Anda di halaman 1dari 12

Minerals Engineering 66–68 (2014) 13–24

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

The effect of saline water on mineral flotation – A critical review


Bo Wang ⇑, Yongjun Peng ⇑
School of Chemical Engineering, The University of Queensland, St. Lucia, Brisbane, QLD 4072, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Due to scarcity of fresh water and stringent regulations on the quality of discharged water, more and
Received 20 December 2013 more flotation plants have to use groundwater, sea water or recycle water with a high concentration
Revised 7 April 2014 of electrolytes. Although a number of studies have been conducted to investigate the effect of saline
Accepted 12 April 2014
water (or salt solutions) on mineral flotation, effective ways to solve the problems encountered in min-
Available online 6 May 2014
eral flotation plants using saline water are currently not available. This paper presents a review of pub-
lished articles addressing the effect of saline water on the interfacial phenomena taking place in the
Keywords:
flotation process, such as surface wettability, bubble-particle collision and attachment, mineral particle
Saline water
Hydration layer
interactions and frothing. This review provides an overall picture of the current status of studies in this
DLVO theory area and pinpoints directions of future research to address different problems associated with using
Bubble coalescence saline water in mineral flotation.
Particle interaction Ó 2014 Elsevier Ltd. All rights reserved.
Froth stability

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.1. Principles of mineral flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2. Flotation practice using saline water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2.1. Base metal sulphides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2.2. Coal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2.3. Potash . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2. Water structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1. Pure water structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2. Saline water structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.1. Water structure model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.2. Effect of ions on water structure in flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3. Effect of saline water on particle properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1. Hydration layers on particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2. Electrical double layers around particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.1. DLVO theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.2. Wetting film between particles and bubbles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.3. Particle aggregation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4. Effect of saline water on bubble properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1. Inhibition of bubble coalescence in salt solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2. Froth stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

⇑ Corresponding authors. Tel.: +61 7 3365 7156; fax: +61 7 3365 3888.
E-mail addresses: bo.wang@uq.edu.au (B. Wang), yongjun.peng@uq.edu.au (Y. Peng).

http://dx.doi.org/10.1016/j.mineng.2014.04.017
0892-6875/Ó 2014 Elsevier Ltd. All rights reserved.
14 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

Notation

A Hamaker constant, erg C bubble coalescence rate, m3 s1


A131 the complex Hamaker constant between particles 1 sep- cs/A the surface energies between solid-air
arated by medium 3 cs/W the surface energies between solid-water
a the radius of spherical particles cW/A the surface energies between water–air
B retarded van der Waals coefficient, ergcm wd the surface potential,
C the hydrophobic interaction constant j the reciprocal Debye
c concentration of surfactant, mol/L q density, kg/m3
ct transition salt concentration, mol/L r surface tension, kg/s2
D0 the decay length s bubble contact time, s
g gravitational constant hB buoyancy driven collision rate, m3 s1
H the distance separating the two particles hLS collision rate due to laminar shear, m3 s1
h film thickness between coalescing bubbles, m hT collision rate due to turbulence, m3 s1
R radial coordinate of bubble column, m
Rd radius of contact area between bubbles Subscripts
rb rubble radius, m b bubble
Rg gas constant f final
t coalescence time, s g gas
T temperature i, j particle i, j
l liquid
Greek letters
e energy dissipation rate per mass, m2/s2

1. Introduction by a bubble it must first collide with the bubble. The collision effi-
ciency (Ec) quantifies the process of collision, and is totally con-
1.1. Principles of mineral flotation trolled by hydrodynamics in the flotation cell. The particle size
and density, bubble size and velocity all affect the value of the col-
Mineral flotation which enables the selective separation of fine lision efficiency. Following collision, only hydrophobic particles
particulate valuable minerals from waste or gangue minerals tend to attach to the bubbles. The attachment efficiency (Ea) quan-
exploits the difference in surface wettability of valuable and gan- tifies the process of attachment and is controlled by pulp chemis-
gue minerals and is recognised as an effective method for the con- try and mineral surface chemistry. Collector adsorption, activation
centration of valuable minerals. In the practice of mineral flotation, and oxidation may produce a mineral surface that is coated with
a pulp of solid particles in water is conditioned with a small hydrophobic regions, facilitating the attachment of the particle
amount of reagents with air being purged into the pulp to form to the bubble. The detachment efficiency (Ed) quantifies the pro-
bubbles that can collide with particles. Collector is normally added cess of detachment. For the bubble-particle aggregate to be trans-
to render one mineral (usually the valuable mineral) selectively ported to the pulp/froth interface at the top of the pulp, it must be
hydrophobic, while frother is added to help the formation of small stable against detachment by the shear forces operated in the
bubbles in the pulp and a froth layer on the top. flotation cell. The bubble-particle detachment efficiency is also
It is established that the flotation rate constant, k, and the bub- controlled by the hydrodynamics in the flotation cell and surface
ble surface area flux, Sb, follow a linear relationship (Grano, 2006): hydrophobicity. The sub-processes of collision, attachment and
detachment together control the intrinsic collection efficiency
1 (Dai et al., 2000):
k¼ Sb Ecoll ð1Þ
4
where collection efficiency, Ecoll, refers to the intrinsic efficiency of Ecoll ¼ Ec  Ea  ð1  Ed Þ ð2Þ
particles captured by individual bubbles. This term embraces col-
lision, attachment and detachment as sub-processes in the overall Once the particles move from the pulp to the froth, the froth
collection process as shown in Fig. 1. For a particle to be collected provides a second chance for selective separation of minerals by
allowing the drainage of hydrophilic gangue minerals back into
the pulp. A metastable froth is a prerequisite for the successful
transport of mineral laden bubbles from the pulp/froth interface
over the lip of the flotation cell and to the concentrate launders.
It is desirable that the mineralised froth breaks down (i.e., the bub-
bles collapse) once it reports to the concentrate launder. The over-
all flotation efficiency is governed by the particle recovery in both
pulp and froth phases.
Dissolved ions in solutions may alter the water structure, parti-
cle surface wettability and colloidal interactions between bubbles
and particles and therefore have a positive or negative effect on
mineral flotation. Before reviewing the past and recently published
articles related to the impact of saline water on interfacial pro-
Fig. 1. Collision, attachment and detachment processes controlling the collection cesses in mineral flotation, flotation plants using saline water is
efficiency in mineral flotation. summarised.
B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24 15

Table 1
Some of the base metal sulphide flotation plants using saline water (Drelich and Miller, 2012).

Project Company Country Water source


Batu Hijau Newmont Indonesia Sea and fresh water
Las Luces Minera Las Cenizas Chile Sea water
Michilla Antofagasta Chile Sea water
KCGM Barrack/Newmont Australia Saline
Mt Keith BHP Billiton Australia Saline
Raglan Xstrata Canada Saline
Texada Closed Canada Sea water
Tocopilla Closed Chile Sea water
Esperanza-under development Antofagasta Chile Sea water

1.2. Flotation practice using saline water any pre-treatment. Sea water is pumped 145 km from the Pacific
Ocean to a 60,000 m3 pool at the mine site located at an altitude
In an effort to minimise fresh water use, many mine sites have of 2300 m above sea level (Castro, 2012).
introduced water re-use, underground water or seawater as a con- Other important cases are the Batu Hijau Concentrator (New-
ventional practice. One of the consequences is a concomitant mont, operating from 2000) located at the Indonesian island of
increase in the salinity of water on the sites and subsequently in Sumbawa, and Çayeli Bakır Is _ßletmeleri A.S
ß . (CBI) in Turkey. The
flotation. The major commodities that are using saline water in flo- Batu Hijau Concentrator uses sea water for processing a gold-rich
tation are summarised below. phorphyry copper ore (chalcopyrite-bornite) (Castro, 2012), while
CBI processes a complex Cu–Zn sulphide ore using dissolved metal
ions and sulphide ions, mainly in the form of SO2 2
4 and S2O3 (Bıçak
1.2.1. Base metal sulphides et al., 2012).
A great number of base metal sulphide flotation plants use sal- In South Africa, the recovery of platinum group elements (PGE)
ine water. Some of them are summarised in Table 1. The three through the selective flotation of base metal sulphides also uses
nickel flotation plants (Mt Keith, Leinster mine and Kambalda saline water. Flotation operations can be found in different parts
Nickel Concentrator) in Western Australia operated by BHP Billiton of the Bushveld Complex, e.g., Merensky reef and Platreef. The ionic
use bore water with high ionic strength. Table 2 shows the elemen- concentration varies in all Merensky concentrators, South Africa
tal compositions of the bore water used in the Mt Keith Operation, (Corin et al., 2011; Miettunen et al., 2012; Shackleton et al.,
the largest nickel flotation plant in Western Australia (Peng and 2007; Wiese et al., 2005a; 2005b; Wiese et al., 2007).
Seaman, 2011). The salinity of this water is several times higher
than that of sea water. In the Raglan mine operated by Xstrata
1.2.2. Coal
Nickel (formerly Falconbridge) in Northern Quebec, Canada, salt
In Australia, the coal industry has introduced water re-use as
levels range from 20,000 to 35,000 ppm throughout the year. An
conventional practice due to stringent policy on the amount of sal-
apparent consequence of the high salt content in the Raglan mine
ine water which a mine can discharge into the local river system.
is that the flotation circuit is able to operate without the addition
Table 3 summarises the concentration of major ions occurring in
of frother (Quinn et al., 2007).
most Australian coal mine sites.
In Chile, many copper flotation plants use seawater. One exam-
In Poland, natural water sometimes contains elevated concen-
ple is Las Luces, a copper-molybdenum plant in Taltal, owned by
trations of barium and radium in hard coal mines in Upper Silesia.
the Las Cenizas Mining Group (Grupo Minero Las Cenizas) of Chile.
The main components are shown in Table 4.
In Las Luces, seawater is brought from a distance of 7 km, mixed
Banks et al. (1997) reviewed the discharged water quality from
with tailing dam water and then used in grinding and flotation cir-
coal mine sites in England. The water qualities from selected Brit-
cuits. During the last 15 years the increase in the total dissolved
ish and Svalbardian pumping and abandoned mines and soil tips
solid content of the process water in Las Luces was from approxi-
are shown in Table 5.
mately 36.0 g/L (seawater) to 46.4 g/L (Moreno et al., 2011).
Another large copper plant using seawater in Chile is the Esperanza
Concentrator at Sierra Gorda (Antofagasta Minerals S.A AMSA). 1.2.3. Potash
This plant is processing ore at 95,000 tpd using sea water without Potash is extracted from buried ancient evaporates by under-
ground or solution mining, accounting for most of the potassium
process. Another important source is brine from landlocked water
Table 2 bodies, such as the Dead Sea, Salar de Atacama and Great Salt Lake.
The composition of bore water used in the Mt Keith Operation (mg/L) (Peng and
Seaman, 2011).
In early 1940s, potash was discovered in Saskatchewan, Canada
shown in Table 6 while drilling for oil. The known massive K2O
Na+ K+ Ca2+ Mg2+ Cl SO2
4 deposit is estimated at 67 billion tons, which is more than 40% of
20,000 940 400 5100 32,000 23,000 the world’s known reserves. The deposits locate across the south-
ern plains of Saskatchewan sloping southerly from a 1000 m depth

Table 3
Summary of the concentration of major ions in process water at Australian coal mine sites (Ofori et al., 2009).

Characteristics pH Conductivity Hardness Na+ Mg/l (M) K+ Mg/l (M) Ca2+ Mg/l (M) Mg2+ Mg/l (M) Cl Mg/l (M) SO2
4 Mg/l (M)
uS/cm (CaCO3) Mg/l (M)
Minimum 7.1 243 28 385 (0.017) 3.4 (0.0001) 6 (0.0001) 3 (0.0001) 333 (0.009) 57 (0.0006)
Average 8.0 5939.0 253.0 1178 (0.051) 19 (0.0005) 85 (0.002) 72 (0.003) 1092 (0.031) 1097.6 (0.011)
Maximum 9.2 12,860 570 3100 (0.135) 54 (0.001) 365 (0.009) 180 (0.007) 2360 (0.067) 4800 (0.050)
16 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

Table 4 Table 6
Chemical compositions of water flowing through the old mine working in the Silesia Potash producers in Saskatchewan (Perucca, 2003).
Coal Mine (Pluta, 2001).
Company Location Mining Processing
Chemical Water flowing intoold mine Water flowing out old
Agrium Vanscoy Conventional Flotation
components working mine water
IMC Belle Plaine Solution Mechanical
pH 7.4 6.0 Crystallization
Cl (g/dm3) 48.6 33.3 IMC Colonsay Conventional Flotation
SO2
4 (g/dm )
3
<0.01 0.27 IMC Esterhazy (K1 & K2) Conventional Heavy Media-flotation
Ba2+ (g/dm3) 0.12 <0.01 PCS Allan Conventional Flotation
Fe-total (g/dm3) 0.024 0.034 PCS Cory Conventional Mechanical
226
Ra (Bq/dm3) 23.3 0.3 Crystallization
PCS Patience Lake Solution Natural Crystallization
PCS Lanigan Conventional Flotation
at Saskatoon to more than 1600 m depth at Belle Plaine and up to PCS Rocanville Conventional Flotation
3000 m in depth in North Dakota (Perucca, 2003).
In North America, many researchers (Christenson and
Yaminsky, 1995; Hancer et al., 2001; Laskowski et al., 2003; attributable to water ‘‘structure making’’ or ‘‘structure breaking’’
2007; Miller et al., 1992; Pawlik et al., 2003; Yaminsky et al., has been extensively discussed, developed, and reviewed by
2010) focus on potash flotation investigation and recently, they Ben-Naim (1974), Franks (1972), Wilhelm et al. (1977) since
have achieved significant progress in the areas of soluble salt flota- 1945 when Frank and Evans (1945) gave a new meaning ‘‘ Iceberg’’.
tion chemistry particularly. Frank and Wen (1957), Nemethy and Scheraga (1962) used
The primary production of sodium carbonate (soda ash) in the ‘‘flickering clusters’’, and Glew (1962) used ‘‘clathration shells’’ to
United States comes from the trona deposits of South Western Wyo- explain the water structure. Any of these is consistent with
ming. Trona is sodium sesquicarbonate (Na2CO3NaHCO32H2O), terminology in which the class of inert, nonpolar solutes is referred
containing minor undesirable gangue materials. Since sodium car- to as ‘‘structure makers’’.
bonate (Na2CO3) and bicarbonate (NaHCO3) are the main constitu- Fig. 3 shows an ion surrounded by three concentric regions
ents of trona, most of the studies have focused on the flotation of (Frank and Wen, 1957). The innermost region (region A) is one of
carbonate and bicarbonate salts. Micro-flotation studies with immobilization, the second region (region B) contains the water
carbonate salts (Na2CO3 and NaHCO3) showed that contrary to the that is less ice-like, i.e., more random in organization than ‘‘nor-
strong flotation of NaHCO3 with both anionic and cationic collectors, mal’’, and the third region (region C) contains normal water polar-
Na2CO3 is much more difficult to float (Ozcan and Miller, 2002). ized in the ordinary way by the ionic field which has become
relatively weak. Gurney (1953) proposed that the cause of struc-
2. Water structure ture-breaking was the approximate balance in region B between
two competing orienting influences which act on any given water
2.1. Pure water structure molecule. One of these was the ‘‘normal’’ structural orienting influ-
ence of neighbouring water molecules and the other was the ori-
It has long been known that liquid water possesses distinctive enting influence upon the dipole of the spherically symmetrical
structural features which retain a certain degree of similarity or anal- ionic field.
ogy to ice. Fig. 2 shows a proposed resonance scheme for the hydrogen
bond in water. It considers resonance among the three bond struc- 2.2.2. Effect of ions on water structure in flotation
tures. The + and  signs represent formal charges, and the resonance In flotation, certain ions are considered as ‘‘structure breaking’’,
(with suitable weighting coefficients) of molecule b between struc- such as Cs+ and I which are large inorganic ions and ‘‘structure
tures I and II represent the ordinary partial polarity of the O–H bond. making’’, such as Li+, Mg2+, F and Cl which are small inorganic
The mixing-in of a contribution from structure III constitutes the for- ions and have been classified as shown in Table 7. The reason pro-
mation of a hydrogen bond (Frank and Wen, 1957). posed by Hancer et al. (2001) is that if a collector adsorbs at the salt
interface, it will displace interfacial water or penetrate through the
2.2. Saline water structure structure of water. If the structure of water is strongly hydrogen
bonded due to the presence of structure making anions and cat-
2.2.1. Water structure model ions, it will be hard for collector molecules to reach the surface
It is well known that the water structure may be evoked by and be adsorbed. However, if those ions have a tendency to destroy
the presence of ionic solutes. The idea that ionic solutes are the structure of water and create a condition for the adsorption of

Table 5
Comparison of discharge water qualities from selected British and Svalbardian pumping and abandoned mines and soil tips (Banks et al., 1997).

Discharge (l/s) pH TDS (mg/) Fe (mg/) Al (mg/) Mn (mg/) Zn (mg/) Cu (mg/) Cl (mg/) SO2
4 (mg/)

Pumping coal mines


Kibblesworth (Durham) 740 7.1 3185 0.63 0.056 690
Nicholsons (Durham) 90 7.1 3100 5.8 0.034 1170
Kimblesworth (Durham) 105 7.3 1800 5.0 0.030 380
Tilmanstone (Kent) 7.7 2107 25 795 404
Moorgreen Piper (E.Midlands) 6.9–7.9 <0.1–7.0 3600–10,800
Longyearbyen Mine 3 (Svalbard) c. 0.06 8.2 <0.01 <0.02 0.004 0.055 <0.055 236 7.4
Coal spoil tips
Crook (Durham) 3.5 1000 70 810
Quaking Houses (Durham) 4.1 2314 15 1358
Oatlands (Cumbria) 5.5 287 0.97 5.2 0.05 <0.007 146
Thurcroft (S. Yorkshire) 6.8 18.6 <0.045 2.0 <0.007 511 1327
Longyearbyen Mine 3 (Svalbard) c. 0.1 3.7 1.6 1.8 0.40 0.49 0.014 4.5 77
Sverdrupbyen Mine 1 (Svalbard) c. 0.25 2.7 179 27.5 3.2 1.3 0.168 7.0 1077
B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24 17

Fig. 2. Proposed resonance scheme for the hydrogen bond in water (Frank and Wen, 1957).

the accommodation of the hemimicelle aggregates at the salt sur-


face while structure making salts prevent the nucleation and
growth of these surface phases, as described in Fig. 5.
The interaction between water molecules and the three chloride
salts dissolved in aqueous solutions were studied by measuring the
shift in the hydrogen-bonding of water molecules (Cao et al.,
2011). The results indicate that KCl is a structure breaker salt,
while MgCl26H2O and KMgCl36H2O are structure maker salts,
but the surface charge is not a determining factor in the flotation
of soluble salts.
However, the particular arrangement of ions in the interface is
strongly influenced by ion hydration. Parsons et al. (2009),
Weissenborn and Pugh (1996) indicated that ion hydration (but
not the hydration force) was critical to specific ion effects in bubble
Fig. 3. A simple model for the structure modifications produced by a small ion: (A) coalescence, as it strongly influenced the positioning of ions in the
region of immobilization of water molecules; (B) region of structure breaking; and interfacial region which ultimately determined whether coales-
(C) structurally ‘‘normal’’ water (Frank and Wen, 1957).
cence was inhibited.

collector molecules, the flotation of soluble salt minerals will 3. Effect of saline water on particle properties
become feasible.
It is important to note that the general correlation between bulk 3.1. Hydration layers on particles
water structure and the flotation response of alkali halides sug-
gests that there are same structural phenomena occurring at the In the 1930s, Russian researchers Klassen and Mokrousov
salt surface. As depicted in Fig. 4, the interfacial water structure (1963) first studied the effect of inorganic salts on coal flotation.
at the KI surface has the least stability because the larger anion acts They found that high salt concentrations increased the floatability
as a structure breaker and tends to destroy the interfacial water of naturally hydrophobic minerals even in the absence of organic
structure corresponding to a higher contact angle. On the other collectors. They attributed this phenomenon to the action of the
hand, the KCl and NaCl surfaces exhibit a greater tendency toward salts promoting froth formation. However, after obtaining addi-
surface hydration and, thus, a more stable interfacial water struc- tional data they proposed that the lowering of electrochemical
ture. Contact angles on KCl and NaCl surfaces are therefore smaller. potential of coal particle surfaces by the adsorbed ions reduced
Due to this surface hydrophobicity, structure breaking salts allow the stability of the hydration layers surrounding the particles
resulting in their disruption. The destabilisation of hydration layers
surrounding coal particles increased their surface hydrophobicity.
Table 7
This theory was further supported by Blake and Kitchener
Surface activity and coalescence behaviour in aqueous electrolyte solutions (Hancer
et al., 2001). (1972). In their experiments, a small gas bubble was advanced
towards a polished silica plate under dilute aqueous solutions
F Cl Br I
and equilibrium films were formed on hydrophilic and methylated
Li F NF NF NF hydrophobic silica. As the KC1 concentration, used to adjust the
Structure makers
ionic strength, was increased, films became increasingly unstable
Na NF NF NF NF
K NF F F F towards ambient vibrations, although, in general, films of smaller
Cs NF F F F diameter were more stable than larger films. The higher the ionic
Structure makers strength was, the less the stable films were. Muller (1988, 1990)
Rb NF F F F
proposed his modified hydration shell hydrogen-bond model to
Note. NF, no flotation; F, flotation. calculate the contributions of hydration to the thermodynamics
from the formation of hydrogen bonds between water molecules

Fig. 4. Comparison of interfacial water structure at KI, KCl and NaCl surface (Hancer et al., 2001).
18 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

Fig. 5. Collector adsorption is inhibited at the NaCl surface (structure maker) due to strong interfacial water structure and stabilization of micelles in the bulk solution, whilst
collector adsorption is promoted at the KCl surface (structure breaker) due to weak interfacial water structure and the stability of the hemimicelle surface state (Hancer et al.,
2001).

 
in the hydration shell of nonpolar substances. His simple model Z i qwðxÞ
C i ðxÞ ¼ C i ðBÞ exp  ð3Þ
predicted that hydration-shell hydrogen bonds would have higher kT
bond-breaking enthalpies and entropies than those in pure water
and it manifested itself by the existence of a larger fraction of bro- wðxÞ ffi w0 exp ðjxÞ ð4Þ
ken hydrogen bonds in the hydration shell than in the bulk. Such
where Ci(B) and Ci(x) are the ion concentrations in bulk and at dis-
predictions about the fraction of broken hydrogen bonds were later
tance x from the charged surface, respectively. It is assumed that
confirmed by computer simulations (Laidig and Daggett, 1996;
these are dilute solutions (i.e., w(B) = 0). Ziq is the local density of
Mancera, 1996a; Mancera, 1996b). A series of simulations of solu-
any ion of charge and w is the potential energy (Pashley and
tions of methane in both pure and salt (NaCl) water have been
Karaman, 2004).
reported indicating the existence of an enhanced hydrophobicity
When the sign of the electrical charge on interacting colloids in
in the presence of salt (Mancera, 1998a; 1998b).
the suspension is the same, overlapping of their double layer leads
Arnold and Aplan (1986) observed that the coal flotation was
to repulsion between them. Attraction occurs when interacting
greater in tap water than in distilled water. They hypothesized that
colloids have oppositely charged double layers. If the sign of the
when ions (e.g., Ca2+, Mg2+, Na+, Cl, SO2 4 ) were present, the
charge is the same but the magnitude is different, a weak repulsion
hydrated layer at the surface of the coal particle became instable
will exist at large separation distances, but a strong attraction will
and much thinner. This is because the addition of electrolytes com-
occur when the colloids closely approach one another. Interacting
pressed the electrical double layer (Fuerstenau et al., 1983).
particles and bubbles may act in the same way (Deryagin et al.,
Klassen and Mokrousov (1963) showed that the contact angle of
1982).
coal particles in NaCl solutions was increased with increasing salt
The electrical double layer interaction is solely due to the over-
concentration. This is the basis of their proposition for the mecha-
lapping of the diffusive part consisting of electrolyte ions. With an
nism of dehydration of the hydration sheath around the hydropho-
increase in electrolyte concentration, the density of ions in bulk
bic coal particles. However, Laskowski and Iskra (1970) found to
solutions increases, so the counter-ion in diffuse layer enters into
the contrary that contact angles measured on two methylated sil-
the Stern layer, which makes the diffuse layer thick due to the
ica plates did not change with increasing KCl concentration. Yoon
increase in attraction force and decrease in repulsion force.
and Sabey (1989) did not find an increase in contact angle with
The compression of the electrical double layer in saline water
increasing salt concentration by using either sodium salts or sul-
enhances the thinning and rupture of the wetting film between
phate salts either.
bubbles and particles which are a critical step in the formation of
a stable bubble-particle aggregate, an important phenomenon in
3.2. Electrical double layers around particles flotation. The compression of the electrical double layer in saline
water also promotes particle aggregation in flotation and therefore
3.2.1. DLVO theory affects the recovery of both valuable and gangue minerals.
The DLVO (named after Derjaguin, Landau, Verwey and Over-
beek) theory (Derjaguin and Landau, 1941; Verwey and 3.2.2. Wetting film between particles and bubbles
Overbeek, 1948) describes the forces between charged surfaces 3.2.2.1. Coal flotation. Laskowski (1965) conducted coal flotation
interacting through a liquid medium. The theory considers two using inorganic salts (NaCl and KCl) and proposed that the electri-
main sources of interactions between particles and bubbles in cal double layer around the particles was compressed, resulting in
aqueous solutions, namely, attractive van der Waals forces and the opening of hydrophobic surface sites. The sites may then
electrostatic repulsion due to the electrical double layer. attract bubbles by hydrophobic bonding.
The notion of Debye length (j1) (also referred to as the ‘‘double Yoon (1982), Yoon and Sabey (1989) studied the effectiveness
layer thickness’’) indicates the extent of the diffuse layer which of salts on the flotation of coal using a large number of salts and
forms the surface where the surface potential has fallen to 1/e of compared the results with those of conventional flotation where
original value. Because of the equilibrium between any ion next kerosene collector and Dowfroth 250 frother were used (shown
to a charged surface and the corresponding ions in the bulk solu- in Table 8). It was found that the flotation responses in terms of
tion, the electrochemical potential of an ion at distance x from separation efficiency (defined as combustible recovery-ash
the surface must be equal to its bulk value: recovery) were comparable to those achieved with conventional
B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24 19

Table 8 100
The effect of salts on the flotation of coal (Yoon, 1982).
90
A
Reagents Yield (%) Ash (%) Separation efficiency 80
B
NaCl 50.4 11.0 57.8 70

Recovery (%)
Na2CO3 63.0 14.5 63.0
60
Na2PO4 56.4 12.4 62.5
Na2SO4 65.2 15.6 65.0 50
(NH4)2SO4 63.0 12.7 66.9
40 C
CuSO4 66.7 13.9 68.4
FeSO4 69.3 15.5 66.6 30
Fe2(SO4)3 71.1 18.1 59.7 20
Al2(SO4)3 76.5 21.9 52.2
Kerosence & Dowfroth 250 72.9 18.3 61.0 10
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Concentration (M)
flotation. When the flotation rate constants were compared, the H2O MgCl2 NaCl CaCl2 MgSO4 Na2SO4 LaCl3 LiCl KCl
kinetics of the flotation in the presence of salts was faster than that CsCl NH4Cl NaAc LiClO4 NaClO4 HClO4 HCL H2SO4 H3PO4
of conventional flotation. This was attributed to the compression of
Fig. 6. Coal flotation recovery with different groups of electrolytes (Pugh et al.,
the double layer by the presence of salts, facilitating the bubble-
1997).
particle adhesion process. The best flotation results were obtained
with salts of divalent cations which were found to reduce the elec-
trophoretic mobility to zero over wide pH ranges.
Yang et al. (1988) used an oil agglomeration process to test the 100

recovery of different coal samples in NaCl solutions and found that 90


coal recovery was increased with increasing salt concentration. 80
They attributed the increase in combustible recovery to the com- 70
pression of the electrical double layer which was thinning sur-
Recovery (%)
60
rounding the particles and oil droplets in the salt solution.
50
Medrzycka and Zwierzykowski (1988) studied the removal of
hydrocarbon from oil/water emulsions by flotation tests in the 40

presence of different inorganic salts (NaCl, Na2SO4 and Na3PO4) 30


and found that inorganic salts affect flotation efficiency very 20
strongly. This effect was associated with their influence on the 10
electrical double layers of the bubbles and droplets. The electrical
0
double layer repulsion between bubbles and droplets was higher 0 0.5 1 1.5 2 2.5 3 3.5
in NaCl solutions than in Na2SO4 or Na3PO4 solutions. Debye length (nm)
Li and Somasundaran (1993) investigated the floatability of a MgCl2 Series3 CaCl2 MgSO4 Na2SO4 LaCl3 LiCl KCl CsCl

bituminous coal in different NaCl solutions using a modified Halli- NH4Cl NaAc LiClO4 NaClO4 HClO4 HCL H2SO4 H3PO4

mond tube to delineate the role of the electrostatic interaction


Fig. 7. The relationship between the Debye length (1/j) and the flotation
between bubbles and particles and coal hydrophobicity. As salt performance of graphite in different inorganic electrolytes (Pugh et al., 1997).
concentration was increased, the effective distance for the double
layer interaction was decreased due to the compression of the dou-
ble layers. The electrostatic interaction became the predominant of smaller stable bubbles and also a reduction in the electrostatic
controlling force once the effective distance of the electrostatic interactions between particles and bubbles, and between particles
interaction between bubbles and particles was comparable to that and particles due to the double layer thickness, which is an impor-
of the hydrophobic interaction. However, the reason at high salt tant parameter controlling the flotation process according to the
concentration was not yet clear. They gave a possible explanation modified DLVO theory (Xu and Yoon, 1990b). Xu and Yoon (1990b)
which was similar with Klassen and Mokrousov (1963) showing suggested the decay length which Debye length of the electrolyte
that an increase in salt concentration destabilized the hydration solution approaches (Pugh et al., 1997) was determined from the
layer around coal particles and thus facilitated the drainage of critical film thickness for coagulation as followed (Xu and Yoon,
the water film between the bubble and the particle rendering the 1990b):
coal particles more floatable.
 
Pugh and his co-workers (Paulson and Pugh, 1996; Pugh et al., aA131 eaw2d aCD0 H
VT ¼  þ ln½1 þ exp ðjHÞ þ exp  ð5Þ
1997; Weissenborn and Pugh, 1995; Weissenborn and Pugh, 1996) 12H 2 2 D0
investigated the effect of inorganic electrolytes on coal flotation
(shown in Figs. 6 and 7). From Fig. 6, generally coal recovery where H is the distance separating the two particles, e is the dielec-
increased with concentration and varied according to the cationic/ tric constant of medium 3, a is the radius of spherical particles, wd is
anionic pair which can be classified into three groups according to the surface potential, j is the reciprocal Debye length, A131 is the
their flotation performance. Salts with divalent and trivalent cations complex Hamaker constant between particles 1 separated by med-
or anions, including MgCl2, CaCl2, Na2SO4, MgSO4 and LaCl3 which ium 3, C is the hydrophobic interaction constant, and D0 is the decay
gave high flotation response were classified to Group A. NaCl, LiCl, length.
KCl, CsCl, NH4Cl which gave medium flotation response were classi- Harvey et al. (2002) investigated the floatability of coal in NaCl
fied to Group B. Finally, NaAc, NaClO4, HCIO4, HCl, H2SO4, LiClO4 and and MgCl2 solutions using a modified Hallimond tube to examine
H3PO4 which gave a very low flotation response, even up to concen- the role of the electrical double layer interaction between bubbles
trations as high as 0.3 M electrolyte were classified to Group C. and particles. They found that in low electrolyte concentration
High flotation recoveries were attributed to an increase in the solutions coal flotation was decreased with salt concentration,
bubble-particle collision probability due to the higher concentration but in high electrolyte concentration solutions, coal flotation was
20 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

increased with salt concentration. The observation in low electro- DLVO theory of colloidal stability. In the presence of humic acid
lyte concentration solutions was not expected based on the expla- and NaCl, MgCl2, the nanoparticle suspension was effectively sta-
nation previously. bilized due to steric repulsion which was led by the adsorption
of humic acid on the particle surface. In addition, the increased
3.2.2.2. Other mineral flotation. Collins and Jameson (1976) studied nanoparticle stability was also observed in the presence of humic
the flotation rate of polystyrene latex particles of diameter 4– acid at lower CaCl2 concentration. However, at higher CaCl2 con-
20 lm with the charge in the range of 30–60 mV approximately. centration, the dissolved (unadsorbed) humic acid marcomole-
They found that the actual rate of flotation was strongly influenced cules aggregated through intermolecular bridging, which was
by the particle charge, varying by an order of magnitude in the evidenced by the increase in the scattered light intensity. These
range considered. It seemed that with particles size 4–20 lm, humic acid aggregated in turn bridged the fullerene nanoparti-
the electrical double layers affected flotation rate dominantly. In cles and aggregated together, resulting in enhanced aggregate
the presence of cationic surfactants, both particles and bubbles growth.
became positively charged that tend to increase electrical double Particle aggregation in coal flotation will not be significantly
layer repulsion which must be overcome by the dispersion forces increased with electrolyte concentration up to 1 M NaCl. So it
between the polystyrene/adsorbed layer on the particle and the appears that under 1 M NaCl, the increase in single bubble exper-
adsorbed layer on the bubble. iments cannot result in increasing capture efficiency due to particle
Pushkarova and Horn (2008) investigated surface forces aggregation. But near saturation, the increase in surface tension
between an air bubble and a flat mica surface immersed in aqueous and adhesion is much more significant (10%) and coal flotation
electrolyte solutions using a modified surface force apparatus. The in hyper-saline water is expected to be influenced by greater par-
experiments showed that there was a long-range electrical double ticle aggregation (Harvey et al., 2002). Wang and Peng (2013)
layer repulsion which was not enough to stabilize a thick wetting investigated fine coal flotation using hyper-saline water and de-
film between mica and bubble in water and electrolyte solutions of ionised water and also observed the calculated ENT (the degree
different concentrations. The electrical double layer repulsion indi- of entrainment) of 38 lm particles to be greater than 1. They
cated that the air bubble surface was negatively charged, but the attributed this to the recovery of fine mineral matter particles
charge was very weak. In their experiments, an effect of ions on entrapped in coal particle aggregates. They found that saline water
the properties of water films on the hydrophobic solid was not promoted coal aggregation as a result of hydrophobic interactions
restricted to the compression of the double layer only. and therefore fine gangue entrapment. Pawlik et al. (2004) attrib-
uted the aggregation of hydrophobic coal particles to the hydro-
3.2.3. Particle aggregation phobic forces which occur over a broad pH range, while the
Sivamohan (1990) reviewed the selective aggregation by the aggregation of hydrophilic coal particles to van der Waals interac-
addition of electrolytes. At low and intermediate concentrations tions when the zeta potential values were near zero (e.g., around
of different electrolytes, the repulsive energy due to the surface the iso-electrical point).
charge is long-ranged, however the attractive van der Waals Peng and Seaman (2011) investigated the flotation of slime-fine
energy is short-ranged. The so called secondary minimum in the fractions of Mt. Keith pentlandite ore in de-ionised water and bore
total interaction curve is considered insignificant. Any aggregation water with high ionic strength. The zeta potential between serpen-
due to the secondary minimum is easily broken by gentle stirring. tine and pentlandite particles was close to zero in bore water, so
At high electrolytic concentration, rapid coagulation ensues when the reduced electrostatic attraction mitigated serpentine coating
the energy barrier occurring when any attempt to bring the zeta resulting in improved plentlandite flotation. Meanwhile, in bore
potential to zero at low and intermediate concentrations of differ- water, MgO recovery was decreased as well. Peng and Seaman
ent electrolytes was removed. (2011) attributed this to aggregation of fine serpentine mineral
Xu and Yoon (1990a) studied the coagulation of hydrophobic particles by electrolytes with lower entrainment.
particles (methylated silica). Their work showed that at low elec- Corin et al. (2011) studied the effect of water quality on the
trolyte concentration and under conditions of low shear, the elec- entrainment of naturally floatable gangue (talc) in the flotation of
trostatic repulsive forces could be easily overcome by long range platinum group elements. In this study, a polymeric depressant
hydrophobic attraction leading to weak coagula structures, even Stypres 504 (a modified guar gum) was used to depress talc. They
in cases where the surface potential of the particles was more neg- found that with an increase in the ionic strength, the amount of
ative. Zhou et al. (1996) showed the influence of gas nuclei on talc entrained per unit water decreased. The mechanism proposed
hydrophobic coagulation. The presence of nuclei decreased the is that the coagulative nature of talc increased with increasing in
density and increased the size of coal aggregates formed from the ionic strength because the calcium and magnesium ions were
the coagulation of fine particles. From modelling studies it was adsorbed onto the talc particle surfaces resulting in lower surface
shown that gas nuclei had a volumetric mean diameter between charge and electrostatic repulsion and therefore a reduction in
6 and 17 lm. With regard to the hydrophobic particles, since the entrained mass.
zeta potential in dilute electrolyte solutions was shown to be lower
than the bubbles, it could be anticipated that hydrophobic coagu-
lation occurred fairly readily. In addition, since amorphous graph- 4. Effect of saline water on bubble properties
ite particles probably entrapped gas on the surface, particle
bridging by two gas nuclei probably enhanced this hydrophobic 4.1. Inhibition of bubble coalescence in salt solutions
effect. The capture of hydrophobic aggregates by the bubbles must
make a major contribution to the flotation process (Pugh et al., Marrucci (1969) found that salts can inhibit bubble coalescence
1997). by retarding thinning of the intervening liquid film between bub-
Chen and Elimelech (2007) investigated the effect of monova- ble pairs. At sufficiently high salts concentration, surface tension
lent and divalent electrolyte concentrations on initial aggregation gradient which results from the thinning process immobilize the
kinetics of fullerene nanoparticles in the presence of Suwannee gas–liquid interface between coalescing bubbles. When it occurs,
River humic acid. In the absence of humic acid, the aggregation the time required for coalescence dramatically increases. Eqs. (6),
behaviour of the fullerene nanoparticles in the presence of NaCl, (7) show the critical concentration of electrolyte required to immo-
MgCl2, and CaCl2 was found to be consistent with the classic bilize the gas–liquid interface of coalescing bubbles:
B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24 21

 1=2  2
Br @r
ct ¼ 1:18 Rg T ð6Þ
rb @C
(Prince and Blanch, 1990b)

r 2=3
b
sij ¼ ð7Þ
21=3
(Levich, 1962)
( "  2  #)1=2
dh 8 4c dr 2 2r A
¼ þh þ ð8Þ
dt R2d ql Rg T dc rb 6ph3 Fig. 8. Bubble in pure water (Left) and salt water (Right) (Lessard and Zieminski,
1971).
(Oolman and Blanch, 1986)
1 XXn T   o Craig et al. (1993a) made the intriguing observation that certain
CT ¼ hij þ hBij þ hLS
ij  exp t ij =sij ð9Þ
2 i j electrolytes had no effect on bubble coalescence, even at very high
concentrations (0.5 M). He assessed the inhibition of bubbles to
(Prince and Blanch, 1990a) coalescence in electrolyte solutions by the application of a combin-
Film thinning times for bubbles with immobile interfaces are on ing rule based on the nature of the cationic/anionic pair. This rule
the order of seconds. Even the incorporation of transport rates alle- (Table 10) enabled one to predict whether or not the electrolyte
viating the surface tensions gradients and giving partial mobility to would inhibit coalescence of gas bubbles in the electrolyte solu-
the gas–liquid interface result in coalescence time of the order of tions. Christenson and Yaminsky (1995) observed surface activity
hundreds of milliseconds, as well (Nicodemo et al., 1972). The cal- and coalescence behaviour in electrolyte solutions (Table 11).
culated contact time in turbulent dispersions by Eq. (8) is approx- The specific effects of electrolytes on bubble coalescence were
imately 10 ms, however bubble coalescence is not envisaged for thought to be related to their effects on the water structure and
systems where the salt concentration is sufficient to immobilize hence the hydrophobic interactions. Henry and Craig (2008),
the gas–liquid interface. When salt concentrations are lower than Henry and Craig (2010), Henry et al. (2007) measured the effects
the transition value, Eq. (9) may be used to calculate thinning of electrolytes on bubble coalescence in non-aqueous solvents
times. Eq. (8) has been numerically integrated for various bubble and aqueous electrolyte solutions. They proposed a mechanism
sizes with the initial and final film thickness previously described. for coalescence inhibition in which some electrolyte combinations
The numerical solutions provide coalescence time which may use modify the hydrodynamic boundary condition at the air–water
Eq. (9) to calculate coalescence rates for bubbles in electrolyte interface. They suggested that bubble coalescence inhibition was
solutions. driven by non-equilibrium concentration gradients of solute at
Marrucci and Nicodemo (1967) measured the average bubble the interface.
size in a bubble column in the presence of a number of different Craig et al. (1993b) suggested that the coalescence in pure
electrolytes, KCl, KOH, KNO, KI, K2SO4, CuSO4, K3PO4, AlCl3 and water might also be caused by a strong hydrophobic attractive
Co(NO3)2, at a number of superficial gas velocities. Their conclusion force which opposed to the hydrodynamic repulsion existing
was that the electrolytes increased the electrical repulsive forces at between the colliding bubbles. The reduction in the attraction
the bubble surface, inhibiting coalescence between bubbles. between bubbles occurred with the adsorption of certain ion pairs
Lessard and Zieminski (1971) investigated the effects of inorganic at the air/water interface which reduced this hydrophobic force.
electrolytes, AlCl3, MgSO4, Na2SO4, CaCl2, MgCl2, NaCl, LiCl, and This was explained by the local influence of ions on the water
NaBr, on bubble coalescence and interfacial gas transfer in aqueous structure which was someway related to the hydrophobic force.
solution (shown in Table 9). The coalescence experiments con- Interactions between frother and saline water have been also
sisted of contacting a number of pairs of bubbles and evaluating proposed. With increasing frother concentration, surface tension
the coalescence percentage as a function of solute concentration. of the solution decreases but the ability of frother to decrease sur-
They found the existence of a sharp transition concentration which face tension was increased in electrolyte solutions (Castro et al.,
enabled a comparison of the effectiveness of the salts. The concen- 2013). Castro et al. (2013) studied the interaction of MIBC with
tration resulting in 50% coalescence was defined as the transition NaCl. They found that at a concentration of about 120 ppm MIBC,
concentration, at which coalescence was sharply reduced. These a critical point occurred. For MIBC concentration exceeding this
concentrations correlated well with ionic entropy of solution and point, the surface tension of NaCl containing MIBC was lower than
with the self-diffusion ability of water in solution (see Fig. 8). the surface tension of water containing MIBC. Below this point the
effect of surface inactive compound (inorganic salts) dominated
over surface active compound (MIBC). However, above this MIBC
Table 9
concentration, the surface tension began to decrease and the effect
Bubble coalescence transition concentration in
different electrolytes (Lessard and Zieminski, was dominated by MIBC. This point was called the surface tension
1971). switch point (s.t.s.p.).
Salt Transition concentration,
M (50% coalescence) 4.2. Froth stability
MgSO4 0.032
AlCl3 0.035 Froth structure and froth stability are known to play a signifi-
MgCl2 0.055
CaCl2 0.055
cant role in determining the mineral grade and recovery achieved
NasSO4 0.061 from a flotation operation. Froth stability is mainly dependant on
LiCl 0.16 frother (type and concentration) and amount and nature of the
NaCl 0.175 suspended particles, in particular, particle hydrophobicity and size
NaBr 0.22
(Johansson and Pugh, 1992; Schwarz and Grano, 2005). However,
KCl 0.23
there are other parameters such as quality of process water, gas
22 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

Table 10
A combination rule of electrolytes on the inhibition of bubble coalescence (Craig et al., 1993b).

b a a a a a a b a
Anions H+ Mg2+ Na+ Ca2+ K+ NH+4 Cs+ Me4N+ Li+
 p
a OH  I.Sol I.Sol
p p p p p p
a Cl  
p p p
a Br  Unavail
p p p p p
a NO 3 
b ClO3 Unavail Unavail  Unavail I.Sol Unavail I.Sol Unavail Unavail
p p p p
a SO2
4  I.Sol
p
b ClO4   Unavail I.Sol  I.Sol Unavail
p p
b CH3COO     
p
a Oxlate2  I.Sol I.Sol I.Sol I.Sol Unavail Unavail
p p
Combining Rules: aa or bb gives , ab or ab gives , I.Sol = Insufficiently soluble Unavail = Salt unavailable, addition of salt: prevents coalescence Has no effect on
coalescence .

Table 11
80
Surface activity and coalescence behaviour in aqueous electrolytes solutions
(Christenson and Yaminsky, 1995). MgCl2
NaCl
Electrolyte dc/dc  Dc/c(1 M) (dc/dc)2 Transition concn

Cumulative Recovery (%)


60
NaCO3
CaCl2 3.2 10 Yes
NaCl 1.6 2.6 Yes
(CH3)4NCl 0.6 0.4 No Increasing frother dosage
NaClO3 0.55 0.3 No 40
CH3COOK 0.45 0.2 No
HCl 0.3 0.1 No
HNO3 0.8 0.6 No
(COOH)2 0.85 0.7 No 20
HClO4 1.6 2.6 Yes
CH3COOH 3.2 10 Yes

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
dispersion and particle contact angle that can affect froth stability. Dynamic Froth Stability (min)
The factors affecting froth stability are being extensively studied
Fig. 9. Correlation between coal recovery and dynamic froth stability in the
(Barbian et al., 2005; Melo and Laskowski, 2006). A comprehensive presence of frother at 0.1 M salt concentration (Kurniawan et al., 2011).
review on froth and froth stability in flotation has been published
by Subrahmanyam and Forssberg (1988), Farrokhpay (2011).
Generally, the action of frothers is believed to result from showed the lowest froth stability that was similar to the recovery
retarding coalescence (King, 1982). However, it is well documented response (shown in Fig. 9). They explained the experimental
that many salts inhibit bubble coalescence (Craig, 2004; Craig results based on the salt transition concentration. When using NaCl
et al., 1993b; Laskowski et al., 2003; Lessard and Zieminski, and MgCl2 solutions, coal recovery, dynamic froth stability and
1971; Marrucci and Nicodemo, 1967). Inorganic ions appear to froth bubble size changed with the electrolyte addition below
slow inter-bubble film drainage. There are, however, significant the transition concentration, while beyond the transition concen-
differences between inorganic salts and frothers. High salt concen- tration, the changes were not significant.
trations are required for coalescence inhibition (Craig et al., 1993b; Corin et al. (2011) investigated the effect of ionic strength of
Lessard and Zieminski, 1971) compared to a few parts per million plant water on valuable mineral and gangue recovery in a platinum
of frother (King, 1982). Inorganic salts generally tend to increase bearing ore from the Merensky reef, South Africa. With increase in
surface tension while frothers reduce it; and inorganic ions are the ionic strength of the system, there was an increase in froth sta-
usually not able to form froth (foam) in two-phase (water–gas) bility, resulting in an increase in mass pulls and water recoveries.
systems. Lekki and Laskowski (1975) observed that only in the
presence of hydrophobic particles would salt solutions form a sta-
ble froth. They noted that inorganic electrolytes fall into the cate- 5. Conclusions
gory of surface inactive agents while frothers are surface active.
Quinn et al. (2007) studied why the Raglan concentrator The importance of saline water in flotation is being increasingly
(Xstrata Nickel) did not employ frother (MIBC) compared to a typ- recognised. There is an agreement that saline water enhances the
ical frother system and focused on gas dispersion (bubble size and flotation of coal and other valuable minerals. However, there is
gas holdup) and froth overflow rate. In two-phase tests, the results not an agreed mechanism proposed to explain the increased flota-
revealed that bubble size reduced, gas holdup which correlated tion by saline water. Three mechanisms, destabilisation of hydra-
with ionic strength increased and froth formation was limited in tion layers surrounding the particles, inhibition of bubble
salt solutions. At an ionic strength (0.4 M NaCl) the increase in coalescence and the compression of the electrical double layer
gas holdup was comparable to ca.10 ppm MIBC. In three-phase around the particles, have been proposed to explain the increased
tests, for a sulphide ore, bubble size and froth overflow rate were mineral flotation in saline water primarily on coal flotation, but are
also comparable between 0.4 M NaCl and 10 ppm MIBC. contradictory in some circumstances. Limited studies have been
Kurniawan et al. (2011) investigated the froth properties in coal conducted to investigate the effect of saline water on the base
flotation using MgCl2, NaCl and NaClO3 solutions in the absence metal mineral flotation as well as the recovery of gangue minerals.
and presence of Dowfroth 250. They found that in the presence The properties of flotation reagents in saline water have not been
of Dowfroth 250, MgCl2 gave the most stable froth, while NaClO3 studied.
B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24 23

Acknowledgements Henry, C.L., Craig, V.S.J., 2008. Ion-specific influence of electrolytes on bubble
coalescence in nonaqueous solvents. Langmuir 24 (15), 7979–7985.
Henry, C.L., Craig, V.S.J., 2010. The link between ion specific bubble coalescence and
The authors greatly appreciate financial support from Austra- hofmeister effects is the partitioning of ions within the interface. Langmuir 26
lian Coal Association Research Program (ACARP), and discussions (9), 6478–6483.
Henry, C.L., Dalton, C.N., Scruton, L., Craig, V.S.J., 2007. Ion-specific coalescence of
and suggestion from Frank Mercuri and John Gartlan at Xstrata
bubbles in mixed electrolyte solutions. J. Phys. Chem. C 111 (2), 1015–1023.
Coal and Ian Brake, Ben Cronin and Susan Watkins from BHP Billi- Johansson, G., Pugh, R.J., 1992. The influence of particle size and hydrophobicity on
ton Mitsubishi Alliance. the stability of mineralized froths. Int. J. Miner. Process. 34 (1–2), 1–21.
King, R.P., 1982. Principles of Flotation. South African Institute of Mining and
Metallurgy, Johannesburg.
References Klassen, V.I., Mokrousov, V.A., 1963. An introduction to the theory of flotation,
second ed. Butterworths, London.
Arnold, B.J., Aplan, F.F., 1986. The effect of clay slimes on coal flotation, Part II: the Kurniawan, A.U., Ozdemir, O., Nguyen, A.V., Ofori, P., Firth, B., 2011. Flotation of coal
role of water quality. Int. J. Miner. Process. 17 (3–4), 243–260. particles in MgCl2, NaCl, and NaClO3 solutions in the absence and presence of
Banks, D., Younger, P.L., Arnesen, R.-T., Iversen, E.R., Banks, S.B., 1997. Mine-water Dowfroth 250. Int. J. Miner. Process. 98 (3–4), 137–144.
chemistry: the good, the bad and the ugly. Environ. Geol. 32 (3), 157–174. Laidig, K.E., Daggett, V., 1996. Testing the modified hydration-shell hydrogen-bond
Barbian, N., Hadler, K., Ventura-Medina, E., Cilliers, J.J., 2005. The froth stability model of hydrophobic effects using molecular dynamics simulation. J. Phys.
column: linking froth stability and flotation performance. Miner. Eng. 18 (3), Chem. 100 (14), 5616–5619.
317–324. Laskowski, J.S., 1965. Coal flotation in solution with a raised concentration of
Ben-Naim, A., 1974. Water and Aqueous Solutions. Introduction to a Molecular inorganic salts. Coal Int. (Redhill, Surrey, England: 1994) 211 (5448), 361.
Theory. Plenum Press, New York. Laskowski, J.S., Cho, Y.S., Ding, K., 2003. Effect of frothers on bubble size and foam
Bıçak, Ö., Ekmekçi, Z., Can, M., Öztürk, Y., 2012. The effect of water chemistry on stability in potash ore flotation systems. Can. J. Chem. Eng. 81 (1), 63–69.
froth stability and surface chemistry of the flotation of a Cu–Zn sulfide ore. Int. J. Laskowski, J.S., Iskra, J., 1970. Role of capilllary effects in bubble-particle collision in
Miner. Process. 102–103, 32–37. flotation. Trans. Inst. Min. Met. 79, C1–C6.
Blake, T.D., Kitchener, J.A., 1974. Stability of aqueous films on hydrophobic Laskowski, J.S., Pawlik, M., Ansari, A., 2007. Effect of brine concentration on the
methylated silica. J. Chem. Soc., Faraday Trans. 1: Phys. Chem. Condens. krafft point of long chain primary amines. Can. Metall. Q. 46 (3), 295–300.
Phases 68, 12. Lekki, J., Laskowski, J., 1975. A new concept of frothing in flotation systems and
Cao, Q., Wang, X., Miller, J.D., Cheng, F., Jiao, Y., 2011. Bubble attachment time and general classification of flotation frothers. In: Proceedings of the 11th
FTIR analysis of water structure in the flotation of sylvite, bischofite and International Mineral Processing Congress, Cagliari, Italy.
carnallite. Miner. Eng. 24 (2), 108–114. Lessard, R.R., Zieminski, S.A., 1971. Bubble coalescence and gas transfer in aqueous
Castro, S., 2012. Challenges in flotation of Cu–Co sulfide ores in sea water. In: electrolytic solutions. Ind. Eng. Chem. Fundam. 10 (2), 260–269.
Drelich, J. (Ed.), The First International Symposium on Water in Mineral Levich, V.G., 1962. Physicochemical hydrodynamics. Prentice-Hall, Englewood
Processing. Society for Mining, Metallurgy, and Exploration, Seattle, USA. Cliffs, N.J.
Castro, S., Miranda, C., Toledo, P., Laskowski, J.S., 2013. Effect of frothers on bubble Li, C., Somasundaran, P., 1993. Role of electrical double layer forces and hydrophobicity
coalescence and foaming in electrolyte solutions and seawater. Int. J. Miner. in coal flotation in sodium chloride solutions. Energy Fuels 7 (2), 244–248.
Process. 124, 8–14. Mancera, R.L., 1996a. Hydrogen-bonding behaviour in the hydrophobic hydration of
Chen, K.L., Elimelech, M., 2007. Influence of humic acid on the aggregation kinetics simple hydrocarbons in water. J. Chem. Soc., Faraday Trans. 92 (14), 2547–2554.
of fullerene (C60) nanoparticles in monovalent and divalent electrolyte Mancera, R.L., 1996b. Towards an understanding of the molecular basis of
solutions. J. Colloid Interface Sci. 309 (1), 126–134. hydrophobicity. J. Comput.-Aided Mol. Des. 10 (4), 321–326.
Christenson, H.K., Yaminsky, V.V., 1995. Solute effects on bubble coalescence. J. Mancera, R.L., 1998a. Computer simulation of the effect of salt on the hydrophobic
Phys. Chem. 99 (25), 10420–10420. effect. J. Chem. Soc., Faraday Trans. 94 (24), 3549–3559.
Collins, G.L., Jameson, G.J., 1976. Experiments on the flotation of fine particles: the Mancera, R.L., 1998b. Does salt increase the magnitude of the hydrophobic effect? A
influence of particle size and charge. Chem. Eng. Sci. 31 (11), 985–991. computer simulation study. Chem. Phys. Lett. 296 (5–6), 459–465.
Corin, K.C., Reddy, A., Miyen, L., Wiese, J.G., Harris, P.J., 2011. The effect of ionic Marrucci, G., 1969. A theory of coalescence. Chem. Eng. Sci. 24 (6), 975–985.
strength of plant water on valuable mineral and gangue recovery in a platinum Marrucci, G., Nicodemo, L., 1967. Coalescence of gas bubbles in aqueous solutions of
bearing ore from the Merensky reef. Miner. Eng. 24 (2), 131–137. inorganic electrolytes. Chem. Eng. Sci. 22 (9), 1257–1265.
Craig, V.S.J., 2004. Bubble coalescence and specific-ion effects. Curr. Opin. Colloid Medrzycka, K.B., Zwierzykowski, W., 1988. The effect of the nature of inorganic ions
Interface Sci. 9 (1–2), 178–184. on hydrocarbon flotation. Separat. Sci. Technol. 23 (6–7), 719–729.
Craig, V.S.J., Ninham, B.W., Pashley, R.M., 1993a. Effect of electrolytes on bubble Melo, F., Laskowski, J.S., 2006. Fundamental properties of flotation frothers and their
coalescence. Nature 364 (6435), 317–319. effect on flotation. Miner. Eng. 19 (6–8), 766–773.
Craig, V.S.J., Ninham, B.W., Pashley, R.M., 1993b. The effect of electrolytes on bubble Miettunen, H., Kaukonen, R., Corin, K., Ojala, S., Keiski, R.L., 2012. Effect of reducing
coalescence in water. J. Phys. Chem. 97 (39), 10192–10197. grinding conditions on the flotation behaviour of low-S content PGE ores.
Dai, Z., Fornasiero, D., Ralston, J., 2000. Particle-bubble collision models – a review. Miner. Eng. 36–38, 195–203.
Adv. Colloid Interface Sci. 85 (2–3), 231–256. Miller, J.D., Yalamanchili, M.R., Kellar, J.J., 1992. Surface charge of alkali halide
Derjaguin, B., Landau, L., 1941. Theory of the stability of strongly charged lyophobic particles as determined by laser-Doppler electrophoresis. Langmuir 8 (5), 1464–
sols and the adhesion of strongly charged particles in solutions of electrolytes. 1469.
Acta Physicochim. URSS 14, 633–662. Moreno, P.A., Aral, H., Cuevas, J., Monardes, A., Adaro, M., Norgate, T., Bruckard, W.,
Deryagin, B.V., Dukhin, S.S., Rulev, N.N., 1982. Kinetic theory of the flotation of small 2011. The use of seawater as process water at Las Luces copper–molybdenum
particles. Russ. Chem. Rev. 51 (1), 51–67. beneficiation plant in Taltal (Chile). Miner. Eng. 24 (8), 852–858.
Drelich, J., Miller, J.D., 2012. Induction time measurements for air bubbles on Muller, N., 1988. Is there a region of highly structured water around a nonpolar
Chalcopyrite, Bornite, and Gold in seawater. In: Drelich, J. (Ed.), The First solute molecule? J. Solut. Chem. 17 (7), 661–672.
International Symposium on Water in Mineral Processing. Society for Mining, Muller, N., 1990. Search for a realistic view of hydrophobic effects. Acc. Chem. Res.
Metallurgy, and Exploration, Seattle, USA. 23 (1), 23–28.
Farrokhpay, S., 2011. The significance of froth stability in mineral flotation – a Nemethy, G., Scheraga, H.A., 1962. Structure of water and hydrophobic bonding in
review. Adv. Colloid Interface Sci. 166 (1–2), 1–7. proteins. II. Model for the thermodynamic properties of aqueous solutions of
Frank, H.S., Evans, M.W., 1945. Free volume and entropy in condensed systems III. hydrocarbons. J. Chem. Phys. 36 (12), 3401–3417.
Entropy in binary liquid mixtures; partial molal entropy in dilute solutions; Nicodemo, L., Marrucci, G., Acierno, D., 1972. Bubble pair coalescence in electrolyte
structure and thermodynamics in aqueous electrolytes. J. Chem. Phys. 13 (11), solutions. Quad. Ingn. Chim. Ital. 8, 1.
507–532. Ofori, P., Firth, B., McNally, C., Nguyen, A., 2009. Working Effectively with Saline
Frank, H.S., Wen, W.-Y., 1957. Structural aspects of ion-solvent interaction in Water in Coal Preparation. CSIRO Energy Technology, the University of
aqueous solutions: a suggested picture of water structure. Discuss. Faraday Soc. Queensland.
24, 133–140. Oolman, T.O., Blanch, H.W., 1986. Bubble coalescence in stagnant liquids. Chem.
Franks, F., 1972. Water: a Comprehensive Treatise. Plenum Press, New York. Eng. Commun. 43 (4–6), 237–261.
Fuerstenau, D.W., Rosenbaum, J.M., Laskowski, J., 1983. Effect of surface functional Ozcan, O., Miller, J.D., 2002. Flotation of sodium carbonate and sodium bicarbonate
groups on the flotation of coal. Colloids Surfaces 8 (2), 153–173. salts from their saturated brines. Miner. Eng. 15 (8), 577–584.
Glew, D.N., 1962. Aqueous solubility and the gas-hydrates, the methane-water Parsons, D.F., Boström, M., Maceina, T.J., Salis, A., Ninham, B.W., 2009. Why direct or
system 1. J. Phys. Chem. 66 (4), 605–609. reversed hofmeister series? Interplay of hydration, non-electrostatic potentials,
Grano, S., 2006. Effect of impeller rotational speed on the size dependent flotation and ion size. Langmuir 26 (5), 3323–3328.
rate of galena in full scale plant cells. Miner. Eng. 19 (13), 1307–1318. Pashley, R.M., Karaman, M.E., 2004. Applied Colloid and Surface Chemistry. J. Wiley,
Gurney, R.W., 1953. Ionic Processes in Solution. McGraw-Hill, New York. Chichester, West Sussex, England; Hoboken, N.J.
Hancer, M., Celik, M.S., Miller, J.D., 2001. The significance of interfacial water Paulson, O., Pugh, R.J., 1996. Flotation of inherently hydrophobic particles in
structure in soluble salt flotation systems. J. Colloid Interface Sci. 235 (1), 150– aqueous solutions of inorganic electrolytes. Langmuir 12 (20), 4808–4813.
161. Pawlik, M., Laskowski, J.S., Ansari, A., 2003. Effect of carboxymethyl cellulose and
Harvey, P.A., Nguyen, A.V., Evans, G.M., 2002. Influence of electrical double-layer ionic strength on stability of mineral suspensions in potash ore flotation
interaction on coal flotation. J. Colloid Interface Sci. 250 (2), 337–343. systems. J. Colloid Interface Sci. 260 (2), 251–258.
24 B. Wang, Y. Peng / Minerals Engineering 66–68 (2014) 13–24

Pawlik, M., Laskowski, J.S., Melo, F., 2004. Effect of coal surface wettability on Weissenborn, P.K., Pugh, R.J., 1995. Surface tension and bubble coalescence
aggregation of fine coal particles. Coal Preparat. 24 (5–6), 233–248. phenomena of aqueous solutions of electrolytes. Langmuir 11 (5),
Peng, Y., Seaman, D., 2011. The flotation of slime-fine fractions of Mt. Keith 1422–1426.
pentlandite ore in de-ionised and saline water. Miner. Eng. 24 (5), 479–481. Weissenborn, P.K., Pugh, R.J., 1996. Surface tension of aqueous solutions of
Perucca, C.F., 2003. Potash processing in Saskatchewan – a review of process electrolytes: relationship with ion hydration, oxygen solubility, and bubble
technologies. CIM Bull. 96 (1070), 61–65. coalescence. J. Colloid Interface Sci. 184 (2), 550–563.
Pluta, I., 2001. Barium and radium discharged from coal mines in the Upper Silesia, Wiese, J., Harris, P., Bradshaw, D., 2005a. The influence of the reagent suite on the
Poland. Environ. Geol. 40 (3), 345–348. flotation of ores from the Merensky reef. Miner. Eng. 18 (2), 189–198.
Prince, M.J., Blanch, H.W., 1990a. Bubble coalescence and break-up in air-sparged Wiese, J., Harris, P., Bradshaw, D., 2005b. Investigation of the role and interactions of
bubble columns. AIChE J. 36 (10), 1485–1499. a dithiophosphate collector in the flotation of sulphides from the Merensky reef.
Prince, M.J., Blanch, H.W., 1990b. Transition electrolyte concentrations for bubble Miner. Eng. 18 (8), 791–800.
coalescence. AIChE J. 36 (9), 1425–1429. Wiese, J., Harris, P., Bradshaw, D., 2007. The response of sulphide and gangue
Pugh, R.J., Weissenborn, P., Paulson, O., 1997. Flotation in inorganic electrolytes; the minerals in selected Merensky ores to increased depressant dosages. Miner.
relationship between recover of hydrophobic particles, surface tension, bubble Eng. 20 (10), 986–995.
coalescence and gas solubility. Int. J. Miner. Process. 51 (1–4), 125–138. Wilhelm, E., Battino, R., Wilcock, R.J., 1977. Low-pressure solubility of gases in
Pushkarova, R.A., Horn, R.G., 2008. Bubble-solid interactions in water and liquid water. Chem. Rev. 77 (2), 219–262.
electrolyte solutions. Langmuir 24 (16), 8726–8734. Xu, Z., Yoon, R.-H., 1990a. A study of hydrophobic coagulation. J. Colloid Interface
Quinn, J.J., Kracht, W., Gomez, C.O., Gagnon, C., Finch, J.A., 2007. Comparing the Sci. 134 (2), 427–434.
effect of salts and frother (MIBC) on gas dispersion and froth properties. Miner. Xu, Z., Yoon, R.H., 1990b. A study of hydrophobic coagulation. J. Colloid Interface Sci.
Eng. 20 (14), 1296–1302. 134 (2), 427–434.
Schwarz, S., Grano, S., 2005. Effect of particle hydrophobicity on particle and water Yaminsky, V.V., Ohnishi, S., Vogler, E.A., Horn, R.G., 2010. Stability of squeous films
transport across a flotation froth. Colloids Surfaces A: Physicochem. Eng. between bubbles. Part 1. The effect of speed on bubble coalescence in purified
Aspects 256 (2–3), 157–164. water and simple electrolyte solutions. Langmuir 26 (11), 8061–8074.
Shackleton, N.J., Malysiak, V., O’Connor, C.T., 2007. Surface characteristics and Yang, G.C.C., Markuszewski, R., Wheelock, T.D., 1988. Oil agglomeration of coal in
flotation behaviour of platinum and palladium tellurides. Miner. Eng. 20 (13), inorganic salt solutions. Coal Preparat. 5 (3–4), 133–146.
1232–1245. Yoon, R.H., 1982. Flotation of coal using micro-bubbles and inorganic salts. Min.
Sivamohan, R., 1990. The problem of recovering very fine particles in mineral Congress J. 68 (12), 76–80.
processing – a review. Int. J. Miner. Process. 28 (3–4), 247–288. Yoon, R.H., Sabey, J.B., 1989. Coal flotation in inorganic salt solution. In: Botsaris, G.,
Subrahmanyam, T.V., Forssberg, E., 1988. Froth stability, particle entrainment and Glazman, Y.M. (Eds.), Interfacial Phenomena in Coal Technology. Dekker, New
drainage in flotation – a review. Int. J. Miner. Process. 23 (1–2), 33–53. York, pp. 87–114.
Verwey, E.J.W., Overbeek, J.T.G., 1948. Theory of the Stability of Lyophobic Colloids. Zhou, Z.A., Xu, Z., Finch, J.A., 1996. Effect of gas nuclei on hydrophobic coagulation. J.
Elsevier, New York. Colloid Interface Sci. 179 (1), 311–314.
Wang, B., Peng, Y., 2013. The behaviour of mineral matter in fine coal flotation using
saline water. Fuel 109, 309–315.

Anda mungkin juga menyukai