You are on page 1of 31

 

 
Some physicochemical aspects of water-soluble mineral flotation

Zhijian Wu, Xuming Wang, Haining Liu, Huifang Zhang, Jan D. Miller

PII: S0001-8686(16)30011-2
DOI: doi: 10.1016/j.cis.2016.06.005
Reference: CIS 1663

To appear in: Advances in Colloid and Interface Science

Please cite this article as: Wu Zhijian, Wang Xuming, Liu Haining, Zhang Huifang,
Miller Jan D., Some physicochemical aspects of water-soluble mineral otation, Advances
in Colloid and Interface Science (2016), doi: 10.1016/j.cis.2016.06.005

This is a PDF le of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its nal form. Please note that during the production process
errors may be discovered which could aect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Some physicochemical aspects of water-soluble mineral flotation

PT
Zhijian Wua, Xuming Wangb, Haining Liua, Huifang Zhanga, Jan D. Millerb,*
a
Laboratory of Salt Lake Resources and Chemistry, Qinghai Institute of Salt Lakes,

RI
Chinese Academy of Sciences, Xining 810008, China
b
Department of Metallurgical Engineering, College of Mines and Earth Sciences,

SC
University of Utah, 135 S 1460 E, Salt Lake City, UT 84112, USA

* Corresponding author. Tel.: +1-801-581-5160; fax: +1-801-581-4937. E-mail address:

NU
Jan.Miller@utah.edu (J.D. Miller).

Abstract MA
Some physicochemical aspects of water-soluble mineral flotation including
hydration phenomena, associations and interactions between collectors, air bubbles, and
water-soluble mineral particles are presented. Flotation carried out in saturated salt
D

solutions, and a wide range of collector concentrations for effective flotation of different
salts are two basic aspects of water-soluble mineral flotation. Hydration of salt ions,
TE

mineral particle surfaces, collector molecules or ions, and collector aggregates play an
important role in water-soluble mineral flotation. The adsorption of collectors onto
bubble surfaces is suggested to be the precondition for the association of mineral particles
P

with bubbles. The association of collectors with water-soluble minerals is a complicated


CE

process, which may include the adsorption of collector molecules or ions onto such
surfaces, and/or the attachment of collector precipitates or crystals onto the mineral
surfaces. The interactions between the collectors and the minerals include electrostatic
AC

and hydrophobic interactions, hydrogen bonding, and specific interactions, with


electrostatic and hydrophobic interactions being the common mechanisms. For the
association of ionic collectors with minerals with an opposite charge, electrostatic and
hydrophobic interactions could have a synergistic effect, with the hydrophobic
interactions between the hydrophobic groups of the previously associated collectors and
the hydrophobic groups of oncoming collectors being an important attractive force.
Association between solid particles and air bubbles is the key to froth flotation, which is
affected by hydrophobicity of the mineral particle surfaces, surface charges of mineral
particles and bubbles, mineral particle size and shape, temperature, bubble size, etc. The
use of a collector together with a frother and the use of mixed surfactants as collectors are
suggested to improve flotation.

Key words: Flotation, Water-soluble minerals, Hydration, Hydrophobic interactions,


Electrostatic interactions

1
ACCEPTED MANUSCRIPT

Contents

1. Introduction
2. General froth flotation processes
3. Characteristics of water-soluble mineral flotation

PT
3.1. Flotation in saturated salt solutions
3.2. Wide range of collector concentrations for effective flotation of different salts
4. Hydration phenomena in flotation processes

RI
4.1. Hydration of salt ions
4.2. Hydration of water-soluble mineral surfaces
4.3. Hydration of collector molecules or ions

SC
4.4. Hydration of collector aggregates
5. Collector adsorption at liquid/bubble interfaces
6. Collector association with water-soluble minerals

NU
6.1. Association processes with hydration changes
6.2. Factors affecting association
7. Interactions between collectors and water-soluble minerals
MA
7.1. Electrostatic interactions
7.2. Hydrophobic interactions
7.3. Hydrogen bonding
7.4. Specific interactions
D

8. Association of water-soluble mineral particles with bubbles


TE

9. Conclusions and prospects


9.1. General conclusions and prospects
9.2. Use of frothers
P

9.3. Use of mixed surfactants as collectors


Acknowledgments
CE

References
AC

2
ACCEPTED MANUSCRIPT

1. Introduction

Froth flotation involving physicochemical phenomena at air/water and solid/water


interfaces [1-5], is of considerable significance in many areas including water treatment,
mineral processing, material recycling, etc. [4]. During froth flotation, hydrophobic

PT
particles are concentrated from their aqueous dispersions by attachment at the surface of
air bubbles, resulting in the separation of particles with different hydrophobicity [6, 7].
The selective attachment of mineral particles onto rising bubbles in water suspensions is

RI
the key to froth flotation [8]. However, the surfaces of most minerals are not sufficiently
hydrophobic for direct flotation. Therefore, a selective hydrophobization of the mineral
particles is usually needed, and accomplished, by applying an effective collector [7].

SC
According to their solubility, floated minerals may be divided generally into water-
insoluble and water-soluble minerals. In practical industrial processes and in research the
majority of flotation processes are concerned with water-insoluble minerals. However,

NU
the flotation of water-soluble minerals, such as potash, borax, trona, etc., is also of great
importance, because these minerals are essential for the production of fertilizers,
magnesium metal and oxide, soda ash, baking soda, boron chemicals, etc. [3, 9-14].
MA
Because water-soluble minerals dissolve, their flotation is more complicated than that
of water-insoluble minerals due to the hydration of salt ions, water-soluble mineral
surfaces, collector molecules (ions) and aggregates, collector adsorption at liquid/bubble
interfaces, collector association with water-soluble minerals, association of water-soluble
D

mineral particles with bubbles, etc. Although there have been other review papers about
TE

the flotation of water-soluble minerals, the majority of the published papers are only
concerned with a specific aspect, for example, the role of ion interactions [4], wetting and
spreading [5] on flotation, the effect of saline water [15] on mineral flotation, the
P

significance of froth stability [16] and rheology [17] in mineral flotation, some specific
phenomena and processes related to mineral flotation, including hydration [12, 18, 19],
CE

ion-specific effects [20-23], colloid [12] and surfactant [24] adsorption, etc. In this
review paper, an overview for the flotation of water-soluble minerals is presented, with
the purpose of further understanding flotation mechanisms and improving flotation
AC

operations.

2. General froth flotation processes

In the process of mineral flotation, a pulp of particles in water or in a water solution is


conditioned using a small amount of reagent(s) with air being purged to form bubbles.
Usually, a collector is added to render one mineral more hydrophobic by adsorbing or
attaching onto the mineral particle surfaces, affecting the bubble-particle interactions, and
hence, the attachment time. Sometimes a frother is used to improve the formation of
small bubbles. Upon collision, the more hydrophobic particles attach more easily to the
rising bubbles, which carry the particles to the surface [5, 8, 15]. The flotation rate
constant can be expressed as [15]:

k = SbEcoll/4 (1)

3
ACCEPTED MANUSCRIPT

where Sb and Ecoll are the bubble surface area flux and collection efficiency, respectively.
Ecoll embraces collision, attachment and detachment [15]:

Ecoll = EcEa(1 - Ed) (2)

PT
where Ec, Ea, and Ed are collision efficiency, attachment efficiency, and detachment
efficiency, respectively.
Particle-bubble attachment may be the most important factor determining flotation

RI
efficiency [5, 8]. Bubble surface structure and bubble stability are very important for
successful particle-bubble attachment [16]. As an indicator of the level of inter-particle
interaction, rheological behavior of mineral slurries also influences the flotation process

SC
[17].
For the flotation of KCl particles with an amine as collector, the amine is first melted
by heating up to 70-90oC, and is then neutralized with hydrochloric or acetic acid. The

NU
hot emulsion/dispersion is introduced into the flotation pulp at room temperature.
Because room temperature is much lower than the Krafft point of the utilized amine, a
white precipitate appears immediately when the hot amine emulsion is added to the
MA
potash flotation pulp [25]. The amine precipitates may attach to the KCl particle surfaces.
They may also deposit at the surfaces of the purged air bubbles, followed by spreading
into molecular films. The spreading is fast and efficient in the presence of a frother. The
existence of amine films on the bubble surfaces enables the bubbles to pick up the KCl
D

particles more easily [25].


TE

Due to the low solubility of most collectors in saturated brines, the collector
molecules or collector ions diffusion and subsequent adsorption at the salt mineral
surface does not easily occur. However, the organization and attachment of the collector
P

could be facilitated via air bubbles, which could be covered easily with a layer of the
collector. The flotation of water-soluble salt particles may include at least the following
CE

five processes with some of the processes taking place simultaneously [25]:
(1) Formation of collector colloids/crystals once the collector emulsion is added into the
water suspension.
AC

(2) Attachment and spreading of the collector at air bubble surfaces once air bubbles are
introduced into the suspension.
(3) Attachment of the collector to salt particle surfaces.
(4) Association of the salt particles with collector-coated air bubbles.
(5) Transfer of the salt particles to the surface by the air bubbles.

3. Characteristics of water-soluble mineral flotation

3.1. Flotation in saturated salt solutions

The flotation of water-soluble minerals is carried out in water suspensions usually


saturated with the water-soluble minerals [10, 25, 26]. Dissolved ions from the water-
soluble minerals can cause water structure changes, compression of the electrical double
layer, and inhibition of bubble coalescence [1, 15, 26, 27]. Such dissolved ions also can
change the properties of both air/liquid and solid/liquid interfaces, and therefore, affect

4
ACCEPTED MANUSCRIPT

the interactions between collectors, particles, and bubbles during flotation [26]. In
saturated solutions, such effects are more pronounced.
Due to ion hydration, water hydrogen bonding networks experience substantial
changes. Water in saturated brines should not be treated as a continuous phase. Instead,
some water molecules become part of the ion hydration shells. Ions with different water

PT
affinity have different influences on the water hydrogen bonding networks. Small ions
hold water molecules tightly, while large ions associate with water molecules loosely.
The effects of ion hydration on solution properties have been validated by viscosity

RI
measurements, spectroscopy analyses, and molecular dynamics simulations [12].
The electrical double layer around salt particles is highly compressed due to the high
ionic strength of the saturated brines [10, 15], and therefore, the electrostatic interactions

SC
between the collectors and the charged surfaces of the water-soluble minerals are
suppressed [9]. For dye adsorption onto resins, with increasing ionic strength, more Na+
and Cl- ions can screen the charged sites of the adsorbents, leading to a suppression of the

NU
electrostatic interactions (a reduction in both electrostatic attraction and electrostatic
repulsion) [28]. The compression of the electrical double layer in brines can promote
particle aggregation [15].
MA
It was found that above a certain concentration dependent on the particular electrolyte
chosen (typically ~ 0.1 M), many common electrolytes inhibit bubble coalescence at
moderate concentrations and others have very little effect. The coalescence behavior is
not determined by either the anions or cations but rather the pair of ions [22, 29-31].
D

Based on this consideration, electrolyte ions may be separated into four classes: anions,
anions, cations, and cations [22]. It was found that and assignments, initially
TE

developed empirically as part of combining rules to codify bubble coalescence in


electrolyte solutions, may describe the positioning of ions at the air-water interface. Ions
depleted from the interface are ions, while those accumulated at the interface are ions
P

[29]. According to the combining rules predicting bubble coalescence behavior, with an
CE

or combination, bubble coalescence is inhibited relative to pure water, and


unchanged when an or combination is employed. Up to now, no exceptions have
been found to these rules [22].
AC

The salt action does not stop bubble coalescence, but inhibits it. Coalescence still
occurs in the presence of an electrolyte, but less readily. That is, the electrolytes may not
change the equilibrium interaction forces, but instead, alter the dynamic processes
leading to film thinning and rupture [29]. It was also found that ions have specific effects
on foams. The foam stability increases with small amounts of added salts; the increase
may be quantified by the specific energy of counter-ion adsorption at the air/water
interface [20].
At a sufficiently high salt concentration, the surface tension gradient resulting from
the thinning process immobilizes the gas/liquid interface between coalescing bubbles. In
this case, the time required for coalescence dramatically increases [15]. The influence of
salts on bubble coalescence may also be related to the changes in the viscosity of the salt
solutions, which should be related to the hydration of the salts. It has been reported that
soluble salts in water suspensions can enhance the flotation of coal and some valuable
minerals [15, 32].

5
ACCEPTED MANUSCRIPT

3.2. Wide range of collector concentrations for effective flotation of different salts

For different water-soluble minerals, the collector concentrations for effective


flotation vary considerably. For dodecylamine hydrochloride (DAH) collector, the
concentration of the collector required to float KClO4, K2SO4 and KNO3 is substantially

PT
lower than that required for KCl flotation. The flotation order is
KClO4>K2SO4>KNO3>KCl [12]. KCl flotation commences when the collector
concentration exceeds the solubility limit. All flotation plants processing potash ores and

RI
all lab flotation experiments are carried out at temperatures much lower than the Krafft
temperature of the collector. The Krafft temperature is the temperature below which there
are no collector colloids formed and hydrated collector crystals exist. Instead of micelles,

SC
precipitates composed of hydrated crystals form in the brines [12, 25]. On the other hand,
flotation of KClO4, K2SO4 and KNO3 occurs prior to collector precipitation at DAH
concentrations as much as two orders of magnitude lower than that required for

NU
precipitation [12].

4. Hydration phenomena in flotation processes


MA
In the process of water-soluble salt flotation, substances in water suspensions tend to
form hydration shells or layers, which physically prevent direct contact among the
substances. Structural changes of hydration shells or layers perturb the energy states of
D

the substances and hence affect their properties in the suspensions. Since interactions
TE

among the substances are accompanied by structural changes of the water shells or layers,
it is important to investigate the hydration phenomena and the hydration changes when
hydrated substances approach each other [33]. Such hydration phenomena involve the
P

hydration of salt ions, water-soluble mineral surfaces, dissolved collectors, aggregates,


etc., which are schematically illustrated in Fig. 1, using the hydration phenomena in a
CE

direct KCl flotation system [14] as an example.


AC

6
ACCEPTED MANUSCRIPT

K+
Cl-
Cl-
-
Cl
Hydration of salt ions

Na+ Mg2+

PT
Cl- Cl-
Cl-

-NH3+
-NH3+ Hydration of collector aggregates

RI
Cl-
-NH3+
Cl-
Hydration of collector ions
-NH3+
-NH3+

SC
-NH3+ -
Cl
-NH3+
Cl-
Cl-

NU
Hydration of salt particle surfaces
NaCl

MA KCl

Fig.1. Schematic diagram for the hydration phenomena in a direct KCl flotation system.
D

4.1. Hydration of salt ions


TE

In water solutions, dissolved ions interact strongly with surrounding water molecules,
change the hydrogen bonding networks, and modify the solution properties in an ion-
specific manner, making the system more complex [23, 34, 35]. Water-ion interactions
P

create distorted H-bonded networks, inducing peculiar hydration phenomena [36].


CE

Comprehensive information on the structures of hydrated ions in aqueous solutions is


essential for understanding their roles in flotation processes. For pure water, the chemical
structures of water molecules and weak bonds caused by electrostatic forces and donor-
AC

acceptor interactions between hydrogen and oxygen atoms create favorable conditions for
the formation of intermolecular hydrogen bonds with an O-HO angle close to 180,
leading to the existence of complex intermolecular associations with a general formula of
(H2O)n, where n can vary from 3 to 50 [37, 38].
In salt solutions, water may not be treated as an unperturbed medium simply hosting
the dissolved salts. Due to their asymmetry, water molecules interact in different ways
with cations and anions [39]. It is suggested that, in general, there are two water shells
(layers) surrounding each ion: 1) an inner shell of immobilized water molecules with a
specific orientation; 2) an outer shell of less-ordered water molecules [15, 37]. The
electric field around the cations makes water molecules align in such a way that the
oxygen atoms of the water molecules are directed towards the cations. In the first
hydration shell (layer) of the cations, such an alignment does not favor the formation of
hydrogen bonds between water molecules, which are disrupted either by the steric
hindrance or by the dipole ordering. In the second hydration shell and beyond, water
molecules adjust their orientations to maximize their hydrogen bonding with the
influence of the cations diminishing [40].

7
ACCEPTED MANUSCRIPT

Ion charge densities may govern the interactions between the ions and water
molecules, and a balance of forces between electrostatic interactions (waters dipole
interacting with ions) and hydrogen bonding (water interacting with neighboring waters)
determines the water structures in the hydration layers. Small ions with a high charge
density cause strong electrostatic ordering of nearby waters. In contrast, large ions with a

PT
low charge density lead to surrounding water molecules being largely hydrogen bonded
[41-43]. According to their hydration behavior, salt ions can be chaotropic or
kosmotropic. The discrimination between both types comes from the relative strength of

RI
the ion-water interactions compared to water-water interactions [21]. Chaotropicity
correlates with a low charge density, and large single-charged ions tend to be chaotropic,
which have smaller effects on local hydrogen bonds. On the contrary, kosmotropicity

SC
correlates with a high charge density, and small or multiple-charged ions tend to be
kosmotropic, which interfere strongly with local hydrogen bonds [44]. Binding of
chaotropes to water molecules is weaker than that of water molecules to each other, while

NU
binding of kosmotropes to water molecules is stronger than that of water molecules to
each other [36, 44].
Using the techniques of neutron diffraction with hydrogen isotope substitution, it
MA
was shown that, for the hydration of KF, KCl, KBr, and KI, in the first hydration shells of
both anions and cations, the water structures are strongly affected, with the largest effects
occurring with the highest salt concentrations and the smallest anion. However, outside
the first hydration shells, the water structures are only mildly perturbed [45]. Gas phase
D

infrared vibrational predissociation spectroscopy, x-ray absorption spectroscopy, solution


TE

neutron diffraction, etc also indicate that water molecules are affected over only a short
distance from simple, small ions. And compared to the strength of water-water
interactions, the long range electric fields generated by simple ions must be weak [46].
P

The orientational relaxation time for the water molecules outside the first hydration shell
may be the same as in the bulk solutions [45]. The lifetime of a water molecule in the
CE

hydration shell of a kosmotropic ion such as Li+ or Ca2+ is long. Consequently the
hydration shell may be said to move together with a kosmotropic ion, while a chaotropic
ion may be modeled independently of its hydration shell [47].
AC

Although hydration numbers of ions are operationally defined and values obtained by
different methods are not in good agreement [48], some general trends are relatively clear
[37, 49, 50]. Studies on alkali metal ion hydration using large angle X-ray scattering
(LAXS) and double difference infrared spectroscopy (DDIR) show that Na+, K, Rb+, and
Cs+ ions are all weakly hydrated with only a single water shell. Li+ ions are more strongly
hydrated, probably having a second hydration shell. The influence of Rb+ and Cs+ ions on
water structures is suggested to be very weak, and it is difficult to quantify this effect in a
reliable way [37]. In chloride solutions, Li+, Na+, K+, and Rb+ ions are reported to have a
hydration number of 4.8(3), 5.1(1), 5.3(6), and 6.9(4), respectively [50]. Li+ ions
retain a well-defined hydration shell with a hydration number of 4.8(3), which is almost
independent of ion concentration [50]. There is relatively strong correlation between M-O
bond distances and coordination numbers for the alkali metal ions even though the M-O
interactions are weak and the number of complexes of potassium, rubidium and cesium
with well-defined coordination geometry is very small [37].
Generally speaking, anions have a larger radius and a lower charge density than
cations, and therefore, they have a weaker degree of hydration than cations [18], leading
8
ACCEPTED MANUSCRIPT

to a stronger tendency for anions to accumulate at solution/air interfaces than that of


cations [51]. Models based solely on electrostatic interactions cannot explain ion specific
properties of electrolyte solutions [52]. The interaction of water with anions can be
approximated by a linear hydrogen bond, suggesting that the dominant forces on ions in
water may be short range forces of a chemical nature. This chemical interaction may

PT
involve substantial charge transfer to solvent for strongly hydrated ions, resulting in
delocalization of charge [46]. It has been reported that dispersion interactions may not be
ignored and may play a major part in ion specificity [52]. During hydration of simple ions,
such as Li+, Na+, K+, F-, Cl-, Br-, the interactions between the ions and water molecules

RI
may include electrostatic interactions, covalent bonding, and dispersion interactions. For
different ions, the relative importance of the three kinds of interactions is different.

SC
In water solutions, the oppositely charged ions may associate to form unique entities,
ion pairs. The general association reactions of two ionic solutes may be represented as
[37]:

NU
M+(ion) + X(ion) M X(SSIP) MX(CIP) MX(salt) (3)

The reactions involve two intermediates, the contact ion pair (CIP) and the solvent
MA
separated ion pair (SSIP). No water molecules intervene between the ions in CIP, while
the two ions are held apart by water molecules in SSIP. All electrolytes in dilute aqueous
solutions have a general trend to associate or even form stable neutral ion pairs. Since
cations and anions in solution are generally attracted through electrostatic attractions, the
D

charge magnitude and radius of ions are important for the ion association. On the other
TE

hand, entropy generally opposes the ion association. The stable structures of NaCl(H2O)n
(n 10) clusters have been well characterized at various levels of quantum theories [53].
It was reported that ion pairs are less strongly hydrated than free ions [54].
P

Contact ion pair formation is actually dominated by hydration-dehydration.


Oppositely charged ions with equal water affinity tend to come together in solution to
CE

form contact ion pairs, whereas oppositely charged ions with differing water affinities
tend to stay apart [46]. Two strongly hydrated small ions of opposite charge may
experience a very strong reciprocal attraction. Consequently, they can come together
AC

forming direct ion pairs and expelling or sharing the hydration spheres between them. In
the case of weakly hydrated soft ions, although the electrostatic attraction between them
is much smaller than that between kosmotropes, the hydration shells are so loosely bound
that the chaotropic ions can also form direct ion pairs by expelling the hydration water.
The interaction between a hard and an oppositely charged soft ion is different. The
attraction of the soft ion is weak, and the hard ion will keep its hydration shell. Therefore,
a soft/hard ion pair is always separated by water and cannot form strong ion pairs [21].
For example, it was found that LiCl and CsI form direct ion pairs at the air solution
surface [55].

4.2. Hydration of water-soluble mineral surfaces

The behavior of water molecules near solid/liquid interfaces is very important in


many processes. Water molecules within a few nanometers of a surface behave
differently from those in the bulk liquid, with the water molecules ordering into discrete
layers known as hydration layers [19, 34, 56]. Similar to ionic hydration, the strongly
9
ACCEPTED MANUSCRIPT

hydrated surfaces are hydrophilic. On the contrary, weakly hydrated surfaces are
hydrophobic. That is, surface hydration layers are added next to hydrophilic surfaces,
while no hydration layers are added to hydrophobic surfaces. The surface hydration
layers may prevent certain ions and molecules from penetrating. Strongly hydrated
cosmotropic ions or molecules may penetrate into the surface hydration layers, while

PT
weakly hydrated chaotropic ions or molecules are excluded from the surface layers [47].
The flotation of soluble salts is affected significantly by interfacial water structures
[3]. Compared to structure breaking salts such as KCl, RbCl, and NaHCO3, structure

RI
making salts, including LiCl, NaCl, Na2CO3, MgCl26H2O, KMgCl36H2O, etc,
composed of ions with stronger affinity to water molecules, tend to create stronger
surface hydration and a more hydrophilic surface, as revealed by both contact angle

SC
measurements and bubble attachment time measurements [12]. Recent studies show that
for water molecules near charged solid/liquid interfaces, in addition to a diffuse region
where water molecules do not show specific orientations, there is a compact interfacial

NU
region of water molecules next to the solid surfaces which are not described in classical
electric double layer theories [57].
For simple salts, because the hydration of cations is much stronger than that of anions,
MA
in general, the hydration of the surfaces of simple salt particles may be dominated by the
hydration of the cations.

4.3. Hydration of collector molecules or ions


D
TE

Solutes resisting wetting by water or being dissolved therein are generally regarded as
being hydrophobic [37, 58]. Hydrophobic interactions are very important for protein
folding and assembly, receptor-ligand binding, membrane formation, and cellular
P

transportation [59, 60]. They are also very important for some separation processes
including flotation. The introduction of a hydrophobic solute into water causes an
CE

increase in the free energy of the system, as a result of the structural modifications of the
water molecules around the solute molecules or ions. When two solute molecules or ions
are close enough to each other, the water molecules thus affected by the two solute
AC

molecules or ions are smaller than when they are far apart, resulting in an effective,
water-mediated hydrophobic attraction between the two solute molecules or ions.
Hydrogen bonding between neighboring water molecules is the key element of the
hydrophobic effect, which plays a crucial role in many physical, chemical, and biological
processes [58].
When inserted into liquid water, large hydrophobic solutes break some hydrogen
bonds in their immediate vicinity, resulting in large positive enthalpy of hydration and
hence in a free energy change proportional to the solute surface area. The hydration of
large hydrophobic solutes is expected to be enthalpically driven. Fewer water hydrogen
bonds have to be broken when two large hydrophobic solutes are in contact with their
hydrophobic tails than when they are far from each other, so there is a negative enthalpy
change when such hydrophobic solutes approach each other. The free energy change,
which is dominated by the enthalpy change, will be negative and there will be a
thermodynamic driving force for their attraction [58]. Collectors used in the flotation of
water-soluble minerals are usually surfactants which have a hydrophilic head and a
hydrophobic tail. The hydration state of the whole collector molecule or ion is the
10
ACCEPTED MANUSCRIPT

addition of the hydrophilic hydration of its hydrophilic head and the hydrophobic
hydration of its hydrophobic tail.
On the macroscale, hydrophobic interactions promote the aggregation of oil-like
objects in water by minimizing the interfacial energy [59, 60]. However, the hydration of
small hydrophobic molecules (1 nm) may differ from that of macroscopic objects.

PT
There are intricate relationships between the molecular hydration mechanisms and the
length scale of the molecules. The crossover length scale of microscopic to macroscopic
molecules, critically important to proteins, polymers, etc., was reported to be

RI
approximately 1 nm [60]. Although the extended polymers and proteins span hundreds of
nanometers, their hydration behavior is determined by the size of a single hydrophobic
monomer. As the hydrophobic particle size decreases from the macroscopic to the

SC
microscopic regime, the scaling relationship changes from a dependence on interfacial
area to a dependence on volume [60].

NU
4.4. Hydration of collector aggregates

As discussed in Section 3.2, instead of collector micelles, precipitates composed of


MA
collector solid particles and hydrated crystals form in the brines in the flotation processes
of some water-soluble minerals. Precipitation of surfactants occurs when the product of
the thermodynamic activity of the precipitating surfactant ions and their counter ions
satisfy the solubility product equation below the critical micelle concentration (CMC).
D

However, the structures of the surfactant solid particles and hydrated crystals are not
TE

clear. Therefore, the hydration of surfactant micelles is discussed in this section.


Micelles are self-organized molecular assemblies of amphiphilic molecules that
comprise a hydrophobic core and a hydrophilic shell. The core of a micelle is formed by
P

hydrocarbon chains (hydrophobic tails) with polar and/or charged headgroups


(hydrophilic heads) projecting outward into the bulk solution. Depending on the nature of
CE

the surfactant headgroups, micelles can be ionic or nonionic. The hydrophilic shell of the
micelles can form the Stern layer and the palisade layer when the micelles are ionic and
nonionic, respectively. In ionic micelles, the Stern layer is the interface between the
AC

hydrophobic core surface and the hydrodynamic shear surface, consisting of the ionic
headgroups, a few counter ions, and water molecules associated with these ionic
headgroups. In nonionic micelles, the palisade layer defines the interfacial region
between the hydrophobic core surface and the shear surface comprising uncharged
hydrophilic headgroups and water molecules which are hydrogen-bonded to these
uncharged hydrophilic headgroups [61].
The micelle properties can be affected by surfactant concentration, salt, counterions,
and temperature [62-64]. For the nonionic surfactant, Triton X-100, the aggregation
number and radius of its micelles increase with increasing ionic strength, with the
micelles becoming larger and more diffuse in structure [64]. Changing bromide to
chloride for dodecyltrimethylammonium (DTA+) micelles leads to a small decrease in
aggregation number, and an increase in both CMC and counterion dissociation degree
[63]. For ionic surfactant micelles, headgroup/counterion association reduces the micelle
hydration [54].
The hydration of the micelles of anionic surfactants with -OSO3- headgroups,
nonionic surfactants with poly(ethylene oxide) chains, and cationic surfactants with -
11
ACCEPTED MANUSCRIPT

N(CH3)3+ headgroups was found to be in the order of anionic > nonionic > cationic. The
hydration of the micelles of cationic surfactants with -N(CH3)3+ headgroups are the
weakest, because the hydrophobic methyl groups inhibit the approach of water molecules
close to the headgroup region. The palisade layer of the nonionic micelles with
poly(ethylene oxide) chains as the hydrophilic moieties, can bind more water molecules
than the positively charged -N(CH3)3+ headgroups of the cationic micelles, but less water

PT
molecules than the negatively charged -OSO3- headgroups of the anionic micelles [61].
Due to hydrophobic interactions, the collector aggregates in water may arrange in

RI
such a way that the hydrophobic chains associate inside the aggregates with hydrophilic
head groups being outside to minimize the energy of the system. Because the head groups
of the collectors are usually charged or have high polarity, the surfaces of the collector

SC
aggregates would be strongly hydrated. The hydration of the collector aggregate surfaces
may be stronger than that of some salt particle surfaces.

NU
5. Collector adsorption at liquid/bubble interfaces

The driving force for collector adsorption at liquid/bubble interfaces might be the
MA
hydrophobic effect: The solvation of collector ions or molecules corresponding to the
creation of a cavity perturbs the hydrogen bonding networks of water molecules and costs
free energy. At liquid/bubble interfaces the cost in free energy is drastically reduced [55].
The adsorption of collector molecules or ions at liquid/bubble interfaces is schematically
D

illustrated in Fig.2. For the sake of clarity, when the collector is an ionic surfactant, the
TE

counterions are not shown. With the adsorption of the collector, the hydration layers of
the collector hydrophilic heads may rearrange to form new layers around the outer
surface of the hydrophilic heads of the adsorbed collector.
P

Generally speaking, the addition of inorganic salts might affect the adsorption of
collectors at liquid/bubble interfaces in two possible ways: 1) The collector-collector
CE

repulsion lateral to the interfaces is screened, leading to an increase in adsorbed collectors.


2) More ions adsorb in the interfacial regions leading to reduced or increased surface
charge density [55]. For nonionic collectors there is only a small increase in adsorption
AC

upon salt addition. An increase in temperature results in an increase in the area per
molecule or ion of the collector, probably due to increased thermal motion, with a
consequent decrease in adsorption effectiveness [65].

12
ACCEPTED MANUSCRIPT

Hydration layer of the collector hydrophilic head

Collector hydrophilic head

PT
Collector hydrophobic tail

Bubble

RI
SC
NU
Fig.2. Schematic diagrams for the adsorption of collector molecules or ions at
MA
liquid/bubble interfaces. From left to right, the collector concentration increases.

6. Collector association with water-soluble minerals


D

6.1. Association processes with hydration changes


TE

The association of collectors with water-soluble minerals is a complicated process


which may include the adsorption of collector molecules or ions onto the mineral
surfaces, the attachment of collector precipitates or crystals onto the mineral surfaces,
P

and/or the formation of collector hemi-micelles or solloids on the mineral surfaces


CE

[65-70].
Regardless of collector species, effective flotation of KCl commences when
collector precipitation occurs, suggesting that colloid adsorption may be involved in KCl
AC

flotation, because the crystallization energy of the collector in the brine is of the same
order of magnitude as the adsorption energy [12]. Based on optical microscopy
observations, the adsorption of octadecylamine (ODA) only happens at KCl particle
edges and corners with a higher crystal defect density. The ODA colloid/precipitate
attachment is achieved through the replacement of hydration water molecules at the
negatively charged defect sites (due to the association of Cl- anions or dissolution of K+
cations) with ODA head groups [71].
During the association of collectors with water-soluble minerals, the original
hydration states of the collectors and the mineral surfaces will change. When a hydrated
head group of the collector transfers from the bulk suspension to the liquid/mineral
interfacial region, partial removal of water from the hydration shell around the collector
head group may occur [24]. In order to adsorb to the water-soluble mineral surface, an
incoming collector may need to displace hydration water at the water-soluble mineral
surface [72].

13
ACCEPTED MANUSCRIPT

6.2. Factors affecting association

Dissolved salt ions, suspension pH and temperature, and the chain length of the
collector surfactant hydrophobic tails can affect the association between collectors and
water-soluble minerals.

PT
At high salt concentration, the hydrophilicity of the surfactant head group decreases.
Salt addition causes a decrease in the adsorption of ionic surfactants onto an oppositely
charged adsorbent and an increase in their adsorption onto a similarly charged adsorbent.

RI
These effects are presumably due to the decreased attraction between oppositely charged
species and the decreased repulsion between similarly charged species at a higher ionic
strength. Both the efficiency and effectiveness of adsorption of ionic surfactants onto

SC
similarly charged substrates are increased by an increase in the ionic strength of the
aqueous phase [73].
The pH value usually has an obvious effect on the adsorption of ionic surfactants onto

NU
charged solid substrates. With a decrease in pH value, a solid surface will usually become
more positive, or less negative, due to adsorption of protons onto the substrates, resulting
in an increase in the adsorption of anionic surfactants and a decrease in the adsorption of
MA
cationic surfactants. The reverse is true when the pH value is increased [65]. A change in
pH may affect surfactants, especially those containing carboxylate groups or non-
quaternary ammonium groups. A change in pH also may cause the conversion of
surfactants between their neutral and their ionic species, and affect the electrostatic
D

interactions between the surfactants and the solid substrates. Nonionic surfactants with
TE

polyoxyethylene chains may also be affected by pH change, because the ether groups can
be protonated at low pH, becoming positively charged. Also, positively charged
surfactants may be adsorbed by negatively charged substrates [65]. While KCl floats with
P

amines at pH < 10.5, the flotation of NaCl takes place only when this pH is exceeded.
The negative electrical charge of sylvite particles and the positive charge of the amine
CE

precipitate result in Coulombic attraction which then leads to potash ore flotation [25].
Also because of the acid-base equilibrium of amine, K2SO4 and K2SO4MgSO46H2O
flotation with DAH decreases above pH 10. On the contrary, pH has no significant effect
AC

on SDS flotation due to the fact that the SDS collector is the salt of a relatively strong
acid [12].
An increase in temperature causes decreased hydration of hydrophilic groups,
favoring micellization; disruption of the structures of water molecules surrounding
hydrophobic groups, disfavoring micellization [65]; a decrease in both the efficiency and
effectiveness of ionic surfactant adsorption, a small change compared to that caused by a
change in pH; and an increase in the adsorption of nonionic surfactants with a
polyoxyethylene (POE) chain as the hydrophilic group, attributed to a decrease in
dehydration of the POE group with increasing temperature [65].
Surfactant hydrocarbon chain length plays a major role in determining surfactant
properties. For example, the critical micelle concentration decreases with an increase in
hydrocarbon chain length. There is a good relationship between surfactant adsorption
effectiveness and the efficiency of the surfactant as a collector [74]. Investigations on the
influence of hydrocarbon chain length on alkylmorpholine flotation has shown an
improvement in halite floatability with an increase in hydrocarbon chain length, the
highest halite recovery being obtained using hexadecyl-morpholine at 18-20oC [75].
14
ACCEPTED MANUSCRIPT

7. Interactions between collectors and water-soluble minerals

Adsorption of surfactants onto mineral surfaces is governed by a number of forces.


The net driving force, Gads, was reported to be the sum of a number of contributing

PT
forces [24, 70]:

Gads = Gelec + Gchem + Gc-c + Gc-s + GH + Gw + ... (4)

RI
where Gelec is the term for electrostatic interactions, Gchem is the term for chemical
reaction due to covalent bonding, Gc-c is the term for lateral interactions due to the

SC
cohesive chain-chain interactions among the hydrophobic interactions between the
hydrophobic chains of the previously adsorbed surfactants and the hydrophobic chains of
the oncoming surfactants, Gc-s is the term for the hydrophobic interactions between the

NU
surfactants and the solid minerals, GH is the term for hydrogen bonding, and Gw is
the term for hydration changes of the surfactants and the mineral surfaces during
adsorption.
MA
Adsorption of collector surfactants onto water-soluble mineral surfaces is very
important for the successful flotation of the minerals. To more easily understand the
mechanisms for the adsorption of collector surfactants onto water-soluble mineral
surfaces, some of the terms in Equation (4) have been combined; the driving forces for
D

the adsorption include electrostatic interactions, hydrophobic interactions, hydrogen


TE

bonding, specific interactions, etc.:

Gads = Gelec + Ghydrophob + GH + Gspecific + ... (5)


P

where Gelec is also the term for electrostatic interactions. Ghydrophob is the term for the
CE

total hydrophobic interactions including Gc-c, Gc-s , and Gw in Equation (4). Gw


is included because when hydrophobic interactions take effect, there must be hydration
changes. Gspecific is the term for specific interactions including precipitation reactions,
AC

complexation reactions, etc.

7.1. Electrostatic interactions

Solid particles in aqueous suspensions may develop surface charges due to surface
group ionization, ion preferential dissolution, ion adsorption, isomorphous substitution,
etc. Charged surfaces will attract counterions, forming a layer immediately adjacent to
the surfaces, as well as a second, more diffuse, layer of counterions. These two layers are
usually referred to as the electrical double layer [76]. In systems in which there are ionic
surfactants and charged solid particles, electrostatic interactions usually play an important
or even a governing role in the adsorption process [24]. Electrostatic interactions may
take place in the form of ion exchange. Ion exchange involves the displacement of
similarly charged ionic species from the surfaces of the minerals. The physicochemical
nature of the ion exchange reaction is basically controlled by electrostatic interactions.
In solutions with high ionic strength, the electrical double layers around salt particles
are highly compressed, and the electrostatic interactions between the collectors and the
15
ACCEPTED MANUSCRIPT

charged surfaces of the water-soluble minerals are suppressed [9, 10]. For the adsorption
of organic dyes onto resins, with increasing ionic strength, more Na+ and Cl- ions can
screen the charged sites of the adsorbents, leading to a suppression of the electrostatic
interactions (a reduction in both electrostatic attraction and electrostatic repulsion) [28].
Preliminary results from AFM force measurements indicate that, while short-range

PT
repulsive forces are present between similarly charged hydrophilic surfaces, relatively
long-range attractive forces exist between oppositely charged hydrophilic surfaces at high
ionic strength. These results suggest that although electrostatic interactions are

RI
suppressed by high ionic strength, surface charge plays an important role in influencing
particle interactions even at high ionic strength [13, 77, 78]. In flotation systems there
may be oppositely charged particles, and the electrostatic interactions may cause

SC
coagulation of the oppositely charged particles. For example, both fine sylvite and halite
particles are stable in their own brine, however, when mixed together, they immediately
coagulate. While sylvite particles are negatively charged, halite particles are positively

NU
charged. The coagulation is caused by the electrostatic attraction between the sylvite and
halite particles [25]. MA
7.2. Hydrophobic interactions

Hydrophobic interactions are the driving forces for self-assembly of amphiphilic


compounds into micelles, liquid crystalline phases, membranes, etc, and they are usually
D

an important motive for the adsorption of surfactants at mineral surfaces. The


TE

hydrophobic interactions helpful for surfactant adsorption may include: 1) the direct
hydrophobic interactions between surfactant hydrophobic groups and the hydrophobic (or
less hydrophilic) mineral surfaces; 2) the hydrophobic interactions between the
P

hydrophobic groups of the previously adsorbed surfactants and the hydrophobic groups
of oncoming surfactants. When the hydrophobic interactions take effect, some related
CE

hydration changes may take place. The hydrophobic interactions generally increase with
an increase in the carbon chain length of the surfactants [65]. The hydrophobic
interactions between surfactant alkyl chains and the hydrophobic sites on mineral
AC

surfaces would be a significant factor for the adsorption of surfactants on fully or


partially hydrophobic surfaces [24].
When surfactant concentration reaches a threshold value, analogous to aggregation
in the bulks, surfactant molecules tend to form aggregates at the solid/liquid interfaces,
leading to an obvious increase in adsorption. These aggregates are called hemi-micelles
or solloids. In this case, the driving force for adsorption (Gc-c) results from the free
energy of transferring the hydrocarbon chains from the aqueous suspension into the
hydrophobic interior of the aggregates. Gc-c can be represented as [24]:

Gc-c = - n(CH2)/RT (6)

where n(CH2) is the number of CH2 groups in the hydrocarbon chain and is the energy
of transfer for each -CH2 group. Hemi-micelles are more tightly packed than micelles,
with the hemi-micelles forming at concentrations below the corresponding critical
micellar concentrations [24].

16
ACCEPTED MANUSCRIPT

For the adsorption of ionic surfactants onto mineral surfaces with an opposite charge,
electrostatic and hydrophobic interactions could have a synergistic effect. At first, some
surfactant molecules may adsorb onto the mineral surfaces mainly through electrostatic
attraction, and then more surfactant molecules may adsorb through the hydrophobic
interactions between the hydrophobic groups of the previously adsorbed surfactants and

PT
the hydrophobic groups of oncoming surfactants.
Interfacial water structures affect the hydrophobic interactions between surfactants
and mineral surfaces, and may play a significant role in surfactant adsorption. During the

RI
flotation of soluble minerals, if the salt is a water structure maker, such as LiCl and NaCl,
the water molecules will be strongly bonded at the salt surfaces due to strong hydration of
the salt surfaces, and such water molecules are difficult to be replaced by collector

SC
molecules or ions. Therefore, the adsorption of collector molecules or ions at hydrated
surfaces is difficult due to the lack of effective hydrophobic attraction, and the flotation
of the salts with either cationic or anionic C12 collectors is not possible. Unlike the

NU
structure maker salts, the structure breaker salts such as KCl and RbCl have a tendency to
disrupt the water structures at the salt surfaces, thus collectors may reach the salt surfaces
and effective hydrophobic attraction takes place, allowing for the flotation of the salts [2,
MA
10, 26, 79, 80]. The experimental results using NaHCO3 and Na2CO3 are consistent with
the above statements. The surface tension of the carbonate salts significantly increased as
the concentration of Na2CO3 increased. On the other hand, the increase in the surface
tension of NaHCO3 solutions was not significant. Na2CO3 has a very strong water
D

structure making effect, while NaHCO3 shows a weak water structure making effect. The
TE

hydrophobic interactions between NaHCO3 and the collectors are stronger than those
between Na2CO3 and the collectors. NaHCO3 can be floated with both 12C anionic and
cationic collectors, whereas Na2CO3 does not float with either collector [1, 3].
P

MgCl26H2O and KMgCl36H2O have a structure making effect on water, resulting in


poor flotation with both cationic and anionic collectors. On the other hand, KCl has a
CE

water structure breaking effect and consequently collector adsorption is possible [10].
Structure breaking salts composed of ions with less affinity to water molecules tend to
create a more hydrophobic surface as revealed by both contact angle measurements and
AC

bubble attachment time measurements [12]. Compared to the structure making salts
including LiCl, NaCl, Na2CO3, MgCl26H2O, KMgCl36H2O, etc, structure breaking salts
such as KCl, RbCl, and NaHCO3 are more hydrophobic. Therefore, the hydrophobic
interactions between the collectors and the structure breaking salts are stronger than those
between the collectors and the structure making salts.
In saturated salt solutions, the number of water molecules is usually deficient to
form complete hydration structures around each ion. The hydrophilicity of the surfactant
head groups decreases with increasing ionic strength [73]. The hydrophobic interactions
between the collectors and the salt particles are enhanced by the high ionic strength. For
the adsorption of organic dyes onto resins, the hydrophobic attractions between the dyes
and the resins were also found to increase with increasing ionic strength [28]. For the
adsorption of carboxymethylcellulose onto molybdenite, the ionic strength was also seen
to affect the adsorbed layer properties, with higher ionic strength yielding thicker
adsorbed layers and better coverage [6].

17
ACCEPTED MANUSCRIPT

7.3. Hydrogen bonding

In aqueous flotation systems, there may be three kinds of hydrogen bonds (including
ionic hydrogen bonds [81]): 1) the hydrogen bonds between collectors and water; 2) the
hydrogen bonds between water and mineral surfaces; 3) the hydrogen bonds between

PT
collectors and mineral surfaces, which are helpful for the flotation. Collectors can form
hydrogen bonds with both water and mineral surfaces. Unfortunately, water has a
stronger ability to form hydrogen bonds. The formation of hydrogen bonds between

RI
collectors and mineral surfaces must challenge the strong competition from water itself
[24]. The hydrogen bonds between collectors and minerals, helpful for flotation, could
include: 1) direct hydrogen bonds between collector functional groups and the surfaces of

SC
the water-soluble minerals; 2) indirect hydrogen bonds between collector functional
groups and the water films or layers on mineral surfaces.
Adsorption of a nonionic surfactant such as ethoxylated alcohol and sugar-based

NU
alkyl glucoside on oxides has been proposed to involve direct hydrogen bonding [24].
POE surfactants may adsorb onto silica surfaces via direct hydrogen bonding between
SiOH groups on the silica surfaces and the oxyethylene groups of the POE surfactants
MA
[65]. For the adsorption of boric acid/borate onto composite magnetic particles derived
from Fe3O4 and bis(trimethoxysilylpropyl)amine (TSPA), hydrogen bonds (including
ionic hydrogen bonds) were suggested to play a role in the adsorption [81, 82]. For KCl
and NaCl, it seems that there is no direct hydrogen bonding between KCl or NaCl and the
D

collectors, or else the direct hydrogen bonding is very weak because the ability of Cl to
TE

form hydrogen bonds is much weaker than that of O, N, and F. There may be direct ionic
hydrogen bonds between primary, secondary and tertiary amines and SO42-, CO32-, or
HCO3- groups of the minerals.
P

For indirect hydrogen bonds, it was reported that alkylmorpholine fixation to halite
can occur via hydrogen bonding between the oxygen atoms of morpholine and the
CE

hydrated sodium ions at the halite surfaces [75].


Hydrogen bonding is thought to be relatively weak in comparison with electrostatic
and chemical bonding [70]. The effective formation of hydrogen bonds between
AC

collectors and mineral surfaces, the effect of ionic strength on hydrogen bonding, and the
role of hydrogen bonding on adsorption and flotation is not so clear.

7.4. Specific interactions

During flotation, there may be specific interactions (precipitation, complexation, etc.)


between collectors and minerals [24]. However, for the flotation of water-soluble
minerals, if such interactions exist, they will be hindered by the dissolved species of the
water-soluble minerals, because the dissolved species may also form precipitates and
complexes with the collectors in the water suspensions, instead of on the surfaces of the
water-soluble minerals.
Another specific interaction is an interaction which happens in a process similar to
molecular or ionic imprinting, which is a technique to create template-shaped cavities in
matrices with memory of the template molecules or ions to be used in molecular or ionic
recognition [83, 84]. The selective adsorption of amine at KCl particle surfaces is a result
of the replacement of surface lattice K+ ions with RNH3+ ions. In contrast, RNH3+ ions
18
ACCEPTED MANUSCRIPT

cannot replace surface lattice Na+ ions of NaCl crystals due to the size of Na+, which is
smaller than that of the RNH3+ head group [12, 25, 71]. The replacement of K+ ions with
RNH3+ ions is similar to the molecular or ionic imprinting process. The ionic radii of Na+
and K+ ions are 1.0 and 1.4 , respectively [85], the mean gyration radius of the R-
NH3+ head is 1.34 [86]. Therefore, the RNH3+ head group fits in the KCl crystal but
not in the NaCl crystal. After the replacement of K+ ions with RNH3+ ions, the RNH3+

PT
ions should lose their hydration water.
Unlike biomicromolecules, most collectors used in flotation do not have as many

RI
groups which can form hydrogen bonds. Therefore, hydrogen bonding may not be
important for the interactions between collectors and the water-soluble mineral particles.
Specific interactions (precipitation, complexation, etc.) need very special conditions. So,

SC
electrostatic and hydrophobic interactions would be the most common interactions. And
their relative importance is different depending on the specific system.

NU
8. Association of water-soluble mineral particles with bubbles

Association between solid particles and air bubbles in aqueous suspensions is a key to
MA
understanding froth flotation. The association may include three processes: 1) collision
between air bubbles and particles, 2) attachment of the particles to the bubbles, and 3)
detachment of the particles from the bubbles [87, 88]. Interception was reported to be the
dominant collision mechanism, with gravity and inertia gradually gaining importance as
D

the particle size increases [89]. The attachment of particles to air bubbles is considered to
TE

be the key to successful flotation since this attachment determines the selective separation
between hydrophobic and hydrophilic particles [87, 90, 91]. The particle-bubble
interactions are composed of DLVO and non-DLVO interactions, and a hydrodynamic
P

component [90]. DLVO and hydrodynamic forces prevent a particle and a bubble from
approaching each other, however the presence of hydrophobic surfaces results in
CE

hydrophobic attraction and strong aggregate formation [91].


Particle-bubble attachment may occur when a bubble and a particle closely approach
each other. A wetting thin film between the solid/liquid and air/liquid interfaces is formed
AC

and surface forces govern further film stability. If the net of the surface forces is
attractive, the film becomes unstable followed by rupturing, leading to the formation of a
three-phase contact line and the attachment of the particle to the bubble. The bubble-
particle contact line spreads further across the solid surface to form a stable wetting
perimeter [8]. During the above processes, three steps are important to the attachment: 1)
thinning of the intervening liquid film between the particle and the bubble, 2) rupture of
the film to give a three phase contact nucleus, and 3) expansion of the three-phase
contact line from the nucleus to form a stable wetting perimeter [92, 93].
Since the solubility of long-chain amines in brines is low, the mechanism responsible
for flotation in this particular case may be different from conventional flotation. It was
postulated that the collectors in the potash ore flotation pulp are transported by bubbles
and the adsorption of the collectors onto the bubble surfaces is the precondition for the
association of the water-soluble mineral particles with the bubbles [25].
Hydrophobicity of mineral particle surfaces, surface charge of mineral particles and
bubbles, mineral particle size and shape, temperature, bubble size, etc. can affect the
association between solid particles and air bubbles. The association of water-soluble
19
ACCEPTED MANUSCRIPT

mineral particles with bubbles is affected by both thermodynamic and kinetic factors.
More attention should be paid to the kinetic factors.
It seems that only hydrophobic particles tend to attach to bubbles [15]. When the
collector concentration exceeds its critical micelle concentration (CMC), a bilayer of
collector molecules at particle surfaces is formed, with the outer surface being

PT
hydrophilic, preventing mineral particles from contacting with bubbles. Therefore, the
detected attachment time sharply increases after the CMC and the flotation recovery
decreases [8]. For the quartz-dodecyl ammonium hydrochloride (DAH) system, the

RI
attachment time for the quartz particles decreases with increasing DAH concentration up
to CMC (about 1.510-2 M), and then the attachment time increases dramatically [8].
Bubble attachment time measurements suggest that both the cationic collector, ODA, and

SC
the anionic collector, SDS, adsorb at the KCl surface to create a hydrophobic state. But
neither collector adsorbs at the MgCl26H2O and KMgCl36H2O surfaces [10].
Surface charges of mineral particles and bubbles obviously can also affect the

NU
attachment time. When quartz particles and air bubbles are stabilized in poly oxyethylene
methyl ether (POEME) solution at pH 9.8, the surfaces of both the quartz particles and
the bubbles have the same charge, causing a long attachment time and thus a flotation
MA
depression. However, in the POEME solution at pH 4.5, the quartz particles and the
bubbles are oppositely charged. The attraction between the bubbles and the mineral
particles reduces the attachment time significantly and promotes flotation [8].
Attachment time was found to increase significantly with increasing particle size [8].
D

Particle shape was also reported to be important to the induction time. Angular particles
TE

exhibit an induction time an order of magnitude lower than that of spherical particles [94].
Attachment time decreases considerably with increasing temperature. The smaller the
bubble size, the less the attachment time, which indicates that small bubbles are
P

beneficial to flotation [8].


CE

9. Conclusions and prospects

9.1. General conclusions and prospects


AC

The main physicochemical aspects of water-soluble mineral flotation including


hydration phenomena, associations and interactions between collectors, air bubbles, and
water-soluble mineral particles have been comprehensively discussed. Some general
conclusions and prospects are summarized as follows:
(1) Flotation carried out in saturated salt solutions, and a wide range of collector
concentrations for effective flotation of different salts are two basic aspects of water-
soluble mineral flotation. Dissolved ions from the water-soluble minerals can cause water
structure changes, compression of the electrical double layer, and inhibition of bubble
coalescence. The wide range of collector concentrations for effective flotation results in
complicated associations between collectors, mineral particles, and bubbles due to the
complex physicochemical species of the collectors existing in the suspensions.
(2) Hydration of salt ions, mineral particle surfaces, collector molecules or ions, and
collector aggregates plays an important role in water-soluble mineral flotation. Small ions
with a high charge density cause strong electrostatic ordering of nearby waters, breaking
hydrogen bonds. In contrast, large ions with a low charge density lead to surrounding
20
ACCEPTED MANUSCRIPT

water molecules becoming primarily hydrogen bonded. Compared to structure breaking


salts such as KCl, RbCl, and NaHCO3, structure making salts, including LiCl, NaCl,
Na2CO3, MgCl26H2O, KMgCl36H2O, etc, composed of ions with stronger affinity to
water molecules, tend to have stronger surface hydration and create a more hydrophilic
surface. The hydration state of the whole collector molecule or ion is the addition of the

PT
hydrophilic hydration of its hydrophilic head and the hydrophobic hydration of its
hydrophobic tail. The hydration of the micelles of anionic surfactants with -OSO3-
headgroups, nonionic surfactants with poly(ethylene oxide) chains, and cationic
surfactants with -N(CH3)3+ headgroups was found to be in the order of anionic >

RI
nonionic > cationic.
(3) Adsorption of collectors onto bubble surfaces, association of collectors with water-

SC
soluble minerals, and association of water-soluble mineral particles with bubbles are
three important processes for the flotation of water-soluble minerals. The adsorption of
collectors onto bubble surfaces is suggested to be the precondition for the association of

NU
mineral particles with bubbles. The association of collectors with water-soluble minerals
is a complicated process which may include the adsorption of collector molecules or ions
onto the mineral surfaces, and/or the attachment of collector precipitates or crystals onto
MA
the mineral surfaces. The key to successful froth flotation is the association between solid
particles and air bubbles which is affected by hydrophobicity of the mineral particle
surfaces, surface charge of mineral particles and bubbles, mineral particle size and shape,
temperature, bubble size, etc.
D

(4) The interactions between the collectors and minerals include electrostatic and
TE

hydrophobic interactions, hydrogen bonding, and specific interactions, with electrostatic


and hydrophobic interactions being the common mechanisms. For the association of ionic
collectors with minerals with an opposite charge, electrostatic and hydrophobic
P

interactions could have a synergistic effect, with the hydrophobic interactions between
the hydrophobic groups of the previously associated collectors and the hydrophobic
CE

groups of oncoming collectors being an important attractive force.


(5) Interfacial problems are still the key problems to understanding the mechanisms for
the flotation of water-soluble minerals. Changes of collector aggregation states and
AC

orientations, and changes of hydration of the collectors and the salt particle surfaces
during the association of collectors with salt particles, and the association of bubbles with
salt particles remain to be further investigated. Ingenious techniques should be developed
to characterize the water-soluble mineral surfaces in situ.
(6) One of the problems existing in the present flotation systems for water-soluble
minerals is the low solubility of the collectors. The low solubility of the collectors leads
to poor association between collectors, bubbles, and mineral particles. The use of a
frother may improve the collector solubility and diffusivity, and therefore, may improve
the flotation. Mixed surfactants used as collectors in flotation may exhibit better surface
properties than the corresponding single surfactant and may have a synergistic effect due
to the balance of electric charges and the adjustment of surfactant aggregates.

9.2. Use of frothers

The viscosity of saturated salt suspensions is high and the collector diffusion rate is
low. Frothers can improve collector solubility and dispersion and activate collector
21
ACCEPTED MANUSCRIPT

adsorption at the surfaces of bubbles and mineral particles. They can also reduce bubble
size, prevent bubble coalescence, and stabilize the bubbles. Therefore, frothers can
improve flotation efficiency [25, 27, 95, 96].
Frother addition was reported to enhance alkylmorpholine adsorption at halite
surfaces and to decrease alkylmorpholine consumption considerably [75]. During potash

PT
ore flotation using an amine as the collector, if a frother is mixed with the amine when
the amine is dispersed in a hot, acidified aqueous solution, the spreading of the amine is
dramatically enhanced. The frother acts as an active component in potash flotation [25].

RI
Although the frother has been reported to be an effective additive, its cooperative
behavior and interaction with the collector are not so clear. Furthermore, more frothers
could be selected and tested to improve the flotation.

SC
9.3. Use of mixed surfactants as collectors

NU
Surfactant mixtures have received particular attention and commercial surfactant
systems are usually mixtures of surfactants with different hydrophobic and polar groups.
Mixed surfactants used as collectors in flotation may exhibit better surface properties
MA
than the corresponding single surfactant and may have a synergistic effect due to the
balance of electric charges and the adjustment of surfactant aggregates [7, 24, 65, 70, 97].
Remarkable adsorption enhancement and packing improvement of oppositely charged
surfactants, sodium dodecyl sulfate (SDS) and dodecylamine hydrochloride (DAH) at the
D

air-water interface were observed using sum frequency generation spectroscopy and
TE

tensiometry [98]. For the co-adsorption of mixed cationic-nonionic surfactants at the


surfaces of alumina, silica, and kaolinite, a highly compact mixed monolayer of cationic
and nonionic surfactants is formed at the mineral surfaces [97]. Anionic surfactants
P

themselves do not adsorb on negatively charged silica surfaces, but substantial adsorption
can be induced by co-adsorption of a nonionic ethoxylated alcohol surfactant. Adsorption
CE

of nonionic ethoxylated alcohols on alumina surfaces is negligible, but it can be enhanced


by several orders of magnitude through co-adsorption of an anionic surfactant [70].
Compared to a single surfactant, mixed anionic/cationic surfactants provide better
AC

flotation performance in muscovite processing due to a synergistic effect [99].


Comparison of the operation efficiency of a single alkylmorpholine and an
alkylmorpholine mixture has shown that the mixture results in better flotation activity.
The best halite floatability has been obtained when a mixture of hexadecylmorpholine
and octadecylmorpholine were used [75]. More effort should be made to understand the
behavior and related interaction of mixed surfactants in the flotation of water-soluble
minerals.

Acknowledgments

This work was financially supported by the National Science and Technology
Support Program (2012BAE01B04), National Natural Science Foundation of China
(51403229, 21401209), and the High Technology Research and Development Program of
Qinghai Province (2014-GX-216A).

22
ACCEPTED MANUSCRIPT

References

[1] Ozdemir O, Karakashev SI, Nguyen AV, Miller JD, Adsorption of carbonate and
bicarbonate salts at the air-brine interface, Int J Miner Process, 2006; 81: 149-158.
[2] Hancer M, Miller JD, The flotation chemistry of potassium double salts: schoenite,

PT
kainite, and carnalite, Miner Eng, 2000; 13(14-15): 1483-1493.
[3] Ozdemir O, Karaguzel C, Nguyen AV, Celik MS, Miller JD, Contact angle and
bubble attachment studies in the flotation of trona and other soluble carbonate salts,

RI
Miner Eng, 2009; 22: 168-175.
[4] Ozdemir O, Du H, Karakashev SI, Nguyen AV, Celik MS, Miller JD, Understanding
the role of ion interactions in soluble salt otation with alkylammonium and alkylsulfate

SC
collectors, Adv Colloid Interface Sci, 2011; 163: 1-22.
[5] Gharabaghi M, Aghazadeh S, A review of the role of wetting and spreading
phenomena on the flotation practice, Curr Opin Colloid Interface Sci, 2014; 19: 266-282.

NU
[6] Kor M, Korczyk PM, Addai-Mensah J, Krasowska M, Beattie DA,
Carboxymethylcellulose adsorption on molybdenite: The effect of electrolyte
composition on adsorption, bubble-surface collisions, and flotation, Langmuir, 2014; 30:
MA
11975-11984.
[7] Rybinski W, Schwuger MJ, Adsorption of surfactant mixtures in froth flotation,
Langmuir, 1986; 2: 639-643.
[8] Albijanic B, Ozdemir O, Nguyen AV Bradshaw D, A review of induction and
D

attachment times of wetting thin films between air bubbles and particles and its relevance
TE

in the separation of particles by flotation, Adv Colloid Interface Sci, 2010; 159: 1-21.
[9] Yalamanchili MR, Kellar JJ, Miller JD, Adsorption of collector colloids in the
flotation of alkali halide particles, Int J Miner Process, 1993; 39: 137-153.
P

[10] Cao Q, Wang X, Miller JD, Cheng F, Jiao Y, Bubble attachment time and FTIR
analysis of water structure in the flotation of sylvite, bischofite and carnallite, Miner Eng,
CE

2011; 24: 108-114.


[11] Titkov S, Flotation of water-soluble mineral resources, Int J Miner Process, 2004; 74:
107-113.
AC

[12] Du H, Ozdemir O, Wang X, Cheng F, Celik MS, Miller JD, Flotation chemistry of
soluble salt minerals: from ion hydration to colloid adsorption, Miner Metall Process,
2014; 31(1): 1-20.
[13] Miller JD, Veeramasuneni S, Yalamanchili MR, Recent contributions to the analysis
of soluble salt flotation systems, Int J Miner Process, 1997; 51: 111-123.
[14] Wang X, Miller JD, Cheng F, Cheng H, Potash flotation practice for carnallite
resources in the Qinghai, Province, PRC, Miner Eng, 2014; 66-68: 33-39.
[15] Wang B, Peng Y, The effect of saline water on mineral flotation - A critical review,
Miner Eng, 2014; 66-68: 13-24.
[16] Farrokhpay S, The significance of froth stability in mineral flotation - A review, Adv
Colloid Interface Sci, 2011; 166: 1-7.
[17] Farrokhpay S, The importance of rheology in mineral flotation: A review, Miner
Eng, 2012; 36-38: 272-278.
[18] Valle-Delgado JJ, Molina-Bolvar JA, Galisteo-Gonzlez F, Glvez-Ruiz MJ,
Evidence of hydration forces between proteins, Curr Opin Colloid Interface Sci, 2011; 16:
572-578.
23
ACCEPTED MANUSCRIPT

[19] Ninham BW, Duignan TT, Parsons DF, Approaches to hydration, old and new:
Insights through Hofmeister effects, Curr Opin Colloid Interface Sci, 2011; 16: 612-617.
[20] Sett S, Karakashev SI, Smoukov SK, Yarin AL, Ion-specific effects in foams, Adv
Colloid Interface Sci, 2015; 225: 98-113.
[21] Kunz W, Specific ion effects in colloidal and biological systems, Curr Opin Colloid

PT
Interface Sci, 2010; 15: 34-39.
[22] Craig VSJ, Bubble coalescence and specific-ion effects, Curr Opin Colloid Interface
Sci, 2004; 9: 178-184.

RI
[23] Ben-Yaakov D, Andelman D, Podgornik R, Harries D, Ion-specific hydration effects:
Extending the Poisson-Boltzmann theory, Curr Opin Colloid Interface Sci, 2011; 16:
542-550.

SC
[24] Zhang R, Somasundaran P, Advances in adsorption of surfactants and their mixtures
at solid/solution interfaces, Adv Colloid Interface Sci, 2006; 123-126: 213-229.
[25] Laskowski JS, From amine molecules adsorption to amine precipitate transport by

NU
bubbles: A potash ore flotation mechanism, Miner Eng, 2013; 45: 170-179.
[26] Ozdemir O, Karakashev SI, Nguyen AV, Miller JD, Adsorption and surface tension
analysis of concentrated alkali halide brine solutions, Miner Eng, 2009; 22: 263-271.
MA
[27] Wang L, Modeling of bubble coalescence in saline water in the presence of flotation
frothers, Int J Miner Process, 2015; 134: 41-49.
[28] Hu Y, Guo T, Ye X, Li Q, Guo M, Liu H, Wu Z, Dye adsorption by resins: Effect of
ionic strength on hydrophobic and electrostatic interactions, Chem Eng J, 2013; 228: 392-
D

397.
TE

[29] Craig VSJ, Do hydration forces play a role in thin film drainage and rupture
observed in electrolyte solutions? Curr Opin Colloid Interface Sci, 2011; 16: 597-600.
[30] Craig VSJ, Ninham BW, Pashley RM, Effect of electrolytes on bubble coalescence,
P

Science, 1993; 364: 317-319.


[31] Craig VSJ, Ninham BW, Pashley RM, The Effect of electrolytes on bubble
CE

coalescence in water, J Phys Chem, 1993; 97: 10192-10197.


[32] Wang B, Peng Y, The interaction of clay minerals and saline water in coarse coal
flotation, Fuel, 2014; 134: 326-332.
AC

[33] Yamakata A, Soeta E, Ishiyama T, Osawa M, Morita A, Real-time observation of


the destruction of hydration shells under electrochemical force, J Am Chem Soc, 2013;
135: 15033-15039.
[34] Kilpatrick JI, Loh SH, Jarvis SP, Directly probing the effects of ions on hydration
forces at interfaces, J Am Chem Soc, 2013; 135: 2628-2634.
[35] Nguyen KT, Nguyen AV, Evans GM, Interactions between halide anions and
interfacial water molecules in relation to the Jones-Ray effect, Phys Chem Chem Phys,
2014; 16: 24661-24665.
[36] Bernardina SD, Brubach JB, Berrod Q, Guillermo A, Judeinstein P, Roy P,
Lyonnard S, Mechanism of ionization, hydration, and intermolecular H-Bonding in
proton conducting nanostructured ionomers, J Phys Chem C, 2014; 118: 25468-25479.
[37] Mhler J, Persson I, A study of the hydration of the alkali metal ions in aqueous
solution, Inorg Chem, 2012; 51: 425-43.
[38] Ignatov I Mosin O, Structural mathematical models describing water clusters,
Mathematical Theory and Modeling, 2013; 3(11): 72-87.

24
ACCEPTED MANUSCRIPT

[39] Kunz W, Nostro PL, Ninham BW, The present state of affairs with Hofmeister
effects, Curr Opin Colloid Interface Sci, 2004; 9: 1-18.
[40] Suresh SJ, Kapoor K, Talwar S, Rastogi A, Internal structure of water around cations,
J Mol Liq, 2012; 174: 135-142.
[41] Karpov GV, On the structure of hydration shells of alkali metal ions in aqueous

PT
solutions, Chem Phys Lett, 2005; 402: 300-305.
[42] Hribar B, Southall NT, Vlachy V, Dill KA, How ions affect the structure of water, J
Am Chem Soc, 2002;124: 12302-12311.

RI
[43] Guo T, Wang S, Ye X, Liu H, Gao X, Li Q, Guo M, Wu Z, Competitive adsorption
of Li, K, Rb, Cs ions onto three ion-exchange resins, Desalination & Water Treatment,
2013; 51: 3954-3959.

SC
[44] Nucci NV, Vanderkooi JM., Effects of salts of the Hofmeister series on the
hydrogen bond network of water, J Mol Liq, 2008; 143: 160-170.
[45] Soper AK, Weckstrm K, Ion solvation and water structure in potassium halide

NU
aqueous solutions, Biophys Chem, 2006; 124: 180-191.
[46] Collins KD, Neilson GW, Enderby JE, Ions in water: Characterizing the forces that
control chemical processes and biological structure, Biophys Chem, 2007; 128: 95-104.
MA
[47] Parsons DF, Ninham BW, Surface charge reversal and hydration forces explained by
ionic dispersion forces and surface hydration, Colloids Surf A, 2011; 383: 2-9.
[48] Marcus Y, Concentration dependence of ionic hydration numbers, J Phys Chem B,
2014; 118: 10471-10476.
D

[49] Bankura A, Carnevale V, Klein ML, Hydration structure of salt solutions from ab
TE

initio molecular dynamics, J Chem Phys, 2013; 138: 014501(1-10).


[50] Mason PE, Neutron scattering studies of the hydration structure of Li +, J Phys Chem
B, 2015; 119: 2003-2009.
P

[51] Yang L, Fan Y, Gao YQ, Differences of cations and anions: Their hydration, surface,
adsorption, and impact on water dynamics, J Phys Chem B, 2011; 115: 12456-12465.
CE

[52] Duignan TT, Parsons DF, Ninham BW, Collinss rule, Hofmeister effects and ionic
dispersion interactions, Chem Phys Lett, 2014; 608: 55-59.
[53] Ghosh MK, Re S, Feig M, Sugita Y, Choi CH, Interionic hydration structures of
AC

NaCl in aqueous solution: A Combined Study of quantum mechanical cluster calculations


and QM/EFP-MD simulations, J Phys Chem B, 2013; 117: 289-295.
[54] Geng Y, Romsted LS, Menger F, Specific ion pairing and interfacial hydration as
controlling factors in gemini micelle morphology. Chemical trapping studies, J Am Chem
Soc, 2006; 128: 492-501.
[55] Schelero N, Klitzing R, Ion specific effects in foam films, Curr Opin Colloid
Interface Sci, 2015; 20: 124-129.
[56] Kaggwa GB, Nalam PC, Kilpatrick JI, Spencer ND, and Jarvis SP, Impact of
hydrophilic/hydrophobic surface chemistry on hydration forces in the absence of
confinement, Langmuir, 2012; 28: 6589-6594.
[57] Dewan S, Carnevale V, Bankura A, Eftekhari-Bafrooei A, Fiorin G, Klein ML,
Borguet E, Structure of water at charged interfaces: A molecular dynamics study,
Langmuir, 2014; 30: 8056-8065.
[58] Djikaev YS, Ruckenstein E, Effect of water-water hydrogen bonding on the
hydrophobic hydration of large-scale particles and its temperature dependence, J Phys
Chem B, 2012; 116: 2820-2830.
25
ACCEPTED MANUSCRIPT

[59] Li ITS, Walker GC, Signature of hydrophobic hydration in a single polymer, PNAS,
2011; 108: 16527-16532.
[60] Li ITS, Walker GC, Single polymer studies of hydrophobic hydration, Acc Chem
Res, 2012; 45: 2011-2021.
[61] Das R, Duportail G, Richert L, Klymchenko A, Mely Y, Sensing micelle hydration

PT
by proton-transfer dynamics of a 3-hydroxychromone dye: Role of the surfactant
headgroup and chain length, Langmuir, 2012; 28: 7147-7159.
[62] Lima FS, Cuccovia IM, Horinek D, Amaral LQ, Riske KA, Schreier S, Salinas RK,

RI
Bastos EL, Pires PAR, Bozelli JC, Favaro DC, Rodrigues ACB, Dias LG, Seoud OAE,
Chaimovich H, Effect of counterions on the shape, hydration, and degree of order at the
interface of cationic micelles: The triflate case, Langmuir, 2013; 29: 4193-4203.

SC
[63] Lima FS, Chaimovich H, Cuccovia IM, Buchner R, Dielectric relaxation
spectroscopy shows a sparingly hydrated interface and low counterion mobility in triflate
micelles, Langmuir, 2013; 29: 10037-10046.

NU
[64] Phillies GDJ, Yambert JE, Solvent and solute effects on hydration and aggregation
numbers of Triton X-100 micelles, Langmuir, 1996; 12: 3431-3436.
[65] Rosen MJ: Surfactants and Interfacial Phenomena, John Wiley & Sons, Inc.,
MA
Hoboken, New Jersey, 2004.
[66] Somasundaran P, Zhang L, Adsorption of surfactants on minerals for wettability
control in improved oil recovery processes, J Pet Sci Technol, 2006; 52: 198-212.
[67] Albijanic B, Ozdemir O, Hampton MA, Nguyen PT, Nguyen AV, Bradshaw D,
D

Fundamental aspects of bubble-particle attachment mechanism in flotation separation,


TE

Miner Eng, 2014; 65: 187-195.


[68] ShamsiJazeyi H, Verduzco R, Hirasaki GJ, Reducing adsorption of anionic
surfactant for enhanced oil recovery: Part I. Competitive adsorption mechanism, Colloids
P

Surf A, 2014; 453: 162-167.


[69] Penfold J, Tucker I, Petkov J, Thomas RK, Surfactant adsorption onto cellulose
CE

surfaces, Langmuir, 2007; 23: 8357-8364.


[70] Somasundaran P, Krishnakumar S, Adsorption of surfactants and polymers at the
solid-liquid interface, Colloids Surf A, 1997; 123-124: 491-513.
AC

[71] Cao Q, Du H, Miller JD, Wang X, Cheng F, Surface chemistry features in the
otation of KCl, Miner Eng, 2010; 23: 365-373.
[72] Atkin R, Craig VSJ, Wanless EJ, Biggsd S, Mechanism of cationic surfactant
adsorption at the solid-aqueous interface, Adv Colloid Interface Sci, 2003; 103: 219-304.
[73] Paria S, Manohar C, and Khilar KC, Kinetics of adsorption of anionic, cationic, and
nonionic surfactants, Ind Eng Chem Res, 2005; 44: 3091-3098.
[74] Abdel-Khalek NA, Relationship of structure to properties of some anionic
surfactants as collectors in the flotation process. 1. Effect of chain length, J Chem Eng
Data, 1999; 44:133-137.
[75] Titkov S, Sabirov R, Panteleeva N, Investigations of alkylmorpholines-collectors for
a new halite flotation process, Miner Eng, 2003; 16: 1161-1166.
[76] Jordens A, Marion C, Kuzmina O, Waters KE, Surface chemistry considerations in
the flotation of bastnsite, Miner Eng, 2014; 66-68: 119-129.
[77] Veeramasuneni S, Hu Y, Yalamanchili MR, Miller JD, Interaction forces at high
ionic strengths: The role of polar interfacial interactions, J Colloid Interface Sci, 1997;
188: 473-480.
26
ACCEPTED MANUSCRIPT

[78] Veeramasuneni S, Yalamanchili MR, Miller JD, Interactions between dissimilar


surfaces in high ionic strength solutions as determined by atomic force microscopy,
Colloids Surf A, 1998; 131: 77-87.
[79] Du H, Miller JD, Interfacial water structure and surface charge of selected alkali
chloride salt crystals in saturated solutions: A molecular dynamics modeling study, J

PT
Phys Chem C, 2007; 111: 10013-10022.
[80] Hancer M, Celik MS, Miller JD, The signicance of interfacial water structure in
soluble salt flotation systems, J Colloid Interface Sci, 2001; 235: 150-161.

RI
[81] Meot-Ner M, The ionic hydrogen bond, Chem Rev, 2005; 105: 213-284.
[82] Liu H, Qing B, Ye X, Wu Z, Boron adsorption by composite magnetic particles,
Chem Eng J, 2009; 151: 235-240.

SC
[83] Lv Y, Tan T, Svec F, Molecular imprinting of proteins in polymers attached to the
surface of nanomaterials for selective recognition of biomacromolecules, Biotechnol
Adv, 2013; 31(8): 1172-1186.

NU
[84] Li Q, Liu H, Liu T, Guo M, Qing B, Ye X, Wu Z, Strontium and calcium ion
adsorption by molecularly imprinted hybrid gel, Chem Eng J, 2010; 157: 401-407.
[85] Watanabe Y, Ohnaka K, Fujita S, Kishi M, Yuchi A, Effects of the spaces available
MA
for cations in strongly acidic cation-exchange resins on the exchange equilibria by
quaternary ammonium ions and on the hydration states of metal ions, Anal Chem, 2011;
83: 7480-7485.
[86] Babiaczyk WI, Bonella S, Guidoni L, Ciccotti G, Hydration structure of the
D

quaternary ammonium cations, J Phys Chem B, 2010; 114: 15018-15028.


TE

[87] Nguyen AV, Nalaskowski J, Miller JD, A study of bubble-particle interaction using
atomic force microscopy, Miner Eng, 2003; 16: 1173-1181.
[88] Ata S, Phenomena in the froth phase of flotation - A review, Int J Miner Process,
P

2012; 102-103: 1-12.


[89] Brabcov Z, Karapantsios T, Kostoglou M, Basaov P, Matis K, Bubble-particle
CE

collision interaction in flotation systems, Colloids Surf A, 2015; 473: 95-103.


[90] Nalaskowski J, Nguyen AV, Hupka J, Miller JD, Study of particle-bubble interaction
using atomic force microscopy - current possibilities and challenges, Physicochem
AC

Problems Miner Process, 2002; 36: 253-272.


[91] Gillies G, Kappl M, Butt HJ, Direct measurements of particle-bubble interactions,
Adv Colloid Interface Sci, 2005; 114-115: 165-172.
[92] Yang S, Pelton R, Raegen A, Montgomery M, Dalnoki-Veress K, Nanoparticle
flotation collectors: mechanisms behind a new technology, Langmuir, 2011; 27: 10438-
10446.
[93] Nguyen AV, Schulze HJ, Ralston J, Elementary steps in particle-bubble attachment,
Int J Miner Process, 1997; 51: 183-195.
[94] Verrelli DI, Bruckard WJ, Koh PTL, Schwarz MP, Follink B, Particle shape effects
in flotation. Part 1: Microscale experimental observations, Miner Eng, 2014; 58: 80-89.
[95] Paruchuri VK, Nalaskowski J, Shah DO, Miller JD, The effect of co-surfactants on
sodium dodecyl sulfate micellar structures at a graphite surface, Colloids Surf A, 2006;
272: 157-163.
[96] Kowalczuk PB, Determination of critical coalescence concentration and bubble size
for surfactants used as flotation frothers, Ind Eng Chem Res, 2013; 52: 11752-11757.

27
ACCEPTED MANUSCRIPT

[97] Wang L, Hu Y, Liu J, Sun Y, Sun W, Flotation and adsorption of muscovite using
mixed cationic-nonionic surfactants as collector, Powder Technol, 2015; 276: 26-33.
[98] Nguyen KT, Nguyen TD, Nguyen AV, Strong cooperative effect of oppositely
charged surfactant mixtures on their adsorption and packing at the air-water interface and
interfacial water structure, Langmuir, 2014; 30: 7047-7051.

PT
[99] Wang L, Hu Y, Sun W, Sun Y, Molecular dynamics simulation study of the
interaction of mixed cationic/anionic surfactants with muscovite, Appl Surf Sci, 2015;
327: 364-370.

RI
SC
NU
MA
D
P TE
CE
AC

28
ACCEPTED MANUSCRIPT

Graphical Abstract

PT
RI
SC
NU
MA
D
P TE
CE
AC

29
ACCEPTED MANUSCRIPT

Some Physicochemical Aspects of Water-soluble Mineral Flotation

PT
Highlights

RI
Review water-soluble mineral flotation

SC
Hydration theory of water-soluble mineral surfaces
Interaction between collectors and water-soluble minerals
Interaction between water-soluble minerals and bubbles

NU
MA
D
P TE
CE
AC

30

You might also like