 1989. [59] Peck AS, Raby LH, Wadsworth ME. Trans AIME 1966;238:301.
[18] Fuerstenau DW. Pure Appl Chem 1970;24:135. [60] Pradip. Trans IIM 1987;40(4):287.
[19] Parks GA, deBruyn PL. J Phys Chem 1962;66:967. [61] Raghavan S, Fuerstenau DW. J Colloid Interface Sci 1975;50:319.
[20] Drzymala J, Lekki J, Laskowski J. Colloid Polym Sci 1979;257:768. [62] Natarajan R, Fuerstenau DW. Int J Miner Process 1983;11:139.
[21] Fuerstenau DW, Lai R, Ball B. Electronic Double Layer at Oxide/ [63] Han KN, Healy TW, Fuerstenau DW. J Colloid Interface Sci
Water Interfaces, presented at AIME Annual Meeting, Washington, 1973;44:407.
DC; 1969 (Feb.). [64] Pradip. Trans Indian Inst Met 1988;41(1):15.
[22] Sidorova MP, Lyklema J, Fridricksberg DA. Russ J Colloid Sci [65] Fuerstenau DW, Shibata J. Int J Miner Process 1999;57:205.
1976;38:716. [66] Ananthpadmanabhan KP, Somasundaran P, Healy TW. Trans AIME
[23] deBruyn PL. Trans AIME 1955;202:291. 1977;266:2003.
[24] Grahame DC. Chem Rev 1947;41:441. [67] Fuerstenau DW, Huang P. In: Lorenzen L, Bradshaw DJ, editors.
[25] Stigter D. J Phys Chem 1964;68:3600. Proceedings XXII Int. Mineral Processing Congress (Cape Town).
[26] Lyklema J, Overbeek JThG. J Colloid Sci 1961;16:501. S Afr Inst Min Metall, vol. 2. 2003. p. 1034.
[27] Boequet PE, Sliepcevich CM, Bohr DF. Ind Eng Chem 1956;48:197.
[28] Smith AL. J Colloid Interface Sci 1976;55:525.

You might also like