You are on page 1of 178

i

FLOTATION
PRINCIPLES AND APPLICATIONS

Janusz S. Laskowski
Professor Emeritus
NBK Institute of Mining Engineering
University of British Columbia

Lima
August 2012

ii

Flotation Principles and Applications


1. Basic characteristics of interfaces.
Wettability of solid surfaces.
Origin of electrical charge at solid/liquid interface
(a) Ionic solids (AgI, calcite, barite, fluorite, etc.)
(b) Oxides (quartz, rutile, hematite, alumina, etc.)
(c) Alumino-silicates (clay minerals)
(d) Sulfide minerals
(e) Soluble salts (sylvine, halite)
Effect of electrical charge on solid surface properties. Electrical double layer.
Electrokinetic potential (zeta potential); Examples of measurements.
Anisotropic minerals.
2. Surfactants
Surface activity. Gibbs adsorption isotherm. Critical micelle concentration. Adsorption
from solution. Adsorption of ionic surfactants onto charged surfaces. Effect of surfactants
on wettability of solid surfaces.
3.

4.

Flotation reagents.
Collectors, classification. Thio-collectors. Weak and strong electrolyte collectors.
Colloid chemistry of fatty acids and primary amines. Oily collectors. Modifiers:
activators, depressants, dispersants, pH regulators.
.
Flotation classification of minerals

5.

Flotation Plant Flowsheets.

(a) Sulfide ores (this chapter was prepared in collaboration with Professor Sergio
Castro, Universidad de Concepcion, Concepcion, Chile)
Theory of sulfide flotation. Classification of sulfide ores:
Cu ores, Cu-Mo ores, Cu-Pb-Zn ores, Cu-Ni ores.
Cu-Mo ores: bulk flotation, Cu-Mo selective
Separation methods. Chilean Cu-Mo processing plants. Processing of
Cu-Pb-Zn ores. Processing of Cu- Ni ores in Canada (Inco Flotation Process). Flotation
of Platinum Group Minerals (South Africa), Flotation of native gold and auriferous ores
(South Africa, Australia). Oil agglomeration in ore processing. Coal-gold oil
agglomeration.
(b)

Flotation of oxidized ores of base metals


Oxidized and mixed oxide-sulfide ores of non-ferrous metals.
Processing of Cu oxidized ores. Reagents. Leaching-precipitation-flotation technology.
Processing of lead oxidized ores. Processing of Zn oxidized ores. Recent improvements
in technology of cationic flotation of oxidized zinc ores.

iii

(c)

Oxides
Flotation of quartz and silicates. Flotation of iron ores, pelletization in the iron ore
processing technology. Selective flocculation/reverse flotation (Tilden Mine flowsheet).
Flotation concentrate pelletization (LKAB, Sweden).

(d)

Flotation of industrial minerals.


Problems in separation of these minerals. Flotation of phosphate ores.
Florida Cargo double float process. Flotation of lanthanum (Bastnasite) ores.

(e)

Flotation of soluble salts. Potash ores.


Separation of sylvite from halite in processing of Canadian sylvinite ores.
Use of insoluble long-chain primary amines as a collector. Blinders.
Typical flowsheets of Canadian plants.

INTRODUCTION
1. Separation methods in mineral processing. Fine particles in mineral processing. Flotation.

Figure A. Classification of separation methods.

iv

Flotation is the most important separation method in the group of the physicochemical
methods. It is based on differences in surface properties of treated minerals while the other
methods are based on differences in bulk properties such as density, magnetic susceptibility, etc.

Figue B. Separation of hydrophobic particles from hydrophilic particles in flotation. In


the flotation process hydrophobic particles are picked by bubbles while hydrophilic particles are
left in the pulp. The process is then based on differences in the surface properties of the mineral
particles.
The separation results strongly depend on particle size and all methods can handle only
particles of a given size range.

Grizzly

(a) Sizing

Vibrating Screens
Sieve Bend
Frequency Sieves
Classifying Cyclones
Dense Medium Vessel
Coarse Jig
Dense Medium Cyclones

(b) Cleaning

Water-Only Cyclone
Spirals
Teeter Bed
Froth Flotation
(c) Dewatering

Screens
High Frequency Screen

Vibratory Centrifuge
Screen-Scroll Centrifuge
Screen-Bowl
Centrifuge
Disc Vacuum Filter
0.01

0.1

10

100

Particle Diameter (mm)

Figure C. Optimal particle size ranges for different oparetions.


Unit operations in the mineral processing plant can broadly be divided into four distinct
groups: comminution (that also includes classification), separation (beneficiation, concentration),
dewatering of the beneficiation products, and water clarification (Figure D).

Figure D. Unit operations in mineral processing plant.


The separation processes can only yield satisfactory results if mineral particles in the feed
are liberated. This is achieved by size reduction of the run-of-mine rock down to the liberation
size. In the low grade ores - which are now mined - the valuable minerals are disseminated in the

vi

form of very fine grains and thus to achieve liberation the rock must by crushed and ground
down to this liberation size (Fig. E).

Figure E. Liberation size of the grains of valuable mineral.


Importance of the flotation among all other methods results from the fact that this is the
only process in which fine liberated particles (after crushing and grinding) can be separated from
fine gangue particles with a satisfactory efficiency.
.

4
1. Basic Charecteritics of Interfaces
1. 1. Work of adhesion and work of cohesion
The work of adhesion between two phases (for example, between two immiscible
liquids, or between solid and liquid) is equal to the reversible work required

Fig. 1.1
to separate unit area of the liquid-liquid interface (for a two phase liquid-liquid system)
and form two separate liquid air interfaces (see Fig. 1.1) and is given by the Dupre
equation
Wa = a + b ab

(1.1)

where a and b are the surface tensions of phases a and b, respectively, and ab is the
interfacial tension a-b.
The work of cohesion for a single liquid corresponds to the work required to pull
apart a column of liquid of unit cross-sectional area and is given by
Wc = 2 a

(1.2)

1.2. Wettability
In the case of a solid wetted by a liquid, the wettability is determined by the solidliquid work of adhesion (W SL ) and the work of cohesion of liquid (W LL ). The classical
boundary condition for the hydrophilic-hydrophobic transition is equality of W SL and
W LL . According to Dupres equation (Eq. 1.1), for a solid-liquid system W SL is given by
WSL = SV + LV SL
and

(1.3)

5
WLL = 2 LV

(1.4)

Since there are no experimental methods to determine the surface tension of a


solid ( SV and SL ) equation 1.3 can only be solved with the use of Youngs equation
(Fig. 1.2)

SV = SL + LV cos

(1.5)

Fig. 1.2. Sessile drop on a smooth solid surface.


By combining Eqs. 1.3. and 1.5 one can get
WSL = LV (cos + 1)

(1.6)

Since both liquid surface tension ( LV ) and the contact angle () can experimentally be
determined, equation 1.6 is often used.
Equations 1.4 and 1.6 give

WSL LV (1 + cos )
=
2 LV
WLL

(1.7)

and

WSL
1
(1.8)
WLL
The condition of hydrophobicity follows from Eq. 1.8; only for WSL < WLL
cos = 2

0.
Fig. 1.3 depicts a sessile droplet resting on a flat solid surface; as this figure
shows, the contact angle at the solid/liquid interface is determined by the values of the

6
work of adhesion of the liquid to the solid, and the work of cohesion of the liquid.
The work of adhesion of liquid to solid (W SL ) can be split into several
contributions. For water molecular interacting with a solid, the most important are: van
der Waals dispersion forces contribution (W SL d), hydrogen bond contribution (W SL h) and,
if the solid has an electrical charge, an electrical forces contribution (W SL e). Then

WSL = WSLd + WSLh + WSLe

(1.9)

For the solid which does no have any polar groups and, therefore, does not interact with
water molecules through hydrogen bonding ( WSLh 0 ), and does not have an electrical
charge so that WSLe 0 , substitution of Eq 1.9 to Eq. 1.8 gives
WSLd
cos 2
1
WLL

(1.10)

Fig. 1.4. Graphite and molybdenite crystal lattices.


Using this equation, Laskowski and Kitchener (1969) showed that because of
exceptionally high work of cohesion of water ( WLL = 145 mJ/m2) for all solids
WSLd < WLL ; therefore, all solids would be hydrophobic ( > 0) if they interacted with
water only through dispersion forces. This explains why some minerals (such as graphite,
molybdenite, sulfur, talc) are hydrophobic by nature. This is perfectly in line with Gaudin
who pointed out that all hydrophobic solids present non-polar molecular groups whereas
hydrophilic solids have ionic or dipolar groups capable of undergoing hydration.
By writing Eq. 1.8 in the form
cos = 2

WSLd + WSLh + WSLe


1
WLL

(1.11)

7
it becomes obvious that solid surface properties are strongly affected by the electrical
surface charge.

1.3. Origin of electrical charge at solid/liquid interface


(A). Ionic Solids. In the case of ionic solids such as AgI, BaSO 4 , CaF 2 etc. (soluble salts
also belong to this group but will be discussed separately), the surface charge arises from
the transport across the interface of ions constituting the lattice. Those particular ions that
are free to pass between both phases and therefore establish the electrical charge are
called potential-determining ions (PDI). For AgI, the potential-determining ions are Ag+
and I-, for a solid like calcite, CaCO 3 , the potential-determining ions are Ca2+ and CO 3 2but since their concentration also depends on pH, the potential-determining ions for this
mineral are Ca 2+, CO 3 2- and also H+, OH- and HCO 3 -.

Fig. 1.16. AgI/aqueous solution interface


The surface charge of a solid is determined by the concentration of potential-determining
ions on the solid surface (or in other words is determined by adsorption density of
potential-determining ions). In case of AgI surface, surface charge is given by

s = F (Ag I )
+

(1.25)

where Ag+ is the adsorption density in mol/cm2 of Ag+ ions, and I- is that of I- ion, and
F is the Faraday constant.
AgI is practically insoluble in water. The solubility product of AgI
K SP = [ Ag + ][ I ] = 10 16

(1.26)

It was experimentally determined that in the system in which AgI is in equlibrium


with aqueous solution, the AgI/water interface does not have electrical charge at Ag+ ions
concentration [Ag+] = 10-5.5 mol/dm3. In analogy with pH (pH = -log [H+]), such
concentrations are expressed as pAg+ = 5.5. (pAg+ = - log [Ag+]).
The previous equations is commonly written

pAg+ + pI- = 16

(1.27)

(in the same way for H+ and OH- ions in water, pH + pOH = 14)
So, the point-of-zero-charge (p.z.c.) conditions for AgI crystals, which are in
equilibrium with aqueous solution, are reached when the concentrations of potentialdetermining ions are: pAg+ = 5.5 (and pI- = 10.5, since pAg+ + pI- = 16). In other words,
at p.z.c. the concentration of [Ag+] = 10-5.5. mol/dm3 is 105 times higher than the
concentration of I- ions (since at pzc pI- = 10-10.5 mol/dm3).
For ionic solids the surface potential, o , can be calculated from the equation

o =

RT C M +
ln
zF C Mo +

(1.28)

For the case of AgI this equation at room temperature gives


o = 0.059( pAg + o pAg + ) = 0.059( pI pI o ) Volt

(1.29)

so to sum it up, particles of AgI in aqueous solutions are not electrically charged only
when [Ag+] = 10-5.5 (thus when [I-] = 10-10.5). This shows that the concentration of silver
ions in the solution must be 105 times higher than the concentration of iodide ions for the
AgI particle not to have an electrical charge. This results from much higher hydration
energy of Ag+ ions in comparison with that of I- ions.

Fig. 1.17. Surface potential at silver iodide/aqueous solution

9
Interface.

Table 1.1. The point-of-zero-charge of some ionic solids


(after D.W. Fuerstenau, Principles of Flotation, Johannesburg, 1982, pp. 17-30)

Material
p.z.c.
Barite, BaSO 4
pBa 6.7
Calcite, CaCO 3
pH 9.5*
Fluoroapatite, Ca 5 (PO 4 ) 3 (F,OH)
pH 6*
Fluorite, CaF 2
pCa 3
Hydroxyapatite, Ca 5 (PO 4 ) 3 OH
pH 7
Scheelite, CaWO 4
pCa 4.8
Silver chloride, AgCl
pAg 4
Silver iodide, AgI
pAg 5.6
Silver sulfide, AgS
pAg 10.2
* from the hydrolysis equilibria and solubility data.
(B). Simple oxides (e.g. SiO 2 , SnO 2 , TiO 2 , Fe 2 O 3 , Al 2 O 3 , etc.). Quartz is probably the
most researched mineral, and one of the least impure. Crystallographically, it is built by
three-dimensional network of alternating Si4+ and O2- ions. The Si-O bond is said to have
about 50% covalent character and is certainly strong. Neither of the normal ions can exist
as such in water; quartz has an appreciable solubility amounting to about 10 mg/dm3 at
25 oC, resulting from the reversible reaction SiO 2 + 2H 2 O = Si(OH) 4aq . As the
orthosilicic acid is a weak acid, rise of pH displaces the reaction to the right as silicate
ions are formed and so increases the solubility.
The surface of quartz combines with water, which can be removed by heating
above about 400 oC. Infra-red spectroscopy shows that siloxane groups, -SiOSi- are
converted by water into silanol groups, -SiOH, which are weakly acidic This is why
quartz in water shows a negative zeta potential, reaching 120 mV in very dilute alkali
and falling to nearly zero at about pH 2. The hydroxylated quartz is strongly hydrophilic,
showing zero contact angle and a thick equilibrium film due to the electrical double layer;
but when the surface is dehydroxylated by strong heating it becomes definitely
hydrophobic.
Oxides are amphoteric and their surfaces after hydroxylation show a positive
charged at low pH and a negative at high pH, with a point-of-zero-charge (p.z.c.)
somewhere between.
It is customary to ascribe the origin of the electrical charge at the oxide
surface/aqueous interface to protonation/deprotonation of the surfaced hydroxyls:
-MOH + H+ = -MOH 2 +
(1.30)
-

-MOH + OH = -MO + H 2 O
and at p.z.c.

10

[-MOH 2 +] = [-MO-]

(1.31)

As these reactions reveal, H+ and OH- ions are potential-determining ions for
oxides. Concentration of these ions is measured as pH.
In practice, it is common to characterize the electrical charge of solid particles
indirectly by measuring so-called electrokinetic potential (also know as zeta potential,
since Greek letter, (zeta), is internationally used as symbol for electrokinetic potential).
Next figure shows the zeta potential pH curves for various oxides. The pH
values at which zeta potential is equal zero (=0) is called iso-electric-point (i.e.p.). It is
usually identital (or close) to the p.z.c. As this figure shows why quartz particles in water
are practically always negatively charged, iron oxide particles are positively charged at
pH<6.5 and negatively charged at pH > 6.5. For alumina the i.e.p. 9. This also indicates
that while silica is more acidic alumina is more basic. It also explains, why for some
minerals (such as oxides) pH affects so strongly their flotation properties.

Fig. 1.18
Electrokinetic
potential (zeta
potential) of
oxides vs. pH at
constant ionic
strength.

C). Alumino-silicates (kaolinite, illite, montmorillonite, etc.).

Fig. 1.19. Idealized sketch of the atomic structure of a three-layered clay mineral.
The nature of the clay-water interface is an illustration of the influence of crystal
structure on the surface properties of solids. The basic building units of clay minerals are

11
the sheets formed by linking together Si-O tetrahedral and the two dimensional arrays of
octahedral formed by the sixfold coordination of Al3+ and
Mg2+ with oxygen and hydroxyl groups (see Fig. 1.19). This group has been considerably
studied because of the importance of clays. Their special feature is their laminar crystal
structure; the exposed basal planes differ in topochemistry from the edges of the
crystals. In case of montmorillonite the basal planes behave more-or-

Fig. 1.20. The edge of montmorillonite platelet.


Less as simple silicate (because the hydrated alumina layer is sandwiched between a pair
of silica layers) while alumina is exposed at the edges. As a result of isomorphous
substitution of some silicon (Si4+) for aluminum (Al3+), the face surface carries a
negative electrical charge while =Al(OH) groups at the edges are either positively or
negatively charged depending on pH. In neutral media the faces are negatively charged
but the alumina sites on the edges are positively charged and this is the origin of
coagulation of clays in distilled water. In beneficiation of China clays, a kaolin clay is
first dispersed in alkaline pH, larger impurity particles (quartz, hematite, anatase, etc.)
are removed by classification, and the fine clay particles which constitute the final
product are thickened and filtered after adjusting pH down to acidic range.
(D). Sulfide minerals. Minerals such as galena (PbS), sphalerite (ZnS), chalcocite
(Cu 2 S), chalcopyrite (CuFeS 2 ), etc. are important sources of metals and their surface
chemistry is of vast interest in connection with the froth flotation process.
In principle, such sulfides are in reversible equilibrium with aqueous metal and
sulfide ions, according to a definite solubility products (K SP ) which are extremely low
(for instance, for galena, K SP = 10-28). But in the presence of dissolved oxygen galena
undergoes superficial oxidation with various oxidation products (lead hydroxide, lead
thiosulfate, lead sulfate, ect) deposited on the surface. As sulfides are electronic semiconductors, their oxidation processes take on the character of electrolytic corrosion
reactions, certain (anodic) areas oxidize preferentially while other are cathodic, as in the
rusting of iron.
Anodic:
Cathodic:
Overall reaction:

MeS = Me2+ + So + 2e
1/2O 2 + H 2 O +2e = 2OH-

(1.32)
(1.33)

MeS + 1/2O 2 + H 2 O = Me 2+ + So + 2OH-

(1.34)

12

Presence of various oxidation products (So oxidizes further into S 2 O 3 2-, SO 4 2- , for
example Fe2+ to Fe3+, etc.) on the surface makes identification of PDI ions very difficult.
Oxidation of sulfide minerals can be manipulated by controlling the redox potential (Eh)
of the pulp during processing (grinding and flotation).

Fig. 1.21. Schematic corrosion reactions on sulfide surface.


(E). Soluble salts (e.g. KCl, NaCl, etc.). The lattice ion hydration theory explains the
surface charge of the soluble minerals as in the case of Group A minerals. These minerals
are discussed here separately owing to their industrial importance. Because of their
solubility in water there was no experimental method to measure experimentally the
electrical charge of such crystals in aqueous media, and the theoretical predictions could
only recently be verified by experiment.
Contrary to general belief that solid particles in concentrated electrolyte solutions
cannot be electrically charged, in 1968, R.J.Roman, M.C. Fuerstenau and D.C. Seidel
postulated that sylvite and halite particles carry opposite electrical charge in brine. It was
postulated that while KCl crystals were negatively charged NaCl crystals were charged
positively. The hypothesis was based on the observation that while fine KCl suspensions
in KCl brine and fine NaCl particle suspensions in NaCl brine were very stable, the fine
particles of KCl and NaCl in KCl-NaCl brine strongly aggregated. Yalamanchili and
Miller (1992) studied aggregation in various systems and confirmed these observations.
They also found strong aggregation between fine quartz and NaCl particles in NaCl brine.
The observations suggested that these particles interacted in brine only because of their
electrical charge. The electrical charge of the particles of water-soluble minerals in
aqueous systems has been recently directly measured (J.D. Miller and M.R.
Yalamanmchili, 1992). This was accomplished with the use of the Laser-Doppler
Electrophoresis. In this method the movement of the salt particles (dissolving very
quickly in water) is measured over a very short time (15 sec) in an applied electrical field.
Varying ionic strength does not allow the calculation of the zeta potential values, but the
measured electrophoretic mobility of the particles allows determination of the electrical
charge of such particles. Accompanying figure shows some results (Fig. 1.22).

13

Fig. 1.22. Electrophoretric measurements for sylvite and halite particles [Miller &
Yalamnchili, Langmuir, 8, 1464 (1992)].
According to the lattice ion hydration theory, the sign of the surface charge is
determined by the magnitude of the hydration energy of the respective surface lattice
ions. If the surface cation has a more negative hydration energy than the surface anion,
that is if the escaping tendency of the cation is larger then that of the anion, the surface
acquires a negative charge. The converse is also true.

Fig. 1.23. Dissolution of ionic solid.

According to the lattice ion hydration theory, the sign of the surface charge is
determined by the magnitude of the hydration energy of the respective surface lattice
ions. If the surface cation has a more negative hydration energy than the surface anion,
the cations will have a greater tendency to undergo hydration and be preferentially

14
released to solution; the surface will thus acquire a negative charge. The converse is also
true.

Salts

Sign of Surface Charge


Predicted Experim.

LiF
NaF
KF
RbF
CsF

Negative free energy


of hydration, kcal/mol
Cation
Anion
112.0
109.6
88.4
109.6
71.1
109.6
65.9
109.6
58.2
109.6

+
+
+
+

+
+
+
+
+

LiCl
NaCl
KCl
KCl*
RbCl
CsCl

112.0
88.4
71.1
71.1
65.9
58.2

82.5
82.5
82.5
82.5
82.5
82.5

+
+
+
+

+
+
+
+

LiBr
NaBr
KBr
RbBr
CsBr

112.0
88.4
71.1
65.9
58.2

75.7
75.7
75.7
75.7
75.7

+
+
+

Table 1.2 Sign of the surface charge for selected alkali halides (Miller & Yalamanchili,
1994). *Laser quality KCl.

1.4. Effect of electrical charge on solid surface properties (on wettability)

Reversible Electrode
Silver iodide/water interface constitutes perfect reversible electrode (Fig. 1.16).
The surface electrical charge of such an electrode results from differences in adsorption
of the potential determining ions (in this case Ag+ and I-) and is given by Eq. 1.25. The
surface potential, o , is given by Eq. 1.28 in which a M+ and a M+ o, are the activities of
Ag+ ions in solution, and in solution at the point of zero charge, respectively (Fig. 1.17).
From Eq. 1.28
do = 2.3

RT
dpAg +
F

(1.35)

15

Fig. 1.24. Contact angle vs. pAg+ (potential determining ions) on silver iodide (after R.
Ottewill et al.).
and from Lippmanns equation:
d = do

(1.36)

The substitution of equations 1.25 and 1.35 into 1.36 gives:


d = 2.3RT (Ag + I )dpAg +

(1.37)

Thus as shown by Ottewil et al in 1978, the curve = f (o ) = f ( pAg +) indicates


capillary maximum at:
d
= 2.3RT (Ag + I ) = 0
dpAg +
that is at Ag + = I which is the p.z.c.

(1.38)

16
Ottewill et als experimental results (Fig. 1.24), which illustrate the effect of
elecrtical charge on the wettability of the AgI surface in aqueous solution confirm
entirely this theoretical anaylsis. Since the p.z.c. for this system is around pAg+ = 5.5, the
agreement with experiment is very good indeed.
Thus, the experiments carried out with both polarizable and reversible electrodes
confirm that the solid surface wettability is a maximum at the point of zero charge.

Fig. 1.25. Flotation rate constant in 0.5 M NaCl solutions for three coals vs. the zeta
potential of the coal.
Fig. 1.25 shows the results of so-called salt flotation of three different coals in 0.5
M NaCl aqueous solution (Fuerstenau, Rosenbaum and Laskowski, Coll.Surf., 8, 153
(1983). The flotation tests and the zeta-potential measurements were carried out varying
pH, and the flotation results (the flotation tests were carried out in 0.5M NaCl) are
plotted versus the zeta potential values. At a high electrolyte concentration (0.5 M NaCl)
the double layer is completely compressed and a good agreement between the contact
angle and rate of flotation can be expected under such conditions. The results clearly
show that in these experiments the maximum flotation rate constant was indeed observed
when the coal surface did not carry any electrical charge.

17

1.5. Electrical double layer


Due to the reasons discussed in the preceding section, solid particles in water are almost
always electrically charged. Some ions will be adsorbed on such surfaces, some will be
distributed according to the influence of electrical forces and thermal motion. The former
form an inner region of the electrical double layer (adsorbed layer), the latter a diffused
region. This situation is shown below.

(1.39)

Fig. 1.26. Electrical double layer without specific adsorption of ions (A). Graeme model
of the electrical double layer (in the presence of specifically adsorbing ions) (B).

18

Fig. 1.27. Schematic representation of a diffuse electric double layer.


Distirbution of the electrical potential at the solid/liquid interface (for small
potentials) is given by:
= o exp[ x ]

(1.40)

This relationship shows that at low potentials the potential decreases


exponentially with distance from the charged surface. Please note that at a distance 1 /
the potential drops by a factor of 1/e (e ~2.7). This distance is used as a measure of the

Fig. 1.28. Fraction of double layer potential versus distance from a surface according to
the Debye-Huckel approximation; (a) curves drawn for 1:1 electrolyte at three
concentrations; (b) curves drawn for 0.001 M symmetrical electrolytes of three different
valence types.

19
extension of the double layer ( thickness of the double layer). According to the theoretical
equation it has the value

1 / = [ kT / 2 e 2 N A cz 2 ]1/ 2
For an aqueous solution of a symmetrical electrolyte at 25 oC
cz 2
m-1
= 0. 328 1010 (
)1/ 2
3
mol dm
or
= 3.288 I nm-1

(1.41)

(1.42)
(1.43)

where I is the ionic strength.


Of particular importance in colloid science is the fact that the thickness of the
double layer depends markedly on the ionic concentration as shown below. In aqueous
solution of a 1:1 electrolyte at 25 oC the values of 1 / are: at 10-4 mol/litre 30.4 nm, at
10-3 mol/litre 9.6 nm, at 10-2 mol/litre 3.0 nm.
1.6. Electrokinetic potential (zeta potential) and electrokinetic measurements
Electrokinetic phenomena occur when an electrically charged phase moves with
respect to an adjoining phase. Fig. 1.29 shows distribution of electrical potential in the
double layer surrounding a spherical charged particle. When this particle sediments under
gravity the plane of slip between the moving particle and a stationary liquid is involved;
this slipping plane is depicted as shear plane and is located somewhere between the
adsorbed (Stern) layer which moves with the solid particle and a diffuse part of the
double layer. The exact location of the shear plane is unknown.

20
Fig. 1.29. Distribution of electrical potential () in the double layer region surrounding a
charged particle, showing the position of the Stern potential ( ), zeta potential ( ) and
reciprocal Debye length (1/).
It is generally assumed that the shear plane is located at a small distance further
out from the surface than the Stern plane and that (zeta potential) is marginally smaller
in magnitude than . In colloid chemistry it is customary to assume that = .While
for some solids the surface potential can be calculated using known equations (see
equations 2.28 and 2.29), for many solids it can only be estimated and the experimentally
accessible values of the zeta potential are extremely helpful in such estimations.
Of four electrokinetic phenomena (electroosmosis, electrophoresis, streaming
potential, sedimentation potential) by far the most commonly used to determine the zeta
potential is electrophoresis. This is a study of the movement of charged particles
dispersed in a medium containing electrolyte under the influence of an electric field. In
most cases the microelectrophoretic technique is used in which the movement of particles
in a dilute dispersion is followed microscopically. A number of commercial instruments
are available, some of which involve direct measurements of the electrophoretic mobility,
and others of which are designed to reduce the tedium by using sophisticated optics and
electronics to give average electrophorertic mobility and even mobility distribution.
The electrophoresis data are commonly expressed in terms of the particle
electrophoretic mobility (u E ) i.e., the velocity divided by the potential gradient across the
cell and quoted in units of s s 1 / volt cm 1 . According to Henry, the relation between
electrophoretic mobility and zeta potential depends on the particle radius (a) and the
thickness of the double layer surrounding the particle ( 1 / ) expressed in the product a ,
as given by:
uE =

f (a )
1.5

(1.44)

where is the zeta potential of solid particles, is the permitivity of the medium in
which these particles are immersed (the dielectric constant of a material is equal to the
ratio between its permitivity and the permitivity of a vacuum), and is the viscosity of
the medium.
For very small particles in dilute solution where the thickness of the double
layer( 1 / ) is large, a << 1 and the correction factor f (a ) = 1 (Huckel equation). For
large particles in more concentrated solution and a >> 1 and f (a ) = 1.5
(Smoluchowski equation).
Electrokinetic measurements should be made at constant ionic strength. For
example, in a study of the effect of pH on the electrophoretic mobility of particles whose
surface charge is determined by H+ and OH- ions, both the surface and zeta potential

21
change with pH. But the rate of decay of potential from the surface is also a function of
the ionic strength (Eq. 1.40); hence the zeta potential is dependent on both o and . To
measure the true effect of pH on zeta potential requires that be maintained constant
and this is achieved by conducting the measurements in a solution of indifferent
electrolyte. For example, if this is a symmetrical electrolyte (let say with ions of valency
1, such as KCl or KNO 3 , then since in Eq.1.43
1
(1.45)
I = c i z i
2
Where c i is the concentration of ion i in mol dm 3 , and z is the ion valency, then for the
case of KCl
1
(1.46)
I = {[ K + ](1) 2 + [Cl 1 ](1) 2 }
2
Since [ K + ] = [Cl 1 ] = c , the concentration of the salt, one obtains that for such a case
I = c. So, if a 1-1 symmetrical salt is used as a supporting electrolyte in electrokinetic
experiments then from Eq. 1.43 one calculates that at a concentration of 10-3 mol dm 3
has a value of approximately 0.1 nm-1. Hence, for particles of radius 1 m,
a 100 . Increasing the electrolyte concentration to 10-1 mol dm 3 increases
a 1000 . Thus for such cases the Smoluchowski equation is applicable to calculate the
zeta potential values from the experimentally determined electrophoretic mobility.

Fig. 1.30. Effect of concentration of potential-determining ions on zeta potential at three


concentrations of indifferent electrolyte (example shown is for goethite with KNO 3 used
as indifferent electrolyte) .
Smoluchowskis equation, depending on the units used, has the following forms:

22

10 3
uE
7.81x10 7

(u E is in ms-1 per Vm-1)

(1.47)

= 1.28 x10 3 u E

(u E is in ms-1 per Vm-1)

(1.48)

= 12.8 xu E

(u E is in ms-1 per Vcm-1)V

(1.49)

The figure below shows, after D.W. Fuerstenau, the zeta potential of Al2O3 at pH
6.5 and 10 as a function of concentration of various salts. As seen, Ba2+ adsorbs
specifically onto negativelly charged alumina, while SO42- ions adsorb specifically onto
positively charged alumina.

Fig. 1.31. Zeta potential of alumina at pH 6.5 and pH 10 as a function of NaCl, Na 2 SO 4 ,


and BaCl 2 concentrations (Fuerstenau, 1970).
The zeta potential of goethite plotted vs. pH at various concentrations of NaCl,
and flotation of goethite are shown after Iwasaki et al., 1960.

Fig. 1.32. The dependence of the flotation of goethite on surface charge. The upper
curves show zeta potential as a function of pH at different concentrations of NaCl,
indicating the i.e.p. to be at pH 6.7. The lower curves are the flotation recoveries in 10-3

23
M solutions of dodecylammonium chloride, sodium dodecyl sulfate and sodium docedyl
sulfonate (Iwasaki et al., 1960).

1.7. Examples of electrokinetic measurements in testing flotation


The zeta potential measurements have been heavily utilized in the studies
of the surface and flotation properties of industrial minerals (apatite, calcite, dolomite,
barite, etc). These minerals are difficult to separate because they have similar surface
properties especially in the presence of significant concentrations of dissolved species in
the flotation pulp.

10

11

12

40

40

Potential(mV)

Calcite
Calcite+Water glass
Calcite+ferric silicate hydrosol
20

20

-20

-20

10-2 mol/L KCl

-40

-40

Water glass/ferric silicate dydrosol: 0.375g/L


-60

-60
6

10

11

12

pH
Figure 6. Effect of water glass and ferric silicate hydrosol
on zeta potential of calcite.

Figures 6 and 7 (Figures 1.33 and 1.34), taken from our paper with K. Ding
(Can.Metllurg.Quart., vol. 45, 199-206 (2006), show the effect of water glass (sodium
silicate) on the zeta potential of calcite and dolomite. These measurements were utilized
in this project to study the effect of water glass on surface properties of these two
minerals.
As already indicated, the problems in selective flotation of various industrial
minerals mostly result from their semi-solubility and high concentration of various ions
in such flotation pulps. This was quite extensively studied with the use of electrokinetic
measurements.

24

10

11

12

40

40

Potential (mV)

Dolomite
Dolomite+Water glass
Dolomite+ferric silicate hydrosol
20

20

-20

-20

-40

-40

10-2 mol/L KCl

Water glass/ferric silicate hydrosol: 0.375 g/L


-60

-60
6

10

11

12

pH
Figure 7. Effect of water glass and ferric silicate hydrosol
on zeta potential of dolomite.

As these results demonstrate the zeta potential pH curves for calcite and dolomite are
very similar, and they are both affected by water glass. Water glass was shown by M.C.
Fuerstenau at al [Trans. SME, vol. 241, 319-323 (1968)] to depress very strongly calcite
flotation with oleic acid. As the two figures above indicate the zeta potential curves for
both calcite and dolomite in water glass solutions become very similar to the curves for
quartz (and other silicates). Depression of the anionic flotation of these two minerals by
water glass is then quite understandable. This also indicates that the flotation of these two
minerals with cationic collectors (amines) should be activated by water glass and this was
confirmed (Ding & Laskowski, CMQ, 45, 199-206 (2006).
The following examples, showing results of zeta potential measurements with
calcite and apatite, were taken from the paper by Amankonan and Somasundaran
[Colloids & Surfaces, 15, 335-353 (1985)].

25

Fi. 1.35. Zeta potential vs. pH curve for apatite in water and in calcite supernatant.

Fig. 1.36. Zeta potential vs. pH curve for calcite in water and in apatite
supernatant.

26

Fig. 1.37. Zeta potential pH curves for apatite and calcite in the mixed calciteapatite supernatant (1:1 mixture of these two minerals on weight basis).

Using thermodynamic data, Pradip and Fuerstenau [IJMP, 32, 1-22(1991)]


calculated that barite in solutions of 10-3 M NaCO 3 , at about pH 9 and above, should be
coated by BaCO 3 and via electrokinetic measurements demonstrated that this really
happens.

Fig. 1.38. The electrophoretic mobility of barite and barium carbonate as a


function of pH in sodium carbonate solutions (Pradip and Fuerstenau, 1991).

27

Fig. 1.39. The electrophoretic mobility of bastnaesite, calcite and barite as a


function of pH at 1.0 mM concentration of sodium carbonate (Pradip & Fuerstenau,
1991).
Pradip and Fuerstenau argue that since both gangue minerals (in the Montain Pass
ore), calcite and barite, are positively charged which favors their anionic flotation (with
fatty acids), both sodium carbonate and lignin sulfonate are necessary to depress their
flotation.
1.8. Anisotropic minerals

28
Manuscript in press (in Canadian Mining Journal)

ANISOTROPIC MINERALS IN FLOTATION CIRCUITS*


J. S. Laskowski
Norman B. Keevil Institute of Mining Engineering
The University of British Columbia
Vancouver, B.C., Canada
jsl@apsc.ubc.ca

This paper was presented at the 8th UBC-McGill-UA Symposium on


the Fundamentals of Mineral Processing, Metallurgical Society of CIM,
Vancouver, October 3-6, 2010.

29

ANISOTROPIC MINERALS IN FLOTATION CIRCUITS


ABSTRACT
Anisotropic minerals are important constituents of many ores. This group includes
both valuable minerals (e.g. molybdenite in Cu-Mo ores) as well as gangue minerals
(e.g. talc in platinum bearing sulfide ores in South Africa, graphite in Cu-Ni sulfide ores
in Canada, chrysotile in Ni sulfide ores in Australia, clay minerals in all types of ores,
etc.). Aqueous suspensions of anisotropic minerals exhibit quite different properties from
the properties of the suspensions of isotropic minerals. The presence of anisotropic
minerals in flotation circuits affects flotation process.
KEYWORDS
Anisotropic minerals, clays, talc, molybdenite, flotation, pulp properties, rheology

INTRODUCTION
Mineral crystallo-chemistry as shown by Gaudin et al (1957) - is responsible
for the properties of the solid/liquid interface which is determined by the chemical
composition of the solid and the electrical charge of the solid surface.
Isotropic minerals.
In the case of isotropic minerals, all sides of the crystal are created by breaking the same
bonds with resulting mineral surfaces being homogeneous and having identical electrical
charge, an example being quartz. The new surfaces are created by breaking identical SiO bonds when larger pieces of quartz are crushed. As a result all the new surfaces have
the same composition. This surface hydroxylates to form surface silanol groups (SiOH)
which ionize as in silicic acid . In general, such surface hydroxyls are amphoteric and
become positively charged in acids and negatively charged in alkalis. The charge results
from the following reactions:
-MOH + H+ = -MOH 2 +

(1)

-MOH + OH- = -MO- + H 2 O

(2)

with M standing for metal. As a result, the surface acquires either positive or negative net
electrical charge. The switch-over point (p.z.c., point of zero-charge) varies from one
oxide to another (corresponding to their relative strengths as acid or base). At the p.z.c.
the concentrations of negative and positive charges are identical:
[-MOH 2 +] = [-MO-]

(3)

30
As these reactions reveal, concentration of H+ and OH- ions determines the charge
(and potential at the interface) of oxides and these ions are referred to as potentialdetermining for these minerals.
In all studies on particle-particle interactions in which the effect of attractive
(dispersion) forces and repulsive (electrostatic) forces are discussed (DLVO theory)
isotropic solid particles are used which retain a uniform charge independent of the
distance between the two interacting particles. For two fine quartz particles suspended in
water this is shown in Fig. 1.

Figure 1 - Electrical repulsion between two negatively charged identical particles


suspended in water.

The electrical repulsive forces are calculated based on zeta-potential


measurements carried out in most cases with the use of electrophoresis. In this
experiment, the electrophoretic mobility of the solid particles (Fig. 2) is measured in an
electrical field and this is used to calculate the zeta potential via Smoluchowskis
equation.

Figure 2 - Definition sketch for electrophoresis showing a negatively charged spherical


particle moving in electrical field.
Driving electrical force, E q, where E is the electrical field and q is the electrical
charge of the particles, is opposed by a hydrodynamic friction given by a viscous drag
(given by Stokes law); and electrophoretic friction caused by the oppositely charged ions
moving in the direction opposed to that of the particle. The resulting Somulchowskie
equation allows calculation of the zeta-potential from the measured electrophoretic
mobility of particles a few microns in size.

31
The point is that the spherical isotropic particles are considered in this type of
derivations and the question then arises what would be the result of the electrophoretic
experiment if the particles were anisotropic and had different electrical charge on various
sides as shown in Figure 3. This question has not yet been satisfactorily answered.

Figure 3 - Schematic of the anisotropic particle in electrical field.

Anisotropic minerals
In the case of anisotropic particles the surface charge on different sides of the
crystal is different (as in the case of clays). The definition adopted by Chander at al.
(1975) is as follows: An anisotropic surface consists of two broad types one, which is
formed by the rupture of van der Waals bonds and is hydrophobic and the other, which is
formed by rupture of ionic or covalent bonds and is hydrophilic. The definition used
here is more general and does not necessarily require that one of the crystal sides be
hydrophobic. Examples of this are clay minerals which are typically anisotropic and are
usually hydrophilic.
Clays are the best known examples of anisotropic minerals. However
molybdenite, graphite and talc are also anisotropic minerals and are inherently
hydrophobic. While clays on one side, and molybdenite, graphite and talc on the other,
are very different what they have in common is a sheet-structure (which is also referred
to as laminar crystal structure).
The sheet-structure of clay minerals discussed in this chapter are made up of
layers of silica tetrahedra condensed with gibbsite [Al(OH) 3 ] (kaolinite) or brucite
[Mg(OH) 2 } (talc). Kaolinite, a 1:1 layer silicate, has two types of basal planes, the
tetrahedral Si-O (bottom in Fig. 4) plane and the octahedral Al-OH (upper) plane (Carty,
1999). The 1:1 layers are held together in the crystal by hydrogen bonds. The bottom
tetrahedral plane carries negative electrical charge at all pH values as a result of
isomorphous substitution of some Si4+ by Al3+ The octahedral basal plane of kaolinite, as

32
well as its edges, carries a charge that depends on solution pH (the charge arises from the
presence of amphoteric =AlOH groups on these surfaces). Thus, the topochemistry of
the exposed planes differ quite significantly.

Figure 4 - Structural basic unit cell of kaolinite (1:1 layer alumino-silicate).

Figure 5 - The edges of montmorillonite platelet.


In montmorillonite, a 2:1 layer silicate, the hydrated alumina layer is sandwiched
between a pair of silica layers, and alumina is exposed only at the edges. Thus, for this
mineral while the planes always carry negative electrical charge, only the edges,
depending on pH, may carry either positive or negative charge (Figure 5).
The basic structural unit of talc is shown in Figure 6.

Figure 6 - Basic structural unit of talc (2:1 layer magnesium silicate).

33

As shown by Burdukova et al. (2007), because of the substitution of some Si4+


ions with Al3+ and Ti3+ in talc tetrahedral layers these basal planes exhibit negative
electrical charge. However, the charge at the edges depends on pH. The most important
difference between talc and the clays is that it is inherently hydrophobic. In talc the layers
of silica tetrahedra are held together by van der Waals bonds and the breaking process
proceeds by rupturing these weak bonds. This surface of talc can then interact with water
only through dispersion forces which makes it hydrophobic (Laskowski & Kitchener,
1969). According to Gaudin et al. (1957), native hydrophobicity results when at least
some fracture or cleavage surfaces form by rupture of weak secondary bonds. However
the edges are created by rupture of covalent bonds and such sites spontaneously react
with water to form hydrophilic M-OH sites.
The curled tubular structure of chrysotile is much more complex than the simple
edge/plate structure of clays. In chrysotile [Mg 3 Si 2 O 5 (OH) 4 ], the dimensions of the
silicate tetrahedral layer are about 9% smaller than the corresponding ones in the
octahedral brucite layer. This is illustrated in Figure 7. The imperfect fit of octahedral
layers and the tetrahedral layers causes the crystal structure to bend, curl and form
concentric hollow cylinders. The bending of the sheets is continuous and results in tubes
that give the mineral its fibrous nature (asbestos mineral).

Figure 7 - (A) The curved morphology of chrysotile (Klein & Hurlbut, 1993);)
(B) A simplified structure of chrysotile fiber (Yada, 1971).
As chrysotile has a spiral shape, a tetrahedral-octahedral edge is likely to occur at
the end of each tube as well as run along the length of it (Figure 7B). The tubes curl in a
way to expose the magnesium rich octahedral layer on the outside. Therefore, the surface
charge of this site is likely to be similar to that of brucite (which is positively charged
over a broad pH range). Due to this separation of charge between edges and faces, as well
as the flexibility of the chrysotile fibers, they are able to align themselves in a number of
configurations; this results in a very complex tangled structure with adverse effect on
chrysotile slurry rheology.

34
The crystalo-chemical structure of graphite and molybdenite, two inherently
hydrophobic minerals, are shown in Figure 8. In molybdenite, sheets of molybdenum
atoms are sandwiched between two sheets of sulfur atoms. The sulfur and molybdenum
atoms within the layers are strongly covalently bonded, but the successive layers of sulfur
atoms are held together by weak van der Waals bonds. These bonds provide excellent
cleavage characteristics parallel to the base of the hexagonal crystals producing a
hydrophobic surface (since sulfur does not form hydrogen bonds with water). Similar
situation exists in graphite.

Figure 8 - Crystalo-chemical structure of graphite and molybdenite.

PROPERTIES OF AQUEOUS SUSPENSION OF ISOTROPIC AND


ANISOTROPIC MINERALS
Isotropic minerals
As inspection of Figure 1 reveals, the electrical repulsion forces between interacting solid
particles in water will entirely disappear when these particles do not carry electrical
charge. Practically, the charge is indirectly characterized by the zeta potential
measurements which provide the iso-electric point (i.e.p.), that is the pH (strictly
speaking the concentration of potential-determining ions) at which the zeta potential of
the mineral is equal zero. Around this pH the suspension is very unstable, the particles
aggregate (coagulation) and settle quickly. Since aggregation results in the formation of
the net-work between aggregating particles, the rheological measurements give high
shear yield values for such a case.
Comparison of the experimentally determined yield stress versus pH curves with
the zeta potential-pH curves (since for most systems, potential-determining ions are H+
and OH- ions) can yield very important information on the nature of the particle surface
charge. Figure 9 (after Johnson et al. [8]) shows such a comparison. As this figure
indicates, the yield stress pH curves reveal the maximum exactly at the pH of the i.e.p.
for this mineral. At this point, the van der Waals attraction is not opposed by any
electrical repulsion, the suspension is unstable, it coagulates, and because of the structure

35
that develops between the coagulating particles the rheological measurements provide
high yield stress values. As the pH moves away from the i.e.p., the yield stress values
decrease both in higher and lower pH ranges; in these pH ranges the particles are
stabilized against aggregation by electrical double layers. This behavior is typical for
isotropic minerals.

Figure 9 - Electrophoretic measurements showing that the maximum coagulation occurs


at the iso-electric point of zirconia suspensions (Johnson et al., 2000).
Anisotropic minerals.
Clays. Because of the importance of clays this group has been extensively studied. Figure
10 shows the yield stress values plotted versus pH for kaolinite suspensions at different
volumetric solids content (Johnson et al., 2000).

36

Figure 10 - The yield stress pH curves for kaolinite suspensions at different volumetric
solids content (Johnson et al., 2000).
The rheological measurements in this case do not correlate with the electrokinetic
measurements at all. The iso-electric point for kaolinite determined from the zeta
potential measurements is around 3.5 while the point of maximum coagulation of
kaolinite suspensions lies around pH 5.5.
The lack of correlation raises serious questions about the applicability of the
Smoluchowskie equation to the calculation of zeta potential from the measured
electrophoretic mobility for plate-like anisotropic particles. The case is depicted in Figure
3. The behavior of a plate-like anisotropic particles in an electrical field is unknown and
there is no mathematical model that allows calculation of the zeta potential values from
the electrophoretic mobility of such particles. All such measurements must therefore be
treated as an estimation only.

Figure 11 - The likely zeta potential values for faces and edges of kaolinite (Johnson et
all., 2000).
This discussion, based on a large number of pieces of experimental evidence
(Williams & Williams, 1978; Rand & Melton, 1977; Tombacz & Szekers, 2006), leads to
the conclusion that the zeta potential values for kaolinite alumina edges and the zeta
potential for silicate faces would have been different if it had been possible to measure
them independently. Such an estimation is shown in Figure 11. From this plot it can be
inferred that maximum coagulation in kaolinite suspensions takes place around pH of 5.5,

37
the pH level at which differences between electrical charges of the two different sites of
kaolinite platelets are the largest (Fig. 12).

Figure 12 - Coagulation of clay particles over a pH range of 4 to 6.

Hydrophobic anisotropic minerals. This group includes talc, molybdenite and graphite.
The most characteristic feature of the anisotropy of these minerals is inherent
hydrophobicity of basal surfaces of these particles.
Figure 13 shows the zeta potential values of talc particles calculated from the
electrophoretic measurements carried out as a function of pH (Fuerstenau & Huang,
2003). These values agree very well with many other reported data obtained with the use
of the same experimental technique. They all give the iso-electric point for talc around
pH 2.5.

Figure 13 - Zeta potential of talc as a function of pH (Fuerstaneu & Huang, 2003).

The results obtained using the titration technique, and shown in Figure 14,
indicate that the point-of-zero charge for talc occurs around pH of 7.7 (Burdukova et al.
2007). This does not correlate at all with the position of the i.e.p. for talc (Figure 14). For
comparison this i.e.p. value is also given in Figure 14.

38

Figure 14 - Potentiometric titration curve for talc that identifies the point-of-zero charge
for talc to be around pH 7.7 (Burdukova et al., 2007).

Figure 15 - Casson yield stress curve for talc aqueous suspensions as function of pH
(Burdukova et al., 2007).
Further evidence for the anisotropic properties of talc particles comes from the
rheological measurements. Such experiments indicate that the maximum coagulation of
talc aqueous suspensions occurs around pH 5.5. From these results the model describing
electrical charge distribution on talc faces and edges was derived (Figure 16).

Figure 16 - Proposed charge distribution on the surface of talc particles (Burdukova et al.,
2007).

39
The differences in the surface properties of basal planes and edges of talc
particles, directly measured by Fuerstenau and Houang (2003), clearly show anisotropy
of talc particles. The receding contact angles measured on the cleaved face of a talc
crystal exceeded 60 degrees over a very broad pH range (from 4 to 12), while the contact
angles measured on the edges prepared by cutting the talc crystal were less than 20
degrees. This is even more convincing in Figure 17 which shows the effect of a cationic
collector (dodecylammonium acetate) on the wettability of talc faces and edges (at pH of
4-4.5).

Figure 17 - Effect of dodecylammonium acetate, measured at pH of 4 4.5, on receding


contact angles on the face and edge of a talc crystal (Fuerstenau & Huang, 2003).
As this figure reveals dodecylammonium cations orient differently at the
face/solution and edge/solution interfaces. While adsorption of this collector at edges
makes them hydrophobic, the adsorption on the faces results from the hydrophobic chain
interaction with the hydrophobic surface making this surface hydrophilic at higher
collector concentrations.
The hydrophobic-to-hydrophilic ratio that is the face-to-edge ratio changes for
anisotropic minerals with particle size. As Figure 18 demonstrates, the adsorption of two
polysaccharides (dextrin and guar gum) onto talc strongly depends on the particle size of
talc particles that is on the face-to-edge ratio. For example, adsorption of dextrin onto a 150 +75 m size fraction of talc is more than 2 mg/m2, while for the size fraction of -63 +
53 m it is only about 1 mg/m2 and for the -37 m fraction is a little higher than 0.5
mg/m2 (Rath et al., 1997). For both, dextrin and guar gum, the adsorption expressed per
square meter of the mineral surface clearly decreases with decreasing particle size. Since
the face-to-edge ratio decreases with decreasing particle size these results suggest that the
adsorption takes place mostly on the faces of the talc particles. In a flotation process this
should lead to a depressed flotation of talc.
Since molybdenite, graphite and talc are naturally hydrophobic, it is also possible
to measure wettability of their surfaces; there are quite a few papers on the flotation
properties of these minerals. The problem, however, is that only the faces of these
minerals are hydrophobic, and therefore only these parts of the surface of such particles
are responsible for their natural floatability. Wettability measurements on cleaved

40
graphite and molybdenite surfaces by Arbiter et al. (1975) gave the values around 80
degrees for graphite, and, depending on sample preparation, around 70-90 degrees for
molybdenite. Chander and Fuerstenau (1972) confirmed large differences in contact
angles measured on hydrophobic faces and hydrophilic edges of molybdenite crystal.
Lopez-Valdivieso et al. (2006) reported that while contact angles measured on
molybdenite faces were in the range of 60 degrees, the edges were completely
hydrophilic. Using Atomic Force Microscopy (AFM) Lopez-Valdivieso et al. (2006)
detected micro-crystals and micro/edges on the faces of molybdenite crystal. This
demonstrates that molybdenite faces are highly heterogeneous. These findings are in a
good agreement with those of Komiyama et al (2004) who found terraces and rims of
nanometric size on molybdenite faces and explained the catalytic activity of molybdenite
by the occurrence of such high energy crater structures on basal planes.

Figure 18 - Effect of pH on the adsorption densities of dextrin and guar gum for three
different particle size fractions of talc (Rath et al., 1997).

It is important to note here that the hydrophobic-to-hydrophilic ratio is different


for particles used in contact angle (larger specimens), flotation (fine particles) and zeta
potential (very fine particles) measurements. The particles used in flotation and zeta
potential measurements have a higher ratio of hydrophilic-to-hydrophobic surface than
the specimens utilized in contact angle measurements and therefore these measurements
may provide different correlations (Chander et al., 1975).
Molybdenite flotation. The rate of flotation of very fine particles decreases with the size
of such particles because of the low probability of particle-to-bubble collision. For
molybdenite, decreasing particle size also increases the edges-to-faces ratio making these
particles less hydrophobic. Therefore, the floatability of fine molybdenite sharply
decreases with the particle size (Castro & Laskowski, 1997; Castro et al., 2011). It was
also shown that the floatabilty of these fine particles is very sensitive to the presence of

41
hydrophilic macromolecules (polysaccharides, flocculants, etc.) (Castro & Laskowski,
2004). In the copper-molybdenum industry there are two situations where the use of
flocculants may affect Mo recovery: (i) increased use of recycled water in which the
concentration of residual polymer builds up and (ii) the need for thickening pulps of CuMo bulk concentrate before pumping over a very long distances (between a grinding
plant and flotation plant or between the Cu concentrator and the moly plant).
Inherently hydrophobic minerals can be floated in concentrated solutions of
inorganic salts, even without any other flotation agents (Klassen & Mokrousov, 1963).
This is co-called salt flotation process.
Klassen (1966) showed that while bituminous coals, which are very hydrophobic, float
very well in concentrated electrolytes, the salt flotation of less hydrophobic subbituminous coals and anthracites was not satisfactory. The relationship between coal
surface hydrophobicity and its salt flotation response was confirmed by Fuerstenau et al.
(1983). Since simple inorganic ions cannot render hydrophobic solid surfaces only
highly hydrophobic minerals respond well to the salt flotation process. Flotation of
molybdenite in 0.5 M NaCl solutions was shown to be very good indeed (Castro &
Laskowski, 2011: Castro & Laskowski, 2012). However, in the flotation of Cu-Mo
sulfide ores pyrite must be depressed and this is commonly achieved with the use of
lime. Any change of pH of such a well equilibrated system as sea water leads to
hydrolysis of divalent cations. Since magnesium hydroxide is much less soluble than
calcium hydroxide, increase of pH with the use of lime leads to the formation of
magnesium hydroxy-complexes and hydroxides. It was shown that molybdenite flotation
in alkaline pH is extremely sensitive to the presence of MgOH+ and Mg(OH) species in
the pulp (Castro et al., 2012; Laskowski & Castro, 2012).
Talc flotation. In the plants treating platinum bearing ores in South Africa,
polysaccharides (e.g. guar gum, carboxymethyl cellulose) are used to depress talc. For
some depressants ionic strength of the pulp is very important. For instance, while in the
case of carboxymethyl cellulose (anionic polymer), in the KCl solutions depression is
negligible at 10-3 M it is satisfactory at 10-2 M; in the presence of Ca2+ and Mg2+ ions,
depression is satisfactory even at the 10-3M concentrations (Shortridge et al., 1999)
(flotation tests were carried out at pH 9). Electrolyte concentration (ionic strength) is not
that important when guar gum is used in the flotation tests. As it was later confirmed,
macromolecules of carboxymethyl cellulose coil in the solutions at higher ionic strength
and this seems to increase adsorption of CMC (Pawlik et al., 2003). While guar gums
with larger molecular weight turned out to be stronger talc depressants, this was not
observed with CMC (Shortridge et al., 2000).
Talc is a significant constituent of the gangue in the platinum-bearing sulfide ores.
Talc depression increases the grade of the sulfide concentrates. This picture was however
recently found to be more complex (Robertson et al., 2003). Because of the plate-like
shape and hydrophobicity of talc faces, the presence of talc particles in the froth leads to
dry, large-bubbled froths. Depression of talc and better pulp dispersion leads to more
unstable froths with lower water recoveries and thus reduced entrainment.

42
Graphite flotation. Graphite also frequently appears as a difficult-to-handle gangue in
sulfide ores. INCO has three sulfide producing regions of which the most important are in
the Sudbury area (Ontario) and in Thompson (Manitoba). The Thompson circuit also
includes a graphite separation stage following the copper/nickel separation. The presence
of graphite in the nickel concentrate helps to maintain the reducing potential of the
electric furnace bath. However, graphite is rejected from the copper concentrate by
depressing chalcopyrite and floating off graphite. Several options were considered: the
use of sodium sulfide (which turned out not to be economical due to its high
consumption), the use of cyanides, and the use of thioglycolate (TGA) or
trithiocarbonate. Thioglycollate (HSCH 2 COOH) was patented in 1948 but it was not used
until recently. These agents are much less toxic than cyanides. Initially TGA was applied
to depress copper sulfides and float graphite, trithiocarbonate has been used for the last
couple of years.
Clays. Clays appear in almost all types of ores. They form very fine suspensions in water
which may be very stable resulting in tailings disposal problems. Because of their
anisotropic character, clay particles, depending on pH, may carry both positive and
negative electric charges, and thus can easily interact with all other minerals. This often
leads to a slime coating and inhibited flotation of valuable minerals. Desliming is the
most radical remedy which also requires the use of dispersing agents. In the case of clays
polyphosphates are commonly utilized (e.g. hexametaphosphate) or water glass.
In potash ore flotation plants in Canada, differential flotation of sylvite (KCl)
from halite (NaCl) is carried out in saturated brine with the use of long-chain primary
amines. Desliming process involves selective flocculation of slimes with the use of
polyacrylamide flocculants and flotation of the flocculated slimes with some secondary
branched amines (Chan et al., 1982; Perucca & Cormode, 1999). Since such a desliming
is not complete so-called blinders are still utilized (e.g. guar gum, CMC, starch). By
adsorbing onto remaining slimes the macromolecules of the blinders prevent the collector
from adsorbing onto slimes.
Successful flotation of kaolinite after flocculating it with modified
polyacrylamide was reported by Yuehua et al (2004). Ma et al (2009) showed that
kaolinite can be floated with ether monoamine and ether diamine in NaCl solutions.
Kaolin clay has found many important applications, for instance in manufacturing
of pottery, or as an inert filler. This requires prior beneficiation and removal of impurities
such as anatase, rutile, iron oxides, etc. The size of such particles is below 10 microns.
The best known is ultraflotation in which coarse calcite with addition of fatty acids (e.g.
tall oil) is used as a carrier to piggy-back fine anatase in the conventional flotation
process. Several carrierless processes were also developed. In all these processes fatty
acids are used along with Ca salts (activator) under alkaline conditions (Willis et al.,
1999). The problem is in controlling the concentration of the activator which may induce
clay coagulation when in excess, or may even cause flotation of the clay particles rather
than the colored impurities. Perhaps the most significant improvement was the

43
introduction of hydroxamate collectors patented in 1986 (such as S6493 reagent
manufactured by Cytec).
Slime coating. Perhaps the most instructive example which illustrates industrial
application of dispersants in flotation processes was provided by Jowett and his coworkers. Jowettt et al. (1956) demonstrated that hydrophobic coal particles become
completely hydrophilic in the presence of clays which form a slime coating on coal
surface (Fig. 19).

Fig. 19 - Effect of clays and dispersant on wattability of coal surface in aqueous solution
(Jowett et al., 1956).

Coal hydrophobicity can, however, be restored by simply adding a dispersant.


In the quoted paper, sodium hydrogen phosphate was used (today we would rather use
hexamethaphosphate). Both coal and clay particles were found to develop more negative
zeta potential values in 50-100 mg/L sodium hydrogen phosphate solutions (Fig. 20).
This is then a clear example of electrostatic stabilization; the specific adsorption of
phosphate anions increases the negative charge of the coal and clay particles, bringing
about strong coulombic repulsion which keeps clay particles apart from the coal surface
thereby restoring coal natural floatability.

44

Fig. 20 - Effect of Na 2 HPO 4 on the zeta potential of coal and clay particles (Jowett et al.,
1956).
The problem of slime coating in flotation will be further illustrated by the
example taken from the flotation of nickel sulfide ores in Western Australia; the gangue
in these ores contains serpentine minerals (antigorite, chrysotile, lizardite). The effect of
lizardite (fine-grained silicate) and chrysotile (fibrous silicate) on flotation of pentlandite
was studied by Edwards et al (1980). In processing of Candian ultramafic Nickel ores
with seperpentine gangue Dai et al. (2009) recommended desliming and flotation at a pH
of 10.1 adjusted using soda ash.
Fibrous gangue particles in flotation circuits. As Figure 7 shows, the fibrous chrysotile
tubes curl in a way that exposes the magnesium rich octahedral layer on the outside, and
therefore, the surface charge of this site is likely to be similar to that of brucite. To avoid
problems associated with electrophoretic measurements of anisotropic minerals, electrical
charge of brucite will be used to characterize electrical charge of the chrysotile surface.
Figure 21 shows zeta potential of brucite and silica plotted versus pH.

45

Fig. 21 - Zeta potential of brucite and silica as a function of pH (adapted after Miller et
al., 2007).

Figure 22 shows the zeta potential pH curves for lizardite and pentlandite
(Bremmel et al., 2005). As Figures 21 and 22 indicate, the surface of chrysotile fibres, as
well as surface of lizardite, are both positively charged over a very broad pH range. Over
this pH range, pentlandite particles are charged negatively. It could then be expected that
fine particles of the serpentine minerals will form slime coatings on the surface of
pentlandite. With the use of scanning electron microscope it was possible to demonstrate
that such coatings indeed appear on the surface of pentlandite conditioned in the pulp in
the presence of these minerals. Using small scale flotation experiments with pure
pentlandite, Edwards et al. (1980) demonstrated that at chrysotile-to-pentlandite levels
above about 1:200 there was almost complete depression of the pentlandite flotation
(these ores may contain more than 80% of serpentine minerals). The effect of lizardite
was less severe. With the use of carboxymethyl cellulose, which adsorption makes the
surface of silicate minerals negative, it was possible to eliminate depressing effect of the
slimes.

Fig. 22 - Zeta potential of lizardite and pentlandite particles in 10-3 M KNO 3 as a


function of pH (Bremmel et al., 2005).

46

Closed water circuits in mineral processing plants are almost universal today. In
such circuits electrolyte concentration build up in recycle streams over time resulting in a
high electrolyte concentration of the process water. The discussed phenomena and the
effect of CMC on slime coating may look differently at highly ionic strength of the
process water (Welham et al., 1992). Peng and Seaman (2011) have recently shown that
the effect of CMC strongly depends on the ionization of these macromolecules; this is
expressed as degree of substitution (DS) (charge density of CMC macromolecules
increases with the value of DS). In their flotation tests they used CMC samples with DS
varying from 0.4 to 0.9 and they found that the effect of CMC strongly depends on the
degree of substitution. While the flotation selectivity in de-ionised water increased with
increasing DS, the opposite effect was observed when the flotation tests were carried out
in the Mt. Keith process water. In the process water, that is at a higher ionic strength, the
less anionic CMC (that is with smaller DS values) provided more selective flotation
results.
The major effect brought about by the presence of fibrous gangue particles in the
flotation feed results from the shape of such particles. Rheology of fine particle systems
depends on many factors; particle size, shape and solids concentration influence the
rheological behavior to a great extent. In general, the rheological behavior becomes more
non-Newtonian as the particle size decreases Also, as demonstrated by Barnes et al.
(1989), the rheology of the fine suspensions strongly depends on the shape of the
particles (Fig. 23). The effect of the solids concentration on the suspension rheology is
much more pronounced if the particles are not spherical.
As Figure 23 reveals, at a particular volume solid concentration the suspension
jams up making flow impossible, i.e. at this solid concentration the viscosity tends to
infinity. This concentration is referred to as the maximum packing fraction. Its value
depends on the arrangements of the particles, and this depends on particle shape. While
for spherical particles these are large numbers, for the rods this concentration is about
20% (by volume). Horie and Pinder (1979) used a nylon thread with a diameter of 43.1
m and varying length (from 0.987 to 6.72 mm) and showed that the rheology of such a
suspension strongly depends on the L/D ratio.

47

Figure 23 Dependence of the apparent viscosity (at a shear rate of 300 s-1)
differently shaped particles in water on concentration (Barnes et al., 1989).

of

Figure 24 Rheological properties of the aqueous suspensions of chrysotile and the


nickel sulfidfe ore that contains chrysotile (Kilickaplan et al., 2010).
Some rheological mesurements with chrysotile suspension and the suspension of the
finely ground Ni sulfide ore (from Australia) which contains chrysotile as a gangue are
shown in Figure 24 (Kilickaplan et al., 2010). As this figure demonstrates, for the
suspension of fibrous chrysotile particles the maximum packing fraction is around 1%
(by volume), and for the ore it is around 8% (by volume) which is about 20% by weight.
These results explain why in the case of the tested Ni ore the solids content in the
flotation feed cannot exceed 20% (Merve Genc et al. 2012).

48

REFERENCES
Arbiter, N., Fujii, Y., Hansen, B. & Raja, A. (1975). Surface Properties of Hydrophobic
Solids, Advances in Interfacial Phenomena of Particulate/Solution/Gas Systems;
Application to Flotation Research (P. Somasundaran and R.B. Grieves, eds.), AIChE
Symposium Series, 150, Vol. 71, pp. 176-182.
Barnes, H.A., Hutton, J.F. & Walters, K. (1989). An Introduction to Rheology, Elsevier,
Amsterdam, pp. 120-124.
Bremmel, K., Fornasiero, D. & Ralston, J. (2005). Pentlandite-Lizardite Interactions
and Implications for Their Separation by Flotation. Colloids and Surfaces A,
252, 207-212.
Burdukova, E., Becker, M., Bradshaw, D.J. & Laskowski, J.S. (2007). Presence of
Negative Charge on the Basal Planes of New York Talc, J. Coll. Interf. Sci., 315, 337342. .
Carty,W.M. (1999). The Colloidal Nature of Kaolinite, Am. Ceram. Soc. Bull., pp. 7277.
Castro, S. & Correa, A. (1995). The Effect of Particle Size on the Surface Energy and
Wettability of Molybdenite, Processing of Hydrophobic Minerals and Fine Coal - Proc.
1st UBC-McGill Int. Symposium (J.S. Laskowski and G.W.Poling, eds.), Montreal, CIM
Metallurgical Society, Montreal, pp. 43-57.
Castro, S., Jara, C., Munoz, M. & Laskowski, J.S. (2011). Floatability of Molybdenite in
Sodium Chloride Solutions, to be published.
Castro, S. & Laskowski, J.S. (1997). The Effect of Hydophilic and Hydrophobic
Polymers on Molybdenite Flotation, Proc. 5th Southern Hemisphere Meeting on Mineral
Technology, Buenos Aires, pp., 117-120.
Castro, S. & Laskowski, J.S. (2004). Molydbenite Depression by Shear Degraded
Polyacrylamide Slutions, Particle Size Enlargement in Mineral Processing- Proc. 5th
UBC-McGill Int. Symposium (J.S. Laskowski, ed.), CIM Metallurgical Society, pp. 169178.
Castro, S. & Laskowski, J.S. (2011). Froth Flotation in Saline Water, KONA, No. 29, 415.
Castro, S., Rioseco, P., Zuniga, M. & Laskowski, J.S. (2012), 26th Int. Mineral
Processing Congress, New Delhi.
Chan, S. et al., Canadian Patent 1,211,235/1982.
Chander, S. & Fuerstenau, D.W. (1972). On the Natural Floatability of Molybdenite,
Trans. SIME/SME, 252, 62-68

49
Chander, S., Wie, J.M. & Fuerstenau, D.W. (1975). On the Native Floatability and
Surface Properties of Naturally Hydrophobic Solids, Advances in Interfacial Phenomena
of Particulate/Solution/Gas Systems; Application to Flotation Research (P.
Somadundaran and R.B. Grieves, eds.), AIChE Symposium Series, 150, Vol. 71, pp. 183188.
Dai, Z., Bos, J.A., Quinn, P., Lee, A. & Xu, M. (2009). Flowsheet Development for
Thomson Ultramafic Low-Grade Nickel Ores, Advances in Mineral Procesing Science
and Technology Proc. 7th UBC- McGill-UA Symp. (C.O. Gomez, J.E. Nasset and S.R.
Rao, eds.), CIM Metallurgical Society, Sudbury, pp. 217-228.
Edwards, C.R., Kipkie, W.B. & Agar, G.E. (1980). The Effect of Slime Coatings of the
Serpentine Minerals, Chrysotile and Lizardite, on Pentlandite Flotation. Int. J. Mineral
Processing, 7, 33-42.
Fuerstenau, D.W. & Huang, P. (2003). Interfacial Phenomena Involved in Talc Flotation
and Depression, Proc. 22nd Int. Mineral Processing Congress (L. Lorenzen et al., eds.),
Cape Town, Vol. 2, pp. 1034-1043.
Fuerstenau, D.W., Rosenbaum, J.M. & Laskowski, J.S. (1983). Effect of Surface
Functional Groups on the Flotation of Coal, Colloids & Surfaces, 8, 153-164.
Gaudin, A.M., Miaw, H.L. &. Spedden, H.R. (1957) Native Floatability and Crystal
Structure, Proc. 2nd Int. Congr. Surface Activity, London, Butterworths, Vol. 3, pp. 202219.
Horie, M & Pinder, K.L. (1979). Time-dependent Shear Flow of Artificial Slurries in
Coaxial Cylinder Viscometer With a Wide Gap, Canadian J. Chem. Eng., 57, 125-134.
Johnson, S.B., Franks, G.V., Scales, P.J., Boger, D.V. & Healy, T.W. (2000). Surface
Chemistry Rheology Relationships in Concentrated Minerals Suspensions, Int. J. Min.
Processing, 58, 267- 304.
Jowett, A., El-Sinbawy, H. & Smith, H.G., Slime Coating of Coal in Flotation Pulps,
Fuel, 35, 303- 310.
Kilickaplan, I., Ndlovu, B. & Laskowski, J.S. (2010). Rheology of Aqueous Suspensions
of Needle-like Mineral Particles, Rheology in Mineral Processing Proc. 8th UBCMcGill-UA Symp., (M. Pawlik, ed.), CIM Metallurgical. Society, Montreal, pp. 193-203.
Klassen, V.I. (1966). Coal Flotation, Katowice, Slask,, pp. 104-109 (Translation from
Russian into Polish).
Klassen, V.I. & Mokrousov, V.A. (1963). An Introduction to the Theory of Flotation,
London, Butterworths, pp. 338-341.

50
Klein, C. & Hurlbut C.S. (1993). Manual of Mineralogy, New York, Wiley and Sons.
Komiyama, M., Koyohara, K., Fujikawa, T., Ebihara, T., Kubota T. & Okamoto, Y.
(2004). Crater Structure on a Molybdenite Basal Plane Observed by Ultrahigh Vacuum
Tuneling Microscopy and Its Implication to Hydrotreating. J. Molecular Catalysis A:
Chemical, 215, 143-147.
Laskowski, J.S. & Kitchener, J.A. (1969). The Hydrophobic-Hydrophilic Transition on
Silica. J. Coll. Interf. Sci., 29, 670-679..
Laskowski, J.S. and Castro, S.(2012). Hydrolyzing Ions in Flotation Circuits: Sea Water
Flotation, 13th Int. Mineral Processing Symp., Bodrum, Turkey.
Lopez Valdivieso, A., Madrid Ortega, I., Reyes Bahena, J.L., Sanchez Lopez, A.A. &
Song, S. (2006). Propiedades de la Interface Molibdenita/Solucion Acuosa y su Relacion
con la Flotabilidad del Mineral, Proc 16th Congreso International de Metalurgia
Extractiva, Saltillo, Mexico, pp. 299-310.
Ma, X., Bruckard, W.J. & Holmes, R., (2009). Effect of Collector, pH and Ionic Strength
on the Cationic Flotation of Kaolonite, Int. J. Mineral Processing, 93, 54-58.
Merve Genc, A., Kilickaplan, I. & Laskowski, J.S. (2012). Effect of Pulp Rheology on
Flotation of Nickel Sulphide Ore with Fibrous Gangue Particles, Can. Metallurg. Quart.,
2012, in press.
Miller, J.D., Nalaskowski, J., Abdul, B. & Du, H. (2007). Surface Characteristics of
Kaolinite and Other Selected Two Layer Silicate Minerals, Can. J. Chem. Eng., 85, 617624.
Pawlik, M., Laskowski, J.S. & Ansari, A. (260). Effect of Carboxymethyl Cellulose and
Ionic Strength on Stability of Mineral Suspensions in Potash Ore Flotation Systems, J.
Coll. Interf. Sci., 260, 251-258.
Peng, Y. & Seaman, D. (2011). The Flotation of Slime-fine Fraction of MT. Keith
Pentlandite Ore in De-ionised and Saline Water. Minerals Engineering, 24, 479-481.
Perucca, C.F. & Cormode, D.A. (1999). The Use of Polymers in Potash Beneficiation at
Agrium Potash Plant. Polymers in Mineral Processing Proc. 3rd UBC-McGill Int.
Symp. (J.S. Laskowski, ed.), CIM Metallurg. Society, pp. 393-404.
Rand, B. & Melton, I.E. (1977).Particle Interactions in Aqueous Kaolinite Suspension, J.
Coll. Interf. Sci., 60, 308-336.
Rath, R.K. ,. Subramanian, S. & Laskowski, J.S., (1997). Adsorption of Dextrin and
Guar Gum onto Talc. A Comparative Study. Langmuir, 13, 6260-6266.

51
Robertson, C., Bradshaw, D.J. & Harris, P.J. (2003). Decoupling the Effects of
Depression and Dispersion in the Batch Flotation of a Platinum Bearing Ore, Proc. 22nd
Int. Mineral ProcessingCongress (L.Lorenzen and D.J. Bradshaw, eds), Cape Town, vol.
2, pp. 920-928.
Shortridge, P.G., Harris, P.J. &. Bradshaw, D.J. (1999). The Influence of Ions on the
Effectiveness of Polysaccharide Depressants in the Flotation of Talc. Polymers in
Mineral Processing Proc. 3rd UBC-McGill Int. Symposium (J.S. Laskowski, ed.), CIM
Metallurgical Society, pp. 155-169.
Shortridge, P.G., Harris, P.J., Bradshaw, D.J. & Koopal, L.K., (2000) The Effect of
Chemical Composition and Molecular Weight of Polysaccharide Depressants on the
Flotation of Talc, Int. J. Mineral Processing, 59, 215-224.
Tombacz, E. & Szekeres, M. (2006). Surface Charge Heterogeneity of Kaolinite in
Aqueous Suspensions in Comparison with Montmorillonite, Appl. Clay Sci., 34, 105124.
Wellham, E.J., Elber, L. & Yan, D.S. (1992). The Role of Carboxymethyl Cellulose in
the Flotation of a Nickel Sulfide Transition Ore. Minerals Eng., 5, 381-395.
Williams, D.J.A. & Williams, K.P., (1978). Electrophoresis and Zeta Potential of
Kaolinite, J. Coll. Interf. Sci., 65, 79-87.
Willis, M.J., Mathur, S. & and R.H. Yoon, (1999). Kaolin Flotation: Beyond the
Classical, Advances in Flotation Technology (B.K. Parekh and J.D. Miller, eds.), SME,.
pp. 219-229.
Yada, K. (1971). Study of the Microstructure of Chrysotile Asbestos by High Resolution
Electron Microscopy, Acta Crystallographica, A27, 659-664.
Yuehua, H., Wei, S., Haipu, L. & Xu, Z., (2004). Role of Macromolecules in Kaolinite
Flotation. Minerals Engineering, 17, 1017-1022.

52
Chapter 2 Surfactants

Surfactant is an abbreviation for surface active agent, which literally means


active at a surface. The surface (interface) can be between solid and liquid, between
liquid and air, or between two immiscible liquids (e.g. water and oil). The primary
property of a surfactant in a solution is that its concentration is higher at the interface than
in the bulk of the liquid. Thus the surfactant concentrates at the interface.
2.1. Adsorption from solution - Gibbs adsorption isotherm
The relationship between surface (interface) tension and the adsorption ()
occurring at a surface (interface) is given by the Gibbs adsorption isotherm. For a simple
two-component solution (consisting a solvent and a surfactant)

d = d

(2.1)

Since chemical potential of component i

i = io + RT ln ai

(2.2)

where io and ai are standard chemical potential and thermodynamic activity of


compound i. Because a = fc and since for c0, f1 then for dilute solutions a = c
and thus

d = RTd ln a RTd ln c

(2.3)

and the Gibbs equation is therefore written

1 d
1
d
=
RT d ln c
2.3RT d log c

(2.4)

Since for liquids surface tension can easily be measured, this equation allows calculation
of the adsorption of a given compound at the liquid/gas interface.
The surface tension data is shown schematically as the top curve and the
adsorption densities calculated from Equation (2.4) are shown below (Fig. 2.1). In Region
A there is little change in and hence little increase in adsorption. In Region B,
decreases steadily and due to the minus sign in the Gibbs equation, adsorption ()
increases. In Region C, decreases linearly with log c and thus is constant. At the end
of Region C the system switches to another lower energy state where micelle formation
occurs.

53
In the case of solids the surface tension cannot be experimentally measured. But
adsorption onto a solid can. Gibbs equation allows calculation of the surface tension
changes if the adsorption densities are known.

Fig. 2.1. The variation of surface tension and adsorption with concentration of a
surface-active agent.

Fig. 2.2. gives examples of the most common surface tension vs. concentration
curves.

Curve I is typical for some non-ionic surface-active substances such as, for
example, aliphatic alcohols. Curve II is characteristic for inorganic salts (surface inactive
substances). They dissociate in water and the ions are negatively adsorbed (they are
repelled) from the air/water interface. Since the concentration of ions in the surface layer
is smaller than in the bulk the surface tension increases. Curve III is typical for true
surfactants (such as detergents). For such substances at quite modest concentrations the

54
decrease in surface tension ceases with the formation of micelles. In water, the surface
tension plotted vs. concentration of a surfactant, drops from 72 mN/m down to usually
about 30-40 mN/m. The critical micelle concentration (c.m.c.) is indicated in the figure.

Fig. 2.3. Surface tension concentration isotherms for aqueous solutions of n-aliphatic
alcohols (from C.C. Addison, 1945; and Posner, 1952) and two commercial frothers:
MIBC and DF-200.
Fig. 2.3. shows the effect of concentration of aliphatic alcohols on surface tension.
As can be seen from Fig. 2.3, the fall off of the surface tension is much steeper for
aliphatic alcohols with longer hydrophobic chains; in other words, surface activity of an
aliphatic alcohol increases with the number of carbon atoms in the molecule. In the
homologous series of n-alcohols, the bulk concentration necessary to obtain a given
surface coverage decreases by about three time for every additional - CH 2 - group in the
hydrocarbon chain of the molecule (Traube rule). Alcohols with branched radicals are
less surface-active than the corresponding straight-chain alcohols. The surface activity of
MIBC is comparable with that of n-hexanol.
2.2. Critical Micelle Concentration
Solutions of highly surface-active substances at fairly well defined concentrations
undergo a dramatic change in several physical properties (e.g. conductance, turbidity,
surface tension, etc.). This is shown in Fig. 2.4. The concentration around which the
change takes place is referred to as critical micelle concentration. At concentrations
higher then c.m.c. aggregates formed by surfactant molecules, known as micelles appear
in solution (Fig. 2.5).

55

Fig. 2.4. Physical properties of sodium dodecyl sulfate solutions at 25 oC.

Fig. 2.5. Schematic representation of a micelle (anionic surfactant)


2.3. Adsorption from solutions
The classification system of solution adsorption isotherms proposed by Giles et al
is shown in Fig. 2.6.

Fig. 2.6. Giles et als classification of the solution adsorption isotherms.

56
In the S type isotherm, the convex shape of the curve relative to the
solution-concentration axis means very little adsporption occurs at low concentrations
and increases progressively as the concentration is raised. This type of isotherm is
common for low affinity solutes (adsorbents). Type C is characteristic for physical
adsorption. Type L (and H) represents Langmuirian adsorption isotherm which is
characteristic for strong chemical interactions. Type H is commonly observed in the
measurements involving polymers.
2.4. Adsorption of ionic surfactants onto charged surfaces
The following mechanisms may be involved in the adsorption of an ionized
surfacant onto a charged solid surface:
i. Ion exchange - involves replacement of counterions adsorbed onto the solid
from the solution by similarly charged surfactant ions.

ii. Ion pairing - adsorption of surfactant ions from solution onto oppositely
charged sites unoccupied by counter-ions.

iii. Hydrophobic bonding - occurs when the combination of mutual attraction


between hydrophobic groups of the surfactant molecules and their tendency to escape
from an aqueous environment becomes large enough to permit them to adsorb onto the
solid adsorbent by aggregating their chains (hemi-micellization).
In general, also hydrogen bonding between substrate and adsorbate can contribute
towards interactions leading to the adsorption.
Isotherm of the adsorption of sodium dodecyl sulfonate (anionic collectors) onto
positively charged alumina (pH 7.2), as studied by Wakamatsu and Fuerstenau (1973), is
shown in Figure 2.7. The figure also shows wettability of the alumina surface as a
function of sodium dodecyl sulfonate concentration, and the zeta potential of amulina
particles measured under the same conditions. The adsorption isotherm can be divided
into three regions. At low concentrations, adsorption of sulfonate ions occurs by ion

57
exchange mechanism with the adsorbed counter-ions. The shape of the isotherm in this
region is of a low-affinity type; only electrical interactions are responsible for the.

Fig. 2.7. The adsorption


density, the electrophoretic
mobility, and the contact angle
on alumina as a function of the
equilibrium concentration of
sodium dodecyl sulfonate at pH
7.2, and ionic strength 2 x 10-3
M NaCl (Wakamatsu and
Fuerstenau, 1973).

adsorption. In this region the zeta potential is constant. In the second region, the adsorbed
ions begin to associate, with adsorption increasing markedly due to the enhanced
adsorption energy because G spec becomes effective. The third region is reached when
the zeta potential sign reverses. At concentrations higher than this, the electrostatic
interaction opposes the specific adsorption effects, resulting in a decrease in the slope of
the adsorption isotherm.
The following figure shows schematically the solid/liquid interface in these three
regions.

58
Fig. 2.8. Schematic representation of adsorbed surfactant at a solid/solution interface,
showing adsorption as individual ions and as associated hemi-micelles.
The specific interactions shown in this figure must depend on the hydrocarbon
chain length; since hemi-micellization results from the association of the hydrocarbon
chains, it should take place at much lower concentrations when the length of the
surfactant hydrocarbon chain is increased. Experimental results confirm such a
conclusion (see Fig. 2.9).

Fig. 2.9. Adsorption


isotherms for sodium
alkylsufonates of
different hydrocarbon
chain lengths on
alumina at pH 7.2, 25 oC
and 2 x 10-3 M NaCl
(Wakamatsu and
Fuerstenau, 1973).

In the absence of chemisorption, ionic surfactants charged similarly to the surface do not
adsorb and that is why the flotation of oxides depends so much on pH. However, by
strongly adsorbing in the Stern plane multivalent ions charged oppositely to the surface
can function as a link between the surfactant and the solid (activators). It is obvious that
in the pH ranges in which multivalent cations adsorb onto quartz, its flotation with
anionic collectors will be enhanced.

59
2.5. Effect of solid surface properties on adsorption of surfactants
Figure 2.10 shows schematically a general trend in the mode for adsorption of
non-ionic surfactants onto hydrophobic (A) and hydrophilic (B) surfaces. Because of
strong interactions between hydrophobic chains of the surfactant and a hydrophobic
surface, the surfactant at a hydrophobic solid/water interface orients with polar groups
facing water phase. This renders hydrophobic solid hydrophilic at higher surfactant
concentrations. For hydrophilic solids the orientation may be reverse, especially in the
case of strong specific interactions between polar surfactant groups and solid surface (as
in the case of flotation collectors). However, as Fig. 2.4 reveals, at sufficiently high
surfactant concentrations (critical micelle concentration, cmc) micelles will form in
solution and the solid surface will again become hydrophilic. Of course, at such high
surfactant concentrations flotation should be completely depressed. This was confirmed
for several systems.

Fig. 2.10 Model for adsorption of nonionic polyoxy-surfactants onto hydrophobic (A)
and hydrophilic (B) surfaces, and corresponding adsorption isotherms (Clunie and
Ingram, 1983).
The following two figures (after Dobias et al) demonstrate that the floatability is
completely depressed at collector concentrations exceeding CMC.

60

Fig. 2.11. Flotation recovery of barite as a function of concentration of sodium C 10 , C 12


and C 14 alkylsulfates (B. Dobias, 1986).

Fig. 2.12. Flotation recovery of quartz as a function of concentration of cetylpyridinium


chloride afte 1 and 10 minutes of flotation (B. Dobias, 1986).

61

Chapter 3
Flotation Reagents.
Collectors: their properties, applications. Modifiers: activators, depressants,
dispersants, pH regulators. Frothers: foams, bubble coalescence and foam
stability, flotation froths, effect of frothers on flotation performance, application of
frothers in flotation processes.
3.1. Introduction
The separation in the flotation process is based on differences in the surface
properties of the separated particles. In this process the hydrophobic particles attach to
bubbles when they collide with them and are lifted with them up to the froth layer while
the hydrophilic particles are carried away with the slurry.
Only a few minerals are hydrophobic by nature (graphite, molybdenite, talc,
sulfur, etc) and many different chemical additives, referred to as flotation reagents, must
be utilized in the flotation process. The most important group includes collectors, the
reagents needed to make given mineral(s) floatable. Classification of flotation collectors
is usually based on chemical composition of these compounds. Example of such a
classification is shown in Figure 3.1.(J.S. Laskowski, Can.Metal.Quart., 2010)

FLOTATION COLLECTORS
Non-ionic
Cationic
Amines

Anionic

Hydrocarbons
Kerosene
Diesel oil

Chelating
compounds

Non-thio
compounds
Thio-compounds
Fatty acids
Alkyl sulfonates
Alkyl sulfates

Xanthates
Xanthic esters
Xanthogen formates
Thionocarbamates
Dithiophosphates

Hydroxamates (O-O type)


Hydroxy oximes (N-O type)
Xanthogen formates (S-O type)
Dimethyl glyoximes (N-N type)
Xanthates (S-S type)
Thionocarbamates (N-S type)

Figure 3.1. Classification of flotation collectors.

New collectors are chemical reagents that adsorb at solid/liquid interfaces by


chemisorption (and this improves selectivity). A majority of new developments in
collector chemistries involve a functional group that contains two donor atoms capable
of bonding a metal species and of forming a ring structure known as a chelate.

62
Chelating-type collectors contain a reactive functional group with such ligand atoms as
S, N, O, and also P, in positions capable of bonding the same metal atom through two or
more different ligand atoms to form a heterocyclic ring [P. Somasundaran & D.R.
Nagaraj, in Reagents in the Mineral Industry, IMM, London, 1984, pp. 209-219;
D.W. Fuerstenau et al., in Advances in Coal and Mineral Processing Using Flotation,
SME, 1989, pp. 3-18]. This relatively new group of flotation collectors has been
introduced to a traditional classification of collectors in Figure 2.1. Such a
classification, while more complete, is however not consistent since some sulfurcontaining compounds appear twice, in the group of thio-collectors and also in the
group of chelating collectors.
This controversy calls for a different classification system which should be
based on the mechanism of action rather than on the collector chemical composition.
The basic point in such an approach is the fact that collector is needed at the point of the
particle-to-bubble attachment.
3.2. Role of surfactant in making mineral surface hydrophobic.
Of various reagents employed in flotation process, those which are referred to as
collectors are utilized to render hydrophobic a given mineral. As seen from Youngs
equation, the value of contact angle is determined by the solid/gas, solid/liquid and
liquid/gas interfacial tensions. Since addition of a flotation collector increases contact
angle this must result from the effect the collector has on the corresponding interfacial
tensions.
The relationship between surface (interface) tension and the adsorption ()
occurring at this interface is given by the Gibbs adsorption equation (Eq. 2.1). As this
equation and Fig. 2.1. show, adsorption causes the interfacial tension to decrease.
By writing Youngs equation in the form:

cos =

SV SL
LV

(3.1)

it becomes obvious that since for a large contact angle

SV SL < LV

(3.2)

for initially hydrophilic solid ( = 0 , cos = 1 ) the contact angle can increase (solid
will become hydrophobic) only if, as a result of surfactant adsorption, SV goes down
more than SL . The conclusion is that an increase in contact angle results from
surfactant (in flotation collecto)r addition only if the collector adsorption is greater
at the mineral-air interface than at the mineral-water interface (de Bruyn, Overbeek
and Schuhmann, 1954). This treatment indicates that for better flotation conditions, one
should look at the conditions of adsorption at the liquid/gas interface and solid/gas
interface rather than just that at the solid/liquid interface, as is usually done.
In the 50s, Gaudin et al. established that in most cases flotation of a mineral is
optimum when only a partial (5-15%) monolayer of collector coverage is present.

63
In the penetration theory, developed by Schulman and Leja, the collector adsorbed on
the mineral particle is postulated to interact with a frother carried on a bubble in the
moment of the particle-bubble collision and attachment. They showed that while
frothers are incapable of adsorbing on most minerals they adsorb onto solids pre-coated
by a suitable collector. This, as Fig. 3.2 shows, leads to the situation which agrees very
well with theoretical conditions given by Equations 3.1 and 3.2.
Fig. 3.2. Adsorption
of the surfactant
brought by bubble
onto solid.

Experimental results of Digre and Sandvik (1968) further substantiated the


penetration theory. They showed that while adsorption of amines on quartz was very
low indeed, in the presence of bubbles this adsorption was 2-3 times higher. These
results confirmed that when air is introduced as bubbles the bubble surface also adsorbs
collector molecules according to their surfaced activity (in the case of amines, while
ionized ammonium ions are likely to adsorb onto a solid surface, molecular amine is
more likely to adsorb onto bubbles). By collision between bubbles and mineral particles
(already partially coated by adsorbed collector), the collector is transferred from the
buble to the mineral surface, which is thus rendered hydrophobic. Some direct contact
angle measurements revealed that contact angles increase by repeated contacts between
bubbles and mineral surfaces.
These results are perfectly in line with those of Ter-Minassian-Saraga, and
recent publications by Yaminsky, who indicate that adsorption of quarterly ammonium
salts (e.g. CTAB) on quartz is very low and cannot explain quartz hydrophobicity in
such solutions. Figure 2.12 shows after Somasundaran (1968) adsorption of
dodecylammonium acetate (DDA) at quartz/solution, solution/gas and quartz/gas
interfaces. These results clearly show that while adsorption of DDA at solution/gas and
quartz/gas interfaces is

64

Fig. 3.3. Comparison of adsorption of dodecylammonium acetate at different interfaces


(Somasundaran, Trans. AIME, 1968).
very high, the adsorption at quartz/solution interface is very much lower. A high
adsorption at a quartz/gas interface can only be explained by a high adsorption at a
liquid/gas interface and the transfer of the amine to the solid surface during collisions of
the mineral particles with the bubbles. This is a very important - and very often
overlooked - aspect of flotation fundamentals which have never been sufficiently
explored in refining engineering solutions.
This exercise leads to the conclusion that an increase in contact angle results
from surfactant only if the surfactant adsorption is greater at the mineral-air interface
than at the mineral-water interface. This result indicates that for better flotation, one
should look at the adsorption at the liquid/gas interface and solid/gas interface rather
than just at the solid/liquid interface, as is usually done.

3.3. NEW CLASSIFICATION OF FLOTATION COLLECTORS


Collector molecule can be carried on to the point of particle-to-bubble
attachment either on the surface of solid particles, or on the surface of bubbles, or on
both. In both cases the collector first must be introduced into the flotation pulp. If
collectors are soluble in water they are added to the pulp in the form of aqueous
solution. If these are oils, that is water-insoluble hydrocarbons, they must be
dispersed in water and then added to the pulp in the form of an oil-in-water emulsion.
Amphipathic water-insoluble compounds, that is long chain surfactants, can form
molecular films at the water/gas interface and so they can be introduced to the flotation
system on the surface of bubbles. Thus, from the point of view of the mode of operation
of these agents the most logical would be classification of the flotation collectors into
(J.S. Laskowski, Can.Metal.Quart., 2010):

65
(i) Water soluble compounds,
(ii) Water-insoluble oily collectors,
(iii) Water-insoluble amphipathic compounds which form molecular films at
water/gas interface.
3.3.1. Soluble collector flotation
Diffusion of collector ions or molecules toward solid surfaces and selective
adsorption onto solid surfaces is a main mechanism of action of the soluble flotation
collectors. In order for the flotation to be selective, the collector adsorption must be
selective.
Collectors can both chemisorb and physisorb onto solid particles and the total
adsorption density includes both forms. While chemisorption results from some specific
interactions and therefore is usually very selective, the physical interactions are less
specific and therefore less selective.
3.3.2. Water-insoluble oily collector flotation
In another mechanism, collector appears in the form of droplets (as in the case of
emulsion flotation with oily collector) which collide with mineral particles as a result of
conditioning and can get attached to the solid surfaces. For the attachment to be possible
the solid must to some extent be hydrophobic (this mechanism will be referred to as
insoluble oily collector flotation). Flotation selectivity in this case then results from the
selective attachment of oily droplets to mineral surfaces. As shown by Taubman and
Yanova only hydrocarbons which are insoluble in water can operate in this way.
The model tests described by Zettlemoyer et al. [J.Coll.Interf.Sci., vol. 34, 545548 (1970)], in which a smooth Teflon disc was placed under water and a drop of the
mineral oil (Drakenol-10) was submerged and placed in contact with the Teflon plate
(the contact angle of water on this disc was 108 deg confirming that it was very
hydrophobic) well describes the situation in the flotation process in which oily collector
is added to the pulp containing mineral particles suspended in water. In the Zettlemoyer
et als experiments the oil droplet in water exhibited a finite contact angle of 37 deg.
(measured through oil), i.e. the oil does not displace water from the Teflon. Water
height was then gradually lowered, and when the oil/air interface was established the
water spontaneously receded to the edge of the disc leaving oil droplets on the Teflon
surface (contact angle of 52 deg. measured through oil phase). When the sequence of
the stages was however reversed, oil droplet was placed on a dry Teflon surface and
then the system was submerged in water, the oil contact angle was zero. This probably
results from the contact angle hysteresis. These observations also explain why it is
commonly assumed that oil forms a film on the surface of hydrophobic particles in
flotation

66

Figure 3.4. Attachment of bubble to a hydrophobic particle in the presence of oil


droplet on the solid surface [based on Zettlemoyer et al [21]).

systems. The mechanism described by Zettlemoyer et al agrees very well with the
mechanism postulated by Klassen and his co-workers in a series of publications on the
mechanism of action of water-insoluble oily collectors. This mechanism is sketched in
Figure 3.4.
One noticeable effect of the oily collector is an increased size of floated
particles. According to Melik-Gaykazyan [1965], the location of the oil as a thread at
the line of contact of the three phases remarkably increases the bubble-particle adhesion
force.
In the discussed collector classification, the flotation agents which are weak
electrolytes (e.g. fatty acids, amines) are somewhere in between the soluble and oily
insoluble collectors. Ionization of these agents strongly depends on pH, and while the
molecular forms of these agents are very insoluble in water (solubility of oleic acid is
about 10-8 mole/L), they may be quite soluble when ionized. This is quite obvious on
the domain diagrams of these agents. The fact that the weak electrolyte type collectors,
depending on pH, appear in the flotation system either in the form of emulsion or as a
solution explains some peculiar flotation properties of such agents. It is well established
that the flotation with water-insoluble oily collectors requires long conditioning time.
On the other hand, an adsorption equilibrium is established fairly quickly for an ionized
collector, and so, in the pH zone in which the collector is ionized the effect of
conditioning time should be negligible. By the way, Kulkarni [Ph.D. Thesis, 1975]
observed that while at pH 8 adsorption of oleate on hematite attained its constant value
after less than 2 minutes, at pH of 4.8 the equilibrium adsorption was not reached even
after 4 hours. As shown in Figures 3.5 and 3.6, flotation tests confirm entirely such
conclusions [Laskowski & Nyamekye, IJMP, vol. 40, 245-256 (1993)].

67
Figure 3.5 shows the effect of conditioning with sodium oleate on flotation
response of fluorite. At pH 9.5 the fluorite recovery increases dramatically over the first
few minutes and then remains almost constant. This is very different at pH 4 and thus
while at short conditioning times the flotation in the alkaline solution is better, this
picture changes dramatically when the conditioning time is long enough. Figure 3.6
shows similar results for the flotation of magnesite with sodium oleate.
These results demonstrate that the transportation of the collector species toward
mineral particles is a flotation controlling step in the emulsion flotation. Because of
very low solubility of oleic acid, in the acidic pH ranges the oleic acid droplets appear
in the system and a typical emulsion flotation ensues. Slow diffusion of the collector
droplets is apparently responsible for the pronounced effect of conditioning time on
flotation in the acidic pH range. In the alkaline range the collector is ionized and as a
result of the fast diffusion of such ions the adsorption equilibrium is reached quickly.
Consequently the optimum on the flotation recovery pH curves shifts towards lower
pH values when the experiments are preceded by longer conditioning times.

Figure 3.5. The effect of conditioning time on the flotation of fluorite under acidic and
alkaline conditions at an sodium oleate concentration of 10-5 .

68

Figure 3.6. The effect of conditioning time and oleate concentration on the flotation of
magnesite at pH 3.5 and 10.5.
The studies on reactive oily-bubbles [Z.Xu et al., Ottawa, 2003], that is the air
bubbles coated with a thin film of oily collector, demonstrated again how beneficial it is
when the collector is brought to the point of particle-bubble attachment on the surface of
bubbles. Misra and Anazia [Mineral & Metallugical Processing, vol. 4, 233-236
(1987)] compared the results of coal flotation tests carried out with fuel oil No. 2, which
was either added directly into a Denver flotation cell in an ultrasonically emulsified
form or as aerosol in the stream of air entering the cell. The results indicate that the
introduction of collector on the surface of bubbles produces superior results. Misra and
Anazia also demonstrated that the induction time needed for the oil-coated bubbles to
attach to coal specimen was many times shorter in comparison with the induction time
measured for the oil-free bubbles.

3.3.3. Water-insoluble amphipathic aliphatic collectors


Amphipathic compounds, insoluble in water, form molecular films at the
water/gas interface (also known as Langmuir-Blodgett films).

69
Figure 3.7. Deposition of the Langmuir-Blodgett film on solid surface
The monolayer films of fatty acids were heavily studied. Gaines [Insoluble
Monolayers at Liquid-Gas Interfaces, 1966] demonstrated that with long-chain amines
(e.g. C22 amine) condensed monolayers are formed in a wide range of conditions. The
stability of monolayers depends on solubility; fatty acids form stable monolayers on
water in acidic solutions when they are not ionized, while long-chain amines form
stable monolayers on water in an alkaline pH range. Spreading of monolayer from a
solid requires detachment of molecules into the surface film which occurs only at the
periphery of the crystal. The process, therefore, strongly depends on temperature and
becomes extremely slow below certain temperatures, but increased solubility or
evaporation make the films unstable if temperature is high. The role of transportation of
such compounds, used as collectors in flotation of potash ores, in the form of molecular
films spread on bubbles is especially important in potash ore flotation and will be
further discussed in the chapter dealing with flotation of potash ores.
Arsentiev and Leja [Colloid and Interface Science, Academic Press, 1976, Vol.
5, pp. 251-270] studied the monolayers of amines and fatty acids spread and
compressed on saturated salt solutions. They used hexanol (or 1:1 mixture of benzene
and ethyl alcohol) to dissolve the tested surfactants in the spreading procedure. In the
spreading rate measurements [Leja, in Potash Technology, Pergamon, 1983] it was
shown that while the spreading rates on saturated salt solutions were negligible for pure
long-chain amines, these rates increased dramatically when the amines were mixed with
alcohols. (e.g. hexanol). The tested flotation frothers such as MIBC and DF-250
behaved similarly to hexanol. The spreading rates for C16-C18 amines mixed with
hexanol were in the range from 300 to 400 mm/s. The KCl plates brought from
underneath adhered strongly to the amine films, NaCl plates did not (Fig. 5.8). There
was no adhesion between KCl plates and the branch-chain amine films.

Fig. 3.8. The adhesion tests between the amine film, spread on a saturated NaCl-KCl
brine, and either NaCl or KCl plates (Arsentiev and Leja).
Long-chain primary amines utilized as a collector in potash ore flotation are
practically insoluble in water. The amines are therefore melted by heating up to 70-90
o
C, then they are neutralized with hydrochloric or acetic acids and such a hot emulsion/
dispersion is introduced into the flotation pulp (which is at room temperature).
According to all reported observations a white precipitate immediately appears when
the hot emulsion of amine is added to the potash flotation pulp. Introduction of a hot

70
amine emulsion into a pulp at the temperature which is much lower than the Krafft
temperature for the utilized amines must cause precipitation of the amines.

3.4. Thio-collectors
Table 3.1.. Chemical structure of the main thiol collectors used for sulfide minerals.
Thiol collectors

Chemical structure

Xanthates

S
R O C

Xanthogen formates

Xanthic esters
Ditiophosphates

R O

S
P

R O

Mercaptobenzothiazol

O
C

S
C

CH2CH3
S
O
S
R O C N CH2CH3
H
S
R1 O C S R2
R

Thionocarbamates

S- Na+ (K+)

+
S- Na
N
SH
S

Xanthates account for about 65% of the collectors utilized in flotation of sulfide
ores, with dithiophosphates representing 15% of the total sulfhydryl collectors used.
Introduction of xanthates in flotation dates back from 1924.
Xanthates are available commercially as solutions, powders and pellets.
The dry xanthate powder is highly irritating (it can ignite spontaneously), pellets offer
better storage stability and fewer dusting problems.
The method of preparation of sodium or potassium alkyl xanthates consists of
the dissolution of an alkali hydroxide in the appropriate alkyl alcohol, followed by an
addition of carbon disulfide to the metal alcoholate. The reaction is as follows:
ROH + NaOH + CS 2 ROCSSNa + H 2 O

(3.3)

As this reaction indicates, xanthates decompose in the presence of moisture, giving off a
highly toxic carbon disulfide as one of the decomposition products. The sodium
xanthates are hydroscopic, whereas the potassium salts are not and, therefore, the
potassium xathates are preferred since they do not pose storage problems.
Xanthates are also delivered in the form of aqueous solutions consisting of about
25% wt. xanthate in water at a pH of about 10.

71
Decomposition of ethyl xanthate in aqueous solutions with time is shown in Fig.
2.7. Decomposition is particularly quick in acidic solutions and, therefore, the liquid
xanthates are alkaline solutions. The rate of decomposition of such highly concentrated
and alkaline xanthate solutions is approximately in the % /day range. Higher
homologues are more stable in water than the lower.

Fig. 3.9. Kinetics of decomposition of ethyl xanthate in acid solutions as indicated by


changes in optical density of solutions with time [Iwasaki & Cooke, 1958].

Fig. 3.10. Flotation of chalcocite with potassium salts of various xanthates

72
As Fig. 3.10. shows [K.L. Sutherland and I.W. Wark, Principles of Flotation,
1955], with increasing hydrocarbon chain length xanthates become more powerful.

Fig. 3.11. Effect of oxygen concentration on the flotation of galena [Bessonov and
Plaksin, 1954].
Xanthates are powerful collectors for sulfides only in the presence of oxygen
(under slightly oxidizing conditions).
The solubility of sulfides is so low that these minerals might be assumed inert.
In principle, such sulfides are in reversible equilibrium with aqueous metal and sulfide
ions, according to a definite solubility products (K SP ) which are extremely low (for
instance, for galena, K SP = 10-28). But in the presence of dissolved oxygen galena
undergoes superficial oxidation with various oxidation products (lead hydroxide, lead
thiosulfate, lead sulfate, ect) deposited on the surface. As sulfides are electronic semiconductors, their oxidation processes take on the character of electrolytic corrosion
reactions, certain (anodic) areas oxidize preferentially while other are cathodic, as in the
rusting of iron.
Anodic:
Cathodic:
Overall reaction:

MeS = Me2+ + So + 2e
1/2O 2 + H 2 O +2e = 2OH-

(3.4)
(3.5)

MeS + 1/2O 2 + H 2 O = Me 2+ + So + 2OH- (3.6)

Presence of various oxidation products (So oxidizes further into S 2 O 3 2-, SO 4 2- , for
example Fe2+ to Fe3+, etc.) on the surface makes identification of PDI ions very
difficult. Oxidation of sulfide minerals can be manipulated by controlling the redox
potential (Eh) of the pulp during processing (grinding and flotation).

73

Fig. 3.12. Schematic corrosion reactions on sulfide surface.


As a result of oxidation, elemental sulfur appears on the surface and this renders such a
sulfide hydrophobic (so-called self-induced flotation). It was shown that the potential at
which flotation commences is close to that at which anodic oxidation of the mineral,
occurs. Some minerals (for instance chalcopyrite) are known to float quite well under
such conditions. Since the self-induced flotation results from the presence of elementar
sulfur on the sulfide surface, more intense oxidation leading to the oxidation of So to
S 2 O 3 2- or SO 4 -2 inhibits flotation.
In the flotation systems involving a sulfide mineral and xanthate in aqueous
solution, the cathodic reduction of oxygen
O 2 + H 2 O + 2e = 2OH-

(3.7)

is coupled with several possible anodic reactions:


MeS + 2X- = MeX 2 + So + 2e

(3.8)

but also by the oxidation of xanthate ions to dixanthogen (and its adsorption) at the
mineral/water interface:
2X- = X 2 + 2e
(3.9)
and also reactions in which sulfur is further oxidized into S 2 O 3 2- or SO 4 2-.
As a result of all these reactions either MeX 2 or X 2 (dixanthogen) or more
likely both appear on the mineral surface.
While some sulfides such as chalcopyrite, chalcotite, galena respond very well to
xanthates, some minerals (e.g. sphalerite) do not float with xanthates without previous
activation (Table 2.1). Iron sulfides (pyrite, marcasite, pyrrhotite) float well with
xanthates but only in acidic environment.
Table 2.2. [K.L. Sutherland & I.W. Wark, Principles of Flotation, 1955]

74

Sphalerite which does not float with ethyl-amyl xanthates, can be floated with the
higher homologues (Fig. 3.13).

Fig. 3.13. Effect of selected xanthates on flotation of sphalerite [K.L. Sutherland & I.W.
Wark, Principles of Flotation, 1955]
Dithiophosphates, containing =P(=S)S functional group are manufactured by
American Cyanamide under the trade name of AEROFLOATS (or AERO Promoters)

Their use as metallic and sulfide mineral flotation collectors has a history that is almost
equal to that of the best known and most widely used collectors, the xanthates.
There is a variety of both alkyl and aryl dithiophosphates available: from diethyl,
diisopropyl, diisobutyl, diisoamyl to diaryl obtained by reacting cresylic acid with P 2 S 5 .

75

The dialkyl dithiophosphates have little or no frothing properties until the chain length
reaches about 5 (amyl). The diaryl dithiophosphates display light to moderate frothing
properties owing to the phenols from which they are made.
Dithiosphaphates as thiol collectors have similar applications as xanthates, i.e. as
collectors for metallic and sulfide base metal ores and ores containing precious metals.
The main characteristics of these collectors include:
1. They are particularly effective for improving the recovery of both precious
metals and basic metals;
2. They are more selective collectors than most other thiol collectors, particularly
against pyrite in alkaline environment;
3. Their selectivity is particularly useful in the treatment of polymetallic ores;
4. They can increase flotation rates.
Dithiophoshates (DTP) are weaker collectors in comparison with xanthates and
dialkyl thionocarbamates and they cannot replace a less selective collector type.
The most successful method of application of DTPs is with another collector type to
enhance the collecting properties of both. The two collector types that are most widely
used in conjunction with DTP are xanthates and sodium mercaptobenzothiazole (NaMBT). Such usage frequently results in reduced total collector consumption and/or
improved metallurgical results. Often, the DTP is used as the principal collector in a
rougher flotation stage to maximize selectivity, followed by a scavenger flotation stage
with the use of xanthate to maximize recovery. When DFP is utilized with Na-MBT the
most common practice is to dose both collectors together at the same addition point(s),
and, if they are staged added, to use a constant proportion of each. In neutral to acidic
circuits the DTP/Na-MBT combinations are powerful general-purpose sulfide collectors
(for instance in a bulk flotation process). In alkaline circuits this combination is more
selective.
The collectors primarily designed for sulfide minerals are not necessarily
selective towards precious metals. The thio-collectors used in treating polymetallic
sulfide ores in most cases require a highly alkaline environment for attaining optimum
selectivity. On one hand, in such plants the cost of lime may account to 50% of the
overall reagent costs, on the other high lime dosages may depress gold flotation (and
also flotation of fine molybdenite in the Cu-Mo sulfide ores). Dithiophosphates were
shown to be more efficient in recovering precious metals (e.g. silver) in sulfide
flotation. Monothiophosphates which have the =P(=S)O functional group are claimed
[Nagaraj et al, 1991] to specifically float silver and gold. The chemistry of these
collectors is shown below:

76

The Aerophine 3418A promoter (diisobutldothiophosphinate) is claimed to have a


particular affinity towards silver and silver sulfides. Used with dithiophosphate as a
secondary collector was shown to be advantageous collector combination. [P.A.
Mingione, 1991]. Much better recoveries of Au in the Cu concentrates were obtained
with this collector in the flotation of a massive sulfide ore containing 0.67% Cu an 0.7
g/t Au with the principal gangue being pyrrhotite.

Fig. 3.14.The differences in collector activities between monothiophosphates,


monothiophospinates and dithiphosphates (all iso-butyl homologues) as a function of
pH for a Western U.S. copper ore; flotation timer 7 min [D.R. Nagaraj et al., Processing
of Complex Ores, Pergamon Press, 1989, pp. 157-166].
Dicresyl monothiophosphate and diisobutyl monothiophosphate were found to
significantly increase gold recovery. The results with dicresyl monothiophosphate
(Fig. 3.15.) show that while pyrite recovery was consistently maximum in the acidic pH
region in agreements with Fig. 3.14, gold was selectively floated in the alkaline region.

77

Fig. 3.15. Gold and pyrite recovery with 50 g/t of DCMTP as a function of pH;
Western U.S. primary gold ore [D.R. Nagaraj et al., 1990].
It was later on reported that since many free gold containing ores have a silver content,
the DCMTP collector is applicable to such ores only because of their high silver
content.
Thionocarbamtes are reaction products of primary (or secondary) amines with CS 2 in
the presence of alkali hydroxide:
RNH 2 + CS 2 + NaOH = RNHCSSNa + H 2 O

(3.10)

The alkyl thionocarbamates are colorless liquids, or solids, insoluble in water. Because
of their poor solubility they are commonly added to a mill. But thionocarbamates are
soluble in organic solvents and can be mixed with frothers in any proportion to aid
feeding. Perhaps the best known is Dow Chemicals Z-200 (ethyl isopropyl
thinocarbamate). They are stable in acid circuits and are used over a broad pH range.
Klimpel showed that thionocarbamates are good collectors for flotation of native gold.
Xanthogen formats were developed to replace xanthates in acid circuits. They are used
in copper-arsenopyrite and copper-zinc circuits where moderately low pH flotation is
mandated. Diethyl xanthogen formate formula is given below as an example:
CH 3 CH 2 O(C=S)S(C=O)OCH 2 CH 3
Mercaptobenzothiazole has been manufactured by Cyanamid for years as
supplementary collector, also for oxidized ores.

2.5.Non-thio collectors
These include anionic taffy acids, alkyl sulfonates and sulfates,

78
and cationic amines: primary, secondary and quaternary.
While fatty acids and primary amines are weak electrolytes (their
dissociation depends on pH), the quaternary amines (e.g. dodecyltrimethylammonium
bromide, DTAB) and alkyl sulfates and sulfonates are strong electrolytes (they are
totally ionized irrespective of pH). Fatty acids are important flotation collectors. They
are utilized in flotation of many minerals bearing Me2+ cations (like Ca2+, Mg2+, Ba2+,
etc.). These minerals include apatite, fluorite, calcite, dolomite, scheelite, barite, etc.
Amines, the cationic collectors, are used in flotation of silicates (e.g. quartz)
since these mineral grains assume negative electrical charge in aqueous environment.
Long-chain primary amines are heavily used n flotation of potash ores (as a collector for
sylvite)..
2.6. Oily collectors
This group includes water-insoluble hydrocarbons, known as oils, exemplified
by Diesel oil, kerosene, etc. Because of poor solubility of these compounds in water
they have to be emulsified and their appear in flotation pulps in the form of fine oil
droplets. Emulsifying agents (strong surfactants which adsorb at the oil/water interface
reducing the oil/water interfacial tension) are commonly applied to facilitate
emulsificaction of the oily collectors.
When oil is dispersed into fine droplets in an aqueous suspension of
hydrophobic particles, the oil droplets will tend to attach to the hydrophobic particles. If
conditioning is sufficient and doses of oil are high enough such particles will be
agglomerated by the attached oil droplets. This process is referred to as oil
agglomeration. The oil agglomeration is the most common way of increasing the size
of fine hydrophobic particles. Because of poor probability of collision of very fine
particles with bubbles, the fine particles do not float very well (even if hydrophobic),
and increasing their size by oil agglomeration can dramatically improve their flotation.
Since oil droplets can attach only to the particles that are to some extent
hydrophobic, oily collectors are commonly used in the flotation of inherently
hydrophobic minerals (e.g. molybdenite, graphite, talc, sulfur; also coal will be
classified here in spite of the fact that it is not a mineral).
Attachment of oil droplets to mineral surfaces. Thermodynamic analysis of the
attachment process, shown in Fig. 3.16, leads to formula 3.11.
Gattach = ow (cos 1)

Fig. 3.16. Attachment of oil droplet to solid particle in water.

(3.11)

79

The result is that the attachment of an oil droplet to a mineral particle is likely to
take place only if > 0; for = 0, cos = 1 and Gattach = 0, the attachment is
impossible.
Table 3.3. Characterization of separation processes in which oil is utilized (after Finkelstein)
Technique
Ionic
Oily collector
Oil consumption Conditioning
Method of
collector
(kg/t)
separation
Extender
Yes
Yes
0.05-0.5
Regular
Flotation
flotation
Agglomeration
flotation

Yes

Yes

a few kg/t

Intense

Flotation

Emulsion
flotation

No

Yes

up to a few kg/t

Regular

Flotation

Oil
agglomeration

Yes

Yes

5-10%

Slow/intense
shearing

Sizing

Liquid-liquid
extraction

Yes

Yes

N/A

Intense

Phase separation

This explains why oily collectors are only used in the flotation of inherently
hydrophobic minerals. In the extender flotation process, a collector which will render
the mineral surface hydrophobic is needed first before adding oil.
The first process in table 3.3, extender flotation, utilizes small doses of an oil
(up to a few hundred grams per ton of ore) to enhance the collector, and/or to reduce its
consumption. Extender oils are frequently used to improve flotation of coarse particles.
This process is utilized to float coarse potash ore fractions.
In agglomerate flotation, which was invented to treat very fine particles that are
not recovered efficiently in conventional froth flotation, oil is added to increase the size
of particles by agglomerating them. The method involves conditioning at high pulp
density first with an appropriate collector to render the particles hydrophobic and then
with an oil to bridge particles together. Pulp dilution and conventional froth flotation
follows.
Emulsion flotation is applied to treat inherently hydrophobic solids (coals,
graphite, sulfur, molybdenite, talc). In this process, water-insoluble oil is used as a
collector. Both agglomerate flotation and emulsion flotation use larger quantities of oil
than extender flotation (up to a few kg/t).
The first three processes in Table 2.3 belong to a family of flotation processes,
the following two do not. Liquid phase agglomeration, as the process is termed by
Capes, can be used to beneficiate fine coal by agglomerating and separating by
screening the hydrophobic agglomerates out of the fine hydrophilic gangue.
In all these processes, the second liquid added under appropriate hydrodynamic
conditions must be immiscible with the suspending liquid (e.g. water), and be capable
of displacing this liquid from the surface of the treated hydrophobic particles.

80
The mechanism of bubble attachment to the particle via the oil droplet on its
surface is shown in Fig. 3.4. In the final stage (Fig. 3.4D) the oil recedes into a thread
along the solid/water/air interface (Klassen) increasing remarkably the strength of the
particle-bubble aggregate. This explains why extender oil is commonly applied to
improve flotation of coarse particles.
2.7. Modifiers
A.
B.
C.
D.

Activators
Depressants
Dispersants
pH regulators

Activators. These are readily soluble in water electrolytes whose ions react with the
mineral surface. Their action alters the chemical composition of the mineral surface.
EXAMPLES. Of all sulfide minerals sphalerite exhibits the poorest floatability with
thio-collectors. The activation of sphalerite with copper salts (CuSO 4 ) is an important
practical example. It is generally believed that activation involves exchange between
zinc atoms in the mineral surface and copper ions in solution:
ZnS + Cu2+ CuS + Zn2+

(3.12)

The driving force for this reaction is the lower solubility of copper sulfide from
zinc sulfide. The activated sphalerite behaves in a similar manner to copper sulfide in its
flotation properties. Other metallic ions, such as Pb2+, Ag+ , etc. also activate sphalerite.
Sulfidization. Oxidized non-ferrous metal minerals can be activated by sulfidization. For
example cerussite (PbCO 3 ), smithsonite (ZnCO 3 ), malachite (CuCO 3 ) and other
minerals which belong to this group can be activated with Na 2 S, or NaHS:
PbCO 3 + Na 2 S PbS + Na 2 CO 3

(3.13)

It must be, however, remembered that sodium sulfide and sodium hydrosulfide are
strongly reducing agents and at high concentrations they depress flotation of
sulfides with thio-collectors.

81
Fig. 3.17. Effect of flotation conditions on recovery of malachite sulfidized with 200
mg/L of Na2S.9H2O. Frother (amyl alcohol) conc. 60 mg/L. Curve 1, with decantation
and change of solution after sulfidization; curves 2, with oxygen bubbled through for 25
min under constant agitation and aeration conditions of 120 cm3 O 2 /min; curve 3,
directly after sulfidization [Soto & Laskowski, Trans. IMM, Sec. C., Vol. 82 (1973)].
Sodium sulfide solutions hydrolyze in water and form free H 2 S which is highly
toxic. Sulfidization must therefore be carried out in alkaline solutions.
Fig. 3.18. Concentration diagram
for NaHS/Na2S solutions.

Na 2 S +2H 2 O
H 2 S + 2NaOH

As Figure 3.18 shows, on molar basis, below pH 5 there are no sulfide ions present, and
at pH 5 only 10% of the H 2 S has ionized to HS-. At pH 7, half of H 2 S has ionized to
HS-, while between pH 8 and 11, over 90% of the sulfide ions are HS-. At pH 12 molar
concentrations of HS- and S2- are equal.
Quartz, if pure, cannot be floated with anionic collectors. It will, however,
respond nicely to anionic collectors if it is treated with solutions of heavy multivalent
cations.
Metallic ions in aqueous solutions (e.g. Cu2+, Pb2+, Co2+, etc.) hydrolyze. For
example total concentration of cobalt in the solutions is given by the equation:
Co (T) = Co2+ + CoOH+ + Co(OH) 2(aq) + Co(OH) 2s

(3.14)

where CoOH+ is the first hydrolysis complex, Co(OH) 2aq is cobalt hydroxide dissolved
in the solution, and Cu(OH) 2s stands for the portion which is insoluble when the
Co(OH) 2 concentration exceeds its solubility in water. The proportion between various
forms is determined by pH as shown in Fig. 3,19 [James and Healy, 1972].

82

Fig. 3.19. Hydrolysis products of Co2+ as a function of pH. Experimental adsorption


isotherm for Co2+ at 1.2 x 10-4 M on silica at 25 oC is also shown.
As this figure demonstrates, Co2+ ions do not adsorb onto silica, and only when
CoOH+ complexes appear in the solution the adsorption dramatically increases.
This is a general phenomenon, the hydroxy-complexes resulting from hydrolysis (e.g.
CuOH+, PbOH+, FeOH2+, Fe(OH) 2 +, etc.) are very surface active [T.W. Healy, 1973].
Fig. 3.20, taken from M.C. Fuerstenaus publications, shows that although
quartz, as expected, does not float with anionic collectors, in the particular ranges in
which metallic ions present in the system hydrolyze the flotation becomes very good.

Fig. 3.20. The pH ranges of quartz flotation with sodium alkyl sulfonate (10-4 M)
in the presence of different metallic ions (10-4 M) [after M.C. Fuerstanu and B.R.
Palmer, 1976].

83
Activation with copper salts is used in South Africa to enhance flotation of iron
sulfides bearing platinum group elements. It was also demonstrated that the presence of
Cu ions enhances sulfidization and flotation the oxidized pentlandite/pyrrhotite ores
[Newell & Bradshaw, Minerals Eng., vol. 20, 1039-1046 (2007)].
Depressants.
Table 3.4. Classsification of dispersing and depressing agents
Dispersants
Depressants
Water glass, Phosphates and
NaCN, K 4 Fe(CN) 6 , Na 2 S,
1a Polyphosphates (e.g. hexameta2a
NaHS, Na 2 CrO 4 , ZnSO 4
phospohate, tripolyphosphate)
Dextrin, Carboxymethyl
Dextrin, Starch, Guar Gum,
1b Cellulose, Guar Gum
2b
Carboxymethyl Cellulose,
Polyacrylates (e.g. Cataflot,
Quebracho, Tannic acid.
Dispex, etc.).

Dispersants are used in mineral processing to prevent fine particle aggregation.


Depressants are used to inhibit flotation of a given mineral, thus, the function of a
depressant is the opposite of that of a collector. This can be achieved either by (a)
preventing collector from adsorbing onto a given mineral, or by (b) making the mineral
surface hydrophilic. As seen from Table 3.4, group 1a of dispersing agents and group 2a
of depressing agents contain simple inorganic compounds. Groups 1b and 2b contain
organic polymeric compounds, and in some cases the same polymers are used as either
dispersants or depressants. While the depressants which belong to group 2a prevent
collectors from adsorbing, the depressants in group 2b, large hydrophilic
macromolecules, make the surfaces on which they adsorb hydrophilic.
EXAMPLES of depressants. While sphalerite does not adsorb short-chain xanthates,
sphalerite activated with copper (or lead) ions resulting from very slight dissolution of
copper minerals present in the ore may cause unintentional activation. Sodium cyanide
is added to complex copper ions in solution (cyanide ions interact strongly with cations
of many metals forming insoluble cyanide salts)
2Cu2+ + 4CN- 2Cu(CN) 2 2-

(3.15)

Sodium or potassium cyanides are utilized. They hydrolyze in aqueous solutions,


forming free alkali and hydrogen cyanide (HCN) which is relatively insoluble in water
and extremely toxic (NaCN + H 2 O = HCN + NaOH). Since even traces of HCN are
highly dangerous, cyanides must be used only in an alkaline medium (see Fig. 2.20).

84

Fig. 3.21. Logarithmic concentration diagram for 10-2 M HCN or NaCN.


HCN = H+ + CN-

(3.16)

[ H + ][CN ]
= K a = 4.8 x10 10
[ HCN ]

(3.17)

-log[H+] = -logKa + log[CN-]/[HCN]

(3.18)

at

[CN-] = [HCN]

pKa = pH = -log 4.8x10-10 = 9.32

(3.19)

From 2.23-2.24

[CN ] 4.8 x10 10


=
[ HCN ]
[H + ]

(3.20)

and the [CN-]/[HCN] ratio can be calculated. This ratio is equal 1 (that is the
concentration of CN- is equal the concentration of HCN at pH = 9.32 (eqs. 3.18-3.19)
80% will be in the CN- form ([CN-]/[HCN] = 4) at a pH of 9.92,
and 20% will be in the CN- from ([CN-]/[HCN] = 0.25) at a pH of 8.71.
This is shown in Fig. 2.20.

85

Fig. 3.22. Percentage of CN- in cyanide solutions as a function of pH.


As Figs. 3.21 and 3.22 indicate, the tests with cyanides must be carried out at pH
exceeding 10.
In various quantities, cyanides are capable of depressing the flotation of sulifies
of zinc, copper, iron, silver, nickel, etc.
Sodium sulfide and sodium hydrosulfide are very strong reducing agents and can
entirely prevent xanthates from adsorbing onto sulfides. The are commonly used at high
concentrations to depress copper sulfides in the selective flotation of Cu-Mo bulk
concentrates.
Chromium salts (K 2 CrO 4 , K 2 Cr 2 O 7 ) are important depressants for galena.
In the flotation of Cu, or Cu-Pb sulfide ores containing sphalerite, SO 2 was
found to be a very selective depressant of sphalerite. This was shown to prevent
undesired activation of sphalerite by Cu2+ ions. Compared to cyanides, SO 2 can be used
without a detrimental effect on copper sulfide and precious metals flotation.
Polymeric depressants. Perhaps the best known depressants which belong to this group
are polysaccharides. It is known that dextrin depresses efficiently flotation of coal.
Other polysaccharides, e.g. guar gum and carboxymethyl cellulose, found applications
as depressants of talc and graphite.
OH
C
6

H
H
C4

C5
H

OH

C
3

C2

OH
H
H
C4
OH

C
6

H
C1

OH

OH
H

C5
H

OH

C
3
H

C2
OH

C1

C4
O

C
6

OH
H

C5
H

OH

C
3
H

C2
OH

C1

C4
O

C
6

C5
H

OH

C
3
H

C2

C1

C4

OH

E. Amylopectin and dextrin

C
6

C5
H

OH

C
3
H

OH

C2

H
C1

86

OH
C
6

OH

C5
H

OH

C
3
H

OH

C4
H
OH
H
H
O

C4

C
6
C5
H
OH
C
3
H

OH

H
H
C1

H
H
OH C1
C2
H

C4

C5
H

OH

C
3
H

OH

H
O

C
6

OH
C4

C2

C
6
C5
H
OH
C
3
H

OH
H

H
O

OH C1
C2
H

C4

C
6
C5
H
OH
C
3
H

H
H
C1

C2
O

C
6

C5
H

O
OH C1
C2
H

C4

OH
C
3
H

H
O
OH C1
C2

G. Guar Gum

Fig. 3.23. Chemical formulas of amylopectin (amylopectin and amylose are constituents
of starch), dextrin, and guar gum.
As seen from 3.23, macromolecules of these compounds contain many OH groups and
are very hydrophilic.
As shown by Q. Liu and J.S. Laskowski (1989), dextrin is not adsorbed by
quartz. However if some metallic ions, for example, lead ions, are adsorbed onto quartz
then adsorption of dextrin on such activated quartz is very high. This is shown in Figure
3.24.

Figure 3.24. (a) Effect of pH on adsorption of dextrin on hydrophilic quartz (Q),


methylated quartz (MQ) and lead-activated quartz (PbQ, MPbQ). (b) Adsorption
isotherms at pH 10. Q.Liu and J.S. Laskowski, IJMP, 26, 297-316(1989); 27, 147-155
(1989).

87

Figure 3.25. Effect of pH on adsorption of dextrin on litharge (PbO).


Adsorption mechanisms

Me

Me

OH

+ H2O

HO
Acting as an acid

Acting as a base
48

Figure 3.26. The postulated mechanism with acid-base interactions playing major role.

88

Figure 3.27. Effect of pH on adsorption of dextrin on galena and chalcopyrite.


For copper sulfides, chalcopyrite and chalcocite, the adsorption of dextrin does
not really depend on pH. It is, therefore, possible to select the pH range in which
adsorption of dextrin will be high on galena and negligible on chalcopyrite. This should
allow for selective depression of galena and flotation of copper sulfides.

Figure 3.28. Effect of pH on flotation of chalcopyrite and galena


[Liu & Laskowski, IJMP, Vol. 27, 147-155 (1989)]

89

100

MINERAL RECOVERY IN FROTH, %

pH = 10.7
KEX = 1.25x10-5 mol/L

CHALCOPYRITE

80

DEXTRIN = 0.5 mg/L


60

40

20

GALENA

0
0

1
2
3
4
CONCENTRATION OF CaCl2, 10-3 mol/L

Figure 3.29. Effect of Ca2+ ions on flotation of chalcopyrite and galena in the
presence of dextrin [Liu & Zhang, Minerals Eng., vol. 13, 1404-1416 (2000).

METAL RECOVERY IN FROTH, %

100

80

60

pH = 10.7

CHALCOPYRITE

KEX = 1.25x10-5 mol/L


40

CaCl2 = 2.7x10-3 mol/L


DEXTRIN = 0.5 mg/L
GALENA

20

0
0

0.5

1.5

2.5

CONCENTRATION OF CITRIC ACID, 10-3 mol/L

Figure 3.30. Effect of citric acid on flotation of chalcopyrite and galena in the
presence of dextrin in solutions of CaCl 2 .

90

Figure 3.31. Effect of pH on flotation of Cu-activated sphalerite and galena in


the presence of dextrin [Rath and Subramanian, 1999].

Figure 3.32. Effect of pH on adsorption of various dextrins on heazlewoodite (Ni 3 S 2 )


and chalcocite (Cu 2 S) [Nyamekye & Laskowski, 1991].

91

Figure 3.33. Effect of dextrin on copper concentrate grade in cleaning flotation of


INCO matte (Nyamekye and Laskowski, 1991).

Figure 3.34. Effect of dextrin addition on the grade-recovery curves for rougher
concentrate cleaning of INCO matte with dextrin.
Dextrin was used in Finland, Kotalahti Mine (1957-1982), in processing Cu-Ni
sulfide ore (the bulk chalcopyrite-pentlandite concentrate was selectively floated by
depressing pentlandite at alkaline pH using dextrin).

92
Another depressant in the same group is carboxymethyl cellulose (Fig. 3.35).

Fig. 3.35. Structural formula of carboxymethyl cellulose.


This is an anionic polymer with varying degree of substitution and varying molecular
weight. As Figure 3.33 indicates, three hydroxyl groups in each monomer are available
for the modification reactions, they are on the C-2, C-3 and C-6 positions, respectively.
Comparatively, the reactivity sequence of the three hydroxyl groups follow the order
C-6 > C-2 >C-3. In carboxymethyl cellulose the carboxymethyl groups replace the
protons in some of the hydroxyl groups in the cellulose. The number of these substituted
functional groups per monomer is called the degree of substitution (DS). As ca be seen,
the maximum degree of substitution of carboxymethyl cellulose is 3. In practice it is
almost always lower than 1.
Because of the presence of anionic (carboxylic) groups in the CMC chain, the
CMC macromolecules are very extended in water. At higher concentration of electrolyte
(higher ionic strength) the electrical charge is screened and this results in coiling of the
CMC macromolecules (Fig. 3.36). Adsorption of CMC onto minerals very strongly
depends on the ionic strength of the solution.
CMC is heavily used in South Africa in the flotation of pentlandite-pyrrhotite
sulfide ores containing platinum group elements to depress talc.

Fig. 3.36. Schematic representation of conformation of CMC macromolecules in


aqueous solution. A, in distilled water; B, in diluted electrolyte; C, in more concentrated
electrolyte.

93

Fig. 3.37. Flotation response of galena vs. pH as a function of sodium carboxymethyl


cellulose (R. Jin et al., Minerals & Metal. Processing, Vol. 4, 227-232 (1987).
CMC depresses strongly galena in alkaline pH range (Fig. 3.37) and Chinese
researchers have shown that CMC can also be used to depress galena in the selective
flotation of Zn-Pb sulfide ores.
Dispersants. As shown in Table 3.4, inorganic and organic polymeric compounds are
used as dispersants in mineral processing applications. These agents are used to prevent
fine solid particles from aggregation (coagulation).
Perhaps the most instructive example which illustrates industrial application of
dispersants in flotation processes was provided by Jowett and his co-workers [9].
Jowettt et al demonstrated that hydrophobic coal particles become completely
hydrophilic in the presence of clays which form a slime coating on coal surface (Fig.
3.38).

Fig. 3.38. Effect of clays and dispersant on wattability of coal surface in aqueous
solution [A. Jowett et al., Fueal, Vol. 35 (1956)].

94
Coal hydrophobicity can, however, be restored by simply adding a dispersant, sodium
hydrogen phosphate (today hexametaphosphate would be used). Both coal and clay
particles were found to acquire more negative zeta potential values in 50-100 mg/L
sodium hydrogen phosphate solutions (Fig. 3.39). This is then a clear example of
electrostatic stabilization; the specific adsorption of phosphate anions increases the
negative charge of the coal and clay particles, bringing about strong coulombic
repulsion which keeps clay particles apart from the coal surface thereby restoring coal
natural floatability.

Fig. 3.39. Effect of Na 2 HPO 4 on the zeta potential of coal and clay particles
[A. Jowett et al., Fuel, Vol. 35 (1956)].
In the DLVO (Deryaguin-Landau-Verwey- Overbeek) theory the total energy of
the interaction between two particles is given by:
VT = VE + VA

(3.21)

where V E is the energy of electrostatic repulsion (positive value means repulsion), and
V A is the Van der Waals attraction (negative values means attraction).
If the interacting particles have only small electrical charge, the electrical
repulsion energy, V E (2), is small and the total interaction energy for this case, V T (2)
exhibits no repulsion at any distance between the interacting particles. Such particles
will coagulate when they collide. If, however, the electrical charge of the particles is
increased then the electrical repulsion term will become V E (1) and the summation with
the V A curve gives the V T (2) curve. This curve shows the energy barrier which, if high
enough, will be impossible to overcome by colliding particles which will be repulsed
and eventually will form a stable colloid/suspension. This explains so called
electrostatic stabilization. Fine particle systems can be electrostatically stabilized by
increasing their electrical charge (e.g. by changing pH, electrolyte concentration, etc).
Figs. 3.38-3.41 demonstrate how the fine particle system can be electrostatically
stabilized by adding properly selected dispersant. The dispersants which belong to
Group 1A (Table 2.4) stabilize fine particle systems by electrostatic stabilization.

95

Fig. 3.40. Total interaction


energy curves, VT(1) and
VT(2) obtained by summation
of an attractive curve, VA,
with different repulsion
curves, VE(1) and VE(2).

The most common inorganic dispersants in use in mineral processing are water
glass and polyphosphates (Table 3.4 groups 1A). Solution chemistry of water glass is
not as simple as we have believed for very long. More recent publications indicate that
it undergoes hydrolysis and polymerization which is very sensitive to pH . As a result
various species appear in solution with polymeric species being the most active.
Water glass (sodium silicate) is probably one of the most widely used
depressing/dispersing agents. Its composition can be expressed by the general formula
of mNa 2 OnSiO 2 . The n/m mass ratio, referred to as modulus, vary from 1.6 to over
3.33. Sodium silicate in water undergoes hydrolysis which produces a number of
monomeric [Si(OH) 4 , SiO(OH) 3 -, SiO 2 (OH) 2 2-], polymeric [e.g., Si 2 O 3 (OH) 4 2-,
Si 2 O 2 (OH) 5 -, Si 3 O 5 (OH) 5 3-] and colloidal species. It is rather well established now that
in sodium silicate solution the active species for depression of flotation are polymeric
silicate species and small colloidal silica particles [W.Q. Gong et al., Int.J.Min.Proc.,
Vol. 39, pp. 251-273 (1993)].
Several procedures in which sodium silicate is reacted with different metallic
salts have been patented [V. Mercade, US Patent, 3,915,391 (1975)]. These procedures
are claimed to produce more active sodium silicate dispersant with increased amount of
the polymerized form. Water glasses with higher modulus is also known to be a stronger
depressant/dispersant.
Close similarity of the Ca-containing minerals makes separation of such
minerals by flotation extremely difficult. Although it is known that much lower oleic
acid concentrations are required to initiate flotation of fluorite than apatite or calcite,
these differences are not sufficiently large. It is also well established that calcite is
strongly depressed by starch and tannins. Sodium silicate and gum arabic are known to
depress calcite and dolomite but not apatite. Klassen and Mokrousov claim that small
additions of sodium silicate may even activate flotation of apatite with fatty acids. They
also shown that the flotation activity of sodium silicates can be altered by polyvalent
cations, and that the selective flotation of fluorite from calcite could be very much
improved by the use of sodium silicate reacted with Al3+ salts. Their results seem to
indicate that while sodium silicate does not depress apatite, and only slightly calcite, the
sodium silicate reacted with ferric salts depresses strongly calcite and weakly apatite.
Mg2+ treated sodium silicate depressed strongly calcite and almost not at all apatite.

96

The effect of sodium silicate on selective flotation of apatite from iron oxides
was recently studied by Gong et al. [Int.J.Min.Proc., Vol. 34, 83-102 (1992)]. The main
difficulty in the process was to reduce the iron content in the apatite concentrate. A key
feature turned out to be the use of sodium silicate as a selective depressant of
hematite/goethite. Since a too strong depressant was also affecting recovery of apatite,
they selected sodium silicate of medium modulus (2.06). The pH and time of standing
of the stock solution of sodium silicate played an important role. The silicate solution
stored at pH 9 remained clear and apparently stable for long periods and the selectivity
was the best. Since the addition of polyvalent metal salts into sodium silicate solution
affect the polymerization of the silicate species, these effects were studied as well. They
reported that all prepared silicate-metal salt mixtures, added to the flotation pulp before
collector, strongly depressed apatite and iron oxides. Silicate-CaCl2 mixtures depressed
apatite and iron oxides more strongly that silicate-AlCl3 mixture, but turned out to be
more selective in that they depressed iron oxides more than apatite. Gong et al.
concluded that sodium silicate is a selective depressant towards iron oxides in the
flotation of apatite in the pH range from 7.5 to 11. The selective action of sodium
silicate can be improved by adjusting variables such as the modulus, the pH and the
time of standing of the silicate solution, or by reacting the silicate solution with some
polyvalent metal salt solutions.
M.C. Fuerstenau and Fitzgerald (1989) reported that the sodium silicate treated
with Al3+ salts used as depressant in the flotation of porphyry and oxide copper ore, and
in a scheelite ore, resulted in increased by 1.5 to 5 percent flotation recoveries.
Polyphosphates. Polyphosphates are another important group of dispersing agents.
Linear polyphosphates form anions of the general formula [ Pn O3n +1 ]( n + 2 ) (i.e.
commercial product marketed under the name Calgon); tripolyphosphate and
hexametaphosphe are shown below as an example

The best known cyclic polyphosphate is hexametaphosphate

Another mechanism by which fine particle systems can be stabilized is steric


stabilization. The steric stabilization is a term which is commonly used to describe the

97
stabilizing mechanism involving adsorbed macromolecules. The precise dimensions of
polymer chains vary quite widely from one type of polymer to another at the same
molecular weight, and strongly depends on solvent. For polymers with molecular
weight greater than 10,000 the chain dimensions are comparable to, or in excess, the
range of the van der Waals attraction [D.H. Napper, Polymeric Stabilization of
Colloidal Dispersions, Academic Press, 1983]. Hence, polymer molecules are of the
right dimension to be used to impart colloid stability, provided that they can generate
repulsion. The steric stabilization results from the repulsion between adsorbed layers of
polymer macromolecules which prefer water (its interaction energy with water is
stronger than the interaction energy with the similar macromolecules). Practically it
means that only very hydrophilic polymers can function as steric stabilizers in aqueous
systems.
The steric stabilization can be visualized as creation of a physical barrier around
interacting particles which prevent them from approaching one another to a distance
where there is a significant attractive force (steric stabilization). In some cases both
effects, electrostatic and steric stabilization can operate together.

Fig. 3.41. Stabilization


of mineral suspensions;
(a) electrostatic; (b)
steric and (c) a
combination of
electrostatic and steric

The best stabilizers are amphipathic block copolymers. One of the comonomers
is nominally insoluble in the dispersion medium. The polymer that is nominally
insoluble in the dispersion medium would rather attach to the solid particle. The
nominally insoluble polymer thus serves to anchor the soluble moieties to the solid
particles. The role of the soluble polymer is to impart steric stabilization (A in the
picture below). Any polymer that is soluble in the dispersion medium is effective as a
stabilizing moiety and any polymer that is insoluble in the dispersion medium is
effective as an anchor polymer (B in the picture below). In practical terms when water is
used as a medium this means that the polymer used as dispersant must be strongly
hydrophilic. This topic will be further discussed in the chapter dealing with flocculation.
AAAAAAAAAAAAA
B
B
B
B

block or comb co-polymer


Perhaps the best known synthetic dispersants are polyacrylates of low molecular
weight (e.g. Cataflot manufactured by Pierrefitte-Auby Co), and low molecular weight
copolymers of the polyacrylate type (e.g. Dispex manufactured by Allide Colloids, now
CIBA). These dispersants are especially effective with dolomite/calcite minerals.

98
As shown in Table 3.4, polymeric depressants, which have already been
discussed, are also used as dispersants.
In general, polysaccharides are applied in mineral processing circuits as:
A. Depressants/flocculants in the separation of iron oxides from silicates;
B. Selective modifiers in the flotation of salt-type minerals (e.g. calcite,
dolomite, apatite, etc.);
C. Depressants of talc and graphite in flotation of sulfide ores;
D. Depressants of slimes of water-insoluble minerals in flotation of potash ores;
E. Depressants in differential flotation of polymetallic sulfide ores.
pH regulators. pH determines electrical charge of many minerals and ionization of
some collectors (fatty acids, amines). It affects strongly flotation of almost all minerals.
Lime (calcium hydroxide), sodium hydroxide, sodium carbonate (soda) are utilized to
increase alkalinity, and sulfuric acid to decrease pH.

99

Chapter 4
Table 4.1. Flotation classification of minerals (M.A. Eigeles)

Group

Examples

Flotation properties

A. Inherently
hydrophobic minerals

Graphite, sulfur, talc,


molybdenite, coals

Float easily with waterinsoluble oily


collectors and frothers

B. Sulfides and native


metals

Copper, gold, galena,


chalcopyrite,
pentlandite, sphalerite,
pyrite

Float easily with thiocollectors; depending


on redox conditions
some may exhibit selfinduced floatability

C. Oxidized minerals of Cerusite, anglesite,


heavy metals
cuprite, tenorite,
malachite, chrysocolla,
smithsonite

Float with thiocollectors following


sulfidization; also with
fatty acids

D. Oxides, silicates and


alumino-silicates

Quartz, rutile,
cassiterite, hematite,
alumina, clay minerals

Depending on pH float
with either cationic or
anionic collectors

E. Sparingly soluble
salts

Calcite, dolomite,
fluorite, apatite, barite,
magnesite, scheelite

Float well with fatty


acids and their salts

F. Soluble salts

Sylvite, halite

Long-chain primary
amines are used to float
selectively sylvite

100

Chapter 5a - Flotation of Sulfide Ores.


Theory of sulfide flotation
Development of the theory of thiol collector/sulfide mineral interaction began in
the 1930s. In 1957, Salamy and Nixon put forward the electrochemical theory and
proposed a way in which a neutral surface could be generated. An anodic oxidation
involving the collector (formation of dixanthogen, i.e., a dithiolate) transfers charge to the
mineral and the simultaneous cathodic reduction of oxygen returns this charge to the
solution phase. Most sulfide minerals are semi-conductors with a high conductivity and
hence can readily sustain such a process. In this mechanism the presence of dissolved
oxygen in water is a prerequisite for the flotation of sulfide minerals to take place..
The mixed potential mechanism
There are a number of ways in which a collector can interact with a sulfide
surface and to satisfy the requirements of a neutral surface species, how it had been
emphasized by Cook:
The reaction to form a metal compound can proceed either in a single stage (c), or
via separate processes occurring at different times (d). For example, the oxidation of the
sulfide surface to sulfate (sulfite or thiosulfate) could occur on exposure of the mineral to
air and these species could exchange with a thiol ion when collector is added (see
mechanisms (d) and (e)) (Table 5a.1)
Each of these mechanisms has been identified experimentally to occur in the flotation
systems.

Table 5a.1. Mixed potential mechanisms for the interactions of thiol collectors with
sulfide minerals
.
Comments
Mixed potential mechanisms
The anodic reaction can be Anodic: X X + e
(a)
ads

the chemisorption of a

thiol ion, (X-), as proposed Cathodic: O 2 + 2H 2 O + 4e 4OH


by Wark.
4X + O 2 + 2H 2 O 4X ads + 4OH

Anodic: 2X X 2 + 2e

(b)

Cathodic: O 2 + 2H 2 O + 4e 4OH
4X + O 2 + 2H 2 O 2X 2 + 4OH
Formation of a metal thiol Anodic: 2X + MS + 4H O MX + SO 2 + 8H + + 8e
2
4
2
compound, from the metal

component of the sulfide Cathodic: O 2 + 2H 2 O + 4e 4OH


mineral, as proposed by
2X + MS + 2O 2 MX 2 + SO 24

(c)

101
Taggart.

Anodic: MS + 4H 2 O MSO 4 + 8H + + 8e

(d)

Cathodic: O 2 + 2H 2 O + 4e 4OH
MS + 4O 2 MSO 4
Ion exchange reaction
coupled to mixed potential
mechanisms.

MSO 4 + 2X MX 2 + SO 24

(e)

Consequently, the hydrophobic coating on sulfide minerals could be formed by a


mixed layer of a metal-thiol compound and neutral molecules of dixanthogen. In same
cases the simple formation of dixanthogen explains the collecting action of thiol
collectors, e.g., the case of pyrite with xanthates.
The role of dissolved oxygen and pulp redox potential are key aspects to explain
flotation of sulfides.
Dixanthogen is a dithiolate resulting from the oxidation and dimerization of two
xanthate ions. It is an oily product of low solubility in water.
S S

C-O-CH2-CH3 +2e-

CH3-CH2-O-C

2CH3-CH2-O-C
S-Na+

Ethyl xanthate ions

(5a.1)

Ethyl dixanthogen molecule

Eh at which the oxidation of xanthate ions to dixanthogen molecules starts is given by the
following equation.
= E 0 0.059 log X
Eh

(5a.2)

Where, E is the standard potential for the couple (X 2 /X-); and X- denotes the xanthate
ions concentration. On the basis of eq. (5a.2), the Eh was plotted as a function of the
hydrocarbon chain length. It can be noted that xanthates and dithiophosphates are easier
to be oxidized as the number of carbon atoms in the alkyl chain increases. At the same
time, xanthate ions are easier to be oxidized compared with dithiophosphate ions.

102

0.7
Xanthates
Dithiophosphates

0.6

Eh, V

0.5

0.4

0.3

0.2

0.1

0.0
1

Number of C atoms

Fig 5a.1. Theoretical Eh of thiol ions oxidation as a function of the number of carbon
atoms in the aliphatic chain of alkyl xanthates and dialkyl dithiophosphates (10-4M).

Colassification of sulfide ores


To facilitate discussion the sulfide ores will be classified into:
(i)
Copper ores,
(ii)
Cu-Mo ores,
(iii) Cu-Pb-Zn ores,
(iv)
Cu-Ni ores,
(v)
Ores of Platinum Group Minerals,
(vi)
Flotation of native gold and auriferous ores,
(i)
Copper ores.
The most common copper sulfide minerals treated by flotation are chalcopyrite,
(CuFeS 2 ), followed by chalcocite (Cu 2 S), bornite (Cu 5 FeS 4 ) and covellite (CuS). All the
copper ores contain pyrite (or pyrrhotite), minor quantities of gold and silver, with a
gangue of quartz, silicates, and calcite.
In the case of porphyric copper deposits, primary and secondary copper sulfide
minerals can be recognized. Primary copper sulfide minerals occur at the deep zones of
the deposit, where non-altered species, such as, chalcopyrite, bornite and pyrite are
predominant. However, in the secondary zone, the minerals have been altered by surface

103
water circulation, resulting in sulfide minerals with high Cu content, such as, chalcocite
and covellite.
Although, in general, copper sulfides have similar flotation properties, they are
also very distinct. The selection of thiol collector can be made by mineralogical
considerations. So, for chalcopyrite and bornite, the best collectors are thionocarbamate
and xanthogen formates. This does not means that xanthates can not be used to float
chalcopyrite, but is not the optimum collector. In the case of chalcocite and covellite the
best formulation is a blend between iso-propyl xanthate and a dialkyl dithiphosphate.
In the same line, the behavior of copper sulfide depressants can also be different.
For instance, while chalcopyrite flotation is easily depressed by cyanides, chalcocite
depression requires much higher doses of cyanide. As chalcopyrite contains less than
36% Cu, concentrate grades in processing such ores contain less than 30% Cu.
For the great majority of Cu mills, pyrite must be rejected, so flotation is carried
out at a pH of above 9, generally with the addition of lime to the ball mill (but this
depends on whether the ore contains gold). Conditioning with aeration is helpful to
depress pyrite. Usually a strong collector (amyl xanthate) is used, or the more selective
isopropyl xanthate. The concentrate is upgraded in the cleaners by diluting the pulp down
to around 25% solids, and adding a more selective Cu collector (e.g. ethylisopropyl
thionocarbamate, Z-200, in combination with more lime to increase pH). Scavenger
concentrates are usually reground to improve both Cu recovery and grades.
Alkaline circuits are universal. According to Ronald Crozier (Flotation Theory,
Reagents and Ore Testing, Pergamon Press, 1992) only two plants: Sewell at El Teniente
in Chile and Lepanto in the Philippines employ acid circuits. In 1975 the average copper
concentrator treated ores with 0.782% Cu, of which 85.35% was recovered. Head grades
ranged from a low of half a percent copper in Canada, to a high of 1.455% in Chile, while
recoveries went from a high of 87.7% in Canada to a low of 80.4% in Europe. These
numbers must, however, be treated with caution since, due to the market situation, some
plants treating the lower ore grades shut down temporarily or permanently when the
copper price is low.
The frequency of use of collectors in copper mills (as single collectors or blends)
is given in Table 8b.3 after Crozier.
Table 5a.2.. Chemical structure of the main thiol collectors for sulfide minerals.
Thiol collectors

Chemical structure

Xanthates

S
R O C

Xanthogen formates

S
C

S- Na+ (K+)

O
C

CH2CH3

104
Thionocarbamates

Xanthic esters
Ditiophosphates

S
R O C N CH2CH3
H
S
R1 O C S R2
R O

S
P

R O

Mercaptobenzothiazol

+
S- Na
N
SH
S

Table 5a. 3. Collector use frequency comparison for copper mills,


% of total mills surveyed (after Crozier)

105
Copper sulfide flotation plants in Chile
The most important copper sulfide ore flotation plants in Chile are:
Minera Escondida (Los Colorados y Laguna Seca plants).
Codelco Norte (Chuquicamata plant)
Collahuasi
El Teniente
Andina
El Salvador
Los Pelambres
Candelaria
Las Trtolas
El Soldado
Mantos Blancos
Enami

Operation characteristics of copper ores flotation


As it was mentioned before, most of copper sulphide mills operate at alkaline pH.
Usually pH 10.5 is employed for secondary minerals (chalcocite and covellite); and
pH from 8.5 to 10.5 for primary minerals (chalcopyrite and bornite). Lime
consumption is aprox. 0.5 to 1.5 Kg/tonnes. Solid percent in rougher circuits ranges
from 35%-42%, and usual particle sizing in feed is around 20% +100 Tyler mesh.
Collector doses are around 20-40 g/tonnes; and frother around 10-15 g/tonnes, being
MIBC the most common, alone or in blend with DF-250.
In flotation plants with capacity larger than 50,000 tonnes/day, rougher and
scavenger circuits are usually designed with large volume flotation machines, ranging
from 1,500 to 8,500 ft3, the most common is the 4500 ft3 size for rougher circuits.
Rougher and scavenger retention time are around 25 to 32 min. Re-grinding of
rougher concentrates (around 70% -325 mesh) in vertical mills or ball mills is usually
employed. Te most common cleaner flotation machines are columns which operate at
pH 11-12 to depress pyrite.
When chalcopyrite is the predominant copper sulfide, the final Cu concentrate is
around 28%-30%Cu; however, when is chalcocite the concentrate grade is
significantly higher, ranging from 35%-45%Cu.

106

Fig. 5a.2.The simplest flowsheet for copper sulfide flotation, involving only one
cleaner step with column cells and a wide cleaner-scavenger.

107

Fig. 5a.3. Alternative flowsheet for copper sulfide ore flotation combining
conventional and column cells.

108

Fig. 5a.4. Schematic representation of a flowsheet with only conventional flotation


cells.

109

Fig. 5a.5. Schematic representation of a flowsheet with two column cell stages and
one conventional scavenger circuit.

(ii)
Cu-Mo ores.
The plants treating such ores produce two products: copper concentrate and
molybdenum concentrate. The separation process involves bulk flotation of both
copper sulfides and molybdenite (MoS 2 ) which is followed by a differential flotation
of molybdenite from depressed copper sulfides. Such ores contain 0.4 to 0.6 % Cu
(up to 1.6% in Chile) and 0.02-0.03% Mo. In spite of the very low content of
molybdenite, this mineral (because of its natural hydrophobicity) floats easily, and its
recovery varies from 50 to 70%. An example of Duval Sierrita Concentrator in
Arizona will be used here as an example. Fig. 8b.6 shows the bulk flotation at the

110
Duval Sierrita Plant (primary flotation circuit).

Fig. 5a.6(I). Primary flotation circuit at the Duval Sierrita Concentrator.


The ore contains three sulfides: chalcopyrite, molybdenite and pyrite. The primary
flotation circuit is a conventional roughing, cleaning, and recleaning operation in
which the rejects from both the cleaning and recleaning stages are returned to the
immediately preceding step (Fig. 8b.6). Typical operating conditions are given below:
Rougher flotation
Feed
Tail
Concentrate

0.32% Cu, 0.03% Mo


0.032% Cu, 0.005% Mo
7-9% Cu

111

Cleaner flotation
Rejects
Concentrate

0.2% Cu
18% Cu

Recleaner flotation
Rejects
Concentrate

1-3% Cu
25% Cu, 2-3% Mo

Typical recoveries are:


90% Cu
83% Mo
Flotation densities are normally maintained at:
Roughers
Cleaners
Recleaners

38-40% solids
10-12% solids
5-8% solids

The following reagents are utilized:


Reagent
Alkyl ester of amyl xanthate
Potassium amyl xanthate
Shell aromatic solvent
(ligh hydrocarbon oil)
Lime
MIBC

Addition point

Dosage

ball mill feed


ball mill feed

0.004 kg/t
0.0025 kg/t

ball mill feed


ball mil feed
cleaner reject
pump box

0.02 kg/t
1.25 kg/t
0.025 kg/t

The reagents used for the bulk Cu-Mo flotation are selected not only for
maximum Cu (and Mo) recovery, but also with consideration of their effect in the
subsequent copper-molybdenum differential flotation.

112

Fig. 5a.6(II). The differential flotation circuit at Duval Sierrita Plant.

113
Typical operating parameters in the molybdenum flotation circuit

Rougher flotation feed

2-4 % Mo

Rougher flotation concentrate

4-8% Mo

Rougher flotation tailings

0.15-0.20% Mo, 26% Cu

Cleaner reject

1-3% Mo

Cleaner concentrate

10-20% Mo

First recleaner reject

8-18% Mo

First recleaner concentrate

42-46% Mo

Second recleaner tailings

30-35% Mo

Second recleaner concentrate

46-50% Mo, 3% Cu

Molybdenum section recovery

90-94%

Rougher flotation density

25-39% solids

Cleaner flotation density

15% solids

First recleaner flotation density

10% solids

Second recleaner flotation density

1%

Flocculants were originally used in the molybdenum thickener to control


thickener overflow clarity. This was found to be detrimental to coppermolybdenum separation and its use was discontinued.

114

Surface properties of molybdenite


Molybdenite (MoS 2 ) has a leayered crystal structure and is characterized by inherent
hydrophobicity (similarly to graphite, talc, sulfur, etc.). The atomic structure consists of a
sheet of Mo atoms sandwiched between sheets of sulfur atoms. The Mo-S bonds are
strong covalent bonds whose rupture creates high energy surface sites called edges; but
the interactions between sulfur atoms at the top and bottom of separate sandwich-like trilayers, are week bonds (Van der Waals type), whose breakage forms low energy surface
sites called faces. The natural floatability of molybdenite is conferred by faces sites.
Therefore, the hydrophobicity of a given particle of molybdenite is the result of the
faces/edges ratio. Since molybdenite is an anisotropic mineral fine molybdenite particles
are more hydrophilic than coarse ones as they are haracterized by higher ratio of edges to
faces.
In plant practice, due to its natural floatability, molybdenite is floated only with oil
collector, such as, kerosene or diesel oil. However, when xanthates are used as copper
sulfide collectors, non-polar dixanthogen molecules can be formed on its surface
improving its floatability.

Fig. 5a.7. Side view of the layer structure of molybdenite (IMOA, International
molybdenum Association).

115

Fig. 5a.8. Microphotography of molybdenite particles.

The effect of lime on molybdenite flotation


The floatability of molybdenite depends on pH. It is well known that higher Mo
recoveries are obtained at low pH in the alkaline range. In commercial operations Mo
recovery decreases with pH, particularly when lime is used to adjust pH.
Calcium ion species depress molybdenite flotation in a similar way to depression
of pyrite. Hence, a special care must be taken in the cleaner stages, to avoid loses of Mo
as a result of lime excess.
Activation of molybdenite by sulfuric acid
It is a common practice in a number of moly plants to apply a pre-conditioning
step to the bulk Cu-Mo concentrates (moly plant feed) with sulfuric acid, adjusting the
slurry pH to pH 6.5 - 8.0. This procedure cleans molybdenite surface, leaching the
hydrophylic coating of calcium hydroxide, and restoring molybdenite hydrophobicity,
and its floatability in the subsequent Cu-Mo selective flotation stages. The risk of this
technology is the evolution of lethal H 2 S gas when NaSH is added to acid pulps, which

116
demands the installation of closed hermetic flotation machines, and abatement of the
excess of H 2 S gas emitted to its inner atmosphere.
The effect of polymers on molybdenite flotation
Minerals with natural floatability, including molybdenite, are depressed by
polymers, such as, dextrin, starch, humic acids, and flocculants (polyacrilamides, etc.).
Mechanisms of differential Cu-Mo separation
The fundamentals of the selective flotation of molybdenite involve the use of
strong Cu and Fe sulfide depressants, the conditions that do not affect the natural
floatability of molybdenite.
The most common Cu depressant used in Chilean moly plants is sodium sulfide
(Na 2 S), or sodium hydrosulfide (NaSH). When both are dissolved in aqueous solutions,
the predominant species in alkaline media is HS- ions, as can be see in the
thermodynamic diagram (Fig. 2.16). The HS- and S2- ions are strong reducing agents, i.e.
they are easily oxidized by air, particularly in the presence of mineral particles due to a
catalytic effect. The reducing properties of sulfide ions allow controlling the flotation
process by Eh measurement along the circuits. The oxidation of HS- ions is a side
reaction which increases depressant consumption in the aerated pulps. For this reason, in
a number of industrial plants, nitrogen was introduced to replace air as the flotation gas
phase in moly plants.
The main role of hydrosulfide ions is to induce desorption of thiol collectors from
the surface of Cu and Fe sulfide minerals. The chemical reduction of dixanthogen to
xanthate ions is another important function. Therefore, when desorption of all collector
species from the mineral surface takes place, the clean surface of sulfide mineral is
obtained that loses completely its induced hydrophobicity, and consequently does not
float. Different is the situation of molybdenite, because it exhibits natural floatability.
Oxidation of hydrosulfide ons in flotation of molybdenite
The oxidation of HS- ions usually results in the formation of thiosulfate ions
(S 2 O 3 2-) or sulfite ions (SO 3 2-).

117

2HS- +3H2O

2O2 +4H2O+8e

S2O32- +8H+ +8e-

(5a.3)

(5a.4)

8OH

2HS +2O2

S2O + H2O

HS- +3H2O
3
O2 +3H2O+6e2
3
HS- + O2
2

SO32- +7H+ +6e-

(5a.6)

6OH-

(5a.7)

SO32- + H+

23

(overall reactionl)

(overall reaction)

(5a.5)

(5a.8)

Cu-Mo Differential Flotation


Maximum production of marketable molybdenum concentrate is essential for
many copper processing plants. In the recovery of molybdenite from the bulk Cu-Mo
concentrate standard approach is depression of copper sulfides and pyrite. The list of
copper depressants used at different plants is very long but today the most important is
the use of sodium sulfide and hydrosulfide.
1.
2.
3.
4.
5.

Sodium sulfide and hydrosulfide (Na 2 S and NaHS)


Nokes reagent (P 2 O 5 dissolved in NaOH)
Arsenic Nokes (Anamol D) (As 2 O 3 dissolved in Na 2 S).
Sodium ferrocyanide and ferricyanide
Some plants (in USA) in the past also employed a thermal process to strip the
rougher flotation reagents from the copper concentrate prior to moly recovery.
The prolonged heating of the pulp with steam destroys xanthate collector and
allows molybdenite to be floated using an oily collector and a frother. Fresh water
is used throughout molybdenum circuit.

Sodium sulfide and hydrosulfide.


Two most important chemical reactions affecting HS- and S2- ions stability are
hydrolysis and oxidation.
The depressant action of HS- ions on Cu and Fe sulfides takes place by desorption
mechanisms of the collector attached to the mineral surface. In a simple way, the Kps of
CuS is significantly lower than the Kps of metal-thiol compounds formed on the mineral
surface, and a ion exchange reaction occurs when HS- ions are added to a flotation pulp,
releasing the thiol ions from the mineral surface to the aqueous phase, at the time that a
new coating of CuS is formed on the mineral surface. This desorption reaction depend on
HS- ions concentration in bulk solution, which must be controlled by on-line Eh
measurements. A special care must be taken when the HS- ions concentration decreased
in the pulp over a critical concentration (Eh), because the thiol collector can be readsorbed on the mineral surface and the floatability of Cu and Fe sulfides can be
restored).

118
Nokes reagent
This reagent is also based on sodium hydrosulfide. However, HS- ions are formed
as a result of a chemical reaction between P 2 S 5 and NaOH. Additionally, thiophosphate
ions are produced (mono, di or tri-thiophosphate ions), which reinforce the depressant
action of hydrosulfide ions. The mechanism involved in the chemical reaction of Nokes
reagent (also know as LR-744, and more recently as Tiofos) is described in the following
equations. The amount of HS- ions generated for the reaction is in many cases considered
insufficient, and blend with fresh hydrosulfide ions is successfully used in some plants.
6NaOH + P2S5 2Na3PO2S2 + H2O + H2S + heat

5a.9

H2S + NaOH NaHS + H2O

5a.10

P2S5 + 10NaOH Na3PO2S2 + Na3PO3S + 2Na2S + 5H2O

5a.11

Arsenic Nokes (Anamol-D)


It is the product of reaction of an excess of HS-ions with 10%-20% of arsenic
trioxide. In the final reagent the predominant species are HS ions, and the co-product are
thioarsenate ions, whose depressant role is similar to that described for thiophosphate
ions in Nokes Reagent.
As2O3 + 3Na2S + 2H2O Na3 AsO2S2 + Na3 AsO3S + 4H+

(5a.12)

As2O3 + 3Na2S + 2H2O Na3 AsO4 + Na3 AsOS3 + 4H+

(5a.13)

The most common is the use of Na 2 S (or NaSH) to depress copper sulfides in the
differential of flotation of Cu-Mo bulk concentrate. These compounds are very strong
reducing agents, (S2- and HS- ions are oxidized by oxygen and this reduces the oxygen
concentration in the pulp). In the differential flotation with the use of these agents the
most important is maintaining of the NaSH steady concentration in the system ,
pH must be maintained around 9. In an open system the consumption of NaSH may be as
high as 10 kg/t (or more). The use of flotation columns and closed environment with the
flotation carried out with N 2 allowed the Lornex Plant (now Highland Valley Copper,
B.C., Canada) to reduce the consumption of NaHS by 60%.
The Highmont Valley Plant treats about 90,000 t/d of ore which contains 0.35%
Cu and 0.015 Mo. The bulk Cu-Mo concentrate contains 29-30% Cu and 1.1 % Mo. The
final products are: copper concentrate 30% Cu, 0.03% Mo, and the moly concentrate
that contains 53.5% Mo and 1.5% Cu (which is further upgraded to 54.5% Mo by
leaching with ferric salts and H 2 SO 4 ).

119
Moly plants in Chile
The following are the main moly plants in Chile:
1. Codelco-Chile Divisions:
a). Codelco Norte (Chuquicamata)
b). El Teniente
c). Andina
d). El Salvador
2. Los Pelambres
3. Las Trtolas (Sur Andes)
4. Valle Central
5. Collahuasi
Table 5a.4. Molybdenum production in Chilean plants in 2004.
Plant
Codelco Norte

Tonnes of Mo
24,.271

El Teniente

3,919

Andina

2,980

El Salvador

1,154

Los Pelambres

7,853

Sur Andes (Las Trtolas)

1,706

TOTAL

41,883

(Ref. International Molybdenum Market, COCHILCO-CHILE, August 2005)

Description of flowsheets
In this section, a brief description of selective flotation of molybdenite in Chilean
plants is presented. The moly plants are commonly fed from a Cu-Mo bulk concentrate
thickener. One of the functions of this thickener removal of residual flotation reagents by
partial replacement of process water by fresh water. Frequently, sulphuric acid is added
to lower pH to 7.5 - 8.5 to the thickener discharge, which is the feed to the rougher circuit
of moly plant.
The most used Cu depressant is sodium sulfide and hydrosulfide, in doses ranging
from 4 to 9 Kg/ton of Cu-Mo bulk concentrate. In order to decrease the oxidation of HSions during flotation, and reduce depressant consumption, an inert gas (nitrogen) is
frequently employed as flotation gas. However, nitrogen gas is easier to apply when
molybdenum plant is located closed to a copper concentrate smelter, where nitrogen uses
to be a by-product of the oxygen plant. The alternative technology is the employment of
closed hermetic flotation cells, with recycling oxygen exhausted air as flotation gas (low
oxygen air).

120
The size of flotation conventional cells more frequently used is 500 ft3 for rougher
circuits, and 300 ft3 for cleaner stages (usually the first cleaner). Forced air flotation cells
are recommended for rougher circuits. Moly plants are usually designed with 4-5 cleaner
stages, column cells are frequently used in the last two cleaner circuits.

Table 5a.5. Depressan reagents, doses and gas employed in Chilean moly plants
Molybdenum
Depressant
Doses, Kg/tonne Flotation gas
plant
reagent
alim.
Codelco Norte
Sodium
6.5
Nitrogen
hydrosulfide
El Teniente
Tiofos
8.5
Nitrogen
(Nokes + NaSH)
Andina
Sodium
5.0
Low oxygen
hydrosulfide
air
El Salvador
Anamol
7.1
Nitrgeno
(NaSH + As 2 O 3 .)
Los Pelambres
Sodium
5.0 Kg/ton
Nitrgeno
hydrosulfide
Las Trtolas
Sodium
----Low oxygen
hydrosulfide
air
(H 2 S)
Valle Central
Tiofos
----Low oxygen
(Nokes + NaSH)
air

121
Bulk Cu-Mo
concentrate

C6/C7

H2O

H2SO4

Rougher
H2SO4

Final Cu concentrate
Filter Plant
Re grinding

C8
1stCleaner
H2O
dilution

Scavenger

Thickener tank

2ndCleaner

3rdCleaner

4thCleaner

Re grinding

5thCleaner

40 ft

Final Mo concentrate

Fig. 5a.9. Codelco Norte (Chuquicamata) moly plant.

122
Bulk Cu-Mo
concentrate
P4

NaSH

H2SO4

Rougher

P5

1stCleaner

Scavenger

Final Cu concentrate

P2

2ndCleaner
P6

Scavenger

4thCleaner

3rdCleaner

P7

Re grinding

5thCleaner
P8

Final Mo concentrate

Fig. 5a.10. El Teniente moly plant.

123
Bulk Cu-Mo
concentrate
H2SO4
Final Cu
concentrate

Rougher

Rougher

3rd Cleaner

Scavenger

2nd Cleaner

1st Cleaner

4th Cleaner

Final Mo
concentrate

Fig. 5a.11. Andina moly plant.

124
Bulk Cu-Mo
concentrate
Thickener 2 and 3 (one stdby)

150 ft

Diesel

Bank
4x100ft3

2 Bank
4x100ft3

Col 5

Rougher

Rougher

Scavenger

Anamol

Bank
4x100ft3

st

1 Cleaner

2nd Cleaner

Col 1

Thickener 1

150ft

NaCN

Final Cu
Concentrate

Heating
tank

2nd Cleaner

Col 4

Scavenger

Col 2

Scavenger

Col 3

Purification

3rd Cleaner

Final MoS2
Concentrate

Fig. 5a.12. El Salvador moly plant.

125
Bulk Cu-Mo
concentrate

pH 6.5 6.8

9 cells 300 ft3

Rougher

9 cells 300 ft3

Rougher

Final Cu concentrate

9 cells 300 ft3

1st Cleaner

9 cells 300 ft3

1st Cleaner

70-75%
flow

2 cells 300 ft3

2nd Cleaner

25-30%
flow

3rd
2nd
3rd
Clean. Clean. Clean.
3 cells 300 ft3

Final Mo concentrate

Fig. 5a.13. Los Pelambres moly plant.

126
Bulk Cu-Mo
concentrate

100 ft

H2SO4
pH 7.5
12 x 300 ft3
Rougher
H2SO4

2 x 300 ft3

Final Cu
concentrate

1st Cleaner

2nd Cleaner

1.5 m

18 ft

Final Mo
concentrate

Fig. 5a.14. Las Tortolas moly plant.

127
Bulk Cu-Mo
concentrate
H2SO4

NaSH
Tiofos

Rougher

1st Cleaner

2nd Cleaner

3rd Cleaner

4th Cleaner

5th Cleaner

Final Mo concentrate

Fig. 5a.15. Valle Central moly plant.

Final Cu concentrate

Scavenger

128

(iii)

Flotation of Cu-Pb-Zn ores.

Since both copper and lead sulfides float easily with thio-collectors, a bulk flotation
of Cu-Pb sulfides while depressing the zinc and iron sulfides is most widely used.
Following the bulk flotation the zinc minerals are activated and floated, while the bulk
Cu-Pb concentrate is treated by depressing of either the copper or lead to produce
separate concentrates.
The solubility of sulfides is so low that these minerals might be assumed inert.
However, all the sulfides are reactive with oxygen, with the oxidation products being
appreciably soluble in water. Therefore, copper and lead ions commonly appear in the
pulps of polymetallic sulfide ores. (In some processes copper ions are intentionally
added for activation. Sulfide minerals are floated with the use of tiol collectors. Since
sphalerite responds poorly to thiol collectors its activation with intentionally added
copper sulfate is widely utilized; copper sulfate is also used to activate iron sulfide
minerals. The latter is quite common in South Africa in the flotation of sulfides bearing
platinum group minerals).
The bulk Cu-Pb floatation is performed in an alkaline circuit, usually pH 7.5-9.5.
Alkaline pH is very important since it controls activation phenomena (heavy metal ions
precipitate as hydroxides in alkaline solutions). When this is not sufficient, other
depressants are also required. Lime, in conjunction with depressants such as cyanide and
zinc sulfate, may be added to the mills and bulk circuit. The tests carried out at the
Anglovals Prieska Cooper Mine in South Africa revealed that for efficient
depression/deactivation of sphalerite the mass ratio of zinc sulfate heptahydrate
(ZnSO 4 .7H 2 O) to sodium cyanide (NaCN) should be in a range from 3:1 to 3.5 :1. These
additions are typically of the order of 390 to 510 g/t and 130 to 170 g/t, respectively. A
minimum depressant dosage is in the range of 300 g/t of ZnSO 4 .7H 2 O and 100 g/t of
NaCN. It is essential that the initial depressant conditioning of the ore takes place in a rod
mill at a pH between 6.8 and 8.
The depression with NaCN depends strongly on pH because NaCN hydrolyzes to
form HCN:
Na+ + CN- + H 2 O Na+ + OH- + HCN

(5a.14)

As this reaction indicates, an increase in concentration of OH- ions (increase in


pH) increases concentration of CN- ions (in acidic solutions, see Figs. 2.19 and 2.20,
highly toxic HCN gas appears!!!). CN- forms stable complexes with heavy metals:
Cu2+ + 3CN- [Cu(CN) 2 ]- + [CN] 2

(5a.15)

Apart from the reaction with metal ions in solution, cyanide can also desorb
surface copper from already activated sphalerite (copper xanthates are fairly soluble in
cyanide solutions). Cyanides also form complexes with gold and are commonly in

129
leaching gold ores. Therefore, the use of cyanide in processing complex ores may lead to
lower recoveries of precious metals.
Another option is the use of sulfur dioxide (or sodium sulfite) which is added
during the grinding stage. It was shown that the depressing action of sulfur dioxide
increases in the following order:
Chalcopyrite < chalcocite < covellite < galena < pyrite < marmatite < sphalerite
SO 2 is a considerably softer reagent than cyanides and can be used without a
detrimental effect on copper sulfide and the precious metal flotation. Only when the
precious metals are bound to galena, the flotation may be inhibited to some extent.
Table 5a.6. Methods of separating copper-lead concentrates (B.A. Wills,
Mining Magazine, January 1984).

Cu flotation with SO 2 can either be carried out at a pH of around 7, or can be carried out
at the pH of 10.5 (lime). This reduces quite significantly the Zn content in the Cu
concentrate. Dithiophosphates are often used instead of xanthate (or in combination with
xanthate) in processing of gold-containing sulfide ores.

130

In the differential flotation of the Cu-Pb bulk concentrate, either copper sulfides,
or galena may be floated. The choice of method for separating depends on the relative
abundance of the copper and lead minerals. It is preferable to float the mineral present in
least abundance. Galena is usually depressed when the Pb/Cu ratio is greater than unity,
and vice versa copper sulfides arec depressed when the Pb/Cu ratio is smaller than unity.
Since copper sulfides are depressed with cyanides, the choice of the method also
depends on the mineralogical composition of the copper sulfides; chalcocite and covellite
do not respond to depression by cyanide (for such ores it is better to depress galena with
chromium salts, sodium dichromate).
(iv). Cu-Ni ores
The flowsheet of the Outokumpu Kotalahti Mill (the mine was closed a few years
ago) will be used to illustrate typical copper-nickel separation practice. The ore contains
pentlandite, chalcopyrite and pyrrhotite in a gangue of amphibole, plagioclase, mica and
quartz. The Cu-Ni bulk concentrate is treated with dextrin in alkaline environment to
depress pentlandite and float chalcopyrite. The tailings of the Cu-Ni bulk flotation are
conditioned with copper sulfate to activate the pyrrhotite which contains some nickel and
is added to the nickel concentrate. Amyl xanthate is used as collector. Metallurgical
results are given in Table 5a.7.

Fig. 5a.16. Outokumpu Kotalahti Mill.

131
Table 5a.7. Metallurgical results of the Kotalahti Mill.

The largest producer of nickel in the world, INCO (now VALE), has developed
entirely different technology. It includes production of the Cu-Ni bulk concentrate which
is sent to the smelter. Bessemer matte (known as Inco matte)r, which contains chalcocite
and heazlewoodite (Ni 3 S 2 ), and also some ferromagnetic alloys (source of platinum
elements), after very slow cooling down is crushed and ground to the flotation sizes, the
alloys are separated by magnetic separation (about 10% of the total mass) and chalcocite
is floated off from heazlewoodite in almost saturated solution of lime (pH of about 12).
Diphenylguanidine (DPG) is used as a collector.

DPG is soluble in ether and ethanol, is soluble in dilute acids, but it is insoluble in water.
It is a very powerful and selective collector in the matte separation [Tipman et al.,
FLOTATION A.M. Gaudin Volume, SME, Vol. 1, pp. 528-548].
The INCO flowsheets are shown in Figures 5a.17, 5a.18 and 5a.19.

Fig. 5a.17

132

Fig. 5a.18. Processing of the INCO matte by flotation.

133

Fig. 5a.19. Comparison of the first cleaner grade-recovery with the 50 mm pilot plant
column.

134

(v). Platinum group minerals.


Platinum together with the so-called platinum metals group, occurs almost
exclusively in the native state, particularly in the form of alloys. However, several
sulfides, arsenides, selenides and tellurides have also been identified as platinum-group
minerals. Most of platinum-group minerals are encountered as discrete particles, down to
micron size, associated with base-metal sulfides, such as pyrrhotite, pyrite and
pentlandite. Other modes of occurrence have also been suggested. They consist of
submicroscopic solid solutions, especially of palladium and rhodium within base-metal
sulfides. Since PGMs are associated with the sulfide minerals (e.g. chalcopyrite,
pentlandite and pyrrhotite) the process is in essence a bulk flotation. The flotation
performance of the different sulfides varies, with pyrrhotite being ranked the most
difficult to float. Pyrrhotite recovery is enhanced by activation with copper sulfate.
In the South African region there are three main types of the ores, UG-2 and
Merensky from the Bushveld Complex and the high nickel/high talc ores in the Great
Dyke of Zambia. The UG-2 ore is low in sulfides and high in Cr 2 O 3 , and this causes
problems in the smelting process. The Merensky ore is much higher in sulfides and
generally does not contain much Cr 2 O 3 . Because of high density of chromite, the UG-2
ore is amenable to pre-concentration by gravity methods (e.g. spirals).
The PGM plants are designed to maximize recovery. To achieve maximum
recovery the particles containing PGM are floated as soon as they are liberated; grinding
80% below 75 m is common. Fig. 5a.20 shows a typical circuit with two stages of
milling and two stages of flotation [Goodall, 1995 Colloquium, Mintek].

Fig. 5a.20. Schematic layout of a MF2 circuit.

135

The primary milling circuit produces a gring of 30-40% below 75 m and this is
followed by a flotation stage where the liberated mineral particles are floated. In the
secondary milling circuit primary rougher tailings are milled to 70-80% passing 75 m
and the liberated material is floated again. The use of a flash flotation cell on the primary
mill discharge (Fig. 5a.21) is an interesting option. The objective of this cell is to produce
a high grade-low yield final concentrate from the primary mill discharge. The flash
flotation cell needs high energy inputs to suspend the coarse particles and to aerate the
pulp at the high pulp densities.

Fig. 5a.21. A flash flotation in the primary milling circuit.

Fig. 5a.22. Tailings treatment plant


There are two sources of platinum losses: unliberated PGM particles in the
flotation tailings, and coarse liberated particles that did not float due to oxidation
(pyrrhotic pentlandite). In the tailings treatment plant, the slimes are removed by
desliming in cyclones and the underflow is milled and floated (Fig. 5a.22). The freshly
exposed surfaces of the oxidized particles are then treated with CuSO 4 activator.

136
Fig. 5a.23 shows gravity circuit treating the primary rougher tails that can produce
a saleable chromite concentrate (about 40% Cr 2 O 3 ). However, it is not possible to
remove significant amount of the Cr 2 O 3 without losing some PGM recovery.

Fig. 5a.23. Gravity circuit.


In the UG-2 ores a significant percentage of the PGM are slow floating. This is
mostly due to the presence of oxidized species (pyrrhotite and pyrrhotic pentlandite). In
order to recover these slow floating species a good aeration and high energy input are
required, and the forth must be removed quickly from the cell. These conditions favour
the use of rather small flotation cells.
(vi). Flotation of Native Gold and Auriferrous Sulfides.
Gold occurs in its native (free) form, as gold compounds (e.g. tellurides), in
association with sulfide minerals such as pyrite, marcasite, arsenopyrite, pyrrhotite, in
association with Cu/Pb/Ag/Zn sulfides, and also in association with non-sulfide minerals.
Although gold can appear in association with many different sulfides the most important
are iron sulfides. Perhaps the most common are the mixed ores; in processing such ores
free gold particles and sulfide mineral particles containing gold are recovered. From the
processing point of view the most important properties of native gold are: its high density
(19.2) that enables the use of gravity separation methods, its solubility in cyanide
aqueous solutions that makes cyanidation the most important unit operation in processing
gold ores, and its hydrophobicity that renders free gold particles floatable.
Although pure gold is reported to be hydrophilic, native gold flakes are
hydrophobic under flotation conditions. Since free gold particles are hydrophobic they
have tendency to float if these particles are free of contaminations (e.g. slimes, coating by
iron oxides, etc.).
Klimpels results (1998) illustrate quite convincingly good natural floatability of
free gold; these results show that flotation of the 150 m +75 m particles of gold is

137
quite good using only a strong frother (propylene glycol methyl ether, e.g. DF-250 or DF1012). Lins and Adamian found that free gold particles up to 212 m in size floated very
well with frother only (Aerofroth 65), and that the size range of a good flotation could be
extended up to 850 m with the use of both frother and collector (amyl xanthate).
According to Klimpel gold flotation is inherently a slow rate process, and ionized
water-soluble collectors such as xanthates and dithiophosphates slow down further the
rate of flotation of naturally floating minerals including native gold. The flotation of
native gold ores from the Pacific Rim at a natural pH of 7.3 with six different collectors
revealed that much higher dosages of water soluble collectors were required for optimum
recoveries. Recovery from the collectorless flotation of native gold was better than the
recoveries obtained with small dosages of amyl xanthate or dithiophosphate (12.5 g/t).
The results with ethyl isopropyl thionocarbamate gave clearly superior recoveries and
rates. Klimpel was of the opinion that a two-stage rougher flotation, with only an oily
water-insoluble collector and a frother in the first stage followed by a high reagent dosage
second stage is the best solution.
In sulfide ores, gold may occur as native gold and electrum, or as tellurides such
as calaverite (AuTe2), gold-silver, gold-antimony and gold-copper tellurides. The most
common sulfides with which gold is associated are pyrite and arsenopyrite. In pyrite,
gold contents varying from less than 1 p.p.m. to exceptional values as high as 2515
p.p.m. have been reported. In aresonopyrite significant amounts of submicroscopic gold,
even superior to 1000 p.p.m., have been found. Beside pyrite and arsenopyrite, other
sulfides may also be hosts to the gold. Such associations were found with chalcopyrite,
chalcocite-covellite, pyrrhotite, galena and several more. Silver appears mostly in
association with galena (galena may contain up to 1% silver). The optimal processing
method for recovering gold is determined by many factors such as the mineralogy of the
gold bearing minerals as well as the mineralogy of the associated gangue minerals.
Typically the processing scheme includes a flotation concentration step before or
after a cyanidation step. With optimum cyanide concentration of about 0.05% NaCN,
clean gold particles dissolve at a rate of 3.25 mg/cm2/hr while for silver the rate is about
one-half that of gold. These values indicate that a 44 m gold particles will dissolve in 13
hours and a 149 m particles will take about 44 hours to dissolve. Overall dissolution of
gold in dilute cyanide solutions maybe represented by the equation:

4 Au + 8CN + O2 + 2 H 2 O 4 Au (CN ) 2 + 4OH

(5a.16)

Dissolved oxygen is an essential ingredient in the cyanide leaching step. In actual plant
practice the rate of dissolution is affected by the liberation characteristics of gold.
The flotation concentrates may be treated by conventional smelting and refining
or may be subjected to cyanidation to recover the precious metals value. Regrinding or
roasting prior to cyanidation results in higher recoveries.

138
In the case of refractory ores (those ores that show poor response to conventional
cyanidation techniques), it is more economical to pre-concentrate the pyritic minerals by
flotation followed by roasting* of smaller fraction of concentrates rather than the total
ores. The roasted material is then subjected to conventional cyanidation techniques.

------------------*Roasting is a unit process than can be applied to sulfide ores or concentrates to convert
tem to oxides and/or sulfates. The product from a roasting operation is known as a
calcine. Sulfides are converted to oxides in an oxidizing roast. The process involves
controlled oxidation of metal sulfides in the solid state below their melting points (i.e. at
temperature in the range from 500 to 1000 oC).
------------Since cyanide, sodium sulfide, and lime are known to suppress the flotation of
gold, their use should be avoided in the treatment of gold ores. This suggests to give
preference to the use of sodium dioxide (or sodium sulfite) instead of cyanide for
chalcopyrite-galena separation and the use of soda ash instead pf lime for pH control.
The flotation of auriferous sulfides, particularly pyrite, is of great importance in
gold mines in South Africa and Australia [OConnor and Dunne, Minerals Eng., 4, 1057
(1991)]. In South Africa more than 60 million tones of ore are processed annually in
flotation plants to recovery auriferous pyrite. The flotation process in South Africa is
broadly categorized into two areas: auriferous pyrite flotation, and refractory gold ore
flotation. The former is the most important and accounts for more than 95 % of the gold
ores treated by flotation. The auriferous pyrite flotation involves either the treatment of
residue, both current and reclaimed, or the treatment of freshly ground ores.
The Australian gold ores are classifieds as: copper-gold ores, refractory gold ores
(requiring oxidation of the concentrate), partially refractory gold ores (only requiring fine
grinding of the concentrate) and pyritic gold ores.
Both pyrite and arsenopyrite float well in the pH range 3 to 10. Pyrrhotite float
best in acid circuits. In the plants treating residue from the uranium acid leach circuits it
is usually below 4 (also in the plants that employ acid treatment following cyanidation to
destroy cyanide). The flotation of auriferous pyrite ores in South Africa all takes place at
pH values of about 11. Although in this range pyrite is normally depressed, the addition
of copper sulfate promotes the flotation of pyrite (consumption of copper sulfate is in the
range form 30 to 100 g/t). Flotation at high pH favours depression of talcaceous minerals.
Thio-collectors are utilized to float pyrite, pyrrhotite and arsenopyrite. (e.g. sodium isobutyl xanthate, at about 120 g/t), and also some amount of dithiocarbamates and
dithiophosphates. In Australia, potassium amyl xanthate is mainly used. In the acid
circuits mercaptobenzothiazole is utilized. In South Africa, floatable gangue in depressed
with guar gum or modified guar gum. An alternative option is talc prefloat with the use
of fother only (used at one plant in Western Australia).

139
Forrest et al [Mineral Eng., 14, 227-241 (2001)] showed that free gold could not
be selectively floated from copper sulfides, but could be separated from pyrite. The best
collector for gold recovery was found to be Aero 7249 (a blend of diisobutyldithiophosphate with di-isobutylmonothiophosphate) at pH of 11.5 (in this pH
range pyrite is depressed).
Since some locked gold may still be present in the material (pyrite) resulting from
cyanidation, an interesting option is the use of a cationic collector (amine) to recover this
pyrite and reprocess it by grinding and/or roasting. The use of a cationic collector for the
post-cyanidation flotation of pyrite is preferred for two reasons: (i) Cyanide which is
present in this product depresses flotation with sulfhydryl collectors; (ii) Cyanidation is
carried out in the alkaline solutions (pH of about 11) and that requires the use of sulfuric
acid before pyrite can be floated with sulfhydryl collectors. When amines are used the
presence of cyanides in the system does not inhibit the flotation, and the alkaline pH of
the system is just right for the cationic flotation. Fatty amine acetate (20 g/t) along with
Dow 200 frother is used in such a process at the Venterspost mine in South Africa.
Oil Agglomeration in Ore Processing. Only very hydrophobic particles can be
agglomerated with oil in water. The use of this technique to beneficiate solids which are
hydrophobic by nature is then obvious (e.g. coal). But, non-hydrophobic minerals can be
beneficiated by this method as well; the only difference is that such particles have first to
be rendered hydrophobic with the use of properly selected collectors.
Iron ores are commonly floated with the use of fatty acids. The process is
customarily preceded by desliming. In the 50s, the use of fatty acids (tall oil) fuel oil
mixture was studied in Sweden and this research led to a successful agglomerate flotation
process. The tests revealed that a high solids content in the conditioning stage and long
conditioning time with the emulsified collector were essential. Runolina et al [1960], and
Lapidot and Mellgren [1968] studied this process further and its application to treat
ilmenite ores. It was shown that the oil does not spread spontaneously on wet ilmenite
surfaces; spreading requires a vigorous conditioning at high solids content. It was
observed that the agglomeration of fine particles initially was not selective, and only
after sufficient conditioning hydrocarbon oils coat only hydrophobic mineral particles.
Sulfides are floated with the use of sulfhydryl collectors (e.g. xanthates). Oil
agglomeration of the sulfide particles rendered hydrophobic with xanthate should then be
possible. House and Veal studied oil agglomeration of chalcopyrite, and chalcopyritesilica-pyrrhotite mixtures. The agglomerates were recovered by screening at 250 microns.
Xanthates were utilized as a collector. The oil agglomeration of chalcopyrite was very
fast. At a sufficient impeller speed (over 800 rpm) and at least 60 kg/t of oil (6%) copper
recoveries were high. The oil agglomeration was found to be efficient only below pH of
10. Sphalerite oil agglomeration was possible only after activation with Cu2+ and the use
of amyl xanthate. Pyrite was found to agglomerate rapidly at pH 4. Pyrite was not
agglomerated at a pH of 9.5 at which oil agglomeration of chalcopyirte is still good. This
was confirmed by Kocabab et al who demonstrated that un-oxidized pyrite is oleophobic
in acidic environment. It was shown later on that the oil agglomeration of pyrite could be

140
dramatically improved after sulfurizing the pyrite [Drzymala and Wheelock]. These tests
indicated quite an attractive way of using the oil agglomeration/agglomerate flotation
technology to process further rougher concentrates after additional regrinding.
Valderrrama and Rubio [Int. J. Min. Proc., vol. 52, 273 (1998)] found that depending on
the intensity of conditioning, fine gold particles either adhered to the surface of coarser
particles and floated along with them (carrier flotation), or aggregated among themselves
yielding more floatable species. The carrier flotation was promoted at rather mild shear
energy input (0.5 2. kWh/m3 of pulp), at higher energy inputs (2-3 kWh/m3)
detachment of the fine particles from larger carrier particles was observed, but at even
higher energy inputs (3-4 kWh/m3) the aggregation of fine particles was again promoted.
At the energy input levels exceeding these values the aggregation is severely affected and
much lower recoveries are obtained.
Coal-Gold Agglomeration.
As already pointed out when discussing the oil agglomeration of fine coal, this
process turned out to be very unselective with regard to pyrite if larger dosages of oil
were applied. Under such conditions fine pyrite agglomerated along with fine coal into
the agglomerates. This indicates that the mixed gold ore containing both free gold and
gold associated with sulfides should agglomerate quite well if fine coal and oil are added
to the pulp and are subjected to intense conditioning. Such a process is known as Coal
Gold Agglomeration (CGA). The process was patented in 1986 [US Patent, 4,585,548;

Fig. 5a.24. Illustration of the Coal Gold Agglomeration process.


US Patent 4,597,791]. The process involves wet grinding to achieve liberation, mixing
the slurry (grinding product) with added fine coal and oil, high intensity conditioning of
the mixture to initiate micro-agglomeration, then low intensity conditioning to promote
agglomerate growth and finally recovery of the coal/gold agglomerates by screening. It
is then a typical coal oil agglomeration technology; the only difference is that here fine
coal is added as a carrier to pick free hydrophobic gold and sulfide particles and to help
recover them in the form of the agglomerates sufficiently large to be recovered by
screening. Another modification of this process involves separate preparation of the coal

141
agglomerates and their addition in the second stage to the pulp containing gold/sulfide
particles.
The CGA process was found to be especially appropriate for treatment of: (i)
placer ores with gold too fine to be recoverable by gravity methods, (ii) low grade gravity
concentrates to avoid multiple magnetic/gravity separation stages and amalgamation, (iii)
tailings past or present from gravity concentration plants. The agglomerate grades were
found to be in the range from 2000 to 5000 g Au/t or higher, and after combustion the ash
would contain 2 to 5% of gold. While the CGA performance was found to be similar to
the cyanidation, the CGA process was expected to be considerably cheaper to operate,
and would have an obvious environmental advantage.
Further tests on the CGA process revealed the following: metallurgical coal was
found to be the best to be used as a carrier in the process (this is quite obvious as
metallurgical coals are the most hydrophobic), Diesel oil was shown to be quite adequate,
while oil emulsification could reduce conditioning time it was found that it was not
necessary. In accordance with the first papers (and patents) on the spherical
agglomeration, the oil consumption required for high gold recoveries was found to be in
at least 30% range. Also the agglomerates-to-ore ratio turned out to be very important.
Calvez et al. [Minerals Eng., vol. 11, 803-812 (1998)] pointed out that the recovery of
gold by coal-oil agglomerates depends on three major factors: (i) degree of inter-particle
collisions, (ii) agglomerate strength and stability and (iii) mineralogy (liberation size and
surface properties) of gold and the associated sulfide minerals. The first factor is
governed by shear energy brought about by the degree of agitation, and available
agglomerate surface area and number as dictated by agglomerate-to-ore and oil-to-coal
ratios. The second factor is affected by coal particle size and amount and type of oil,
wherein weaker agglomerates which tend to be abraded during gold recovery stage are
formed with coarser coal particles, lighter oils or over-critical amount of oil. The third
factor which is dictated by the inherent properties of gold and other minerals, can be
altered by surfactants and pH modifiers, collectors and/or depressants to optimize gold
recovery.

Fig. 5a.25. Effect of agglomerate-to-ore ratio and of mixing time on gold recovery at oilto-coal ratio of 0.35. Pulp density 20%, impeller speed 750 rpm [Calvez et al.].

142

As Fig. 5a.25 shows, a high oil dosage of 35% is required in this process, and also
a high ratio of the agglomerated coal-to-ore which clearly depends on the intensity of
conditioning. Since in the later modification of this process the pre-agglomerated coal is
conditioned with the ore, it is important to know what happens with the gold particles
when they get in contact with the agglomerates. The tests reveled that gold particles are
attached at the agglomerate surface as clusters or as individual grains and that with time
of stirring the gold particles become more securely embedded between coal particles and
become an integral part of the agglomerates.
Recovery of the coal/gold agglomerates by flotation is also possible [Moses &
Peterson, Minerals Eng., Vol. 13, 255-264 (2000)].

143

Chapter 5b. Flotation of oxidized ores of base metals


Oxidized and mixed oxide-sulfide ores of non-ferrous metals are the products of
oxidation of sulfide ores. Among processes that lead to the formation of oxidized ores the
most important are:
-

oxidation and dissolution of sulfides, and formation of metal sulfates


which are quite water-soluble;

precipitation from sulfate solutions of various compounds such as carbonates,


oxides and hydro-oxides, phosphates, silicates, etc.;

secondary dissolution and re-precipitation.

One of the characteristic features of oxidized ores is the complicated


mineralogical composition (e.g. more than one hundred minerals of Pb and Zn are
known). The most important are oxides, sulfates and carbonates. These minerals exhibit
different surface and flotation properties, for some of them the flotation processes have
not yet been developed (e.g. chrysocolla). Sulfates present in oxidized ores are commonly
activated by various cations and can be coated by layers of secondary oxidation products.
Therefore, their selective flotation is usually much more difficult than the flotation of
sulfide ores. The oxidized ores are usually heavily weathered, eroded, and may contain
high quantities of iron and fine clay slimes. Such slimes further complicate the
beneficiation. Hence, the separation processes are often preceded by desliming. This also
enables removal of a large quantity of soluble salts and gypsum (CaSO 4 .2H 2 O) which is
formed in SO 4 2- saturated solutions.
Washing. Preliminary washing is always recommended whenever the ore is
characterized by a high content of hydrated iron oxides and clays. By the way, it is
impossible to separate Cu, Pb and Zn minerals from this type of slimes by simple
selective flotation. The washing and desliming may also make possible the use of a dense
medium separation in a preliminary concentration stage prior to flotation. Such treatment
allows removal of as large quantity of gangue before fine grinding and flotation. Two
objectives can be achieved by this: reduction of grinding costs and improvement in
flotation selectivity. Very often the medium (2.65 2.9 g/cm3 range) is obtained from
galena flotation concentrate, and treatment in such a medium allows rejection of about
30-40% of coarse gangue.
Activation. A pulp in the flotation of oxidized ores usually contains a few different
oxidized minerals and sulfides, and these minerals surface and flotation properties differ
widely. In the processing stage just before flotation these surface properties must be
unified as much as possible so that these various minerals can respond to the same
flotation collectors. This is achieved by activation with Na 2 S or NaHS. As a result of the
reactions that take place on the surface of the treated minerals, a sulfide layer is formed
on the surface of the treated minerals as the example below shows:

144
PbCO 3 + Na 2 S

PbS + NaCO 3

(5b.1)

Since sodium sulfide and hydrosulfide are strong reducing compounds they depress
flotation of sulfides (by the way they are commonly used to depress copper sulfides in the
differential flotation of Cu-Mo bulk concentrates) and any overdosage in the activation
stage that leaves some S2- and HS- ions in the pulp strongly affects subsequent flotation.
(when thio-collectors are used).
Copper oxide ores
The main copper ores are sulfides, oxides, or a mixture of sulfides and oxides.
The last type is the worlds most extensive ore bodies; they are disseminated, i.e. they are
composed of minute grains of copper minerals in large bodies of rock. They contain on
the average 2% Cu. The sulfide-oxide ore is usually separated by flotation into two
fractions: a high-grade sulfide concentrate and a low-grade oxide fraction. While the
sulfide concentrate is commonly treated by pyrometallurgical methods, the oxide fraction
is usually treated by hydrometallurgical methods (copper sulfides are insoluble in acids).
Fig. 8c.1 shows the processing flowsheet of the plant treating mixed sulfide-oxide copper
ore.

Fig. 5b.1

145
Reagents in the flotation of copper minerals
All copper sulfide minerals can easily be floated provided that they are only
slightly oxidized. Xanthates are commonly utilized as collectors. To achieve selectivity in
the separation of copper sulfides from iron sulfides, not only must xanthate concentration
be controlled but also lime (pH) and in some cases cyanides must be used. Iron sulfides
are depressed in alkaline solutions especially when lime is used to adjust pH. While this
is true in the flotation of sulfide ores, depression of pyrite is much more difficult in the
presence of copper oxidized minerals. This results from the activation of pyrite by copper
ions. In such cases, the selectivity can be improved by the use of cyanides to complex
copper ions. While oxidized minerals of copper, lead and zinc can also be floated with
xanthates without sulfidization, as it was shown in many studies carried out under
laboratory conditions, full-scale tests usually give poor recoveries. In practice, copper
oxidized ores are floated either with thio-collectors after sulfidization or directly with
fatty acids.
Selection of the flowsheet depends on the content of copper sulfides and the
sulfides-to-oxides ratio, dissemination and mineralogical composition of gangue
minerals. Fine grinding (70-80% below 0.075 mm) is usually required to liberate a major
portion of the oxidized copper minerals. Either a bulk flotation of copper and iron
sulfides is carried out first, or sulfides and oxides of copper are floated together. In the
bulk flotation of sulfides, various xanthates such as ethyl, iso-propyl, butyl and amyl are
used at different plants. The more oxidized the ore the longer chain xanthates are utilized.
At some plants xanthates with dithiophosphates (aerofloats) or with
mercaptobenzothiazol are employed. Selective flotation of the copper and iron sulfide
bulk concentrate is usually carried out at pH > 10 adjusted with lime. Consumption of
lime depends on the pyrite content and is usually in a range from 1 to 5 kg./t. If iron
sulfides are activated by copper and their depression is not efficient, cyanides are also
used (up to 200 g/t) at a pH which should be about 9. Simultaneous flotation of sulfides
and oxides using sulfidization and lime to depress pyrite is employed only when the
degree of oxidation is not very high. Lime (500-1500 g/t) is commonly supplied to a mill
and sodium sulfide (300-1000 g/t) also in a grinding circuit.
If a bulk flotation of sulfides is conducted first, then this stage is followed by a
flotation of oxidized minerals. The oxides can be floated:
-

with the use of fatty acids if the gangue contains silicates and aluminosilicates (but not carbonates);
with the use of thio-collectors after sulfidization;
with the use of non-thio collectors following sulfidization..

Selectivity of the flotation with fatty acids and their sodium salts (soaps) depends
on de-activation of silicates (activated by metallic ions). Water glass, Na 2 CO 3 , and
phosphates are used for this purpose. In some plants fatty acids are employed along with
water-insoluble hydrocarbons (kerosene, Diesel oil, etc.). Flotation efficiency can be
significantly improved by emulsifying fatty acids and oil prior to their use.

146

Flotation of oxidized copper ores with the use of fatty acids is carried out in
ZAIRE. The ore contains up to 5.5% of Cu and about 0.5% of Co. Copper mostly appears
as malachite and the main constituent of a gangue are silicate with a very low content of
dolomite. Sodium carbonate (0.6-0.8 kg/t) and sodium silicate (about 1 kg/t) are used to
disperse slimes. Fatty acids in combination with Diesel oil are employed as a collector.
They are emulsified at 50-55 oC prior to addition in a rougher flotation. Whole pulp is
heated up in conditioning tanks to 32-35 oC. The froth product contains 24-25% Cu and
1-1.4 % Co at their recoveries of 82-65% and 56-60%, respectively.
The flotation process in which fatty acids are utilized is less expensive, however,
it is also less selective and can be used only when malachite is a primary copper mineral.
The process in which sulfidization is used is most common; when the gangue
contains carbonates (calcite, dolomite, etc.), fatty acids cannot be used. In most countries
NaHS or combination of NaHS and Na 2 S are applied for sulfidization; in Russia Na 2 S is
utilized. Since sodium sulfide and hydrosulfide are oxidized by oxygen, in all plants the
sulfidizing mixture is added in stages in rougher and scavenger flotation. Consumption of
sodium sulfides, depending on the content of copper, slimes and soluble salts, varies from
0.3 to 2.0 kg/t. The flotation process is carried out at pH 9. Some authors noticed that
addition of ammonium sulfate led to an increased copper recovery and reduced the
consumption of sodium sulfide by about 30%. Isopropyl, butyl, iso-amyl and amyl
xanthates are used as collectors (0.1 to 0.2 kg/t). Better results were reported with
mixtures of xanthate and dithiophosphate (aerofloat), with 40% of collector being added
to a mill, with the use of longer chain xanthates, and simultaneous use of apolar oils in
addition to thio-collectors. The most often used depressant is water glass. The tests on
flotation of oxidized copper ores in sea water revealed that high recoveries can also be
obtained under such conditions when consumption of a collector is doubled and solid
content in the pulp is not higher than 20-23%.
The overall process efficiency depends first of all on the conditions in the
sulfidization and flotation stages.
Combined flowsheets that also include leaching. About 20% of copper is produced by
leaching in H 2 SO 4 solutions. Since hydrometallurgical unit operations are more
expensive than mineral processing, leaching is utilized only when flotation is impossible.
For instance, the ores in which copper appears in the form of silicate (chrysocolla) cannot
be beneficiated by flotation. Also, when copper appears in association with iron and
manganese oxides, or is highly disseminated the flotation provides poor recoveries. Acid
leaching in all cases provides satisfactory results if the treated ore does not contain
substantial quantities of carbonates (ammoniacal leaching can be applied in such cases).
While leaching of oxide ores or roasted sulfides (roasting is the oxidation of metal
sulfides to give metal oxides and sulfur dioxide, and is carried out in the temperature
range from 500 to 1000 oC) does not require further oxidation, leaching of sulfides is a
slow process since it involves oxidation. Examples of reactions that take place during

147
leaching of oxidized copper minerals in solutions of H 2 SO 4 are given below for tenorite,
brochantite, malachite, azurite, atacamite, cuprite and chrysocolla.

(5b.2)

Dissolution of sulfides and native copper require oxidation and this is provided by
ferric ion
Cu + Fe 2 (SO 4 ) 3 CuSO 4 + 2 FeSO 4

(5b.3)

The product of this reaction, ferrous sulfate, is oxidized back into ferric sulfate in the
presence of oxygen:
4FeSO 4 + O 2 + 2H 2 SO 4 2Fe(SO 4 ) 3 + 2H 2 O

(5b.4)

While chalcocite and bornite can be leached relatively easily, dissolution of chalcopyrite
is more difficult. Leaching is speeded up by increasing concentrations of leaching
reagents, temperature, intensity of stirring, and finer grinding. The leaching can also be
carried out underground (in situ mining)
Heap leaching is the most common. Recoveries of 80 to 85% copper on run of
mine oxide ore are reported at some heap operations.
Hydrometallurgical process usually consist of the following steps:
- leaching of the ore or intermediate products;
- purification of the solutions;
- recovery of the wanted metal either by precipitation (cementation) or by
solvent extraction.
Leaching-precipitation-flotation (LPF) process. This process includes leaching in diluted
H 2 SO 4 solutions, cementation of copper using iron scrap (Cu2+ + Fe = Cu + Fe2+), and
flotation of fine metallic copper. The copper sulfides in this process are recovered along
with cemented copper in the flotation stage (leaching of sulfides without roasting is slow
and recoveries are not high) and, therefore, this process is suitable for processing the
mixed oxidized-sulfide copper ores. The cementation process requires not higher than
0.2-0.3% concentrations of a free sulfuric acid; for efficient flotation the particles of

148
metallic copper should be larger than 0.075-0.1 mm. The cementation in very acidic
solutions may result in larger and more dense copper particles. The size of copper
particles decreases with dilution of copper solutions and increase in pH. The size of iron
particles used to cement copper should not exceed 0.1mm as it also affects the size of
copper particles. Conditioning during cementation should be carried out without aeration
and, therefore, not in flotation cells. Since flotation of the cemented copper (and sulfides)
is carried out in this case from acidic solutions (pH 2.5 4), xanthates cannot be used as
collectors. Dithiophopsphates, or derivatives of xanthates (Minerec) are recommended.
H 2 S can also be used to precipitate CuS.

Fig. 5b.2. The LPF process in the oxide-sulfide copper ore processing flowheet.
There are reports that the LPF process can be modified further to recover
cemented copper together with iron by magnetic separation.

149

Processing of lead oxidized ores


The following are common minerals in the oxidized lead ores: anglesite (PbSO 4 ),
cerussite (PbCO 3 ), hydrocerussite (2Pb(CO 3 ) 2 .Pb(OH) 2 ), and several oxides: litharge
(PbO), plattnerite (PbO 2 ) and minium (Pb 3 O 4 ). Anglesite is the first oxidation product on
the surface of galena. Cerussite, the most common oxidized mineral of lead, is often
interlocked with limonite (a mixture of hydrated hematite Fe 2 O 3 , goethite HFeO 2 , and
lepidocrocite FeO.OH).
Of all lead minerals, galena, cerussite, anglesite and wulfenite (PbMoO 4 ) respond
the best to flotation. Galena is one of the easiest minerals to float using xanthates. It is
depressed by Na 2 S which, on the other hand, is needed for sulfidization to float cerussite
and anglesite (these minerals can also be floated with fatty acids without sulfidization).
According to some authors, it is possible to separate cerussite from malachite. These two
minerals, when sulfidized, float differently in the presence of water glass which depresses
sulfidized malachite.
As an example, the flowsheet of the plant treating mixed sulfide-oxide lead ore
(which contains galena with a high Ag content) is discussed. To obtain high overall Pb
recovery without losing some galena (which is depressed by Na 2 S), the sulfides are
floated first and flotation of oxidized lead minerals follows (Fig. 5b.3).
Long conditioning of cerussite with sodium sulfide may reduce its recovery. It is
better to float out cerussite immediately after sulfidization. In some plants (e.g. in
Morocco, Sardinia), conditioning tanks are not used in the summer time and sodium
sulfide is fed directly to flotation tanks. Short contact of the activator with cerussite ores,
as low as possible aeration in the activation stage, following by a good aeration in the
flotation, reduces consumption of both sodium sulfide and xanthate at high lead
recoveries. Calcium ions in the pulp affect the sulfidization and flotation of cerussite. In
the flotation of an ore with a higher content of anglesite and wulfenite, higher
concentrations of sodium sulfide are needed and stage addition of the activator is not
recommended. A longer conditioning time is better (10-15 min). To avoid a high pH, not
only sodium sulfide but also sodium hydrosulfide are utilized. At lower pH (8-9) that can
be achieved when sodium hydrosulfide is applied, soluble basic carbonate is formed
instead of insoluble calcium carbonate. Addition of ammonium salts (up to 1 kg/t)
increases the solubility of Ca and Mg carbonates and prevents their precipitation on the
surface of sulfidized cerussite. Calcium salts do not exhibit depressing effects in the
flotation of anglesite, but barium ions do. This can be inhibited by application of sodium
carbonate. Sulfidization of all oxidized lead minerals proceeds better in soft water.

150

ig. 5b.3. Processing of mixed sulfide-oxide lead ore

Processing of oxidized zinc ores


Sphalerite does not float without activation, and copper ions are commonly
utilized to activate it for this process. Oxidized zinc minerals include: Smithsonite,
ZnCO 3 , hydrozincite ZnCO 3 Zn(OH) 2 , hemimorfite, Zn 2 SiO 3 (OH) 2 and willemite,
Zn 2 SiO 4 . Two methods are commercially used to beneficiate oxidized zinc ores by
flotation. In the first, the deslimed ore is sulfidized and then is treated with copper sulfate
at a higher temperature. Prior elimination of iron sulfides and slimes is essential. A
powerful frother and longer chain xanthate (e.g. iso-amyl xanthate) are beneficial. Under
such conditions 26-27% concentrate grades at 77-80% Zn recoveries can be obtained.
The ores with a low content of iron oxides are good candidates for such a treatment. The
flowsheet below provides more details.

151

Fig. 5b.4. The flowsheet of the plant treating mixed sulfide-oxide Pb-Zn ore
(total Zn recovery 79-82%, recovery of oxidized Zn 76-78%).
In the second method developed by M. Rey, zinc minerals are floated with the use
of primary aliphatic amines following sulfidization at room temperature. This process is
more selective than the first one with regard to iron oxides and is more efficient in
treating zinc silicates. While en excess of sodium sulfide acts as a depressant in the
flotation with xanthates, and flotation starts only when this excess is consumed,
in the amine flotation of zinc ores, an excess of sodium sulfide has no depressing effect,
and the conditioning is not necessary. .
In all plants treating oxidized zinc minerals, galena and lead oxidized minerals are
floated first and then zinc sulfides and iron sulfides. This reduces consumption of
reagents and increases the grade of oxidized zinc concentrates. Prior to sulfidization, the
ore is deslimed, sulfidization is immediately followed by the addition of amine and a
frother. Too long conditioning during sulfidization affects overall results. Powerful
frothers are needed in the flotation of sulfidized Zn minerals. Successful application of

152
Reys process depends firstly on the content of clay slimes and soluble salts. High
consumption of sodium carbonate, water glass and polyphosphates (or carboxymethyl

Fig. 5b.5. Flowsheet of the San Giovani plant in Italy (Sardinia).

cellulose) are unavoidable. Consumption of these reagents can be reduced by careful


desliming of the ore and the use of soft water. Primary amines (C 12 -C 18) are claimed to
be the best collectors in this process. The first commercial flotation started in Sardinia in
1950 and then in 1954 in Morocco; today this technology is used in other countries (e.g.
Brazil). The concentrates produced in Sardinia assayed 40-45% Zn, they were calcined in
a rotary kiln and brought to 55-56%, then processed by leaching or pyrometallurgy.

153

The major issue in the flotation with amines are slimes. As reported by Russian
scientists [Abramov, Beneficiation of oxidized and mixed ores of non-ferrous metals,
Nedra, Moscow, 1986, pp. 268-269] the effect of slimes can be reduced by applying
amines (instead of ammonium salts) as emulsions with oily hydrocarbons. According to
Rey et al [Quarterly Colorado School of Mines, Vol. 56, No. 3, July 1961] this also
requires emulsifying agents (e.g. ethoxylated long-chain alcohols). Abramov reported
that mixing and emulsifying all the components together (with sodium sulfide) results in
the product with much better collecting properties. The emulsion is much more efficient
and its application along with good blinders (carboxymethyl cellulose, water glass,
hexametaphosphate, etc.) may allow for direct flotation without desliming. Beneficial
effect of such emulsions was recently confirmed by Carvalho et al [Proc. 12th Balkan
Mineral Processing Congress, Delphi, 2007]. The emulsion, which they tested with very
good results, had the following composition: amine : diesel oil : MIBC = 1 : 0.16 : 0.4.
They showed that with this emulsion (utilized along with either 1000 g/t of water glass or
1000 g/t of hexamataphosphate) it was possible to significantly decrease consumption of
sodium sulfide.

154
Chapter 5c. Oxides
Flotation of quartz and silicates
The silicates are the most abundant minerals in the earths crust. They are not
only important commercial minerals, but also are generally the gangue minerals in other
ores.
Examples of silicate minerals concentrated by flotation includes quartz (SiO 2 ),
mica (there are many minerals belonging to this group, one of them is muscovite),
feldspar (examples belonging to this group include orthoclase, microcline, albite), garnet
which includes iron rich alumino-silicates, etc.
Quartz has been heavily used in many fundamental studies and its surface
properties and floatability are well known. Pure quartz does not float with anionic
collectors, but it can easily be floated with cationic collectors (e.g. dodecylamine) in
alkaline solutions.
Flotation is extensively used in processing quartz-feldspar-mica ores (known as
the North Carolina pegmatites). Separation of the three minerals is accomplished in three
stages. The first stage recovers the mica, the second removes contaminating minerals, and
the third stage separates the feldspar from the quartz.
Prior to flotation, the ore is subjected to scrubbing and after scrubbing the ore is
deslimed in cyclones at 150-200 mesh. Mica is commonly floated in acid circuits (one
company is floating in acid circuit using fatty acids as collector). Mica floats easily using
primary amines at the pH lower than 3 (sulfuric acid). Fuel oil and a frother are added to
enhance flotation of coarse mica particles. Conditioning time needed for mica is short,
less than one minute. Normally only one stage of flotation is needed to produce an
acceptable mica concentrate.
The second stage of flotation is used to remove impurities ahead of the quartzfeldspar separation. The major contaminants are garnet and biotite, the minerals that
contain iron. The pH in this flotation remains acidic. Petroleum sulfonate is used as
a collector, and also a frother. Less than 5% of the total ore is rejected in this step,
however, this step is very important since it must eliminate the contaminates which
otherwise would lower the quality of the final products. The materials remaining after the
mica and iron float are essentially a quartz and feldspar mixture. Both quartz and feldspar
are used in the manufacture of glass. If both quartz and feldspar are to be made as
individual products, the third flotation stage is used to make the separation. Both minerals
are saleable products. Prior to the quartz-feldspar flotation, the tailings from the iron
floatation is dewatered in cyclones, which remove the reagents from the previous
flotation stages and discard any remaining fines. Following dilution with fresh water,
hydrofluoric acid (HF) is added (as quartz depressant) to the pulp and the pH is lowered
to about 2.5. Amine is used as the collector for the feldspar, which is floated away from
the quartz. Likewise the mica flotation, this flotation is very selective and requires only a
rougher flotation to meet quartz specifications.
A very pure quartz is needed by electronic industry. The price of such a quartz
can be as much as $10,000 per ton. Little is known about further purification of the quartz
concentrates for the needs of the electronic industry; this technology is proprietary.

155
In other type of quartz ores (with quartz produced for the glass industry), only
iron contaminants are removed by flotation using petroleum sulfonate collector in an acid
environment.

Flotation of iron ores


Flotation of iron ores involves separation of iron oxides from silica (an other
gangue minerals) and as summarized by Iwasaki [Mining Engineering, June 1983] there
are a few options to achieve it:
(i)
(ii)
(iii)
(iv)

Anionic flotation of iron oxides using either fatty acids (RCOONa) at


pH of 7, or alkyl sulfates (RSO 4 Na) at pH of 4;
Anionic flotation of silica using fatty acids with activation by Ca2+
in alkaline pH (of about 11);
Cationic flotation of iron oxides with primary amines in acidic
environment and in the presence of F- (pH of 4);
Cationic flotation of silica with primary amines at pH of 7 with dextrin
(or starch) as iron oxides depressant. .

Fig. 5c.1. Electrophoretic


mobility, adsorption and
flotation recovery of
magnetite as a function of
pH. [I. Iwasaki].

156

The electrical nature of interaction between ionic collectors and oxides is well
understood. As Fig. 5c.1. shows, the isoelectric point of magnetite is located at pH 6.5.
Below this pH magnetite surface is positively charged and above this pH is negatively
charged. Dodecylammonium chloride, a cationic collector, is adsorbed above the i.e.p.
and makes magnetite floatable above that pH. Sodium dodecylsulfate, an anionic
collector, is adsorbed below the i.e.p. and makes magnetite floatable below that pH. The
iso-electric points of other iron oxide minerals, hematite, goethite, are also situated
around pH 6.5. As it follows, the iron ores can either be floated using anionic or cationic
collectors and either in acidic or alkaline solutions. In practice, however, selection of the
flowsheet critically depends on the subsequent use of the iron concentrate.

Fig. 5c.2. Electrophoretic


mobility of hematite (H) and
quartz (Q); flotation of hematite
and quartz with 10-4 M dodecylamine; flotation of hematite and
quartz with 10-4 M sodium
dodecylsulfate.

In pyrometallurgical operations that require gases to flow through a charge of


solid materials, it is important that the material is in a lumpy condition and contains a
minimum of fine particles. This ensures that the charge is permeable so that gases can
easily pass through it. Since the iron oxide concentrates are fine they must be pelletized
before any subsequent use. In the pelletization process, the fine material is rolled either in
a drum or on an inclined disc (Fig. 5c.3). The pelletization process is initiated by water
droplets that act as nuclei, bonding between particles is achieved through the capillary
forces. These forces critically depend on wettability of the pelletized particles. The
particles must be hydrophilic for the process to be efficient (Fig. 5c.4).

157

Fig. 5c.3. Disk pelletizer.

Fig. 5c.4.
Contracting capillary
forces in a green
pellet weakened by
an encased
hydrophobic surface.

There is, therefore, an important link between the flotation process and the
pelletization process. As analyzed by Gustafsson and Adolfsson [Proc. 20th IMPC,
Aachen, 1997, Vol. 3, pp. 377-390], the green pellets strength is directly determined by
capillary pressure:
pk = S w cos

(1 e)
e

158

where:

is the surface tension of warter;


S is the specific surface of concentrate;
w is the volume density of concentrate;
is the contact angle:
e is the pore, void volume

and so it critically depends on the surfactants remaining in the pulp (surface tension)
and wettability of the solid particles (flotation product). The surface tension should be as
high as possible (that means as close to that of water as possible), and the contact angle of
0 would be the best. It is obvious that flotation concentrate cannot meet such
requirements.
Although it is possible to remove fatty acid coatings from the iron oxide surfaces
[I.Iwasaki, Trans. SME, Vol. 241, 304-312 (1968)] it is obvious that it is preferable to
use a reverse flotation in which the iron minerals are recovered as hydrophilic tailings.
The principles of such processes will be further discussed using the Tilden Mine
flowsheet and the flowsheet of the LKAB (Sweden) plants.
The flowsheet of the Tilden Mine selective flocculation/flotation plant is shown in
Fig. 5c.5.

Fig.5c.5. Schematic
representation of the
Tilden Mine selective
flocculation/reverse
flotation process.

159

Fig. 5c.6. Adsorption of corn starch on hematite (H) and quartz (Q) [Balajee &
Iwasaki, Trans. AIME, Vol. 244, 401 (1969]).
As simplified flowsheet (Fig. 5c.5) shows, after very fine grinding (83% below 25
m) in the presence of dispersants (either water glass or polyphosphates at pH 10.5-11)
the hematite is flocculated with addition of corn starch (0.11 kg/t). The yield of slimes
(overflow from thickener) is about 25-30%; this pre-concentration stage increases iron
content in the feed from about 35% up to 44%. The thickener underflow is further treated
with additional amount of starch (0.67 kg/t) to depress better iron oxides remaining in the
flotation feed.
In the reverse flotation that follows, quartz is floated with ether primary amine
while depressed/flocculated hematite reports to the tailings. Rougher flotation is followed
by a few cleaning stages. Typical metallurgical results are given in the table below.
Table 5c.1 Tilden Plant Metallurgy
Ore

Slime Tails
%wt

Float
Feed
%Fe

%Fe

%Fe

34.7

14.7

Concentrate
%Fe

32.1

44.1

64.0

Float Tails

%SiO2 %Fe Rec.


4.89

71.1

%wt.
29.3

%Fe
18.0

The process is based on differences in adsorption of starch onto hematite and


quartz. As Figure 5d.6 demonstrates, the adsorption density for starch onto hematite is
more than ten times higher than onto quartz; this difference further increases with
increasing pH. The starch adsorbed onto hematite also plays a role of a blinder which in
the following reverse flotation stage prevents hematite from flotation when amine is
utilized to float quartz.

160

Fig. 5c.7. Flowsheet of the LKAB processing plants.

The Swedish iron ores in Kirunavaara contain about 60% iron. Because the ores
contains some amount of apatite (and also calcium carbonates) flotation is also used, and
Atrac 1562E a mixed anionic collector (which contains main collector, a co-collector and
a foam regulator) is applied. Flotation is carried out from slightly alkaline solutions using
water glass as dispersant and magnetite depressant.
The reverse flotation is much more difficult in processing of lower grade
complex ores. Since fine particles tend to float by entrainment, higher content of gangue
minerals in the iron ores and fine grinding affects iron recovery and the grade of the
concentrate (flotation tailings). Because of increasing consumption of the utilized
reagents improved flotation technologies are needed. Brazilian researchers claim [Pereira
et al., Proc. 23rd IMPC, Istanbul, 2006, Vol. 1, pp. 644-648] that the consumption of
amine in the reverse flotation of iron ores can be reduced by using the amine-diesel oil
emulsion (amine:oil = 4:1); this combination also requires an emulsifier (such as Tergitol
TMN-10) which is added at 5% to the oil. The recent paper from the same school [Turrer
& Peres, FLOTATION 09, Nov. 2009] investigates the use of various depressants in the
reverse cationic flotation of iron ores in Brasil. The tested depressants are compared to
the standard starch. As Fig. 5d.8 reveals, a comparable results were obtained with guar
gum and one of the tested carboxymethyl celluloses (CMC5).

161

70

pH
10.0

pH
9.5

C oncentr ate silica content ( % )

Ir on r ecover y ( % )

60

pH
10.5

50

40

30

pH
10.0

pH 10.5

pH
9.5

1
20
0
180

320

60

600

180

g/t

320

60

g/t

carboxymethylcellulose (red line), guar gum, starch


Figure 5c.8. Effect of pH on flotation performance of Starch, CMC5 and GG in
the cationic reverse flotation of iron ore [H.D.G. Turrer and A.E.C. Peres,
FLOTATION 09, Cape Town, Nov. 2009].

600

162
Chapter 5d. Flotation of Industrial Minerals
Industrial minerals (also referred to as salt-type minerals) are composed of
divalent cations such as Mg2+, Ca2+, Ba2+ etc. and anions such as F-, CO 3 2-, SO 4 2-, WO 4 2-,
PO 4 3-. Examples of the minerals of commercial importance are apatites
[Ca 10 (PO 4 ) 6 (OH,Cl,F) 2 ], Fluorite [CaF 2 ], calcite [CaCO 3 ], barite [BaSO 4 ], magnesite
[MgCO 3 ] and scheelite [CaWO 4 ].
Separation of these minerals from other salt-type minerals is very difficult. As the
figure below shows all these minerals float very well with fatty acids and the differences
between their flotation characteristics are not sufficient for selective separation.

Fig. 5d.1. Illustration of similarities in the flotation behavior of some calcium minerals
(calcite, fluorite, and several appetites from different sources).
Chemisorption of fatty acids on salt-type minerals resulting in the formation of
calcium (magnesium, or barium) soaps is well documented.
Poor selectivity in the flotation of these minerals with fatty acids requires the use
of modifying agents. The two figures quoted after Somasundaran and Hanna [in
Flotation A.M. Gaudin Memorial Volume, AIME, 1979, Vol. 1, pp. 197-272]
demonstrate the effect of starch and tannin on flotation of fluorite, barite and calcite.

Fig. 5d.2. Correlation of


adsorption of starch on
fluorite, barite, and
calcite with the
depression of their
flotation using oleic acid
as collector.

163

Fig. 5d.3. Correlation of adsorption of tannin on fluorite, barite and calcite with
depression of their flotation using oleic acid as collector.
As seen from Fig. 5d.1, fluorite has a high affinity towards oleic acid and floats
easily with this collector even at exceptionally low oleic acid concentrations. This
mineral is not depressed by starch (Fig. 5d.2. On the other hand flotation of calcite can be
easily depressed by both starch and tannin.
Fluorite appears with many gangue minerals including quartz, calcite, barite, and
sulfides. The use of fatty acids is common is its flotation along with various depressants.
For example, in the selective flotation of fluorite from silicates in addition to fatty acids
water glass is used to prevent activation of quartz by calcium ions. The process is carried
out in alkaline solution (Na 2 CO 3 ).
Flotation of industrial minerals will be further discussed using examples of the
flotation plants treating phosphate ores in Florida, flotation of scheelite ores, and flotation
of bastneasite ores (source of lanthanum elements).
Flotation of Phosphate Ores.
Although more than 200 minerals contain at least 1% P 2 O 5 (0.4364 % P), the
only important ore mineral in phosphate rocks is apatite. However, appatite is better
described as a group of minerals rather than as a single mineral. There are different types
of appetites, the most common is fluoroapatite Ca 5 (PO 4 ) 3 F. In sedimentary deposits most
common is francolite, (Ca,Mg,Na) 10 PO 4(6-x) (CO 3 ) x F 0.4x (F,OH) 2 .
There are three major types of phosphate deposits: (i) sedimentary, (ii) igneous
and (iii) guano. Sedimentary deposits are the most important, they contribute about 80%
of the world production. Igneous phosphate ores account for 15% of the production.
Guano deposits represented a very important source of phosphate in the past.

164
Phosphate ore beneficiation depends upon the type of deposit. It is useful to
consider the following general classification:
(i)

(ii)

(iii)

Sedimentary phosphates
- siliceous gangue
- carbonate gangue
- mixed gangue
Igneous phosphates
- non-carbonate gangue
- mixed gangue
- carbonate gangue
Guano phosphates

Sedimentary phosphate ores with preponderant carbonate gangue are by


far the most widely represented in the world and constitute more than 2/3 of the reserves.
Those, with few exceptions, are not being presently exploited.
Igneous deposits with non-carbonate gangue are found in feldspathoic and
pyroxenitic rocks. The major apatite present is well crystallized fluoroapatite.
Gangue minerals are alkaline feldspars, ferro-magnesian silicates, zircon and
sphene. The most important of all igneous deposits mined today that fall in this category
is localized in the Kola peninsula in Russia. The output from this deposit is over 10
million tonnes per year of concentrates assaying 39% P 2 O 5 ,
From an average feed grade of 18% P 2 O 5 . The ore is processed by direct anionic
flotation consisting of one roughing, two scavenging and three cleaning operations. The
collector used is a combination of raw tall oil and distilled tall oil (110 g/t), oxidized
petrolatum and secondary oil tar are used as auxiliary collectors (75 g/t). Sodium silicate
is used as a dispersant/depressant (45 g/t). Overall recovery is high in comparison with
other phosphate plants, 94% P 2 O 5 . There is no desliming step.
The igneous phosphate ore with essentially non-carbonate gangue at Tapira
deposit in Brazil (it contains less than 3% calcite) is beneficiated by flotation with sodium
oleate at pH 9. Corn starch causticized with sodium hydroxide is utilized to depress iron
oxides and iron silicates. From an average feed grade of 8% P2O5 (mainly as
hydroxyapatite), a concentrate assaying 36% P 2 O 5 is produced. Overall recovery is 60%;
the major losses occur in desliming (hydrocyclones with d 50 of about 38 microns). The
output of this mine is about 1 million tones of concentrates per year.
In the category of igneous apatite with mixed gangue a very good example
is deposit at Phalaborwa (South Africa) processed at the Foskor plant.. The ore is
contained in an ingenous complex composed mostly of pyroxenite but whose core is
carbonatite coated with a serpentine-magnetite-apatite rock which is called foskorite.
The ore is mined for copper, titaniferous magnetite, accessory minerals with uranium and
thorium. Apatite from foskorite ore assays in average 10% P 2 O 5 and contains titaniferous
magnetite, calcium and magnesium carbonates, and olivine/serpentine as principal
gangue minerals. The ore from pyroxenite contains 6% P 2 O 5 and gangue mostly
composed of pyroxene, phlogopite and vermiculite. The third type of the ore is contained
in the tailings of the Phalaborwa Mining Company washing plant which recovers copper

165
sulfides by flotation with potassium amyl xanthate after elimination of magnetite through
magnetic separation. At the Foskor plant, conditioning at elevated temp. of 40 oC
involves the following reagents: 500 g/t of sodium hydroxide (pH 8 or above), 1000 g/t of
sodium silicate as a depressant, tall oil as frothing collector (200 g/t) and 50-100 g/t of
nonylphenoltetraglycol ether as a carbonate depressant. When the third type of the ore is
treated, an additional 100-150 g/t of gum Arabic is used as carbonate depressant and
froth controller. Recoveries range from 75 to 85% of phosphate for the two first types of
ores. The output of the Foskor plant is over 3 million tones per year of the concentrate
assaying 36% P 2 O 5 .
In the third category of igneous phosphates apatite with dominantly carbonate
(calcite) gangue processing plant at Jacupirange (Brazil) is the first in the world to treat
an apatite-calcite ore by flotation. The ore contains about 5% P2O5. Calcite is a chief
gangue but dolomite and iron oxides are also present. The process involves grinding in
rod mills in closed circuit with magnetic separators and hydrocyclones (desliming at 20
microns) and anionic flotation of apatite with fatty acids (200 g/t). Causticized corn
starch is used as calcite depressant. Overall recovery is 80%. Overall recovery is 80%.
The losses in the desliming operation account for 12% of the total losses. The concentrate
assays 36% P 2 O 5 . The annual output is 0.3 million tones of the phosphate concentrate
and 0.5 million tones of calcite tailings.
The best example of sedimentary phosphate ores with dominant siliceous gangue
is the Florida (Bone Valley) phosphate industry. The ore contains chiefly francolite, sand
and clays. As the mining of Florida phosphate gradually moves to the south and/or west
of the historically significant Bone Valley, there is a reduction in the production of
pebble phosphate (+1 mm) that can be produced by washing, from about 50% of total
product to about 20%. There is also a greater probability of experiencing a contamination
of the matrix with dolomite.
Clays are separated by classification (they are 100% passing 150 mesh screens).
Coarse phosphate (pebble)can be concentrated on screens. The remaining feed between
16 and 150 mesh, a combination of sand and phosphate, is processed by flotation. For
ores with low carbonate content, two main flotation techniques are used: (i) direct
flotation of phosphates with fatty acids under moderately alkaline conditions and with
modifiers to depress gangue minerals, and (ii) reverse flotation with amines to float the
silicate minerals. The beneficiation of Florida phosphates is achieved by anionic flotation
of phosphate followed by cationic flotation from the acid scrubbed and deslimed flotation
concentrates.
The rougher flotation with fatty acids and fuel oil is not particularly selective, thus
after rougher flotation there is excessive sand left with the phosphate (quartz in the
presence of calcium ions floats with fatty acids).
This concentrate must be purified in a cleaner flotation circuit with amines. The
upgrading is accomplished in three steps. The concentrate is first scrubbed with sulfuric
acid to remove the reagents. The slurry, after scrubbing, passes through a counter-current
wash to reject the acidic, reagentized water. The clean feed is then reacted with amine.
The quartz flotation follows. Most plants sell at least two final products: flotation
concentrate and pebble.

166
The industry follows so-called Lawvers four laws:
(i)
Thoroughly remove clays from flotation feed;
(ii)
Make good, sharp size separations;
(iii) Conduct the rougher reagent conditioning at high percent solids (about
75% solids); longer conditioning time also helps;
(iv)
Feed all process operation at stable rates.

Fig. 5d.4. Florida washer plant.

167

Fig. 5d.5. Schematic of flotation plant.

The Florida phosphate industry utilizes the Cargo double flot process. In the
first step, anionic flotation, tall oil (375-750 g/t) and fuel oil (750-1500 g/t) are used.
pH is slightly alkaline (8-8.2). The most important is a high intensity conditioning at as
high solids loading (about 75%) as possible prior to the flotation. The concentrates from
this flotation are first scrubbed with sulfuric acid to remove the rougher reagents before
the cationic flotation is carried out in a separate flotation circuit.

168
R.O.M. Matrix
82.9% - 16 mesh
12.8 % P 2 O 5
1 x 107 t/year
WASHER Pebble (5.4% -16 mesh,

30.8 % P 2 O 5 , 1.7x106 t/y)


98.3% - 16 mesh
9.3 % P 2 O 5
Slimes
94.2 % -150 mesh
8.3% P2O5
2.8 x 106 t/y

Tailings
97 % - 16 mesh
2.5 % P 2 O 5
4.2 x 106 t/y

DESLIMING

97.4% -16 mesh


0.8% -150 mesh
9.8 % P 2 O 5
5.5 x 106 t/y

FLOTATION PLANT

Fig. 5d.6. Schematic flow diagram for central Florida phosphate.

Concentrate
98.7 16 mesh
0.2% - 150 mesh
33.0 % P 2 O 5
1.3 x 106 t

169
Chapter 5e. Flotation of Soluble Salts
Potassium, along with phosphorus and nitrogen, is a major growth nutrient. As a
commercial source of potassium, potash ores form basis for the manufacture of fertilizers
for the agriculture industry. Kramer and Schubert quoted in 1991 that the world
production of potassium salts amounts to 30 million tones per year of equivalent K 2 O.
Of many known potassium-bearing minerals, the most commercially important is
sylvite (KCl). The most important type of potash ores is sylvinite which contains mostly
sylvite and halite (NaCl). In the Saskatchewan sylvinities, sylvite and halite are
accompanied by carnalite (KClMgCl 2 6H 2 O) in certain localities, with brine-insoluble
matter being represented by clay minerals, dolomite and anhydrite; on average, the
content of insolubles does not exceed a few percent of the total rock mass.
60% to 80% of the potassium ores are processed by flotation. Since two major
minerals (sylvite and halite) are water-soluble, the flotation process is carried out in
NaCl-KCl saturated brine (at room temperature about 7 mole/L of NaCl+KCl solution).
In the selective flotation of sylvite from halite long-chain primary amines are used as a
collector to float sylvite.
Because long-chain primary amines are insoluble in water they are utilized in the
form of aqueous dispersions in potash flotation. The amine is heated up to about 80 oC
and melted it is partly neutralized with acetic (or hydrochloric acid) and dispersed in
water. The resulting emulsion is used as a collector.
As Fig. 5e.1 shows (Roman, Fuerstenau, Seidel, Tran. AIME, 1968), flotation
recovery commences only when the concentration of the collector (long-chain amine)
exceeds the precipitation limit.

Fig. 5e.1. Relationship between recovery of KCl and amine concentration (Roman,
Fuerstenau and Seidel, Trans. AIME, 1968).

170

This was confirmed by Miller, Yalamanchili (1994); while KCl starts floating when
dodecylamine starts precipitating, the flotation of NaCl does not float with cationic
collector (amine) at all. By measuring contact angles at hexadecane drops on KCl surface
in saturated brine they showed that only contact angles as small as 7 degrees were
obtained at dodecylammonium chloride concentrations below its solubility limit; the
contact angles as large as 52 degress were measured at the amine concentrations
exceeding the solubility limit.

Fig. 5e.2. Flotation recovery of KCl and NaCl as a function of dodylammonium chloride
concentration (Miller ands Yalamnchili, 1994).
The reverse situation developed when working with anionic collector (Fig. 8g.3), NaCl
floated and KCl did not.

Fig. 8g.3. Flotation recovery


of NaCl and KCl as a
function of sodium laurate
concentration (Miller and
Yalamanchili, 1994).

171

As reported by Schulman and Rogers in the 50s, long-chain amines are


practically insoluble in a saturated NCl-KCl brine. Surface tension measurements on
saturated brine with dodecylammonium chloride crystals added did not show any frothing
and only a decrease of surface tension by a few mN/m after a week. This leads us to the
observation made by Leja (1983) that in quiescent environment no contact angle or
pick-up was obtained after amine was deposited on the surfaced of unstirred brine.
However, after throughout stirring for a few minutes KCl crystals were picked up and
contact angles developed on the KCl surfaces. This prompted the author to conclude that
since long-chain amines are practically insoluble in brine the underlying mechanism at
such a high ionic strength is likely to be different from that in conventional flotation
which involves diffusion of the dissolved collector species and selective adsorption.
Figure 5e.4 shows the flotation results published by Schubert in 1967. These
curves demonstrate that flotation of sylvite and halite with long-chain amines depends on
pH. While KCl floats in a broad pH range up to 11, flotation of NaCl only begins when
this pH is exceeded. Comparison with Fig. 5.11 reveals that the flotation recovery of both
KCl and NaCl particles revolves around pH 10-11 in the same way as the electrical
charge of the colloidal amine precipitate. Since long-chain amines are insoluble in brine,
they must appear in such a system in the form of colloidal species (or as film around
bubbles as postulated by Leja). With the discovery made by Millers group (University of

Fig. 5e.4. Effect of pH on the flotation of halite and sylvite, and on the adsorption of
long-chain primary amines of these minerals (Schubert, 1967).
Utah) that KCl crystals exhibit negative electrical charge in water, while NaCl is charged
positively, it is apparent that electrical interactions between KCl crystals and colloidal
amine particles play an important role in the potash flotation. Of course, while electrical
interactions seem to be important, they cannot explain all the experimental observations.
One of the most obvious is that only primary straight-chain amines float sylvite well.

172
Only crystallochemical conformity of solidified amine film with the lattice of KCl crystal
can explain such findings. And most telling, industry uses longer chain amines at higher
summer temperatures and this again confirms that an insoluble amine in this system acts
as a collector.
The sylvinite ores may contain up to 10% of insoluble minerals (dolomite
anhydrite, illite, etc.). They appear in the pulp in the form of fine slimes. These particles
have a very large specific surface areas and adsorb substantial amount of the collector.
One of the characteristic features of potash flotation is the use of so-called blinders.
These are modifying (depressing/dispersing) agents. Guar gum, CMC or starch are
commonly used. Figure 5e.5 shows how they operate. A good blinder is expected to
prevent an amine collector from adsorbing onto fine gangue particles (slimes). Without
the proper use of a polymeric blinder, the high surface area of the gangue can prevent the
coarse particles from getting an enough share of the collector to gain floatability.

Fig. 5e.5. Functions of amine and blinder in potash flotation.


The benefits resulting from the use of a blinder in potash flotation are clearly seen
from Fig. 5e.6. In order to get over 80% KCl recovery 1 kg/t of Octadecylamine (ODA)
is needed if used without t a blinder. In such a case a slime-stabilized thick froth appears
which is very difficult to handle and which renders poor sylvite recovery. Curve 2 shows
that CMC is a very good blinder: (1) in order to get the same 84% recovery only 100 g/t
of ODA is needed at about 350 g/t of CMC; (2) the recovery of insoluble slimes
diminishes noticeably. PAM has good blinding abilities (it reduces the amine
consumption needed for high KCl recoveries) but it also increases flotation of the
insolubles. Ligninosulfonate is a poor blinder, it does not improve KCl flotation, but it is
a good slime depressant/dispersant as it reduces flotation of the insolubles.

173

Fig. 5e.6. Effect of blinders on flotation of sylvinite ore with octadecyl amine.
Please note: stands for recovery, n.res. stands for insoluble residue.
Curve 1, flotation of sylvinite ore containing 21% KCl, 72% NaCl and 4-5% of clays
with octadecyl amine only (ODA); Curve 2, flotation with 100 g./t of ODA and varying
dosage of CMC; Curve 3, the same as 2 but using polyacrylamide (PAM); Curve 4, the
same as 2 but using lignsulfonate; Curve 5, the same as 2 but using a proprietary cationic
polymer VA-2 (V.A. Arsentiev et al., 16th Int. Mineral Processing Congress, Stockholm,
1988).
In the case of the Saskatchewan sylvinites, liberation is substantial in some cases
at 9.5 mm and virtually complete in all cases at 1.2 mm. Coarse (+0.8 3 mm) and fine (0.8 + 0.1 mm) fractions are conditioned with reagents separately. Because of the presence
of slimes in the potash ore flotation pulps, and their detrimental effect on sylvite flotation,
one of the most important operations in this flotation process is desliming which follows
scrubbing at high solids content (70%). This operation can include only mechanical
desliming (classification in cyclones), or both mechanical desliming and desliming by
selective flocculation/flotation of slimes prior to the sylvite flotation. In this latter process
polyacrylamide (PAM) is used as a flocculant and secondary amines as a collector to
float off slimes. While only long chain (C 18 -C 22 ) primary amines are used in the
flotation of the sylvite fine fraction, in addition to the amine also extender oil is used in
the flotation of the coarse fraction. Another characteristic feature of the Saskatchewan
flotation operations is the use of longer chain primary amines in the summer time.

174

Fig. 5e.7. Simplified flowsheet of potash flotation with mechanical desliming.

Fig. 5e.8. Flowsheet which includes flocculation/flotation desliming.

175

Fig. 5e.9. The flowsheet with flocculation/flotation desliming as a major desliming


operation.

You might also like