Sei sulla pagina 1di 30

Helsinki University of Technology

Mechanical Process Technology and Recycling

FUNDAMENTAL CONCEPTS IN FLOTATION

Vesa Kirjavainen
Preface
Flotation has been developed for more than hundred years and it contains expertise
from many fields including control and automation, mineralogy, surface chemistry,
inorganic and organic chemistry, just to mention a few. Therefore, it may be difficult
for students to outline the essential features of flotation. Because there are several
different applications of flotation, the purpose of this contribution was to collect basic
phenomena together and present them in the way that it would be easy to see the
connections between them. This is not an exhaustive scientific contribution but rather
lecture notes or a view and vision, which could be different. This view can also
become different with time and development and it should, therefore, be rewritten.
For those who are interested in more details, some books and contributions are
presented in the literature part.
Contents

1. INTRODUCTION............................................................................................................................... 1

2. PROPERTIES OF WATER .............................................................................................................. 2

3. INTERFACES ..................................................................................................................................... 5
3.1. WATER/AIR INTERFACE ............................................................................................................... 6
3.2. INTERFACES OF SOLIDS ................................................................................................................ 8
3.2.1. Wetting............................................................................................................................... 9
3.3. ELECTRICAL DOUBLE LAYER - ZETA POTENTIAL ...................................................................... 11
4. FLOTATION REAGENTS ............................................................................................................. 14
4.1. FROTHERS .................................................................................................................................. 14
4.2. COLLECTORS .............................................................................................................................. 14
4.2.1. Thiol (sulphydryl) collectors........................................................................................... 15
4.2.2. Fatty acids or carboxylic acids....................................................................................... 16
4.2.3. Sulphates and sulphonates.............................................................................................. 16
4.2.4. Cationic collectors .......................................................................................................... 17
4.2.5. Amphoteric and chelating agents ................................................................................... 17
4.3. REGULATING AGENTS ................................................................................................................ 17
4.3.1. Slurry pH ......................................................................................................................... 17
4.3.2. Sulphide minerals............................................................................................................ 18
4.3.3. Silicate minerals .............................................................................................................. 19
4.3.4. Oxide and salt type minerals........................................................................................... 20
5. EFFECT OF FINENESS ON FLOTATION................................................................................. 20
5.1. STABILITY OF COLLOIDS ............................................................................................................ 22

LITERATURE
1

1. Introduction
Flotation is a separation method where solid particles are separated by gas bubbles,
usually by air bubbles, in an aqueous medium. For flotation, it is required that the
surfaces of the floatable particles are water-repellent. Air is introduced into the pulp
so that the hydrophobic particles can be attached onto the air bubbles and rise through
the pulp to the surface forming a mineralized froth layer, which can then be removed.
Wetted, hydrophilic particles do not stick to air bubbles and remain in suspension.
The process is usually known as froth flotation.

Flotation is widely used in mineral processing but it is also applied in de-inking in


paper industry, water treatment and recycling. Due to its history, most terms are
related to mineral industry. For example, when the valuable minerals are separated
into the froth product, the process is called direct flotation. The froth product is
generally called concentrate and the product of non-valuable gangue minerals tailing.
If the gangue minerals are floated, the process is called indirect flotation. Then the
unfloated product is naturally the concentrate.

Most materials and minerals are hydrophilic and wettable in aqueous medium. This is
due to the electrical charges on the surfaces, either positive or negative. The surface
charges attract polar water molecules, as well as hydrogen and hydroxyl ions. For
flotation, the hydrophilic surfaces must be modified hydrophobic by chemicals. The
chemicals used for this purpose are called collectors. However, the surface structure
of some minerals, such as talc, molybdenite and graphite, is inherently neutral and
hydrophobic. Therefore, these minerals can be floated without a collector.

By blowing air into the pulp one can generate bubbles for flotation. However, the
surface tension between air and water is high (7.2 10-4 Nm-1) and after rising to the
surface the bubbles are broken, and they are not capable to carry particles attached to
them. To avoid this, a surface-active chemical is added into the pulp so that a stable
froth bed is formed. Chemicals used for this purpose are called frothers.

The purpose of flotation is usually to separate one or more minerals from the other
minerals, and the selective adsorption of collector on floatable minerals is a
prerequisite for successful separation. The selectivity of collector adsorption is often
contributed by regulating agents. They can either enhance or prevent the adsorption of
collector on a particular mineral. Therefore, they are usually called activators or
depressants.

Thus, flotation reagents can be classified into the three main categories according to
their usage: collectors, frothing agents and regulating agents. However, this
classification is not conclusive. For example, there are collectors, which have the
properties of frothing agents and frothing agents, which have collecting power. In the
same way, an activator may be a depressant under other circumstances, and vice
versa. In addition, it is to be emphasized that chemicals used in flotation vary widely
in composition including organic and inorganic compounds, acids and alkalis, salts of
various compositions, water-soluble substances and materials that are practically
insoluble in water.
2

2. Properties of water
Water is the phase in which all the main processes take place in flotation. The
processes that affect the surface characteristics of particles in water include
dissociation of dissolved species, hydration, and the adsorption of ions and flotation
reagents. Therefore, it is important to know the properties of water.

Water is a polar compound and water molecules interact with each other by attractive
forces called van der Waals forces. These forces are closely related to the polar
structure of molecules. As well known, positive charges repel positive charges and
negative charges repel negative charges, but positive charges attract negative charges.
As a consequence, dipole molecules tend to take a position so that attraction between
molecules occurs. This phenomenon is called Keesom orientation. Brownian motion
disturbs the orientation. A dipole can also induce a dipole moment in another
molecule causing attraction between the molecules. The phenomenon is called Debye
induction.

Nonpolar molecules do also interact with each other. In all atoms and molecules the
continuous motion of negative electrons creates rapidly fluctuating dipoles with the
positive nucleus causing attractive forces. These forces are called dispersion or
London-van der Waals forces. The interaction of nonpolar molecules is caused by the
dispersion forces but with polar molecules, such as water, also the Keesom orientation
and the Debye induction forces occur. The total van der Waals interactions between
two atoms or molecules are given by the sum of those due to orientation, induction
and dispersion forces. The orientation interaction is significant only if dipole moment
is high. The induction interaction is always small. According to Coulomb’s law
electrostatic forces vary inversely with the second power of the distance between two
charges. The interaction due to van der Waals forces is much weaker. The forces
decay inversely with the sixth power of the distance between molecules. When atoms
or molecules come very close to each other their electron clouds repel each other.
Therefore, the resultants of van der Waals forces contain both attraction and repulsion
terms.

In addition to van der Waals forces, hydrogen bonding is characteristic for water
(Fig.1). Hydrogen bonding occurs when an atom is attracted by rather strong forces
to two atoms instead of only one, so that it may be considered to act as a bond
between them. In water the hydrogen atom is covalently attached to the oxygen (about
470 kJ/mol) but has additional attraction (about 23.3 kJ/mol) to a neighbouring
oxygen atom of a water molecule. The bond is partly (about 90%) electrostatic and
partly (about 10%) covalent (Isaacs et al., 2000, Suresh and Naik, 2000). Typically
hydrogen bonding occurs where the partially positively charged hydrogen atom lies
between partially negatively charged oxygen and nitrogen atoms, but is also found
elsewhere, such as between fluorine atoms in HF2, and between water and smaller
halide ions F-, Cl- and Br-.

Water appears to be a simple compound but it has many specific properties. Water
molecules form an infinite hydrogen-bonded network with localized and structured
clustering. According to the present view, these clusters may contain 4-12 water
molecules but much larger clusters have been suggested to occur (Chaplin, 2000). The
lifetime of the structured clusters is very short, pico seconds in magnitude. Several
3

models have been suggested for the structure of liquid water but no model is able to
describe all the anomalous properties of water so far.

H H
O
H H H
O
O
H H H O

O H

H
Figure 1. Schematic representation of hydrogen bonding of water molecules.

Ions destroy the natural hydrogen bonded network of water. If the energy of inter-
action between an ion and water dipoles is greater than the mutual attraction of water
dipoles, the ion will be hydrated. Water molecules are oriented around the ion
forming new structures. The degree of hydration depends on the size and valence of
the ion. Anions are hydrated more strongly than cations of the same size because
hydrogen atoms of water can approach closer than oxygen atoms of water. Ions that
exhibit weaker interactions with water than water itself are known as structure-
breakers, whereas ions that interact strongly with water are known as structure-
makers.

Small ions are strongly hydrated creating local order and higher local density. Large
monovalent ions, such as Cs+ and I-, are very weakly solvated. Their surface charge
density is low and they may be pushed on by strong water-water interactions. Their
translational movement is high. Single atom ions may also be found in clathrate
structures, where the lattice of water contains cavities that are capable of enclosing
molecules without any bonds between them. Smaller ions, such as Rb+ and K+, cause
the partial collapse of clathrate structures, through puckering, increasing the local
mobility of water molecules. The smallest ions hold strongly to the first shell of their
hydrating water molecules and hence there is less localized water molecule mobility.
Divalent and trivalent ions are always more strongly solvated than monovalent ions.
In the primary hydration shell the water molecules are most restricted in their motion
but the effect does not end there. In the second hydration shell the water molecules are
freer to rotate and exchange with bulk water, and so on. The position of hydrated
water molecules on anions and cations is different and so is their ability to form
hydrogen bonds. Altogether, the properties of water depend on all the ions and their
characteristics.

Ions and their hydrations affect the properties of water, such as viscosity, in many
ways. Hydration is an exothermic process. During the formation of the internal layers
of hydrated sheaths there is a considerable quantity of heat evolved. During the
formation of the subsequent layers the amount of heat gradually decreases. If
temperature is increased, the hydration of ions decreases. This is explained so that the
rotational movement of water molecules hinders their orientation.
4

The increase of orientation and the stability of the oriented dipoles decrease the
solubilizing properties of water. The solvent action of water is closely connected with
the hydration of the dissolved ions. If the dipoles of water are already polarized then
the hydration of new ions by these water molecules is hindered. For the same reason,
the conditions for the diffusion of ions become more difficult in polarized water. In
addition, hydrated layers can prevent the adsorption of reagents on particles.

The ionic composition of water is determined by the solubility of particles. Minerals


are soluble if their hydration energy exceeds the lattice energy. Ion hydration energy
increases as the valence of the ion increases and the ionic radius decreases. Also, the
energy of crystal lattice increases. However, hydration energy increases much more
slowly as ion valence increases than does crystal lattice energy. Therefore, an increase
in valence is accompanied by a great reduction in solubility. This is why the sulphides
and oxides of bivalent metals are relatively insoluble in water. The rate of dissolution
depends on the nature of the mineral, the temperature and pH of the pulp, the intensity
of agitation, particle size and the specific surface of the particles, and the ionic
composition of water.

The specific surface of particles determines the overall area that is in contact with
water and, consequently, the number of ions that are transferred into the solution per
unit of time. The intensity of agitation determines the movement of ions away from
the surface of particles. These factors affect the kinetics of dissolution.

The temperature and pH of the pulp do not affect only the kinetics of dissolution but
also the equilibrium concentrations of the dissolved substances. In most cases, the
solubility of minerals increases with an increase in temperature. This is due to the
higher vibrational energy of the constituents in the crystal lattice, and at the same time
due to the decreased forces between the ions, which facilitates the penetration of
water into the lattice.

Ionic equilibrium and solubility are important characteristics of solutions and of


chemical reactions that occur in water. A chemical reaction in solution is possible
when, on collision of ions, molecules are formed in which forces of cohesion between
atoms exceed the forces of hydration. A requirement for a reaction to proceed is a
removal of ions from the solution in the form of weakly dissociated molecules or
nearly insoluble substances, i.e. as precipitate or as gas. At the high ionic
concentration of a weakly soluble substance the solubility decreases. The solubility of
minerals depends on complexes that are formed in each particular case. If the solution
contains similar ions than the mineral, the solubility is decreased. The presence of
foreign ions increases the solubility of minerals.

Natural waters contain different impurities, such as ions, dissolved gases and organic
substances. Typical ions occurring in natural waters are chloride (Cl-), sodium (Na+),
potassium (K+), calcium (Ca2+), magnesium (Mg2+), iron (Fe3+/Fe2+), sulfate (SO42-),
bicarbonate (HCO3-) and carbonate (CO32-). The solubility of minerals may be quite
variable and often becomes significant, causing also reactions with flotation reagents.
Salts, for example, form in many cases insoluble complexes with flotation reagents. It
is difficult to evaluate the solubility of minerals in different conditions. The solubility
of minerals is usually given in pure or distilled water. However, the solubility of
minerals is, as described above, dependent on ions dissolved in the pulp. For example,
5

the solubility of barite increases by factor 14 in the presence of 10-g/l magnesium


chloride. For the same reason, when evaluating the results of laboratory tests, it is
important to remember that the composition of process water may be different.

3. Interfaces
Flotation depends on complex phenomena that occur at the interfaces of solid, liquid
and gas phases. In the bulk phase every atom of a solid and molecule of a liquid is
surrounded by other atoms or molecules. The surface of a solid or liquid phase differs
from the bulk phase so that there is not a balance between the inter-molecular forces
at the surface. At the surface there are cohesive forces on one side only. The atoms or
molecules that reside at the surface have more energy than those in the bulk phase and
can interact with atoms or molecules that approach their sphere of influence. This
excess of energy is called surface energy.

However, the surface energy is not localized in the surface monolayer but is spread
over the so-called surface layer, which attains its maximum at the boundary that
separates the phases. The thickness of the surface layer is of the order of 10-8-10-7 cm.
Surface energy is dependent on the cohesive forces of the bulk phase. As temperature
increases, the atoms in a solid vibrate more decreasing the cohesive forces and the
surface energy.

Surface energy is often given as energy per unit area using ergs per square centimeter
(erg2/cm) or in SI units joules per square meter (J/m2). For liquids, however, surface
energy is usually expressed using an equivalent quantity called surface tension. Then,
the units are dynes per centimeter (dyn/cm) or Newtons per meter (N/m). Surface
tensions reported in dyn/cm and mN/m have the same numerical value. The surface
energies of liquids can be determined accurately by several methods but an accurate
method for measuring the surface energy of solid materials has not been developed.

The mutual attraction of molecules in a homogeneous phase (for example, in water) is


called cohesion. The work of cohesion is defined as the reversible work, per unit area,
which is required to break a column of liquid into two parts, creating two new
equilibrium surfaces, and to separate the parts beyond the range of their forces of
interaction. For liquids the work of cohesion can be evaluated on the basis of surface
tension (g) by
WC = 2g. (3.1)

The attraction of molecules at the interface of two phases (for example mineral and
water) is called adhesion. The work of adhesion is defined as the work, per unit area,
required to separate the phases so that the original interface disappears and two new
interfaces (solid/vapour and liquid/vapour) are created. It is equal to the sum of the
surface energies of both phases taken separately, minus the interphase surface energy
on their interface. The work of adhesion for two phases is usually written by using
surface tensions, and so-called Dupré equation is

WA = g1 + g2 - g12. (3.2)
6

3.1. Water/air interface


In pure water the interaction of molecules is mainly based on van der Waals forces. At
the water/air interface the cohesive forces acting on the surface molecules exceed
those from the gas side and the resultant draws the surface molecules into the liquid
phase (Fig. 2). The resultant is formed of two components. One is perpendicular to the
surface, and the other is directed tangentially to the surface preventing its area from
increasing. The tangential component is called surface tension.

Figure 2. Surface layer at the water/air interface.

Surface tension tends to diminish the surface area of the liquid and to make new
surface one has to work against the surface tension. For a one-component system at
constant temperature and pressure, the work required, per unit area, is equal to the
increase of Gibbs free energy. Thus

dws = gdA. (3.3)

Since the reversible work is equal to dG

Ê G ˆ
= g = Gs , (3.4)
Ë A ¯ T , P

where Gs is the increase of Gibbs free energy per unit area. Thus surface tension
represents the work necessary to form unit area of new surface or the increase of
Gibbs free energy corresponding to the formation of unit area of surface.

The thermodynamic properties of a phase are defined separately for the surface and
the bulk phase so that they are additive. We can consider for example a homogeneous
liquid phase surrounded by a surface area A. If the enthalpy and entropy of the bulk
phase are denoted by Hb and Sb, and Hs and Ss represent the enthalpy and entropy of
the surface per unit area, the thermodynamic quantities can be given by
H = Hb + HsA, (3.5)
S = Sb + SsA, (3.6)
G = Gb + GsA, (3.7)
where
Gb = Hb -TSb (3.8)
and
Gs = Hs -TSs. (3.9)
7

Thus, the thermodynamic properties of surfaces are defined as excesses of the


measured thermodynamic properties that are due to the presence of the surface. Since
Ê Gs ˆ Ê g ˆ
= = -S s, (3.10)
Ë T ¯ P Ë T ¯ P

after substituting equations (3.4) and (3.10) in equation (3.9), and remembering that
surface energy Es is often approximated by surface enthalpy Hs, for the total surface
energy can be written
dg
Es = g - T . (3.11)
dT

The surface layer of pure water is relatively thin due to the short range of van der
Waals forces. Only the first shell or two of nearest neighbours are of importance in
interaction for a molecule. Further away from the surface a molecule experiences
rather symmetrical forces. The surface is, however, in a turbulent state. When the
liquid is in equilibrium with its vapour phase, there is a balanced traffic of molecules
hitting and condensing on the surface. According to the gas kinetic theory, it can be
calculated that there are 1.2x1022 molecules that strike 1 cm2 area of the surface per
second at room temperature. Each square centimeter of liquid water entertains
approximately 1.2x1022 arrivals and departures per second. In addition, corresponding
to the area of a single water molecule (100Å2), the traffic is 1.2x107 per second, and
the lifetime of a molecule on the surface is on the order of a tenth microsecond.

Ions and chemicals tend to change the surface tension of water. In general, many
organic substances lower its surface tension and inorganic salts slightly raise it. Gibbs,
who was the first (1876) to study the adsorption and its thermodynamics at the
solution/gas interface, deduced that the concentration of organic substances in the
surface layer is higher than in the bulk phase, whereas the concentration of inorganic
salts is lower than in the bulk phase. Therefore, the organic substances are said to be
positively adsorbed and the inorganic salts negatively adsorbed. The positively
adsorbed substances are generally called surface active. The amount of the adsorption
in dilute solutions is usually given by the Gibbs equation (theGibbs adsorption
isotherm)
c dg
G=- (3.12)
RT dc

where R is the gas constant, T the absolute temperature, c the concentration (activity)
of the solute and g the surface tension at the air/solution interface. As can be seen
from equation (3.12), G is positive if surface tension increases, and negative if surface
tension decreases.

Surface active substances are compounds that contain one or more polar groups, such
as -OH, -COOH, -NH2 and -SO3H, which are readily solvated, and a nonpolar
hydrocarbon part. This kind of compounds, such as alcohols and short-chained fatty
acids, are soluble in water when the affinity of water molecules to the polar group(s)
is strong enough to draw the polar part into the solution. At the air/water interface
surface-active molecules tend to be oriented so that the polar group is surrounded by
water molecules and the nonpolar part is headed towards the air. Surface activity and
the kinetics of adsorption are dependent on the hydrocarbon part and the structure of
8

the molecule. According to the Traube rule, which is applicable for dilute solutions,
for identical homologues each additional CH2 group decreases the concentration
required to give a certain surface tension approximately by a factor of 3. At elevated
temperatures the adsorption of surface-active substances on the interfaces decreases.
Thus, differing from the bevaviour of pure water, the surface tension increases with
increasing temperature.

3.2. Interfaces of solids


The structure of solid materials is crystalline or amorphous. Minerals are usually
crystalline. Amorphous are such substances as some fused oxides, bitumen and
plastics. The surfaces of crystalline solids are always very heterogeneous in character.
On the atomic scale crystal surfaces contain terraces, ledges and kinks whether metals
or covalent or ionic crystals are in question. All solids also possess imperfections such
as point defects and dislocations, and whenever a surface is created, it incorporates
some of these imperfections in addition to ledges and kinks. When a simple crystal is
considered, there are differences between the energy of atoms depending on their
position in crystal lattice (Fig. 3). In the body of crystal lattice the atoms expend all
their energy in reacting with the neighbouring atoms. In the outermost layer the atoms
react only with the atoms beneath the surface and part of their energy remains free.

Figure 3. Dependence of atom bond saturation


on the position in crystal lattice:
A - all bonds are saturated within C D
the lattice,
B - on the surface certain extent of
bond energy is unsaturated,
C - on the edge the reactive power of A B
atoms is increased,
D - at apex the atoms have the greatest number of
unsaturated (free) bonds and are most active.

In solid substances atoms are in practice immobile, and surface properties depend on
the immediate past history of the solid. A cleavage surface created in vacuum has
different chemical and physical properties from ground or polished surfaces made in
the atmosphere. Any heat treatment and reactions with gases or adsorption of organic
molecules produce different surfaces. Broken bonds on the surface give the surface
atoms excess of energy, surface energy, which is theoretically equal to half of the
cohesion energy (bond energy). Immediately after breaking the surface atoms undergo
some rearrangement strengthening the bonds with the atoms below the surface. This
consumes the surface energy. Some surface energy is lost in heat radiation and in
electrical discharges. The adsorption of gas molecules on the surface vill also satisfy
broken bonds consuming surface energy. The result is the surface energy associated
with the new surface. Surface energy is characterized by broken bonds and the
number of broken bonds at terrace < at ledge < at kink. Thus, the probability of
adsorption of molecules from the neighbouring phase increases in the same order.
Due to anisotropy of crystalline solids surface energy depends also on
crystallographic orientation.
9

The reactivity of particles depends on the occurrence of active sites and heterogeneity
on the surfaces. When new surface is produced by mechanical means, such as
grinding, abrasion, cutting, etc., the surface layer is damaged, and the thickness of the
layer may be a few ångströms or hundreds of microns depending on the method used.
In grinding different phenomena may occur. For example, in addition to particle size
reduction, new crack systems (micro cracks) are created, the density of dislocation
population may change and polymorphism, surface activation and mechano-chemical
reactions may occur (Lin et al., 1975).

Only a percentage of the energy consumed in grinding is used to create new surface.
Therefore, temperature and pressure tend to increase on particle surfaces. As a
consequence, physical and chemical properties of particles may change. Long dry
grinding is found to convert the surface of quartz amorphous. Dry grinding may also
cause polymorphic transformation, for example changing calcite to aragonite. Carta et
al. (1973) have found that calcite and fluorite changed their semiconducting
characteristics from n type to p type. A prolonged dry grinding has also been found to
cause decomposition of some carbonates to oxides. However, although grinding may
cause changes in surface-phase properties, such changes usually require high energy
input and very fine grinding. In addition, surface products, such as amorphous quartz,
are easily dissolved by water. Thus, the effect of comminuting on the surface
properties of particles in the size range (> 10 mm) usually applied in flotation has, at
least so far, turned out to be less important.

In considering flotation, it is to be emphasized that alteration of particle surface


properties due to grinding is not restricted to the factors described above but depends
also on the grinding method. Grinding is usually carried out in steel mills using steel
grinding media. During grinding, there is always some wearing of the mill liners and
the grinding media. A part of the abraded metal tends to be stuck onto the particles. In
wet grinding of sulphide ores, the abraded metal is in galvanic contact with sulphide
particles affecting the electrochemical reactions necessary for flotation, for example,
the interaction of thiol collectors. The galvanic effect can reduce pulp potential so
strongly that it creates difficulties in recovery and selectivity in sulphide flotation.
Particularly, processing of finely disseminated ores, that require very fine grinding for
mineral liberation, is often a very complex task.

3.2.1. Wetting
As shown in the previous section the surfaces of particles are characterized by
unsaturated bonds, which determine their reactivity with water and the ions dissolved
in it. Wetting is an example of the surface reactions. It may be considered as the
adsorption of water ions and molecules on the particle-water interface. The final
reaction at the interface depends on the free surface energy of the solid particle and on
the energy of reaction shown by the water ions and molecules among themselves.

Ionic and covalent bonds are typical for minerals. They are strong bonds and give a
high free energy and polar characteristics for new surfaces. As water is itself a polar
liquid and the molecules have a considerable dipole moment, this is the reason why
most minerals are wettable. Weak van der Waals bonds (molecular bonds), which also
occur in minerals, give low surface energy and nonpolar characteristics after fracture
producing surfaces that are hydrophobic in nature.
10

Such minerals as graphite, molybdenite, sulfur and stibnite create nonpolar surfaces.
For example, graphite and molybdenite have a layered structure. In these layers atoms
form covalent bonds but between the layers there are only van der Waals bonds.
When particles cleavage along these layers they form nonpolar and hydrophobic
surfaces.

Hydrated skins are formed when the bond energy between the water dipoles and the
surfaces of particles is greater than the strength of the bond between the water dipoles.
The adsorption of water shows characteristics of physical and chemical adsorption.
Physical adsorption is nonspecific, reversible and has characteristics arising from the
van der Waals forces. On the other hand, hydrogen bonding and high dipole moment
give water the character of chemical adsorption. Moreover, adsorption of water is an
exothermic reaction like any adsorption process reflecting the strength of the bond.
Considering flotation, some investigations have shown (Frumkin, 1938) that floatable
particles can be hydrated. Extremely thin hydrated layers were found to exist on the
solid surface under the air bubble. Thus, it is obvious that hydrophobic dry surfaces
change gradually to hydrophilic surfaces with the increasing thickness of the water
layer, and very thin water films do not necessarily prevent air bubble attachment. The
properties of thin water films are different from the properties of water in bulk phase.

Wetting can take place by different mechanisms, where the driving forces or the free
energy changes are different (Rosen, 1978). In spreading wetting water is in contact
with the solid surface and water-saturated air is in equilibrium with the surface. Water
displaces air, and a solid/air interface is replaced by a solid/water interface. At the
same time, a water/air interface is formed. The free energy change (G on a unit-area
basis) in spreading wetting, when G1 and G2 correspond to the situation before and
after the wetting, can be written

-G = G1 - G2 = gSV - gSL - gLV (3.13)

where gSV, gSL and gLV are the solid/vapour, solid/liquid and liquid/vapour interfacial
tensions, respectively. Another wetting type, where a particle is completely immersed
into the liquid, can be called immersional wetting (Rosen, 1978). In this case, if so-
called Young’s equation is applied, for the surface free energy change can be written

-G = G1 - G2 = gSV - gSL = gLV cosq (3.14)

where q denotes contact angle, which means the angle between the tangents of air
bubble and solid particle in the contact point of the three-phase system. It can be seen
that for contact angles 0-90o cos q is between 1 and 0 which means that wetting is
spontaneous. If the contact angle is greater than 90o one has to make work to wet the
solid. In all wetting processes reduction of the interfacial tension between solid and
the wetting liquid (gSL) increases wetting. Also, the reduction of the surface tension of
the liquid (g LA) improves wetting. Because the surfaces of floatable particles are
usually hydrophilic chemicals are needed for their flotation. The equations presented
above show that, in addition to collectors, also the role of frother is important in
flotation.
11

3.3. Electrical double layer - zeta potential


At the surface of any phase there is always a separation of positive and negative
charges, and due to polarization effects and adsorption of H+ and OH- ions a pH-
dependent surface potential is formed on the surface of a particle in water. The
charges on the surface attract opposite charges so that the potential decreases when
the distance from the surface increases. In this way so-called electrical double layer is
formed at the solid-liquid interface. Since the electrical neutrality of the system must
be maintained, the net charge on one side of the interface is balanced by equal net
charge of opposite charge on the other side of the interface.

The first model for the behaviour of the charges was outlined by Helmholtz (1879)
and Perrin (1904). They proposed that the electronic charges of a metal were balanced
by counter charges in the electrolyte as if they were forming a charged parallel plate
condenser along the surface (Fig. 4). This simple model is applicable only to metal-
electrolyte system of high salt concentration.

+ -
CHARGED SURFACE + - LIQUID
+ -
+ -
+ -
+ -

Figure 4. Helmholtz model of the electrical double layer.

The model of the electrical double layer was developed independently by Gouy
(1910) and Chapman (1913). They suggested that the charges form a diffuse
continuum of ions (Fig. 5). They considered a uniformly charged flat surface and the
ions were assumed to be point charges distributed according to the Boltzmann
distribution. The concentration of the counterions and the potential fall rapidly at first
close to the surface and then gradually with the increasing distance. The dielectric
constant of the solvent was assumed to have the same value throughout the diffuse
part.

CHARGED LIQUID
SURFACE P
O
T
E
N
T
I
A
L

0 DISTANCE FROM SURFACE

a) b)
Figure 5. Gouy-Chapman model of the electrical double layer.
12

This model was useful for low charge densities only and for distances not too close to
the surface because it neglected the ionic diameters of the charges and treated them as
point charges.

Stern (1924) developed further the model of the electrical double layer. He replaced
point charges with ions of finite size, which can approach the surface at a minimum
distance. He also combined the Gouy-Chapman model of the diffuse layer with the
Helmholtz model of the compact part of the layer close to the surface so that in the
Stern model two layers are in series. If these layers are regarded as capacitors, the
capacity of the double layer C is related to the other two by
1 1 1
= + (3.15)
C CH CG- C

where CH is the capacitance of the Helmholtz layer and CG-C is that of the Gouy-
Chapman layer. Stern also introduced the concept of specific adsorption of ions
within the Helmholtz portion of the double layer. The plane of these specifically
adsorbed unhydrated ions is often called the inner Helmholtz plane (IHP). Specifically
adsorbed ions are typically unhydrated anions. More weakly adsorbed hydrated ions
form the outer Helmholtz layer (OHP). These are usually hydrated cations. The Stern
model is represented schematically in figure 6.

When a particle moves in a liquid it is cut along an imaginary surface that is called the
surface of shear. Within this surface the fluid is in a stationary state. Thus the surface
of shear forms a sheath that envelops the particle, and it moves along a certain
quantity of the surrounding liquid with it and its contained charge. The potential on
this surface is called electrokinetic potential or zeta potential, shortly z-potential.
According to the generally accepted assumption, the shear plane is located at the outer
edge of the inner part of the double layer, i.e. at the outer Helmholtz plane. Some
workers do not accept the distinction to Helmholtz layers and call them just the Stern
layer. Then the shear plane should be placed some distance into the diffuse layer.

IHP OHP

P
O
T Diffuse layer
E
N
T
I
A
L

Figure 6. Schematic representation of modified Stern model, IHP = Inner Helmholtz


plane, location for unhydrated specifically adsorbed ions; OHP = Outer Helmholtz
layer, location for hydrated adsorbed ions.

Zeta potential is an important tool because it can be measured by various methods, for
example, by applying the electrophoretic mobility of particles. The data can be used
in many fields of colloid science and surface chemistry, for instance, in studying the
adsorption of ions and the stability of colloids in flotation.
13

Crystal lattice ions are called potential determining ions to distinguish them from
ions, which are not expected to have a special interaction with the surface. Such ions
are called indifferent ions. Between these categories are specifically adsorbed ions
that interact with some special way with the surface. Specifically adsorbed ions can be
recognized by their ability to change the sign of z-potential (Fig. 7). Indifferent ions,
which do not have a special interaction with the surface, are said to be adsorbed
physically and can only reduce the value of z-potential asymptotically to zero.

IHP OHP

P
O
T
E 0
N
T
I
A
L

Figure 7. Charge reversal due to specific adsorption of counterions.

The structure of the electrical double layer depends on the characteristics of the ions
and their concentration. In general, in the presence of an electrolyte electrical effects
have shorter ranges than in its absence and, as a consequence, the electrical double
layer is compressed. If the concentration of specifically adsorbed ions is high, the
diffuse layer may disappear. With the increasing concentration of indifferent ions z-
potential is reduced approaching zero and the diffuse layer becomes thinner. The
point (or pH) where z-potential is equal to zero is called isoelectric point (IEP). At
this point the concentration of indifferent ions does not have an effect on the value of
z-potential. Potential determining ions and chemically adsorbing ions affect also the
surface potential. The point (or pH) where surface potential is equal to zero is called
the point of zero charge (PZC). When it is changed also isoelectric point changes.

The thickness of the diffuse part of the electrical double layer can be evaluated
mathematically using so-called Debye length (1/k). It is the distance from the surface
into the solution within which the effective electrical interaction with the surface is
considered to occur. In the mathematical form the effective thickness is
1/ 2
Ê ˆ
1 Á eRT
= ˜ (3.16)
k Á 4pF 2 Â Ci Zi2 ˜
Ë i ¯

where e is the dielectric constant of the solution, R is the gas constant, T is the
absolute temperature, F is the Faraday constant, and Ci is the molar concentration of
any ion with valence Zi in the solution. As can be seen, the effective thickness is
inversely proportional to the valence of the ions and to the square root of their
14

concentrations. It is also directly proportional to the square roots of the absolute


temperature and the dielectric constant of the medium.

4. Flotation reagents
Collectors, frothers and regulating agents are normally used in flotation processes for
good recovery and selectivity. Typical chemicals, their properties and usage are
described in the following section. Other reagents, for example, dispersants,
flocculants and emulsifiers are also used in many processes but they are considered
here only if they are an essential part of flotation, for example, with collector addition.

4.1. Frothers
Frothers are surface-active chemicals and added in froth flotation to form froth bed on
the pulp so that floatable particles can be removed from the surface. Some collectors,
however, such as amines and sulphonates, are also surface active and actual frothers
may not be needed with them. Thiol collectors, which are used in sulphide flotation,
are not surface active enough and frothers are used in their flotation. Thus, actual
frothers are nonionic chemicals that do not act as collectors.

In the early days of flotation, pine oil and cresylic acid were used as frother. More
recently, typical frothers used in flotation are aliphatic alcohols (MIBC, 2-ethyl
hexanol), alkoxy-type materials (TEB), polyglycol ethers and polyethers. Frothers
decrease also the size of air bubbles generated improving the efficiency of flotation.
Generally accepted guidelines for frother selection are:
 After the froth product is removed, it must break readily for further treatment.
 The stability of the froth must be such that separation of the valuable minerals
from the gangue minerals is obtained in the froth.
 Sufficient volume and stability of froth is formed at low concentrations.
 The frother must possess a low sensitivity to changes in pH and dissolved salt
concentrations.
 Frother must not have collecting power for gangue minerals.
 Frother is readily dispersible.
 Frother is relatively cheap and easily available.

The action of bubbles and froth bed is studied widely. Some essential features can be
presented together by saying that the entrainment of hydrophilic particles decreases in
thick froth bed due to the drainage with water. Froth and bubbles are obviously
important factors in developing flotation in future.

4.2. Collectors
Most particles are hydrophilic and the surfaces are made hydrophobic for flotation by
using collectors. They are usually heteropolar, organic compounds that have an ionic
end, which is adsorbed at the mineral surface, and an organic group, which gives the
hydrophobic surface to the mineral. Research work has shown that the best selectivity
in mineral separation is often achieved by using particular collectors for mineral
groups, for example, thiol collectors for sulphides and amines for silicates. Therefore,
it is important to know the properties of collectors already when a flotation procedure
is tested.
15

The characteristics of flotation depend on the chemical properties of the collector and
on its adsorption mechanism. These include micelle formation, Kraft point and the
dissociation of collector reagents. Organic compounds tend to form larger structures,
called micelles, when so-called critical micelle concentration (CMC) is exceeded.
Kraft point tells the temperature above which the micelles are formed. The
dissociation of the collector depends on the compound and pH, and it has a significant
effect on the adsorption of collectors. Collectors are usually classified according to
their ionic end to anionic and cationic or non-ionic groups. Properties and features of
typical collectors are described in the following sections.

4.2.1. Thiol (sulphydryl) collectors


It is natural to start from this group because froth flotation was developed in the
beginning of 1900-century, particularly, by the flotation of sulphide minerals with
xanthates, which are also known as dithiocarbonates. They are still the most common
collectors used in sulphide flotation. The first work was done on a trial-and-error basis
but soon detailed research was focused on sulphide collectors (Aplan, F.F. and
Chander, S., 1988). Another important collector group used for sulphides is
dithiophosphates. The structural formulae of these collectors are given below.

S R O S Na
_ _ _ P
R O C S Na
R O S
Xanthate or
dithiocarbonate Dithiophosphate
(R = hydrocarbon chain)

Xanthates and dithiophosphates are the most important thiol collectors and together
they account roughly for 70-80% of the collectors used in sulphide flotation. Now it is
known that their working mechanism is mainly electrochemical, being general for
thiol collectors. They form disulphides when oxidized, if the redox potential (or
mixed potential) is high enough for the formation. In the case of xanthates the
products are called dixanthogens. At lower potentials metal xanthates are formed.
Both surface compounds result usually in the flotation of particles. Dixanthogen is
also an important collector, which is manufactured industrially.

Xanthates are made from an appropriate alcohol, carbon disulphide and an alkali
metal hydroxide, and they were among the first water-soluble collectors. The length
of the hydrocarbon chains is relatively short being generally 2-5 carbon atoms. A long
chain means a strong bond between the collector molecule and a metal atom and,
therefore, long chains also have strong collecting power. Xanthates are popular
collectors, which have good stability at alkaline pH, but they decompose easily in
strongly acid solutions. Commercial xanthates contain also often inorganic impurities
but they can be purified by dissolving them in dry acetone and removing the
impurities by filtration.

Dithiophosphates may be prepared in an analogous manner compared with xanthates


and they are similar to the xanthates in many respects. However, their metal
compounds are more soluble than are the analogous metal xanthates.
Dithiophosphates are also stabile at low pH but decompose in strongly alkaline
16

solutions. These factors give the metallurgist possibility to choose a suitable collector
or to combine them. Other thiol collectors used in flotation of sulphides and metals
are, for example, thiocarbamates, mercaptans, thioureas and mercaptobenzothiazoles.

4.2.2. Fatty acids or carboxylic acids


These reagents represent the group of anionic collectors, which do not include thiol
collectors, probably, due to their electrochemical nature. The properties of the anionic
collectors are well described, for example, by S. K. Mishra (1988) and the related
mechanisms have been explained more recently by J. Laskowski (1999).

Fatty acids (Ralston, 1948; Mehrotra and Bohra, 1983) are well known long-chained
compounds, which form soaps with metals. The pKa values of the compounds used as
collectors are 4.7± 0.5. The collectors form droplets at acid solution that can have
negative or positive surface charge depending on the compound and pH. At alkaline
pH the collector molecules are dissociated into the solution and form soaps with
metals, particularly, when unsaturated bonds are present in the carbon chain. This
means that their collecting power is mainly physical at acid pH but changes more
chemical in alkaline solution. Physical adsorption, however, depends also on the
length of the hydrocarbon chain. In addition, carboxylates form micelles and can be
surface active enough so that frothers are not always needed with them.

The collectors are typically used in the flotation of salt type minerals and oxide
minerals. Usually, fine material is first separated by desliming. Then slurry is
efficiently conditioned with the collector at high pulp density for the adsorption. Fuel
oil is often used to disperse the collector and decrease its consumption. High
temperature in conditioning also improves the adsorption process. Because the
chemical adsorption of the collectors decrease in acid solution pH can be lowered
during cleaning stages. Fatty acids are also used in de-inking with calcium ions that
improve its adsorption on negative sites.

4.2.3. Sulphates and sulphonates


These reagents are anionic collectors that are strong electrolytes and in ionic form at
acid pH, in wide range compared with fatty acids, for example. They can form
insoluble compounds with metals, and their adsorption mechanism have been studied
in detail (Cases, J.M., 1970; Wakamatsu, T., and Fuerstenau, D.W., 1973).

At low concentrations sulphates and sulphonates are usually adsorbed as counter ions
on the particles by electrostatic interaction. When collector ions associate together at
higher concentration near the surface, by van der Waals forces, the phenomenon is
called hemi-micelle formation. In addition, chemical interaction with metals on the
particle surfaces may occur. The electrical double layer surrounding the particles,
which is dependent on pH, influences also these mechanisms. (Fuerstenau, M.C. and
Palmer, B.R., 1976)

The collectors are used in flotation of many minerals, which include silicates, oxides
and salt type minerals. Desliming is usually necessary in their flotation and neutral
oils or alcohols are used to increase the efficiency and to reduce the consumption.
Petroleum sulphonates, which are by-products in oil industry, have been found to be
selective collectors for some minerals at acid pH, particularly, if they contain iron.
17

4.2.4. Cationic collectors


Amine collectors are long-chained compounds and the only essential group in this
category (Smith, R.W., 1988). They are weak electrolytes and in ionic form at acid
pH. In neutral solution the amount of aqueous molecules increases. Amine collectors
precipitate in alkaline solution and form micelles and droplets, which have negative
surface charge at high pH (Laskowski, J., 1999).

The adsorption of the collectors is found to be physical and it depends on the form of
the compounds present in the solution. The floatability of particles is often highest at
pH 9-10 where the droplets contribute to the adsorption of amine collectors. Amine
compounds are also surface-active and frothers are not always needed with them.

Amine collectors are typically used in silicate flotation and slime separation is usually
required. Sometimes, however, when material losses are to be avoided, increasing the
collector dosage may do it. Using neutral oils and alcohols can also increase the
effectiveness of amine collectors. For the physical adsorption of amine collectors, the
conditioning time is usually short and the collector can often be added directly to the
cells. For the same reason, the selectivity of separation between silicate minerals is
often low.

4.2.5. Amphoteric and chelating agents


Amphoteric collectors have two oppositely charged functional groups in the molecule.
They are cationic at acid pH and anionic at alkaline pH, and between them, in the
isoelectric region, positive and negative charges may be equal. These collectors and
their chemistry are known for some time but, however, they have not gained wide
acceptance in industry (Smith, R.W., 1988).

Another group of reagents that has been tested as collectors are chelating agents.
Their chemistry is also well known and these compounds are tested for some minerals
but they may not be efficient collector reagents (Fuerstenau, D.W., Herrera-Urbina, R.
and Laskowski, J., 1988; Nagaraj, D.R., 1988). Research is still focused on collectors
but their effectiveness and selectivity can also be improved by using regulating
agents.

4.3. Regulating agents


The purpose of the regulating agents is to improve or prevent the action of collector
reagents and, therefore, they are called activators or depressants. A regulating agent
can be sometimes an activator but in another case a depressant. Some typical
regulators are considered in the following part.

4.3.1. Slurry pH
The amount of hydrogen and hydroxyl ions is controlled by pH-regulation of the pulp,
which is probably the most common control method in flotation. As well known, the
scale between pH-values is logarithmic. Consequently, relatively small ion amounts
have significant effects in neutral pH-region. In addition, pH is stabile in processes
due to the large volumes but it can vary easily during small-scale tests. This means
that the results may not be the same, for example, in a flotation test at neutral pH if an
acidic peak has taken place during the test. Therefore, careful pH-regulation is
important in flotation tests. There are also other factors in flotation that can be
18

different in processes compared with small scale testing, for example, mixing,
recycling, residence time, etc., but their effect is not considered here.

Lime, soda ash, caustic soda and sulphuric acid are common pH-modifiers but many
other chemicals are used. The reagents have always their characteristics, for example,
as dispersants, and the overall effect of pH-control is very complicated. The reagents
can have effect on collectors and on particles, on their solubility and on the surface
properties. In addition, the soluble ions and their compounds can behave as regulating
agents. As a consequence, similar ore types behave often in a similar way but it is to
be emphasized, not in the same way.

4.3.2. Sulphide minerals


Electrochemical reactions are important in sulphide flotation, in the adsorption of
thiol collectors and in the oxidation of the semiconducting minerals (Aplan, F.F. and
Chander, S., 1988; Rao, S.R., Moon, K.S. and Leja, J., 1976). Galvanic interactions
play also essential role in sulphide flotation. As the adsorption of thiol collectors take
mainly place by electrochemical mechanisms, their action depends on the redox
potential, or on the mixed potential, in the pulp. Redox potential or electrochemical
potential is often denoted by Eh. Thus, any factor that change redox potential can have
an effect on the adsorption of the collectors. It is well known that atmospheric oxygen
is usually present in flotation pulps and it takes part in electrochemical reactions. It is
consumed, for example, in the oxidation of grinding media making the environment
reducing, and its amount depends on slurry aeration. These are factors that may not be
called regulating agents but they are essential regulators in sulphide flotation.

Electrochemical potential can be measured with different electrodes but it is usually


given on the hydrogen scale. Various oxidizing and reducing agents may be added to
control it with pH in flotation. The chemicals include several compounds of sulphur,
in its different oxidation states (Chander, S., 1988). Such reagents, however, have
their chemical characteristics that affect the floatability of particles, in addition to the
electrochemical potential, and this makes the results complicated and difficult to be
interpreted.

Sulphur ion is a common regulator, which can be added as sodium sulphide. It is a


reducing agent and it is also used in the flotation of oxidized minerals for
sulphurization. The working mechanisms are well known. The ion can be either a
depressant or an activator, depending on its usage. As depressant it makes the pulp
reducing and its compounds compete with thiol collectors for adsorption. Various
sulphoxy compounds are common reagents used in sulphide separation. For example,
sulphur dioxide is an efficient depressant used in flotation of sulphide minerals when
they occur finely disseminated in pyrite rich, complex matrix (Bulatovic, S.M. and
Wyslouzil, D.M., 1985). Other sulphur containing regulating reagents include various
sulphites and acids but their working mechanisms are not known in detail.

Cyanide is an efficient depressant that has been traditionally used in sulphide flotation
but the environmental regulations have decreased its usage. It forms metal-cyanide
complexes with metals and ferro- and ferricyanide compounds have been concluded
to be effective depressants in sulphide flotation. The depressing mechanism of
cyanide is, however, a debatable subject. It has been used as a depressant mainly for
iron, zinc and copper sulphides.
19

Cyanide has no effect on lead sulphides and, therefore, other means have been
developed to depress, for example, galena, which is one of the most important lead
minerals in sulphide ores. Chromate (K2CrO4) and dichromate (Na2Cr2O7) ions are
frequently used as galena depressants. They form hydrophilic films on galena surfaces
(Shimoiizaka, J. et al., 1976). Dichromate has also been used to depress copper
minerals with cyanide. In addition, cyanide and chromate have been used as
alternatives to depress either copper or lead in the separation after bulk flotation.

Zinc salts are often used as depressants for zinc sulphides. Zinc ions form hydrophilic
compounds, which are adsorbed on the particles. As zinc is not a very effective
depressant alone, the salts are used together with other reagents, for example, in
combination with cyanide or sulphur compounds. In addition to reagent dosages, also
the stages where the reagents are added to processes are important to achieve the
optimal metallurgical results (Tveter, E.C. and McQuiston, Jr., F.W., 1962). High
selectivity can often be achieved in sulphide separation if lime and depressants are
added already to the grinding mill.

Calcium ion has also an effect on sulphide flotation. It is usually present in processes,
for example, due to lime additions. It can act as a depressant or an activator depending
on process variables but its working mechanisms are not fully understood
(Schreithofer, N. et al., 2000). Other sulphide depressants include dextrin, which have
been used efficiently to depress pentlandite and pyrrhotite in copper separation after
bulk flotation. In the case of molybdenite concentration, copper sulphides have been
depressed with Nokes reagent, which is a mixture of P2S5 and NaOH.

In addition to depression, also activation is important in sulphide flotation. Adding


metal ions, which change place by ion-exchange mechanism with less noble metal
ions on sulphide surfaces, usually carries it out. After this thiol collectors can easily
be adsorbed on the sites. It is well known, that sphalerite is not easily floated with
thiol collectors without activation. It is usually done by adding copper sulphate to the
pulp. Also other metals can activate sulphides. The efficiency of activation depends
also on pH. Unintentional activation can also be detrimental for flotation. This can
take place, for instance, if the ore contains soluble copper minerals, which can
activate sulphides causing unselectivity in separation. In addition to the metal
activation, also sulphurization of oxidized minerals is an applied method in sulphide
flotation.

4.3.3. Silicate minerals


Silicates are non-conductors and only slightly soluble in water. Silicate flotation is
usually carried out with cationic collectors. The collector molecules are adsorbed
physically on the minerals and the adsorption depends on the potential (zeta potential)
on the shear plane between the particles and liquid. Therefore, electrical double layer
is in essential role in silicate flotation. These are different characteristics compared
with sulphide minerals, and attention should be paid accordingly. In addition to pH,
various ions and ionic compounds can be used to control zeta potential and as
regulating agents.

Characteristics of silicate flotation have been considered, for example, by R.M.


Manser (1975). Regulating agents include both inorganic and organic compounds.
20

Particularly, the effect of fluoride (silicofluoride, hydrofluoric acid, etc.) has been
described in detail. The usage of hydrofluoric acid (HF) as quartz depressant is a
known application in the flotation of pegmatite ores. Other common modifiers in
silicate flotation are sodium silicate or water glass, polyphosphates (Calgon), tannin
and starch products.

Sodium silicate is an efficient dispersant and depressant which can be used for
dispersing fines and slimes. Polyphosphates react also with metallic cations and can
remove them from particle surfaces, preventing this way unintentional activation.
Organic compounds form hydrophilic surfaces on particles or give them steric
hindrances against aggregation. In addition, multivalent cations, especially calcium
and ferric ions activate silicate minerals with anionic collectors.

4.3.4. Oxide and salt type minerals


In the flotation of salt type minerals the concentration of soluble ions and their
compounds is high in the pulp. Their adsorption depends on pH and modifies the
surface properties of particles. Anionic collectors are frequently used in the
processing of these ores (Tanaka, Y. et al., 1988). Because cations, or their surface-
active hydroxides, activate silicate minerals with the collectors, selective separation of
minerals is often difficult. Therefore, efficient depressants are needed in the
separation. Starch and tannin are common modifiers used with salt type minerals, and
also polyvalent metal salts. Other regulating agents are similar to those used with
silicate minerals, including water glass, soda ash, ammonia, etc. Although evident, it
is to be emphasized that flotation of the ores is often very complex due to many ions,
which can be adsorbed on particle surfaces making the properties similar.

Flotation of oxide minerals can often be carried out using anionic collectors, for
example, with fatty acids. A prerequisite can be that first other minerals, such as
sulphides, are to be separated because they would be floatable with the collector.
Many other physical methods, for instance, magnetic separation and gravity
separation, can be combined with flotation (Yang, D.C., 1988). Regulating agents in
the flotation are usually relatively simple, including pH-regulators (lime, soda ash,
caustic soda), dispersants (starch, dextrin) and depressants (Quebracho, tannic acid).

5. Effect of fineness on flotation


Flotation is usually connected to mineral processing where it started more than one
hundred years ago. The increasing demand to treat finer materials has strongly
determined the development of flotation technology. With minerals the particle size is
usually between 10-500 mm, however, depending on the ore type. For example,
sulphides are usually floated in finer size (10-200 mm) than silicates, which are lighter
minerals. When the effect of particle size on flotation is investigated, rate constants
are considered in various sub-processes; in collisions of bubbles and particles, in
adhesion of particles on bubbles and in detachment of particles from bubbles.

With increasing fineness flotation becomes more difficult, and too fine grinding is
usually avoided. However, the mineralogical characteristics of ores are often such that
mineral liberation requires fine particle size and high fineness may not be avoided.
Difficulties in flotation of fines are in principle caused by two basic factors, which
affect flotation in various ways. The mass of particles becomes negligible and the
21

specific surface increases. In addition, particles tend to become reactive, particularly


after grinding.
The experience has shown that the floatability of particles is often best in the
intermediate size. For sulphides this size range is about 10-100 mm. In this size range
particles usually float fast and good recovery and high selectivity can be achieved. In
coarser fractions flotation of particles is normally slower and the recovery remains
low compared with the intermediate size. Gravity affects more strongly coarse
particles than fine ones, causing detachment of particles from air bubbles. Therefore,
they require highly hydrophobic surfaces for high recovery. Thus, anything that
decreases the hydrophobicity of particles, for example, too low collector dosage, can
usually first be detected as a low recovery of coarse particles.

Flotation rate of particles decreases also in fine size. This is explained so that the
probability of collision between particles and air bubbles decreases. Due to their
negligible mass, particles start to behave as a part of the liquid phase, flowing with
water independent of their hydrophobicity. In this way, small particles can be
recovered mechanically in thin water layers by so-called entrainment mechanism.
Fine particles can also be entrapped between floatable particles. When fine gangue
minerals are in question mechanical recovery deteriorates the selectivity of flotation.

High fineness increases the specific surface of the material and the need of reagents.
Fine fractions have a high adsorption capacity and can consume a significant part of a
reagent so that there is not enough left it for coarser particles. Surface properties and
electrochemical characteristics can be different for fine particles. They contain edges,
corners and various imperfections, which increase the surface energy of particles
making them reactive. This can affect flotation. Fine particles are oxidized and
dissolved, increasing the concentration of ions in the pulp. Ions and their compounds
may affect the interaction between collectors and minerals. They can enhance or
prevent the adsorption of collector molecules. In sulphide flotation the adsorption of
dissolved cations can cause unintentional activation of sulphide minerals and
nonselective flotation.

Fine particles can also be attached on coarse particles as slime coatings. These
coatings can be detrimental to flotation. The grade of the concentrate may decrease if
the fine particles are valuable minerals and they are attached on gangue particles.
When the slime coating consists of gangue particles, they can prevent the attachment
of air bubbles and cause low recoveries.

The detrimental effects of fineness can be restrained in different ways. Desliming is


obviously the most common of them. The finest part of particles is separated, usually
by classification, before flotation. This is a typical process, particularly, when
minerals are floated with a physically adsorbed collector. Desliming, however, results
also in losses of valuable minerals, and it makes the process less economical. In some
cases, it is possible to avoid this by increasing the collector dosage.

Slime separation is often a prerequisite for flotation also when chemically adsorbed
collectors are used. The efficiency of chemisorption can be increased at elevated
temperatures. It is possible that at a low concentration collector is chemisorbed on
valuable mineral only and selective flotation can be achieved. It has been
recommended that the collectors should be introduced as early as possible, for
22

example, in grinding. This way it is also possible to avoid unselective collector


adsorption due to oxidation of particles.
As the probability of collisions between particles and bubbles decreases with particle
size, flotation technology has been developed taking into account this point of view.
According to a known principle, flotation of fine particles requires lot of small air
bubbles in a relatively quiescent environment, so that enough time is given for bubble
attachment on particles. Fine bubbles have been produced by electrolysis in
electroflotation and by dissolved air in vacuum flotation. In the latter, the air bubbles
are expected to nucleate preferentially on the floatable particles, so that the collision
step between bubbles and particles could be eliminated. Vacuum flotation has been
applied in water purification processes, but for minerals the mentioned methods have
not been used.

The flotation of fine particles is improved in column flotation. The flotation is carried
out in high and circular columns, which do not contain the mechanical rotor-stator
system used in conventional cells for pulp agitation and introducing the air into the
suspension. In column flotation the principle of counter current flow is applied. Air
bubbles are blown from the bottom of the column and rise continuously through the
downward flowing slurry. The feed is injected about one third of the height of the
column from the top. In this way advantageous circumstances have been created for
the bubbles to be attached on the particles.

In addition, entrapped gangue particles are efficiently washed away from the
mineralized bubbles in the flow. The froth bed is thick (>50cm) in column flotation
compared with conventional flotation, and it is washed continuously. Washing water
is spayed in a suitable manner on the froth and, when flowing through the froth bed, it
removes the hydrophilic particles back to the pulp. Washing is also used with
conventional cells in cleaning or recleaning stage when fine materials are processed.

The floatability of fine particles may be improved if the floatable particles form larger
aggregates. This can also increase the flotation rate of fine particles. If aggregation is
achieved by using some inorganic reagent or acid or alkali, particles tend to form
compact aggregates, and it is called coagulation. If organic compounds hold particles
together, by polymers, they usually form open and loose aggregates, and the process
is called flocculation. In selective flocculation the purpose is to flocculate a particular
mineral whereas other minerals remain in dispersion state. Aggregation depends on
added reagents and surface forces. These are considered more in the following
section.

When colloidal particles need to be separated, so-called carrier flotation can be used.
In this application, the colloid particles are attached on some coarse mineral, a carrier
mineral, which can then be floated with a suitable collector. Carrier flotation requires
coagulation of colloids onto the large particles, which can also be added to the pulp.
With this method the purpose is rather colloid separation than selective flotation.

5.1. Stability of colloids


In processing of fines separation is affected by the state of particles in the suspension.
In flotation dispersion of particles is usually advantageous because it improves the
selectivity of separation. In the interaction of fine particles due to their negligible
mass, surface forces become significant and they can easily be aggregated. This can
23

cause difficulties in flotation. Sometimes, on the other hand, the aim is to agglomerate
fine particles by selective flocculation to improve their floatability.

In pulp particles interact by forces of different origins. These forces can be attractive
or repulsive in nature. In the interaction of macroscopic bodies in water dispersion
force is the only significant term of the van der Waals forces. Between similar
particles it is always attractive. Between dissimilar particles it can be either attractive
or repulsive.

Most particles are electrically charged in water. The charge depends on the
dissolution of the particle and hydrolysis of the surface species, and subsequent pH
dependent dissociation of hydroxyl groups. Adsorption of ions and their complexes
from the bulk can affect the charge. In this way the electrical double layer is formed.
The charge in the double layer can be either negative or positive and, consequently,
the forces between particles due to the charges can be attractive or repulsive.

The well-known DLVO-theory is based on the above mentioned forces. The theory
was developed independently by two groups of scientists (Dragoon and Landau, 1941;
Verwey and Overbeek, 1948) to explain the stabilization of colloidal dispersions on
the basis of repulsion caused by overlapping of the double layers surrounding the
dispersed particles.

More recent studies have shown that when particles come very close to each other,
say, at a few nanometers separation, forces due to van der Waals and double layer
interaction often fail to explain their behaviour. Thus, it is concluded that other type
of interactions or forces can also occur. The additional forces can be attractive or
repulsive or oscillatory in nature, and they can be much stronger than the two DLVO
forces at small separations.

The non-DLVO forces can partly be explained by the "rigid" interaction between
atoms, molecules and particles. Water molecules between particles within highly
restricted space can create oscillatory forces. The forces are also known as solvatation
or hydration forces, or structural forces. Hydration forces depend also on the physical
and chemical properties of the surfaces, whether they are hydrophilic or hydrophobic,
smooth or rough, etc. In addition, hydrophobic surfaces cause long-range attractive
forces between the surfaces. The origin of hydrophobic forces is not fully understood
at present. Other non-DLVO interactions may arise from the disruption of the water
hydrogen-bonding network between surfaces, from electrostatic ion-binding and ion-
correlation effects, and from molecular "bridging" effects.

Steric forces arise from the overlap of adsorbed layers of large molecules, especially
polymers. These forces can be either repulsive or attractive depending whether the
outermost layer prefer to be in contact with water or not. If the solvent power of water
for the adsorbed layers is high enough, the interaction between layers will be
repulsive, resulting in dispersion of particles. If the solvent power of water for the
layers is minimal, there is a tendency of the adsorbed molecules on one particle to
penetrate into the layer on another particle promoting aggregation of particles.

In polymer bridging a long chained adsorbate induces aggregation by attaching to two


or more particles. Polymers can provide such bridging between particles particularly
24

under conditions where particles are not totally coated by the polymeric species.
Adsorption of polymers and the effectiveness of the adsorbed polymer in promoting
flocculation are dependent on polymer properties such as polymer weight, nature and
concentration of functional groups and configuration. Flocculation depends also on
mineral properties and solution properties, such as pH, ionic strength and solvent
power for the polymer, as well as on solid concentration and aging of the polymer
solution.

Acknowledgements
The views and discussions are gratefully acknowledged.

LITERATURE

Adamson, A.W. and Gast, A.P., 1997. Physical Chemistry of Surfaces, 6th edition, A
Wiley-Interscience Publication, John Wiley & Sons, New York.
Aplan, F.F. and Chander, S., 1988. Collectors for Sulfide Mineral Flotation, In
Reagents in Mineral Technology, P. Somasundaran and Brij M. Moudgil, eds.,
Marcel Dekker, Inc., pp. 335-369.
Bulatovic, S.M. and Wyslouzil, D.M., 1985. Selection of Reagent Scheme to Treat
Massive Sulphide Ores, In Complex Sulphides, Processing of Ores, Concentrates
and By-products. A.D., Zunkel, R.S., Boorman, A.E., Morris and R.J., Wesely,
eds., The Metallurgical Society, Inc., Warrendale, PA, USA, pp. 101-137.
Carta, M., Ciccu, R., DelFá, C., Ferrara, G., Ghiani, M. & Massaci, P., 1974.
Improvement in Electric Separation and Flotation by Modification of Energy
Levels in Surface Layers, Proceedings of the 10th International Mineral Processing
Gongress, London, M. J. Jones, ed., IMM, London, pp. 349-376.
Cases, J.M., 1970. On the Normal Interaction Between Adsorbed Species and
Adsorbing Surface. Trans AIME 247, pp. 123-127.
Chander, S., 1988. Inorganic Depressants for Sulfide Minerals, In Reagents in
Mineral Technology, P. Somasundaran and Brij M. Moudgil, eds., Marcel Dekker,
Inc., pp.429-469.
Chaplin, M.F., 2000. A Proposal for the Structuring of Water, Biophys. Chem. 83,
pp.211-221.
Chapman, D.L., 1913. A Contribution to the Theory of Electrocapillarity, Philos.
Mag. 25, pp. 475-481.
Derjaguin, B.V., 1937. The Theory of Interaction of Particles Carrying Double
Electrical Layers and of the Aggregative Stability of Hydrophobic Colloids and
Disperse Systems, Bull. Acad. Sci. URSS, Classe Sci. Math. Nat. Ser. Chim., 5,
pp. 1153-1164.
Derjaugin, B.V. and Landau, L.D., 1941. Theory of the Stability of Strongly Charged
Lyophobic Sols and of the Adhesion of Strongly Charged Particles in Solutions of
Electrolytes, Acta Physicochim. URSS14, pp. 633-662.
Frumkin, N.A., 1938. On the Phenomena of Wetting and the Adhesion Bubbles, 1,
Zh. Fiz. Khim. 12, pp. 337-345.
Fuerstenau, D.W., 1980. Fine Particle Flotation, In Fine Particles Processing, Vol. 1,
P. Somasundaran, ed., The American Institute of Mining, Metallurgical, and
Petroleum Engineers, Inc., New York 1980.
25

Fuerstenau, D.W., Herrera-Urbina, R. and Laskowski, J., 1988. Surface Properties and
Flotation Behavior of Chrysocolla in the Presence of Potassium Octyl-
hydroxamate, In Froth Flotation, S.H., Castro, and J. Alvarez, eds., Elsevier
Science Puplishers B.V., Amsterdam, pp. 245-260.
Fuerstenau, M.C. and Palmer, B.R., 1976. Anionic Flotation of Oxides and Silicates,
In Flotation, A.M. Gaudin Memorial Volume, Vol 1, M.C., Fuerstenau, ed., AIME,
pp. 148-196.
Glembotsky, V.A., Klassen, V.I., and Plaksin, I.N., 1972. Flotation, Translated by
R.E. Hammond, edited by H.S. Rabinovich, Primary Sources, New York.
Gouy, G., 1910. Constitution of the Electric Charge at the Surface of an Electrolyte, J.
Phys. 9(4), pp. 457-467.
Gibbs, J.W., 1931. Collected Works of J. Willard Gibbs, Longmans, Green & Co.
New York.
Helmholtz, H.L.F., 1879. Wiss. Abh. Phys. Tech. Reichsanst. 1, p. 925.
Hunter, R.J., 1981. Zeta Potential in Colloid Science, Academic Press, London.
Hunter, R.J., 1993. Foundations of Colloid Science, Volume I and II, Clarendon
Press, Oxford.
Isaacs, E.D., Shukla, A., Platzman, P.M., Hamann, D.R., Barbiellini, B. and Tulk,
C.A., 2000. Compton Scattering Evidence for Covalency of the Hydrogen Bond in
Ice, J. Chem. Phys. Solids 61, p. 261.
Israelachvili, J., 1992. Intermolecular & Surface Forces, 2nd edition, Academic
Press., London.
King, R.P., (Ed), 1982. Principles of Flotation, South African Institute of Mining and
Metallurgy, Monograph series No. 3, Johannesburg.
Klassen, V.I. and Mokrousov, V.A., 1963. An Introduction to the Theory of Flotation,
Translated by J. Leja and G.W. Poling, Butterworths, London.
Laskowski, J.S., 1999. Weak Electrolyte Collectors, In Advances in Flotation
Technology, B.K., Parekh, and J.D., Miller, J.D., eds., Society for Mining,
Metallurgy, and Exploration, Inc., pp. 59-82.
Leja, J., 1982. Surface Chemistry of Froth Flotation, Plenum Press, New York.
Lin, I.J., Nadiv, S. and Grodzian, D.J. M., 1975. Changes in the State of Solids and
Mechano-Chemical Reactions in Prolonged Comminution Processes, Minerals Sci.
Engng. vol 7, no. 4. Oct., 1975, pp. 314-336.
Manser, R.M., 1975. Handbook of Silicate Flotation, Warren Spring Laboratory.
Stevenage, England.
Mehrotra, R.C. and Bohra, R., 1983. Metal Carboxylates, Academic press, Inc.,
London.
Mishra, S.K., 1988. Anionic Collectors in Nonsulfide Mineral Flotation, In Reagents
in Mineral Technology, P. Somasundaran and Brij M. Moudgil, eds., Marcel
Dekker, Inc., pp. 195-217.
Nagaraj, D.R., 1988. The Chemistry and Application of Chelating or Complexing
Agents in Minerals Separation, In Reagents in Mineral Technology, P.
Somasundaran and Brij M. Moudgil, eds., Marcel Dekker, Inc., pp. 257-334.
Pauling, L., 1948. The Nature of the Chemical Bond, 2nd edition, Cornell University
Press, New York.
Perrin, J., 1904. Mécanism de l'éctrisation de contact et solutions colloidales, J. Chim.
Phys. 2, pp. 601-651.
Ralston, A.W., 1948. Fatty Acids and their Derivatives, Wiley, New York.
26

Rao, S.R., Moon, K.S. and Leja, J., 1976. Effect of Grinding Media on the Surface
Reactions and Flotation of Heavy Metal Sulphides, In Flotation, A.M. Gaudin
Memorial Volume, Vol. 1, M.C., Fuerstenau, ed., AIME, p. 509.
Rosen, M.J., 1978. Surfactants and Interfacial Phenomena, A Wiley-Interscience
Publication, John Wiley & Sons, New York, pp. 174-199.
Schreithofer, N., Kirjavainen. V., Heiskanen. K. and Ahveninen, R., 2000. Effect of
Ca2+ and S2O32- on the Flotation of Pentlandite in Real Ore System, In Proceedings
of the XXI International Mineral Processing Congress, P., Massacci, ed., Vol B,
Elsevier, Amsterdam, p. B8b-40.
Smith, R.W., 1988. Cationic and Amphoteric Collectors, In Reagents in Mineral
Technology, P. Somasundaran and Brij M. Moudgil, eds., Marcel Dekker, Inc.,
pp.219-256.
Wakamatsu, T., and Fuerstenau, D.W., 1973. Effect of Alkyl Sulfonates on the
Wettability of Alumina, Trans AIME 254, pp. 123-126.
Stern, O., 1924. The Theory of the Electrolytic Double Layer, Z. Elektrochem. 30,
pp.508-516.
Suresh, S.J. and Naik, V.M. 2000. Hydrogen Bond Thermodynamic Properties of
Water from Dielectric Constant Data, J. Chem. Phys. 113, pp. 9727-9732.
Tanaka, Y., Katayama, N. and Arai, S., 1988. Reagents in Phosphate Flotation, In
Reagents in Mineral Technology, P. Somasundaran and Brij M. Moudgil, eds.,
Marcel Dekker, Inc., pp. 645-682.
Tveter, E.C. and McQuiston, Jr., F.W., 1962. Plant Practice in Sulfide Mineral
Flotation, In Froth Flotation-50th Aniversary Volume, D.W., Fuerstenau, ed.,
AIME, New York, pp. 382-426.
Verwey, E.J.V. and Overbeek, J.Th.G., 1948. Theory of the Stability of Lyophopic
Colloids, Elsevier, Amsterdam.
Wakamatsu, T., and Fuerstenau, D.W., 1973. Effect of Alkyl Sulfonates on the
Wettability of Alumina, Trans AIME 254, pp. 123-126.
Yang, D.C., 1988. Reagents in Iron Ore Processing, In Reagents in Mineral
Technology, P. Somasundaran and Brij M. Moudgil, eds., Marcel Dekker, Inc.,
pp. 579-644.

Potrebbero piacerti anche