Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

minerals

Review
Collecting Agent–Mineral Interactions in the Reverse
Flotation of Iron Ore: A Brief Review
Guixia Fan 1 , Liguang Wang 2 , Yijun Cao 1 and Chao Li 3, *
1 School of Chemical Engineering, Zhengzhou University, Zhengzhou 450001, China; cumtfgx@126.com (G.F.);
yijuncao@126.com (Y.C.)
2 School of Chemical Engineering, The University of Queensland, Brisbane, Queensland 4072, Australia;
liguang.wang@uq.edu.au
3 Henan Province Industrial Technology Research Institute of Resources and Materials, Zhengzhou University,
Zhengzhou 450001, China
* Correspondence: c.li@zzu.edu.cn

Received: 28 May 2020; Accepted: 28 July 2020; Published: 30 July 2020 

Abstract: Froth flotation has been widely used in upgrading iron ores. Iron ore flotation can be
performed in two technical routes: direct flotation of iron oxides and reverse flotation of gangue
minerals with depression of iron oxides. Nowadays, reverse flotation is the most commonly used
route in iron ore flotation. This review is focused on the reverse flotation of iron ores, consisting of
reverse cationic flotation and reverse anionic flotation. It covers different types of collecting agents
used in reverse iron ore flotation, the surface characteristics of minerals commonly present in iron
ores (e.g., iron oxides, quartz, alumina-bearing minerals, phosphorus-bearing minerals, iron-bearing
carbonates, and iron-bearing silicates), and the adsorption mechanisms of the collecting agents at the
mineral surface. The implications of collecting agent–mineral interactions for improving iron ore
flotation are discussed.

Keywords: iron ore; reverse flotation; cationic collector; anionic collector; surface adsorption

1. Introduction
Steel is an indispensable material for the construction industry, shipbuilding, railway construction,
motor vehicle manufacture, bridge building, machinery manufacture, and many other engineering
applications. Steel is made mainly from iron, which is one of the most abundant elements on Earth.
Iron is extracted primarily from iron ores. The iron ores mainly include oxides and hydroxides
such as magnetite [Fe3 O4 ], hematite [Fe2 O3 ], goethite [FeO(OH)], and limonite [FeO(OH)·nH2 O] [1].
The primary gangue mineral in iron ores is quartz. In addition to quartz, iron-bearing silicates
(e.g., amphiboles and pyroxenes), carbonates, clays (e.g., kaolinite), and gibbsite are also commonly
present in iron ores [1,2]. Iron ore beneficiation aims to eliminate the harmful elements in iron
ore concentrate, which could impose a detrimental effect on ironmaking. Table 1 summarizes the
allowed contents of the harmful elements in the iron ore concentrate (in China) and their effects on
ironmaking/steelmaking.

Minerals 2020, 10, 681; doi:10.3390/min10080681 www.mdpi.com/journal/minerals


Minerals 2020, 10, 681 2 of 22

Table 1. The allowed content of the harmful elements in iron ore concentrate [3] and their detrimental
effects on ironmaking.

Detrimental Effect on the Mechanical


Elements Magnetite Concentrate Hematite Concentrate
Properties of Iron/Steel
Si (%) ≤4.2 ≤5.6 Decreased toughness and weldability [4]
Increased brittleness of steel and decreased
S (%) ≤0.50 ≤0.30
weldability and corrosion resistance [5]
Increased hardness and brittleness and
P (%) ≤0.10 ≤0.10
decreased ductility [6]
Al (%) ≤1.1 ≤0.8 Decreased creep resistance [7]

The amount of high grade and easy-to-process iron ores are in continual decline and beneficiation
of iron ores is in increasing demand. A challenge is to upgrade the low-grade iron ores with complex
mineralogy and fine grain size, which require the beneficiation to be carried out at fine or ultrafine size
fractions [8]. A versatile method of beneficiation of fine and ultrafine particles is froth flotation. In the
flotation of iron ores, the difference in flotation rates between iron oxides and gangue minerals needs
to be enlarged, which can be achieved by changing the surface hydrophobicity of certain minerals
using various reagents, such as pH modifiers, depressants, activators, and collectors.
Since the early stage of technology development for iron ore flotation in the 1930s, the following
two technical routes of iron ore flotation have been developed: (i) direct flotation of iron oxides and (ii)
reverse flotation of gangue minerals by depressing iron oxides. The route of reverse flotation is currently
in widespread use in iron ore flotation practice. The majority of studies on reverse flotation of iron ore
focused on reagents and reagent scheme [9–13]. The reverse flotation route can be classified into reverse
cationic flotation and reverse anionic flotation, based on the collector type. The reverse anionic flotation
(flotation of gangue minerals using anionic collectors that mainly include fatty acids) was developed
in the early 1960s [14]. At present, reverse anionic flotation is mainly applied in China to upgrade
iron ores [15–17]. The most popular flotation route in the iron ore industry worldwide is reverse
cationic flotation [1,17,18]. The cationic collectors have evolved from fatty amines in the early industrial
applications to ether amines with relatively high solubility in water at the present [9]. Apart from using
a cationic or an anionic collector alone, the use of reagent mixtures has been increasingly popular in the
reverse iron ore flotation practice. The mixtures can be (i) anionic/anionic collectors, (ii) anionic/cationic
collectors, or (iii) ionic/non-ionic combinations [10]. The presence of the co-surfactants in solution
can enhance the adsorption of the collectors at the solid/water interface and improve the mineral
hydrophobicity [11–13].
The performance of the reverse flotation of iron ores is largely governed by the interactions
between collectors and minerals, which are complex. A common task in the reverse flotation of iron
ores is to separate quartz from iron oxides as quartz is often the major gangue mineral in iron ores.
In some iron ores, removal of non-quartz gangue minerals is also essential, which has been the subject
of several studies [16,19–23]. These non-quartz gangue minerals include alumina-containing minerals,
phosphorus-containing minerals, iron-bearing carbonates, and iron-bearing silicates. The presence of
these gangue minerals not only complicates the flotation system, but also imposes a detrimental effect
on downstream steel-making processes. The surface properties of these gangues minerals are different
from that of quartz, so they are often treated separately in the flotation process.
To date, several review papers [1,9,10,14,18,24,25] either discuss the collector–mineral interactions
in a single flotation route or cover the collectors used in different iron ore flotation routes, but neglect
the fundamental interactions between collectors and minerals. There are still no reviews dedicated to
the interactions between collectors and minerals in different iron ore flotation routes. In particular,
the present work reviewed the interactions between quartz to non-quartz gangue minerals and collectors,
aiming at facilitating the process of complex iron ores. It is expected that a fundamental understanding
of the interactions between collectors and mineral surfaces will contribute to the design of appropriate
flotation collectors and collector regimes for iron ore beneficiation. As only reverse cationic and anionic
Minerals 2020,10,
Minerals2020, 10,681
x FOR PEER REVIEW 33 of
of 22
22

present work focuses on reviewing the fundamental collector–mineral interactions in these two
flotation routes are currently used in industry, the present work focuses on reviewing the fundamental
routes.
collector–mineral interactions in these two routes.
2. Collectors
2. Collectors for
for Quartz
Quartz
Asquartz
As quartzisisthethe main
main gangue
gangue mineral
mineral in ores,
in iron iron this
ores,section
this section
focusesfocuses on the interactions
on the interactions between
collectors and quartz in aqueous solution. The interactions between collectors and non-quartzand
between collectors and quartz in aqueous solution. The interactions between collectors non-
gangues
quartz
will gangues will
be discussed be discussed
in Section 3. in Section 3.

2.1. Anionic
2.1. Anionic Collectors
Collectors
Fatty acids
Fatty acids are
are widely
widely used
used asas aa collector
collector in in the
thereverse
reverseanionic
anionic flotation
flotation of ofiron
ironores.
ores. The
The most
most
popular anionic collectors used in iron ore flotation practice are oleic acid
popular anionic collectors used in iron ore flotation practice are oleic acid and its soaps [16,26]. and its soaps [16,26]. The
oleate
The in aqueous
oleate solution
in aqueous formsforms
solution different species
different and which
species speciesspecies
and which dominate is dependent
dominate on pH.
is dependent
Table 2 summarizes the equilibrium constants for oleate aqueous species
on pH. Table 2 summarizes the equilibrium constants for oleate aqueous species at a total ionic at a total ionic concentration
of 1 × 10−2 mol/L.
concentration Figure
of 1 × 1 shows
10−2 mol/L. the species
Figure 1 showsdistribution
the species diagram of oleate
distribution diagram as aoffunction
oleate asof pH. The
a function
oleate
of is insoluble
pH. The in acidic in
oleate is insoluble pHacidic
region, pH existing in the form
region, existing in theofform
oleic of acid
oleic droplet, emulsion,
acid droplet, and
emulsion,
and insoluble film [27]. The solubility of the oleate ions increases as the pH is raised to alkaline. AtpH
insoluble film [27]. The solubility of the oleate ions increases as the pH is raised to alkaline. At a a
value
pH above
value above 11,11,
thethe
oleate species
oleate speciesexist
existmainly
mainlyininthe theform
formof of oleate
oleate ion (RCOO−−)) and
ion (RCOO and oleate dimer
oleate dimer
((RCOO)222−
((RCOO) 2−). Oleate ions are the functional species that can interact with the quartz surface activated
). Oleate ions are the functional species that can interact with the quartz surface activated
by polyvalent
by polyvalent metal metal cations
cations(the
(the details
detailsare arediscussed
discussedbelow).
below). This
This is
is why
why reverse
reverse anionic
anionic flotation
flotation ofof
ironore
iron oreusing
usingoleicoleic acid
acid or or
its its soaps
soaps is usually
is usually performed
performed in a strongly
in a strongly alkaline alkaline environment
environment (i.e.,
(i.e., above
above pH 11). Fuerstenau and Cummins [28] concluded that, when oleate
pH 11). Fuerstenau and Cummins [28] concluded that, when oleate was used as collector, the suitable was used as collector, the
suitable pH value was between 11 and 12 in
pH value was between 11 and 12 in the flotation of quartz. the flotation of quartz.

Table2.2.Equilibrium
Table Equilibriumconstants
constantsfor
foroleate
oleate species
species in in aqueous
aqueous solution
solution [29][29] (total
(total ionicionic = 1 ×=
concentration
concentration
1 −2
10 × 10 −2 mol/L).
mol/L).

EquilibriaEquilibria Constants
Constants
HOl ⇌ H+ + Ol−
HOl
H+ + Ol−
pK =pK 4.95
0
a = 4.95
HOll
H+HOl l ⇌
− H+ + Ol− pK pK= 12.55
0
+ Ol sp = 12.55
2− − ⇌ Ol 0
2Ol−
Ol2Ol D = 4.00
2
logK logK
= 4.00
−−
HOl + Ol−HOl

HOl + Ol2 ⇌ HOl logK logK AD = 4.75
= 4.75
+ + 0
Na +2(l)H⇌ +2Ol sp = 19.00
NaHOl2(l)
− pK
NaHOl Na+ + H+ +2Ol− pK = 19.00

-3
cmc

-4
Liquid oleic acid region

-5
log [species]

Oleate dimer
-6 Oleate ion

-7 Acid-soap complex
Soluble oleic acid
-8

-9
2 4 6 8 10 12

pH
Figure 1. Species distribution diagram of oleate as a function of pH (total concentration = 1 × 10−2 mol/L)
Figure
(after 1. Species distribution diagram of oleate as a function of pH (total concentration = 1 × 10-2
[29]).
mol/L) (after [29]).
Minerals 2020, 10, 681 4 of 22

Minerals 2020, 10, x FOR PEER REVIEW 4 of 22


At pHs 11–12, the oleate ions cannot, however, directly adsorb onto the quartz surface. At this
At pHs
pH range, the11–12,
quartzthe oleateisions
surface cannot,charged
negatively however, directly
[30], so the adsorb ontoofthe
adsorption quartz
oleate ionssurface.
on the At this
quartz
pH range,
surface the quartz
would surface
be resisted byisthe
negatively charged
electrostatic [30], so the
repulsion. Theadsorption of oleate
quartz needs to beions on the
firstly quartz
activated
surface would be resisted by the electrostatic repulsion. The quartz needs to be
by adsorption of multivalent ions to reverse its surface charge from negative to positive. The mostfirstly activated by
adsorption of multivalent ions to reverse its surface charge from negative to
prevailing cation used for this purpose is Ca2+ (often sourced from lime). The species of calcium positive. The most
prevailing
ions cation
existing used forare
in solution thisalso
purpose is Ca2+ (often
pH-dependent sourced
[31,32]. from2lime).
Figure shows The species
that, at theof pH
calcium
rangeions
of
existing in solution are also pH-dependent [31,32]. Figure 2 shows that,
2+at the
11–12, the dominant species of calcium existing in the solution are Ca and Ca(OH) . It has beenpH range + of 11–12, the
dominantthat
reported species of calcium
the Ca(OH) existing
+ is the in that
species the solution
can adsorbare onto
Ca2+ the
andnegatively
Ca(OH)+. Itcharged
has been reported
quartz that
surfaces,
the Ca(OH)
forming
+ is the species
SiO-Ca(OH) [31]. that can adsorb onto the negatively charged quartz surfaces, forming SiO-
Ca(OH) [31].
-3
Ca2+
log concentration, M

-4

Ca(OH)+

-5

Ca(OH)2(s)

-6
10 11 12 13 14
pH

Figure 2.2. Species


Figure Speciesdistribution
distributiondiagrams of Ca
diagrams
2+
of Caas
2+ aas
function of pHof(total
a function concentration
pH (total = 1 × 10=−31 mol/L)
concentration × 10-3
(after [26]).
mol/L) (after [26]).

The adsorption of oleate ions on the activated quartz surface is considered chemisorption via
The adsorption of oleate ions on the activated quartz surface is considered chemisorption via
forming covalent bonding. The adsorption can be described by the general electron donor/electron
forming covalent bonding. The adsorption can be described by the general electron donor/electron
− functional group is
acceptor model (i.e., covalent bonding model); that is, the oxygen in the COO
acceptor model (i.e., covalent bonding model); that is, the oxygen in the COO− functional group is the
the electron donor and the calcium is the electron acceptor [33,34]. Equations (1) and (2) show the
electron donor and the calcium is the electron acceptor [33,34]. Equations (1) and (2) show the
reactions between the oleate ions and the SiO-Ca(OH) on the quartz surface [26]:
reactions between the oleate ions and the SiO-Ca(OH) on the quartz surface [26]:

SiO–(Ca(OH)
SiO–Ca OH) + + RCOO →
RCOO → SiOCaOOCR OH−
SiOCaOOCR ++OH (1)
(1)

2SiO– (Ca(OH)
2SiO–Ca OH) + + (RCOO)
(RCOO )2− → 2SiOCaOOCR + 2OH −
2 → 2SiOCaOOCR + 2OH
(2)
(2)

The formation of SiOCaOOCR, a precipitate, is non-reversible. The presence of the hydrocarbon


tail ofThe
theformation of SiOCaOOCR,
oleate molecule on the quartz a precipitate, is non-reversible.
surface renders The presence
the quartz surface of the hydrocarbon
hydrophobic.
tail of the oleate molecule on the quartz surface renders the quartz surface hydrophobic.
A problem associated with using oleic acid in flotation is that these reagents exhibit low solubility
A problem
and often require associated with using
heating to enhance theiroleic acidwhich
activity, in flotation is that
significantly these reagents
increases the process exhibit low
cost [35].
solubility and often require heating to enhance their activity, which significantly
Changing oleic acid (or fatty acids) to its soap by the addition of caustic soda has been commonly increases the process
cost [35].
used Changing
to increase oleic solubility
collector acid (or fatty
and acids) to itsSome
efficiency. soap attempts
by the addition
have been of caustic
made tosoda has fatty
modify been
commonly
acids used totheir
by changing increase collector
molecular solubility
structures forand efficiency.
increased Some and
solubility attempts have
activity. been made
Fuerstenau to
and
modify fatty acids by changing their molecular structures for increased
Jia [36] showed that the addition of a second polar group to the fatty acid molecule could not onlysolubility and activity.
Fuerstenau
increase the and Jia [36]
solubility showed
owing that
to the the addition
formation of a second
of hydrogen bondspolar groupthe
between to the
polarfatty acidand
group molecule
water
could not only increase the solubility owing to the formation of hydrogen
molecules, but also enhance the collector adsorption on the mineral surfaces owing to the increasebonds between the polar
in
group and water molecules, but also enhance the collector adsorption on the mineral
electrostatic attraction. Ogata, et al. [37] found that the saturated fatty acids modified by introducing a surfaces owing
to the increase
chlorine atom to intheelectrostatic attraction.
α-carbon position couldOgata,
improve et al. [37]
their found that
solubility. More the saturated
recently, Zhu,fatty
et al.acids
[38]
modified by introducing a chlorine atom to the α-carbon position could improve
introduced a bromine atom to the α-carbon position of lauric acid to improve its solubility, and the their solubility.
More recently,
results suggestedZhu, et the
that al. [38]
newlyintroduced
synthesizeda bromine
collectoratom to the α-carbon
exhibited position
good activity at aofrelatively
lauric acid to
low
improve its solubility, and the results suggested that the newly synthesized collector exhibited good
activity at a relatively low flotation temperature (15 °C). Similarly, Luo, et al. [39] synthesized a new
collector by introducing a bromine atom to the α-carbon position of decanoic acid, and they found
Minerals 2020, 10, 681 5 of 22

Minerals 2020, 10, x FOR PEER REVIEW 5 of 22


flotation temperature (15 ◦ C). Similarly, Luo, et al. [39] synthesized a new collector by introducing a
that theatom
bromine collector
to thealso exhibited
α-carbon goodofselectivity
position at 16and
decanoic acid, °C they
withfound
improved solubility.
that the collectorAlthough these
also exhibited
newly modified ◦
anionic collectors have better solubility and selectivity
good selectivity at 16 C with improved solubility. Although these newly modified anionic collectors (even at low flotation
temperatures),
have none of
better solubility andthem have been
selectivity applied
(even in flotation
at low the iron ore flotation practice,
temperatures), none probably
of them have because
beenof
the highinsynthesis
applied the iron ore cost.
flotation practice, probably because of the high synthesis cost.
Theeffect
The effect ofof ions in process
process water
wateron onreverse
reverseanionic
anionic flotation
flotation hashas
alsoalso
attracted attention.
attracted The
attention.
presence
The presenceof excessive
of excessivepolyvalent
polyvalentcations
cationscould be detrimental
could be detrimental for for
reverse anionic
reverse flotation
anionic [40,41].
flotation For
[40,41].
example, the concentration of Ca 2+,2+mainly sourced from lime addition, can be accumulated to an
For example, the concentration of Ca , mainly sourced from lime addition, can accumulated to an
excessivelevel
excessive levelininthe
theprocess
processwater.
water.Ca 2+
Ca also
2+ also may
may get
get adsorbed
adsorbed on on the
the ultrafine
ultrafineparticles
particlesof ofhematite,
hematite,
renderingitsitssurface
rendering surfacepositively
positively charged
charged to to accommodate
accommodate thethe adsorption
adsorption of fatty
of fatty acids,
acids, thus thus resulting
resulting in
the loss of iron along with the ultrafine quartz. A large amount of cations will precipitate with OH–, –
in the loss of iron along with the ultrafine quartz. A large amount of cations will precipitate with OH
, which in turn hinders the flotation process. Only a couple of studies placed an emphasis on the role
which in turn hinders the flotation process. Only a couple of studies placed an emphasis on the role of
of anions
anions in plant
in plant water.
water. It was
It was found
found thatthat an anion
an anion in water
in water has ahas a greater
greater depression
depression than than a cation
a cation at theat
the same
same chargecharge
in theinflotation
the flotation of hematite,
of hematite, and anand an increase
increase in the in the valence
valence of anionof leads
aniontoleads
a dropto aindrop
the
in the adsorption
adsorption of oleate ofon
oleate on hematite
hematite [42,43]. [42,43].

2.2.
2.2.Cationic
CationicCollectors
Collectors
Amines
Aminesare arewidely
widelyusedusedasascollectors
collectorsininthethereverse
reversecationic
cationicflotation
flotationofofiron
ironores.
ores.Because
Becauseaa
cationic collector carries a positively charged headgroup in water, the adsorption
cationic collector carries a positively charged headgroup in water, the adsorption of the collector of the collector on ona
mineral
a mineralsurface
surfaceis governed
is governed by the magnitude
by the magnitude andandsignsign
of theof surface charge
the surface of theofmineral,
charge which
the mineral, can
which
be measured
can be measuredby a zeta
by apotential metermeter
zeta potential or zetaorprobe [44,45].
zeta probe [44,45].
The
Thezeta
zetapotential
potentialofofquartz
quartzininaqueous
aqueoussolution
solutionisispH-dependent.
pH-dependent.Quartz Quartzhydrolyses
hydrolysesininaqueous
aqueous
solution
solutionto toform
form hydroxyl
hydroxyl at the surface
surface [30].
[30]. The
Thesurface
surfacehydroxylation
hydroxylation achieves
achieves a maximum
a maximum at
at the
the isoelectric
isoelectric point
point (IEP)
(IEP) [46],[46],
a pHa pH
valuevalue at which
at which the charge
the net net charge
at theatsurface
the surface is zero.
is zero. BelowBelow
the IEP theof
IEP of quartz,
quartz, the surface
the surface hydroxyl hydroxyl protonates
protonates and theand the surface
quartz quartz surface
becomesbecomes
positivelypositively
charged,charged,
whereas
whereas
above the above the hydroxylated
IEP, the IEP, the hydroxylated quartz deprotonates
quartz deprotonates and becomes and becomes charged.
negatively negatively As charged.
shown in
As shown
Figure in Figure
3, the IEP of 3, the IEP
quartz of aqueous
in an quartz insolution
an aqueous solution
is about 2. Theisreverse
about 2. The reverse
cationic flotationcationic
of iron
flotation of iron ores
ores is normally is normally
performed performed
at weak at weak
alkaline pHs, wherealkaline pHs, where
the surface theissurface
of quartz negativelyof quartz
charged.is
negatively
Note that charged. Note that
hematite surface canhematite surface can also
also be hydrolyzed be hydrolyzed
in aqueous solutionin aqueous
[47,48]. solution
Figure 3 shows[47,48].
that
Figure
the IEP 3 shows that the
of hematite IEP of hematite
is between 6 and 7,is between 6that
indicating andthe 7, hematite
indicatingsurface
that theis hematite surfacecharged
also negatively is also
negatively
at normal charged
pHs of the at normal
flotationpHs of the flotation
operations operations
(e.g., 8–10.5), but to(e.g., 8–10.5),
a lesser butthan
extent to a the
lesser extent
quartz than
surface.
the
Thequartz surface.
cationic The cationic
collector molecules collector
wouldmolecules wouldadsorb
preferentially preferentially adsorbinstead
onto quartz onto quartz
of oninstead of
hematite,
on hematite,when
especially especially when a for
a depressant depressant
hematitefor hematite is used.
is used.

30

20 Hematite

10 Quartz
Zeta potential (mV)

-10

-20

-30

-40

-50
0 2 4 6 8 10 12
pH

Figure 3. Zeta potential of quartz and hematite as a function of pH (after [49]).


Figure 3. Zeta potential of quartz and hematite as a function of pH (after [49]).
Minerals 2020, 10, 681 6 of 22
Minerals 2020, 10, x FOR PEER REVIEW 6 of 22

AA cationic collector
collectoroften
often forms
forms different
different species species in aqueous
in aqueous solutions,
solutions, and whichandspecies
whichdominate
species
dominate is dependent
is dependent on pH. For on pH. For example,
example, dodecylamine
dodecylamine (DDA),(DDA),the most thecommonly
most commonly used collector
used collector in the
in the development
early early development stage stage of reverse
of reverse cationiccationic
flotationflotation
of iron ofores,
iron may
ores, be
may be present
present in various
in various forms
forms such
such as RNH + , (RNH
as 3RNH 3+, (RNH) 2+3,)2RNH
2+, RNH ·RNH
2 ·RNH + 3, +,RNH
RNH 2 (neutral
(neutral molecule),
molecule), and RNH 2 precipitates
precipitates in
in
3 2 2 3 2 2
the
thesolution
solutionphase,
phase,depending
dependingon onpH pHand and concentration
concentration[50]. [50]. Table
Table33summarizes
summarizesthe theequilibrium
equilibrium
constants
constantsfor
fordodecylamine
dodecylamine aqueous aqueous species
species at total ionic concentration of 5 × 10−2−2mol/L.
× 10 mol/L.As Asdepicted
depicted
in
inFigure
Figure4,4,the
the ionic
ionic forms
forms RNH33+ and + and(RNH3)
(RNH3)22+ 2+
dominate
dominate at the
at pH
the pH range
range of
of2 2to
to9 9and
and the
theneutral
neutral
2
molecule
moleculeRNHRNH2 2precipitates
precipitatesat at pH
pH 10.
10. TheThe concentration
concentration of ion-molecular complex RNH ·RNH33++
RNH22·RNH
exhibits
exhibitsaamaximum
maximumvalue valueat at pHpH 10.5.
10.5. At AtpHpH >>10.5, 10.5,thetheprimary
primaryspecies
speciesare
are RNH
RNH22molecule
moleculeand and
RNH
RNH22precipitation.
precipitation.Filippov,
Filippov,etetal. al.[1]
[1]noted
notedthat thatthe themost
mosteffective
effectivecationic
cationiccollectors
collectorsare arehydrolyzed
hydrolyzed
reagents
reagentsthat
thatpresent
presentbothbothionic
ionic andandmolecular
molecular species
species in the aqueous
in the aqueous phase. Hence,
phase. the optimum
Hence, the optimumpH
for
pHDDA as aascollector
for DDA a collectorshould
should be be
10.5,
10.5,where
where thetheamount
amountofofion-molecular
ion-molecular complex
complex RNH22·RNH ·RNH33++
reaches
reachesthe
themaximum.
maximum.

Table 3.
Table Equilibriumconstants
3. Equilibrium constantsfor
fordodecylamine
dodecylaminespecies
speciesininaqueous
aqueoussolution
solution[40]
[40] (total
(total ionic
ionic
concentration = 5 × 10 −2 mol/L).
concentration = 5 × 10 mol/L).
−2

EquilibriaEquilibria ConstantsConstants
RNH2
RNH ⇌ RNH2(so)
RNH22(so) pK = 4.69pK0so = 4.69
RNH2 +⇌HRNH
RNH3+
RNH + 2 + H+ pK = 10.63pK0a = 10.63
(RNH3⇌
2RNH3+ −
2RNH )22+(RNH ) pK = −2.08

pK0a = −2.08
+ RNH + RNH 2− ⇌ (RNH+ ∙ RNH )
RNH + RNH
(RNH ·RNH )
− pK = −3.12
pK0 = −3.12
3 2 2 3 AB

-3

-4 RNH3+

-5 RNH2(s)
logC

-6 (RNH3+)22+

-7

-8

-9
0 2 4 6 8 10 12 14
pH
Figure 4. Species distribution diagram of dodecylamine as a function of pH (total concentration =
Figure 4. Species
5 × 10−5 distribution
mol/L) (after [40]). diagram of dodecylamine as a function of pH (total concentration = 5 ×
10-5 mol/L) (after [40]).
The floatability of quartz also depends on the amount of amines adsorbed at the surface. At low
The floatability
concentrations of quartz
(Figure alsoamine
5A), the depends ionson theadsorbed
are amount of at amines adsorbed
the quartz at the
surface, surface.
resulting At low
primarily
concentrations (Figure 5A), the amine ions are adsorbed at the quartz surface, resulting
from electrostatic forces, and the hydrophobic amine tails increase the floatability of quartz [51,52]. primarily
from
When electrostatic forces, and theishydrophobic
the amine concentration amine tails
increased, reaching increasehemimicelle
the critical the floatability of quartz [51,52].
concentration (CHC),
When the amine
saturated concentration
monolayer is increased,
hemimicelles are formedreaching
at thethe critical hemimicelle
solid–liquid concentration
interface (see Figure 5B).(CHC),
At this
saturated monolayer hemimicelles are formed at the solid–liquid interface (see Figure
concentration, the floatability of quartz reaches a maximum and the zeta potential is reversed from 5B). At this
concentration, the floatability of quartz reaches a maximum and the zeta potential is
negative to positive, under which concentration the Stern layer acts against further adsorption [53]. reversed from
negative
Above the to CHC,
positive,
theunder
aminewhich concentration
adsorption is driven the Stern
by the layer actsbetween
association against further adsorption
hydrocarbon chains[53].
[54].
Above the CHC, the amine adsorption is driven by the association between hydrocarbon
The polar head of the amine ions that associate with those already adsorbed on the surface might chains [54].
The
staypolar
awayhead
fromofthe
thesurface
amine ions
owingthattoassociate with those
the repulsion betweenalready adsorbed
the ionic onand
heads, the surface
a bilayermight stay
is formed
away from the surface owing to the repulsion between the ionic heads, and a bilayer is formed by
tail–tail hydrophobic interaction (Figure 5C). This type of adsorption of the ions can lead to a
Minerals 2020, 10, 681 7 of 22
Minerals 2020, 10, x FOR PEER REVIEW 7 of 22

by tail–tail hydrophobic interaction (Figure 5C). This type of adsorption of the ions can lead to a
hydrophilic quartz surface, thus decreasing quartz recovery in flotation [51,53,55]. A further increase
hydrophilic quartz surface, thus decreasing quartz recovery in flotation [51,53,55]. A further increase
in the concentration of the amine to its critical micelle concentration (CMC) will allow micelles to
in the concentration of the amine to its critical micelle concentration (CMC) will allow micelles to form
form in the bulk solution (see Figure 5D). Any further increase in the amine concentration will not
in the bulk solution (see Figure 5D). Any further increase in the amine concentration will not affect the
affect the collector adsorption at the quartz surface. In short, the concentration of a cationic collector
collector adsorption at the quartz surface. In short, the concentration of a cationic collector in flotation
in flotation needs to be controlled at an appropriate concentration, and an overdose may be
needs to be controlled at an appropriate concentration, and an overdose may be detrimental for quartz
detrimental for quartz flotation. For dodecylamine hydrochloride (DAC), a typical cationic collector,
flotation. For dodecylamine hydrochloride (DAC), a typical cationic collector, the literature reports
the literature reports that its CMC varies
−2
between 1.25 × 10−2−2mol/L and 1.38 × 10−2 mol/L [56,57].
that its CMC varies between 1.25 × 10 mol/L and 1.38 × 10 mol/L [56,57].

Figure 5. Speculated schematic model of a cationic collector/surfactant adsorption at the quartz/water


Figure 5. Speculated schematic model of a cationic collector/surfactant adsorption at the quartz/water
interface over concentration. (A: concentration < CHC; B: concentration = CHC; C: CHC < concentration
interface over concentration. (A: concentration < CHC; B: concentration = CHC; C: CHC <
< CMC; D: concentration = CMC)
concentration < CMC; D: concentration = CMC)
The molecular structure of a flotation collector is an important factor affecting its flotation
The molecular
performance. Primary structure
alkylamine of acollectors
flotation used
collector
at theis early
an important
stage of ironfactor
ore affecting its flotation
flotation technology
development were abandoned in 1960s owing to their low solubility and activity [1]. The lowtechnology
performance. Primary alkylamine collectors used at the early stage of iron ore flotation solubility
development
in water and low were abandoned
critical micellein 1960s owingoftoalkylamines
concentration their low solubility
result in lowandadsorption
activity [1].densityThe lowat
solubility in water and low critical micelle concentration of alkylamines
the quartz surface. Quaternary ammonium salts exhibit higher solubility and better selectivity result in low adsorption
density
than at the quartz
alkylamine collectorssurface.
[58]. Quaternary
Quaternary ammonium
ammonium salts salts inexhibit
waterhigher
are in the solubility
form ofand ionsbetter
at a
selectivity than alkylamine collectors [58]. Quaternary ammonium
wide range of pH and are, therefore, less sensitive to pH [59]. However, quaternary ammonium salts in water are in the form of
ions result
salts at a wide range oflow
in relatively pHadsorption
and are, therefore,
density atless thesensitive
quartz/waterto pHinterface,
[59]. However,
primarily quaternary
because
ammonium salts result in relatively low adsorption density at the quartz/water
of the strong electrostatic repulsions between the collector ions at the quartz/water interface [60]. interface, primarily
because of and
Fuerstenau the strong
Modi [61] electrostatic
showed thatrepulsions
effective between
flotation the
ofcollector
corundum ions at therequire
would quartz/water interface
a concentration
of the trimethyldodecylammonium salt (a quaternary ammonium salt) 10 times higher than that ofa
[60]. Fuerstenau and Modi [61] showed that effective flotation of corundum would require
concentration of
dodecylamine. the trimethyldodecylammonium
Currently, ether amines and their salts salt (a quaternary
have been widely ammonium salt) 10intimes
used as collector higher
the reverse
than that of dodecylamine. Currently, ether amines and their salts have
cationic flotation practice for upgrading iron ores. According to Araujo, et al. [9] and Fuerstenau been widely used as collector
in the
and Jiareverse
[36], thecationic
addition flotation practice
of a second for group
polar upgrading
to the iron ores. According
molecule to Araujo,
of the primary et al.amines
aliphatic [9] and
Fuerstenau and Jia [36], the addition of a second polar group to the molecule
increases the collector adsorption density at the mineral surfaces. The ether amines with the presence of the primary aliphatic
amines
of one –NH increases the collector adsorption density at the mineral surfaces. The ether amines with the
2 functional group are named ether monoamines. In addition, a second –NH2 functional
presence
group couldof be
oneinserted
–NH2 functional group are
into the molecular named forming
structure, ether monoamines.
ether diamines. In addition, a second
It is expected –NH2
that ether
functional group could be inserted into the molecular structure, forming
diamines have higher solubility and selectivity compared with ether monoamines because the former ether diamines. It is expected
thattwo
has ether diamines have
hydrophilic higher
moieties in solubility
one molecule and selectivity
[62]. Araujo, compared with
et al. [9] ether monoamines
reported that ether diaminesbecause
the former
would exhibithassuperior
two hydrophilic
selectivitymoieties
in coarseinsilicate
one molecule
particle [62]. Araujo,
flotation. Someet al. [9] reported
studies suggested thatthat
ether
a
diamines would exhibit superior selectivity in coarse silicate particle
mixture of monoamines and diamines may be more efficient for the flotation of iron ores with a wider flotation. Some studies
suggested
particle sizethat a mixture[1,63,64].
distribution of monoamines and diamines may be more efficient for the flotation of iron
oresSome
with aprogress
wider particle size distribution
for developing new types [1,63,64].
of cationic collectors was made in the past several years.
Some progress for developing new
Huang, et al. [53] synthesized a cationic gemini collector types of cationic collectors was made in bromide
dimethyl-dodecyl-ammonium the past(EBAB).
several
years. Huang, et al. [53] synthesized a cationic gemini collector dimethyl-dodecyl-ammonium
bromide (EBAB). Gemini collectors contain two hydrophilic head groups (functional groups) and two
hydrophobic tails covalently linked through a spacer [65]. Figure 6 shows a schematic model of a
Minerals 2020, 10, 681 8 of 22

Gemini collectors
Minerals 2020, 10, x FORcontain two hydrophilic head groups (functional groups) and two hydrophobic
PEER REVIEW 8 of 22
tails covalently linked through a spacer [65]. Figure 6 shows a schematic model of a cationic gemini
collector
cationic adsorption at theadsorption
gemini collector quartz surface.
at theThe EBAB
quartz exhibited
surface. The higher solubilityhigher
EBAB exhibited and selectivity
solubilitythan
and
dodecylamine
selectivity than hydrochloride.
dodecylamine Similarly, Weng, etSimilarly,
hydrochloride. al. [66] developed
Weng, etanal. ester-containing
[66] developed quaternary
an ester-
ammonium
containing collector
quaternary M-302. They reported
ammonium thatM-302.
collector M-302 showed better collecting
They reported that M-302 power and higher
showed better
solubility
collectingcompared
power and with dodecylamine
higher hydrochloride.
solubility compared Note that EBAB
with dodecylamine and M-302 could
hydrochloride. Notebe used
that EBABat
neutral pH, which is another advantage over the other collectors normally used.
and M-302 could be used at neutral pH, which is another advantage over the other collectors normally However, no industrial
applications
used. However, of theno two collectorsapplications
industrial have been reported yet. collectors
of the two Recently, ionic
haveliquids (ILs) as collector
been reported in the
yet. Recently,
reverse iron ore
ionic liquids flotation
(ILs) have also
as collector been
in the tested.iron
reverse ILs ore
are aflotation
group ofhave
saltsalso
having
been poorly coordinated
tested. ions,
ILs are a group
and are in liquid state at a temperature below 100 ◦ C or even at ambient temperature. Sahoo and his
of salts having poorly coordinated ions, and are in liquid state at a temperature below 100 °C or even
co-authors
at ambienthave tested quaternary
temperature. Sahoo and ammonium-based
his co-authors have ionictested
liquidsquaternary
(Aliquat-336 and Tricaprylmethyl
ammonium-based ionic
ammonium salicylate)
liquids (Aliquat-336 asTricaprylmethyl
and flotation collector of quartz salicylate)
ammonium [67–69]. They found collector
as flotation that the ofionic liquids
quartz [67–
exhibited stronger adsorption at the quartz surface via electrostatic adsorption
69]. They found that the ionic liquids exhibited stronger adsorption at the quartz surface via compared with DDA
or cetyltrimethylammonium
electrostatic adsorption compared bromide (CTAB).
with DDAThose ionic liquids can be usedbromide
or cetyltrimethylammonium at a wide range of
(CTAB). pH
Those
and
ioniccan even can
liquids exhibit greatatselectivity
be used a wide rangeat aofneutral
pH and pH.canCurrently, the great
even exhibit studyselectivity
of using ionic liquidspH.
at a neutral as
collector
Currently, forthe
iron ores of
study is using
still ationic
the early
liquidsstage and no ILs
as collector for are
ironbeing used
ores is stillinatthe
theindustrial
early stageiron
andore no
flotation operations.
ILs are being used in the industrial iron ore flotation operations.

Figure6.6.Schematic
Figure Schematicmodel
modelof
ofaacationic
cationicgemini
geminicollector
collectoradsorption
adsorptionatatthe
thequartz
quartzsurface.
surface.

ItItisisnoteworthy
noteworthythat thatthe
thepresence
presenceof ofcations
cationshashasaadetrimental
detrimentalimpactimpacton onquartz
quartzflotation
flotationwith
with
cationic 2+ 2+ are2+
cationiccollectors.
collectors.TheThecommonly
commonly present
present metal cations
metal in iron
cations ore flotation
in iron such such
ore flotation as Caas Ca and2+ Mg
and Mg
either sourced
are either fromfrom
sourced hard hard
waterwater
used used
at theatprocessing plantsplants
the processing or released from gangue
or released from gangueminerals (e.g.,
minerals
calcite and Mg-bearing siderite) [70]. The concentration of cations can be
(e.g., calcite and Mg-bearing siderite) [70]. The concentration of cations can be very high aftervery high after accumulation
in the process water
accumulation in theand significantly
process water anddepress the flotation
significantly of quartz
depress when cationic
the flotation collectors
of quartz whenare used.
cationic
As shown in Figure 7, the adsorption of metal cations on quartz surface can
collectors are used. As shown in Figure 7, the adsorption of metal cations on quartz surface can reverse the surface charge
from
reversenegative to positive,
the surface charge which
fromprevents
negativethe adsorption
to positive, of cationic
which prevents amine on the quartz
the adsorption of surface
cationicowing
amine
to
onelectrostatic
the quartz repulsion
surface owing[43,71].
to Hence, the concentration
electrostatic of metal
repulsion [43,71]. cations
Hence, theinconcentration
process waterof should
metal
be closely monitored. The concentration of cations can be reduced by increasing
cations in process water should be closely monitored. The concentration of cations can be reduced by pH to hydroxylate
the cations and
increasing pH to form precipitates
hydroxylate the[72]. Different
cations and formfrom cations, anions
precipitates exhibit a promotive
[72]. Different from cations, effect on
anions
the reverse flotation of iron oxides, in improving the recovery of Fe and
exhibit a promotive effect on the reverse flotation of iron oxides, in improving the recovery lowering the SiO content
2 of Fe andin
the concentrate. This is probably owing to the formation of the inner and
lowering the SiO2 content in the concentrate. This is probably owing to the formation of the inner andouter sphere binuclear or
polynuclear
outer sphere surface complexes
binuclear on mineralsurface
or polynuclear surfaces at near-neutral
complexes pH values,
on mineral providing
surfaces less suitable
at near-neutral pH
leaving groups for detachment into the water, thus successfully inhibiting
values, providing less suitable leaving groups for detachment into the water, thus successfully the dissolution of metal ions
with a higher 2−
inhibiting thevalence [43]. of
dissolution In metal
addition,
ionsanions
with aathigher
a higher valence
valence (e.g.,
[43]. In SO 4 ) tend
addition, to have
anions at aa higher
more
significant effect −
valence (e.g., SOthan anions
42−) tend with aa more
to have lowersignificant
valence (e.g., Cl than
effect ) [73].
anions with a lower valence (e.g., Cl−)
[73].In summarizing the above two sections, one can see that cationic collectors adsorb on the mineral
surface via physisorption (electrostatic interaction), while anionic collectors adsorb on the mineral
surface mainly by chemisorption. The electrostatic interaction occurs by means of the positively
charged cationic collector adsorbing on the negatively charged mineral surface. The chemisorption of
anionic collectors on mineral surfaces is governed by forming covalent bonding between the anionic
collectors and the multivalent ions adsorbed on the mineral surface. The adsorption of these two
types of collectors at the mineral surface is pH-dependent. The collectors for reverse iron ore flotation
Minerals 2020, 10, 681 9 of 22

were advanced via modifying the collector structure, with the aim of improving their solubility
and selectivity.
Minerals 2020, 10, x FOR PEER REVIEW 9 of 22

Figure 7. Effect of cations on the adsorption of amines on the quartz surface.


Figure 7. Effect of cations on the adsorption of amines on the quartz surface.
2.3. Collector Mixtures
In summarizing the above two sections, one can see that cationic collectors adsorb on the mineral
The use of collector mixtures has become common in flotation practice. Different mixtures
surface via physisorption (electrostatic interaction), while anionic collectors adsorb on the mineral
of collectors were tested to remove quartz from other minerals. These mixtures mainly include
surface mainly by chemisorption. The electrostatic interaction occurs by means of the positively
anionic/anionic collectors, anionic/cationic collectors, and ionic/non-ionic combinations [10].
charged cationic collector adsorbing on the negatively charged mineral surface. The chemisorption
The application of mixed anionic collectors in reverse iron ore flotation has been long
of anionic collectors on mineral surfaces is governed by forming covalent bonding between the
investigated [74–76]. The use of mixture of anionic collectors aims to increase the collector solubility
anionic collectors and the multivalent ions adsorbed on the mineral surface. The adsorption of these
in order to reduce the collector consumption. As mentioned in Section 2.1, oleic acid and its soaps
two types of collectors at the mineral surface is pH-dependent. The collectors for reverse iron ore
have low solubility, posing a limitation on their use in iron ore flotation practice. The solubility
flotation were advanced via modifying the collector structure, with the aim of improving their
of oleic acid and its soaps could be improved by adding some hydrophilic functional groups to
solubility and selectivity.
their molecular structures. Another approach to improving their solubility is simply adding some
anionic collectors that have better solubility in water. In general, for the fatty acids with the same
2.3. Collector Mixtures
carbon atom number, the solubility increases with the increasing number of double bonds in their
The use
molecules (i.e.,ofdegree
collector mixtures has become
of unsaturation) [76]. Rama common
Murthy in and
flotation practice. [74]
Mallikajunan Different mixtures
used soaps made of
collectors
from Bombax were tested tooilremove
malabarica and sharkquartz
liver oilfrom otherbarium-activated
to float minerals. Thesequartz. mixtures Bombaxmainly include
malabarica
anionic/anionic
oil contains 43%collectors,
oleic acid anionic/cationic
and 31.3% linoleic collectors,
acid. The and ionic/non-ionic
shark combinations
liver oil contains [10]. acid,
24.9% palmitic
11.2%The application
palmitoleic acid,of11.1%
mixed anionic
stearic acid, collectors
19.6% oleicinacid, reverse iron ore
and 22.3% flotation
gadoleic acid.hasThebeen
degree long
of
investigated [74–76].
unsaturation of Bombax Themalabarica
use of mixture oil isofhigher
anionicthan
collectors
that ofaims to increase
the shark the collector
liver oil. solubility
These researchers
in order
found thatto the
reduce
soapthe collector
from Bombax consumption.
malabarica As mentioned
oil was superior in toSection 2.1, from
the soap oleic the
acidshark
and its soaps
liver oil
have
in lowflotation.
quartz solubility,Lin, posing
et al.a[75]
limitation
developed on their use incollector
an anionic iron oreRA-315,
flotationwithpractice.
the mainThe components
solubility of
oleic acid
being fattyandacids itsand
soaps could
abietic be improved
acids. The abieticbyacids adding
havesomea much hydrophilic
higher degreefunctional groups to than
of unsaturation their
molecular
fatty acids.structures.
This collector Another approach
exhibits good to improving
solubility andtheir solubility
selectivity andishas
simplybeenadding
widelysomeusedanionic
in the
collectors
reverse that have
flotation better
of iron oresolubility
in China.inWei,
water. In [76]
et al. general,
usedfor the fattycotton
a purified acids with
seed thefattysame
acidcarbon atom
as collector
number, the solubility increases with the increasing number of double
in the reverse flotation of an iron ore. The purified cotton seed fatty acid contains oleic acid andbonds in their molecules (i.e.,
degree of
linoleic unsaturation)
acid. The purification[76]. Rama
processMurthy
aimedand Mallikajunan
to increase [74] used
the content soaps acid
of linoleic madeinfrom Bombax
the collector
malabarica
blend as theoil and shark
degree liver oil to of
of unsaturation float barium-activated
linoleic acid is higherquartz.
than thatBombax
of oleicmalabarica
acid. Wei, oiletcontains
al. [76]
43% oleicthat
reported acid
thisand 31.3%exhibited
collector linoleic acid.
good The shark atliver
selectivity oil contains
ambient 24.9% palmitic acid, 11.2%
temperature.
palmitoleic acid, 11.1%of stearic
The effectiveness acid, 19.6% oleicmixtures
using anionic–cationic acid, andin 22.3% flotationgadoleic
has also acid. Therecognized.
been degree of
unsaturation of Bombax malabarica oil is higher than that of
Vidyadhar and Hanumantha Rao [11] found that using the mixture of a cationic collector the shark liver oil. These researchers
found that the soap from Bombax
(tallow-1,3-diaminopropane) and malabarica
an anionic oil was superior
collector (sodium to dodecyl
the soap sulfonate)
from the shark could liver oil in
achieve
aquartz
superb flotation.
separation Lin,ofetfeldspar
al. [75] from
developed
quartzanatanionic collector
pH 2. Wang, RA-315,
et al. with thethat
[12] observed mainthecomponents
mixture of
being fatty
sodium acids
oleate andand abietic acids.acetate
dodecylamine The abietic
couldacids havethe
enhance a much higherof
separation degree
muscoviteof unsaturation
from quartz than
at
fatty
pH 10.acids.
Huang, This et collector exhibits good
al. [77] investigated solubility and
the adsorption selectivity and has been widely
of dodecyltrimethylammonium bromideused in the
(DTAB),
reverse flotation of iron
dodecylpyridinium ore in(DPB),
bromide China.sodium
Wei, et dodecylbenzenesulfonate
al. [76] used a purified cotton seed and
(SDBS), fattysodium
acid as collector
dodecyl
in the reverse flotation of an iron ore. The purified cotton seed fatty
sulfate (SDS) on silica gel at pH 5.6 in individual aqueous solutions, and DTAB–SDBS and DPS–SDS acid contains oleic acid and
linoleic acid. The purification process aimed to increase the content of linoleic acid in the collector
blend as the degree of unsaturation of linoleic acid is higher than that of oleic acid. Wei, et al. [76]
reported that this collector exhibited good selectivity at ambient temperature.
The effectiveness of using anionic–cationic mixtures in flotation has also been recognized.
Vidyadhar and Hanumantha Rao [11] found that using the mixture of a cationic collector (tallow-1,3-
diaminopropane) and an anionic collector (sodium dodecyl sulfonate) could achieve a superb
separation of feldspar from quartz at pH 2. Wang, et al. [12] observed that the mixture of sodium
oleate and dodecylamine acetate could enhance the separation of muscovite from quartz at pH 10.
Huang,2020,
Minerals et al. [77] investigated the adsorption of dodecyltrimethylammonium bromide (DTAB),
10, 681 10 of 22
dodecylpyridinium bromide (DPB), sodium dodecylbenzenesulfonate (SDBS), and sodium dodecyl
sulfate (SDS) on silica gel at pH 5.6 in individual aqueous solutions, and DTAB–SDBS and DPS–SDS
binary mixed solutions. It is expected that the cationic collectors should be strongly adsorbed on
binary mixed solutions. It is expected that the cationic collectors should be strongly adsorbed on
negatively charged silica gel, but no significant adsorption of anionic collectors should occur owing to
negatively charged silica gel, but no significant adsorption of anionic collectors should occur owing
electrostatic repulsion. However, in mixed collector systems, the adsorption of cationic and anionic
to electrostatic repulsion. However, in mixed collector systems, the adsorption of cationic and anionic
collector ions would increase simultaneously, and the surface excess of cations could match that of
collector ions would increase simultaneously, and the surface excess of cations could match that of
anionic ions. Figure 8 illustrated a possible reason that the anionic collector ions are co-adsorbed
anionic ions. Figure 8 illustrated a possible reason that the anionic collector ions are co-adsorbed
specifically with cations as ion pairs at the non-charged sites of quartz through the van der Waals
specifically with cations as ion pairs at the non-charged sites of quartz through the van der Waals
interaction [10]. An increased adsorption density of collectors on the quartz surface would result in
interaction [10]. An increased adsorption density of collectors on the quartz surface would result in
enhanced hydrophobicity of the quartz surface.
enhanced hydrophobicity of the quartz surface.

Figure 8. Co-adsorption of cationic and anionic collectors at the quartz surface.


Figure 8. Co-adsorption of cationic and anionic collectors at the quartz surface.
The metallurgical results for the reverse flotation of iron ores can also be improved using a
combination of ionic collectors
The metallurgical results (anionic or cationic
for the reverse collectors)
flotation of ironandores
non-ionic
can also surfactants,
be improved such using
as fattya
alcohols [13,78,79]. Non-ionic alcohols cannot adsorb on the quartz surface,
combination of ionic collectors (anionic or cationic collectors) and non-ionic surfactants, such as fattywhile their presence could
enhance
alcohols the adsorption
[13,78,79]. of ionic collector.
Non-ionic alcohols Vidyadhar,
cannot adsorb et al. on
[79]the
observed
quartzthat the presence
surface, while of 1-dodecanol
their presence
enhanced
could enhancethe adsorption
the adsorption of theofdodecylamine
ionic collector.ions at the quartz
Vidyadhar, et al.surface at pH 6–7
[79] observed thatand
theresulted
presenceinof a
higher quartz recovery. Filippov et al. [13] found that the addition of
1-dodecanol enhanced the adsorption of the dodecylamine ions at the quartz surface at pH 6–7 and a C13-rich iso-alcohol to ether
diamine
resulted increased
in a higher thequartz
quartz recovery.
flotation recovery
Filippovfrom 52[13]
et al. to 84% at pH
found that10. the
In addition,
additionFilippov,
of a C13-richet al. [78]
iso-
observed that adding iso-alcohols to ether diamine could also improve
alcohol to ether diamine increased the quartz flotation recovery from 52 to 84% at pH 10. In addition,the floatability of Fe-bearing
mica silicates
Filippov, et al.at pH
[78]8,observed
while use thatof theadding
amine alone exhibited
iso-alcohols to little
etherselectivity
diamine on the silicates.
could also improveIt is likely
the
that the adsorption
floatability of ionic
of Fe-bearing collector
mica silicatesis enhanced by the
at pH 8, while usepresence of nonionic
of the amine surfactantlittle
alone exhibited owing to both
selectivity
the hydrophobic
on the silicates. Itchain–chain
is likely that interaction and theof
the adsorption reduction of electrostatic
ionic collector is enhanced repulsion
by thebetween
presence ionic
of
head groups that are shielded from each other by the nonionic surfactant
nonionic surfactant owing to both the hydrophobic chain–chain interaction and the reduction of molecule [10,80]. Figure 9
shows a schematic
electrostatic repulsionof coadsorption
between ionicof cationic
head groupscollector
that areand alcohol
shielded fromon the
eachsurface
other byof the
quartz. The
nonionic
phenomenon may also be explained by Leja–Schulman’s penetration
surfactant molecule [10,80]. Figure 9 shows a schematic of coadsorption of cationic collector and theory [81]; that is, a diffused
monolayer
alcohol on of thecollector
surfacemolecules
of quartz.is The formed at the quartz
phenomenon maysurface,
also and in the meantime,
be explained the insoluble
by Leja–Schulman’s
non-ionic surfactant (e.g., fatty alcohols) forms the diffused molecular
penetration theory [81]; that is, a diffused monolayer of collector molecules is formed at the monolayers at the air/water
quartz
interface. As soon as the air bubble contacts the quartz surface, the surfactant
surface, and in the meantime, the insoluble non-ionic surfactant (e.g., fatty alcohols) forms molecules at the air/waterthe
interface can penetrate
diffused molecular the diffused
monolayers collector
at the air/watermonolayer
interface. atAs
thesoon
quartz surface
as the and adsorb
air bubble strongly
contacts onto
the quartz
the quartz
surface, thesurface, greatly
surfactant increasingatquartz
molecules recovery interface
the air/water by increasing can the local surfactant
penetrate the diffused concentration
collector
and local hydrophobic character of the quartz surface. It is, therefore,
monolayer at the quartz surface and adsorb strongly onto the quartz surface, greatly increasing expected that adding alcoholic
frothers may enhance
quartz recovery the flotation
by increasing performance,
the local surfactantalthough these frothers
concentration and local are normally not
hydrophobic needed in
character of
iron ore flotation.
the quartz surface. It is, therefore, expected that adding alcoholic frothers may enhance the flotation
performance, although these frothers are normally not needed in iron ore flotation.
Minerals 2020, 10, 681 11 of 22
Minerals 2020, 10, x FOR PEER REVIEW 11 of 22

Figure 9. Coadsorption of cationic and nonionic surfactant at the quartz surface.


Figure 9. Coadsorption of cationic and nonionic surfactant at the quartz surface.
3. Collectors for Non-Quartz Gangue Minerals
3. Collectors for Non-Quartz Gangue Minerals
3.1. Alumina-Bearing Minerals
3.1. Alumina-Bearing
The alumina-bearing Mineralsminerals present in the iron ore include gibbsite, kaolinite, montmorillonite,
illite, diaspora, and corundum. Gibbsite (Al(OH)3 ) and kaolinite (Al2 Si2 O5 (OH)4 ) are the two main
The alumina-bearing minerals present in the iron ore include gibbsite, kaolinite,
alumina-bearing minerals present in iron ores [82]. Removal of kaolinite and gibbsite from iron ores is
montmorillonite, illite, diaspora, and corundum. Gibbsite (Al(OH)3) and kaolinite (Al2Si2O5(OH)4) are
aimed at reducing the content of Si and Al in the flotation concentrate.
the two main alumina-bearing minerals present in iron ores [82]. Removal of kaolinite and gibbsite
Separation of gibbsite from iron oxides such as hematite by means of flotation is difficult owing
from iron ores is aimed at reducing the content of Si and Al in the flotation concentrate.
to little difference in surface characteristics between gibbsite and hematite. Generally, hematite and
Separation of gibbsite from iron oxides such as hematite by means of flotation is difficult owing
gibbsite have a similar crystal structure. The cations are trivalent and have common chelation
to little difference in surface characteristics between gibbsite and hematite. Generally, hematite and
characteristics. In addition, the bond distances of Fe–Fe and Al–Al are very similar (i.e., 2.850 Å and
gibbsite have a similar crystal structure. The cations are trivalent and have common chelation
2.852 Å, respectively) [83]. Hence, gibbsite and hematite have a similar surface charge and almost
characteristics. In addition, the bond distances of Fe–Fe and Al–Al are very similar (i.e., 2.850 Å and
identical complexation characteristics. It is, therefore, challenging to separate gibbsite from iron oxides
2.852 Å, respectively) [83]. Hence, gibbsite and hematite have a similar surface charge and almost
via flotation. Thella, et al. [84] argued that the direct production of iron ore concentrate by flotation from
identical complexation characteristics. It is, therefore, challenging to separate gibbsite from iron
iron ores
oxides viacontaining
flotation. high
Thella, alumina (mainly
et al. [84] argued gibbsite)
that thewas challenging,
direct production owing to complex
of iron mineralogy.
ore concentrate by
Desliming
flotation fromwasiron
indispensable
ores containingbeforehigh
flotation
alumina to remove
(mainlyultrafine
gibbsite)gangue particles and
was challenging, to increase
owing to complexiron
grade. It wasDesliming
mineralogy. found that wasthe combination
indispensable of classification
before flotation andto reverse
remove cationic flotation
ultrafine gangue using monamine
particles and
as collector could achieve a high grade iron concentrate having 64.5%
to increase iron grade. It was found that the combination of classification and reverse cationic Fe, 2.2% Al 2 O 3 , while the Fe
recovery was only 28.7% [84]. Kumar, et al. [85] reported that selective
flotation using monamine as collector could achieve a high grade iron concentrate having 64.5% Fe, flocculation for alumina-rich
iron ore
2.2% Al2slimes
O3, while using
thecarboxymethyl
Fe recovery was cellulose as flocculant
only 28.7% could facilitate
[84]. Kumar, et al. [85] thereported
separation thatofselective
gibbsite
from hematite.
flocculation forThe experimental
alumina-rich results
iron ore show
slimes that a finalcarboxymethyl
using concentrate of 64.4% Fe, 4.2%
cellulose Al2 O3 , andcould
as flocculant 1.9%
silica with a yield of 56% could be obtained from a feed containing
facilitate the separation of gibbsite from hematite. The experimental results show that a final56.5% Fe, 7.0% alumina, and 4.9%
silica [85]. of 64.4% Fe, 4.2% Al2O3, and 1.9% silica with a yield of 56% could be obtained from a feed
concentrate
Removal
containing 56.5%of kaolinite
Fe, 7.0% in iron oreand
alumina, flotation is possible,
4.9% silica [85]. but not straightforward. There are multiple
interactions between kaolinite and depressant
Removal of kaolinite in iron ore flotation is possible, and/or collector
but[2].
notItstraightforward.
was found that selectiveThere are depressing
multiple
of hematite against kaolinite using starch is feasible, subject to strict
interactions between kaolinite and depressant and/or collector [2]. It was found that selectivepH control. Ma and Bruckard [86]
observed that starch adsorption on kaolinite would significantly decrease
depressing of hematite against kaolinite using starch is feasible, subject to strict pH control. Ma and with the increasing pH from
7 to 10.5. Starch
Bruckard showedthat
[86] observed an extremely low affinity
starch adsorption towards kaolinite
on kaolinite at pH 10.5, where
would significantly decrease iron oxides
with the
can be well depressed by starch. As reverse cationic flotation of iron
increasing pH from 7 to 10.5. Starch showed an extremely low affinity towards kaolinite at pH 10.5,ores is normally performed at
alkaline pH around 9–10.5, where iron ores can be well depressed while
where iron oxides can be well depressed by starch. As reverse cationic flotation of iron ores is kaolinite is not, it is possible to
separate kaolinite
normally performed from atiron ores ifpH
alkaline kaolinite
aroundcould be rendered
9–10.5, where iron hydrophobic
ores can be using
wella depressed
suitable collector.
while
It was,
kaolinite is however,
not, it is found
possible thattothe amine collectors
separate kaolinite normally
from iron used
ores in if
reverse iron could
kaolinite ore flotation would
be rendered
interact with using
hydrophobic kaolinite in a way
a suitable different from that with quartz. Amine collectors lead to effective
collector.
kaolinite
It was,flotation in acidic
however, found solutions,
that thewhile
amine little adsorption
collectors of collectors
normally used in occurs at alkaline
reverse iron orepH [19,21].
flotation
Figure 10 shows the zeta potential of kaolinite as a function of
would interact with kaolinite in a way different from that with quartz. Amine collectors leadpH. The IEP of kaolinite is at pHto
4.2 [19,87]. Above its IEP, the negative zeta potential of kaolinite increases
effective kaolinite flotation in acidic solutions, while little adsorption of collectors occurs at alkaline with the increasing pH,
hence
pH the adsorption
[19,21]. of collector
Figure 10 shows on kaolinite
the zeta potentialincreases.
of kaolinite However, the flotation
as a function of pH. Theof kaolinite decreases
IEP of kaolinite is
at pH 4.2 [19,87]. Above its IEP, the negative zeta potential of kaolinite increases with the increasing
pH, hence the adsorption of collector on kaolinite increases. However, the flotation of kaolinite
Minerals2020,
Minerals 2020,10,
10,681
x FOR PEER REVIEW 12of
12 of22
22

decreases with an increase in collector adsorption. Hu, et al. [19] investigated the anomalous flotation
with an increase
behavior in collector
in kaolinite solutionadsorption. Hu, et amine
using dodecyl al. [19] (DDA)
investigated the anomalous
as collector based onflotation
crystal behavior
structure
in kaolinite solution using dodecyl amine (DDA) as collector based on crystal
considerations and particle aggregation phenomena. Although the silica (001) and alumina (001) structure considerations
and
basalparticle
planes aggregation
of kaolinite are phenomena.
negatively Although
charged, DDA the silica
has a(001) andinteraction
stronger alumina (001) withbasal planes
the (001) planeof
kaolinite
than with arethenegatively charged,
(001) alumina DDAwhich
plane, has a stronger interactiontowith
can be attributed the the (001) plane
difference than
in the with theat
structure (001)
the
alumina
(001) and plane,
(001which
) planes.can The
be attributed to the difference
self-aggregation betweenin(001the )structure
faces and at the
the(001)
edgeand (001)and
planes planes.
the
The self-aggregation
adsorption of DDA atbetween
the silica(001)
(001)faces
planeand thethe
cause edge planestoand
kaolinite the adsorption
aggregate hydrophobic,of DDA andatgood
the
silica (001) plane
floatability cause theinkaolinite
is achievable to aggregate
acidic solution. hydrophobic,
In alkaline solution,and
the good floatability
kaolinite particlesis are
achievable
dispersed. in
acidic
With thesolution.
presence In alkaline
of DDA,solution, the kaolinite
hydrophobic particles
aggregation are dispersed.
appears to occur inWith the presence
alkaline solution of DDA,
between
hydrophobic
the (001) planes aggregation
owing toappears to occur
the adsorbed in alkaline
DDA, and thussolution between the
the hydrophilic (001)
(001 planes
) faces are owing
exposed to and
the
adsorbed DDA, and thus the hydrophilic (001) faces are exposed and flotation
flotation is not achievable [19]. As silica needs to be removed in an alkaline solution when an amine is not achievable [19].
As silica as
is used needs to be removed
collector, in antoalkaline
it is difficult removesolution
kaolinitewhenand an amine
silica is used
together inas collector,
iron it is difficult
ore flotation using
to remove
amine kaolinite
collectors. It and
maysilica togetherhowever,
be possible, in iron oretoflotation usingremove
sequentially amine collectors.
kaolinite and It may befrom
silica possible,
iron
however, to sequentially remove kaolinite and silica from iron oxides,
oxides, namely removing kaolinite at acidic pHs and removing silica at alkaline pHs. namely removing kaolinite at
acidic pHs and removing silica at alkaline pHs.

10

0
Zeta potential (mV)

-10

-20

-30

-40
0 2 4 6 8 10 12
pH
Figure 10. Zeta potential of kaolinite as a function of pH (after [13]).
Figure 10. Zeta potential of kaolinite as a function of pH (after [13]).
Ammonium quaternary salts can be used as collector for kaolinite in reverse iron ore flotation.
Ammonium
As mentioned quaternary
before, salts can
quaternary be used as collector
ammonium for kaolinite
salt collectors are lessin reverse iron ore
influenced by flotation.
pH [59].
Rodrigues, et al. [88] showed that effective separation between kaolinite and hematite wasRodrigues,
As mentioned before, quaternary ammonium salt collectors are less influenced by pH [59]. achieved
et al. [88]
using DTABshowed
at thethat
pHeffective
range ofseparation
4 to 10. Note between
that, atkaolinite
this pHand hematite
range, was achieved
the quaternary using DTAB
ammonium salt
at the pH range of 4 to 10. Note that, at this pH range, the quaternary ammonium
can also be used as collector for quartz removal, so it is possible to simultaneously remove kaolinite salt can also be
used as
and quartz.collector for quartz removal, so it is possible to simultaneously remove kaolinite and quartz.
The current
The current understanding
understanding of of the
the interaction
interaction between
between anionic
anionic collector
collector and
and kaolinite
kaolinite is is limited.
limited.
According to the work of Xu et al. [89], sodium oleate, the prevailing anionic collector
According to the work of Xu et al. [89], sodium oleate, the prevailing anionic collector used in reverse used in reverse
anionic flotation,
anionic flotation, has
has strong
strong affinity
affinity with
with Al Al sites
sites at
at kaolinite
kaolinite surface
surface atat pH
pH 8–9.
8–9. AtAt this
this pH
pH range,
range,
however, the amount of sodium oleate adsorbed on the quartz surface is
however, the amount of sodium oleate adsorbed on the quartz surface is low. It might be possible low. It might be possible to
remove kaolinite and quartz separately at different pHs. More studies
to remove kaolinite and quartz separately at different pHs. More studies are needed to explore the are needed to explore the
possibilityof
possibility ofusing
usinganionic
anioniccollector
collectorto toremove
removekaolinite
kaolinitefromfromiron
ironores.
ores.
The effectiveness of using a temperature-sensitive
The effectiveness of using a temperature-sensitive polymer, poly polymer, poly (N-isopropyl
(N-isopropyl acrylamide)
acrylamide)
(PNIPAM), as a process aid in the flotation of kaolinite was demonstrated
(PNIPAM), as a process aid in the flotation of kaolinite was demonstrated by Li and Franks by Li and Franks [90].[90].
The
polymer
The polymeracts acts
as dual-function
as dual-function flocculant
flocculant andand collector.
collector.TheThe polymer
polymer preferentially
preferentially adsorbs
adsorbs on
kaolinite via hydrogen bonding to cause flocculation of kaolinite particles
on kaolinite via hydrogen bonding to cause flocculation of kaolinite particles and to render the and to render the particle
surface surface
particle hydrophilic at roomattemperature,
hydrophilic room temperature, whereas at a temperature
whereas higher higher
at a temperature than the polymer’s
than critical
the polymer’s
critical solution temperature (approximately 32 C), adsorption of the polymer molecules inducesa
solution temperature (approximately 32 °C), ◦adsorption of the polymer molecules induces
hydrophilic/hydrophobic transition. A high flotation recovery of kaolinite particles was achieved at
Minerals 2020, 10, 681 13 of 22

a hydrophilic/hydrophobic transition. A high flotation recovery of kaolinite particles was achieved


at 50 ◦ C, which was attributed to flocculation and increased surface hydrophobicity, thus obviating
the need for the addition of a conventional flotation collector. Given the function of flocculating
the fines, this polymer should exhibit good performance for the ores carrying a significant amount
of fine particles. However, this polymer requires heating during flotation, which will increase the
operating cost.

3.2. Phosphorus-Bearing Minerals


Phosphorus is a harmful element in steel-making, causing product defects such as increased
hardness and brittleness and decreased ductility [6]. The phosphorus in iron ore deposits occurs
primarily as apatite (general molecular format Ca5 (PO4 )3 (F, Cl, OH)). In what follows, the interactions
between collectors and apatite in the separation of apatite from iron ores via flotation are discussed.
In anionic flotation of apatite, fatty acids (e.g., oleic acids) and their soaps are primarily used
as collector [91–93], often in conjunction with hydrocarbon supplements (e.g., kerosene and fuel oil),
to reduce the collector consumption [94]. Su, et al. [95] separated apatite from magnetite with a modified
fatty acid collector (Atrac-1562) at pH 8.5–9.0 and at a pulp temperature of approximately 20 ◦ C.
Kou, et al. [96] carried out phosphate flotation using a refined tall oil fatty acid at a dosage of 0.45 kg/t
at pH 10 with a 9:8 (by mass) concentration ratio of fatty acid to diesel. Cao, et al. [97] employed a
mixed collector (i.e., 54 wt.% oleic acid, 36 wt.% linoleic acid, and 10 wt.% linolenic acid) for apatite
flotation at pH 9.5 and at a pulp temperature of approximately 23 ◦ C. The interaction between the
carboxyl group of the fatty acids and Ca(OH)+ ions exposed at the mineral surface was considered
the mechanism of attaining the flotation of the apatite [98], which is the same as the adsorption
mechanism of oleic acid with activated quartz (see Section 2.1). Furthermore, given that apatite
is a sparingly soluble mineral, it has been reported that its dissolution accounts for the floatability
of apatite. Finkelstein [99] argued that, immediately after leaving the mineral lattice, Ca2+ ions
interact with the oleate molecules. The formed calcium oleate then precipitates and renders the
mineral surface hydrophobic. Horta, et al. [98] found that the apatite more dissolvable in water
could provide more Ca2+ ions and exhibit better floatability. It was concluded that, in general,
igneous apatite (e.g., fluorapatite, Ca10 (PO4 )6 F2 ) bears better solubility than sedimentary apatite
(e.g., carbonate-fluorapatite, (Ca,Na,Mg)10 (PO4 ,CO3 )6 (F,OH)2 ) [98].
The use of oleic acid in flotation requires relatively high temperatures, thus the added operating
cost associated with using oleic acid is a concern [100]. At ambient temperature, alkyl hydroxamic acid
mixed with alcohol is an effective collector for phosphate flotation (Miller et al. [94]). The flotation
response demonstrated a very weak pH dependence for the hydroxamic acid collector, and the natural
pH was found to be satisfactory for apatite flotation in most cases. Although the newly developed
collector showed improved performance for phosphate flotation, it has found limited use in industry
owing to its high cost [97].
Flotation of apatite using cationic collectors was also studied. Figure 11 shows the zeta potential
of apatite as a function of pH. The IEP of apatite is 5.4 [22]. Soto and Iwasaki [101] investigated the
effect of pH on apatite flotation using octadecylamine and found that apatite flotation was insensitive
to the pulp pH. They noted that the electrostatic adsorption may not fully account for the flotation
phenomena; flotation of apatite below its IEP with octadecylamine was attributed to the chemical
interaction between the collector and the mineral surface and, above its IEP, the collector would
be adsorbed on the negatively charged apatite surface through electrostatic attraction. A similar
observation was made by Moudgil and Ince [22] when separating apatite from dolomite using
dodecylamine as collector. More recently, Nunes, et al. [6] reported that wavellite [Al3 (PO4 )2 (OH,
F)3 ·5H2 O], a secondary phosphate mineral, exhibited around 100% floatability at pH 8.2 with Flotigam
EDA (an industrial amine collector) and at pH 9.8 with octylamine. They attributed the adsorption of
amines on wavellite to both chemical and electrostatic interactions.
Minerals 2020,
Minerals 2020, 10,
10, 681
x FOR PEER REVIEW 14 of
14 of 22
22

10

Zeta potential
0

-5

-10

-15
2 4 6 8 10 12
pH
Figure 11. Zeta potential of apatite as a function of pH (after [16]).
Figure 11. Zeta potential of apatite as a function of pH (after [16]).
3.3. Iron-Bearing Carbonates
3.3. Iron-Bearing
Siderite (FeCO Carbonates
3 ) is the most abundant non-oxide iron-bearing carbonate in iron ores [1]. As its
theoretical
Siderite grade
(FeCO is 3only
) is the 47.46%, siderite isnon-oxide
most abundant generally iron-bearing
treated as ganguecarbonate mineral
in ironand needs
ores to be
[1]. As its
removed from the concentrate. Note that iron ores associated with siderite
theoretical grade is only 47.46%, siderite is generally treated as gangue mineral and needs to be or other carbonate minerals
are referred
removed to asthe
from refractory ores [102–104].
concentrate. Note thatThe ironbeneficiation of siderite
ores associated with via the combination
siderite or other carbonateof flash
magnetizing roasting to
minerals are referred andasflotation
refractory has been
ores reportedThe
[102–104]. in China [105], but
beneficiation no industrial
of siderite via theapplication
combination is
commissioned yet.
of flash magnetizing roasting and flotation has been reported in China [105], but no industrial
Sideriteiscannot
application be readilyyet.
commissioned removed together with quartz in the reverse flotation of iron ores using
anionic collectors
Siderite cannotsuch beasreadily
fatty acid, despite
removed that fatty
together acids
with havein
quartz been
thewidely
reverseused in theof
flotation flotation of
iron ores
quartz (see Section 2.1) or many carbonates [106]. The presence of siderite
using anionic collectors such as fatty acid, despite that fatty acids have been widely used in the can impose a detrimental
effect on separating
flotation of quartz (see quartz
Sectionfrom iron
2.1) or oxides when fatty[106].
many carbonates acidsTheare presence
used as collector.
of sideriteAs cana salt-type
impose a
mineral, siderite has a relatively high solubility. The dissolved mineral
detrimental effect on separating quartz from iron oxides when fatty acids are used as collector. species can reach a high
As a
concentration. Figure 12 shows the species distribution diagram of siderite
salt-type mineral, siderite has a relatively high solubility. The dissolved mineral species can reach in aqueous solutions [16].
a
The
highdissolved mineral
concentration. species
Figure 12can
showsundergo severaldistribution
the species reactions such as hydrolysis,
diagram of sideriteadsorption,
in aqueous and surface
solutions
and
[16].bulk precipitations,
The dissolved mineral which might
species caninhibit
undergo selective
severalinteractions
reactions such between the collector
as hydrolysis, and other
adsorption, and
minerals [107–109]. Luo, et al. [16] used sodium oleate, calcium chloride,
surface and bulk precipitations, which might inhibit selective interactions between the collector and starch as collector,
and
activator, and depressant,
other minerals [107–109]. Luo, respectively,
et al. [16]inused
the flotation of a mixture
sodium oleate, calcium of chloride,
hematite, and siderite,
starch andas quartz at
collector,
pH 11.4. They found that siderite adversely affected the floatability of quartz,
activator, and depressant, respectively, in the flotation of a mixture of hematite, siderite, and quartz which was attributed to
the adsorption
at pH 11.4. They of found
CaCO3that precipitations at the quartz
siderite adversely affectedsurface. As shownofinquartz,
the floatability Figurewhich
12, thewasconcentration
attributed
of 2− increases significantly with the increasing pH. The appearance of CaCO precipitations was
to CO
the3 adsorption of CaCO3 precipitations at the quartz surface. As shown3 in Figure 12, the
caused by reaction of Ca 2+ from the hydrolysis of CaCl and CO 2− from dissolved species of carbonate
concentration of CO 32− increases significantly with 2the increasing 3 pH. The appearance of CaCO3
minerals.
precipitationsStarch
was can adsorb
caused byon the CaCO
reaction of Ca precipitated
3 2+ at the quartz
from the hydrolysis of CaClsurface
2 andandCO3then
2− from depress the
dissolved
quartz [110].
species of carbonate minerals. Starch can adsorb on the CaCO3 precipitated at the quartz surface and
thenItdepress
is possible to separately
the quartz [110]. remove siderite and quartz from iron oxides using anionic collectors.
Sis and Chander [111] observed that, in anionic flotation, at acidic or neutral pH 5–7, siderite had good
floatability, probably because siderite dissolves Fe2+ species below pH 7.6 [46]. Anionic collector can
interact with the dissolved Fe2+ species and precipitate at the siderite mineral surface. The anionic
flotation of quartz is, however, operated at strongly alkaline pH [24,35,112]. A two-step process can
thus be used to remove these two impurities at different pHs. Zhang [113] demonstrated the two-step
separation of siderite and quartz from hematite, in which siderite was floated under a neutral pH
condition in the first step using sodium oleate as collector, and then the concentrate was refloated via
reverse flotation with sodium oleate as collector under a strong alkaline condition to separate quartz
from hematite.
Minerals 2020, 10, 681 15 of 22
Minerals 2020, 10, x FOR PEER REVIEW 15 of 22

Figure 12. Species distribution diagram of siderite in aqueous solution at 25 °C (after [5]).

It is possible to separately remove siderite and quartz from iron oxides using anionic collectors.
Sis and Chander [111] observed that, in anionic flotation, at acidic or neutral pH 5–7, siderite had
good floatability, probably because siderite dissolves Fe2+ species below pH 7.6 [46]. Anionic collector
can interact with the dissolved Fe2+ species and precipitate at the siderite mineral surface. The anionic
flotation of quartz is, however, operated at strongly alkaline pH [24,35,112]. A two-step process can
thus be used to remove these two impurities at different pHs. Zhang [113] demonstrated the two-
step separation of siderite and quartz from hematite, in which siderite was floated under a neutral
pH condition in the first step using sodium oleate as collector, and then the concentrate was refloated
via reverse flotation with sodium oleate as collector under a strong alkaline condition to separate
quartz from Figure 12. Species distribution diagram of siderite in aqueous solution at 25 ◦ C (after [5]).
hematite.
Figure 12. Species distribution diagram of siderite in aqueous solution at 25 °C (after [5]).
Scarce information on flotation separation of siderite from iron oxides using cationic collector is
Scarce information on flotation separation of siderite from iron oxides using cationic collector is
available. Figure 13toshows
It is possible the zeta
separately potential
remove of siderite
siderite as afrom
function
iron of pH, from which one can see
available. Figure 13 shows the zeta potential of and quartz
siderite as a function oxides
of pH, using anionic
from which collectors.
one can see
that
Sis the
and IEP
Chander of siderite
[111] is
observed around
that, in 7. Ignatow
anionic [46]
flotation, atemployed
acidic or a cationic
neutral pH collector
5–7, siderite (i.e.,
had
that the IEP of siderite is around 7. Ignatow [46] employed a cationic collector (i.e., dodecylpyridinium
dodecylpyridinium
good floatability, chloride,
probably DPCl) siderite
because to float siderite
dissolves atFe
various
2+ pHsbelow
species and achieved
pH 7.6 flotation
[46]. Anionicrecoveries
collector
chloride, DPCl) to float siderite at various pHs and achieved flotation recoveries of 60% at a pH
of 60%
can at a pH
interact above 12. At this Fe
pH2+ range, DPCl precipitate
would be adsorbed onto negatively charged
Thesiderite,
above 12. Atwith
this the
pH dissolved
range, DPCl species
would beand
adsorbed ontoat the siderite
negatively mineral
charged surface.
siderite, matchinganionic
the
matching
flotation the
of characteristics
quartz is, however, of electrostatic
operated at interaction.
strongly alkaline Ignatow
pH [46] proposed
[24,35,112]. A an
two-step adsorption
process can
characteristics of electrostatic interaction. Ignatow [46] proposed an adsorption mechanism of DPCl
mechanism
thus siderite;
be used oftoDPCl withthese
siderite; that is, theations were adsorbed through an exchange +reaction
with thatremove
is, the ions twoadsorbed
were impurities different
through pHs. Zhang
an exchange [113]
reaction demonstrated
involving the DP theiontwo-
and
involving
step the
separationDP + ion and FeOH species at the siderite surface. Abido [114] also observed that siderite
of siderite and quartz from hematite, in which siderite was floated under a neutral
FeOH species at the siderite surface. Abido [114] also observed that siderite had a strong interaction
had
pH acondition
strong interaction
in the atfirstwith dodecylamine at pH as 10.5.
with dodecylamine pHstep
10.5.using sodium oleate collector, and then the concentrate was refloated
via reverse flotation with sodium oleate as collector under a strong alkaline condition to separate
quartz from hematite.60
Scarce information on flotation separation of siderite from iron oxides using cationic collector is
available. Figure 13 shows 40
the zeta potential of siderite as a function of pH, from which one can see
that the IEP of siderite is around 7. Ignatow [46] employed a cationic collector (i.e.,
Zeta potential (mV)

dodecylpyridinium chloride, DPCl) to float siderite at various pHs and achieved flotation recoveries
20
of 60% at a pH above 12. At this pH range, DPCl would be adsorbed onto negatively charged siderite,
matching the characteristics of electrostatic interaction. Ignatow [46] proposed an adsorption
mechanism of DPCl with 0 siderite; that is, the ions were adsorbed through an exchange reaction
involving the DP ion and FeOH species at the siderite surface. Abido [114] also observed that siderite
+

had a strong interaction -20 with dodecylamine at pH 10.5.

60
-40
0 2 4 6 8 10 12
pH
40
Figure Zetapotential
13.Zeta potentialof
ofsiderite
sideriteas
asaafunction
functionof
ofpH
pH(after
(after[37]).
[37]).
Zeta potential (mV)

Figure 13.
20
3.4. Iron-Bearing Silicates
Removal of the iron-bearing silicates from iron ores is challenging because of the presence of both
0
metallic cations and Si in the crystal lattice of these iron-bearing silicates. These cations and Si sites
may have different affinities with reagents present in the flotation system. Manser [115] categorized the
-20 by structure: orthosilicates, pyroxene, amphiboles, and framework silicates.
silicates into four groups
It was reported that the orthosilicates float well with anionic collectors; pyroxenes float with this
collector in some cases,-40but there is no flotation of amphiboles or framework silicates; and the flotation
properties of silicates with0 cationic2collectors4 are reversed,
6 8 orthosilicates
with 10 12 pyroxenes having
and
pH
less floatability than the amphiboles and framework silicates (see Table 4) [115].
Figure 13. Zeta potential of siderite as a function of pH (after [37]).
Minerals 2020, 10, 681 16 of 22

Table 4. Floatability of different silicates by structure in the presence of cationic and anionic collector
(after [115]).

Silicate Class
Collector Type
Ortho- Pyroxene Amphibole Framework-
Anionic Good Poor Nil Nil
Mediocre Good Very good
Cationic
(sensitive to pH) (not sensitive to pH)

A group of typical iron-bearing silicates in iron ores is amphibole, with general chemical
composition NaCa2 (Mg,Fe,Al)5 (Al,Si)8 O22 (OH)2 (the proportions of Na, Ca, Fe, and Mg substitute for
one another in the crystal structure). Amphiboles have been investigated through flotation to evaluate
the feasibility to separate them from iron oxides [13]. In what follows, we will discuss the flotation
behavior of iron-bearing silicates mainly using amphiboles as an example.
Amphiboles can be removed by the use of mixture of amine collectors or an amine collector
plus alcoholic surfactants. Note that there is a relatively large uncertainty for the IEP of amphiboles.
Severov, et al. [23] noted that the IEP of amphiboles is related to a substitution of Al3+ for Si4+ as well
as distribution of Mg2+ in the crystal structure. The cations of Na, Ca, Fe, and Mg can also substitute
for one another in the crystal structure. Filippov, et al. [13] reported that amphiboles could achieve a
collector adsorption degree similar to quartz over a broad alkaline range of pH, owing to the presence
of Si site at the surface. However, starch adsorption at the surface of amphiboles inhibits their recovery
(owing to the presence of Fe site at the surface) when primary monoamines are used as collector
of amphiboles. Hence, a denser adsorption layer of collector is needed to render the silicate surface
hydrophobic enough to achieve high flotation recovery. It has been discussed in Section 2.3 that the
use of a proper mixture of surfactants could result in higher adsorption density and more hydrophobic
mineral surface than the use of a single collector. Filippov, et al. [13] found that the mixtures of ether
diamine (1,3-Propanediamine of chain lengths 10 C) with primary monoamine (DDA) or with alcohols
(C13-rich) could render the surface of an amphibole hydrophobic, even in the presence of starch,
resulting in effective flotation of the silicates with high quality magnetite concentrates (SiO2 content <
1.0% and iron content up to 70.3%). There is little information available about the reverse flotation of
iron-bearing silicates using anionic collectors. Note, however, that anionic collectors have been used
in direct flotation to separate iron oxides from iron-bearing silicates. For example, Mei, et al. [116]
achieved flotation separation of aegirine (NaFe2 SiO6 ) from hematite with ammonium hexafluorosilicate
((NH4 )2 SiF6 ) as depressant for aegirine and NaOl as collector for hematite at pH 4–5. They proposed
that the depressant ions SiF6 2− interact with aegirine surface via chemisorption.

4. Recommendation for Future Work


Flotation will remain a key process solution for the beneficiation of iron ores in the fine or ultrafine
size fractions in the foreseeable future. From the aspect of the collector–mineral interactions, more work
needs to be done in water chemistry and reagent development. Water chemistry studies should
consider the effects of the quality of recycled water in flotation performance. There can be soluble
salts with polyvalent metal cations, residual collector, or depressant contained in the recycled water.
A better understanding of the interactions between various minerals and new collectors and their
mixture is needed. With the rapid depletion of easy-to-process iron ores, there is a pressing need to
develop novel solutions to the flotation of iron ores containing multiple gangues. The gangue minerals
may exhibit similar surface properties to iron oxides. Thus, it is necessary to develop a more selective
depressant for an efficient separation. In addition, different gangues may preferentially interact with
different collectors. It is likely that mixed collectors could exhibit better adaptability to the presence of
various gangue minerals and remove the gangues more effectively than single-collector schemes.
Minerals 2020, 10, 681 17 of 22

5. Conclusions
Reverse cationic and anionic flotation routes are currently used in the iron ore beneficiation industry.
Anionic collectors such as fatty acids adsorb at the mineral surface via chemisorption. Cationic collectors
such as amines adsorb at the mineral surface via electrostatic interaction. For both cationic and
anionic collectors, only a certain type of species can interact with quartz. Hence, the pH needs to be
carefully controlled, at which the desired collector species reaches its maximum concentration. For the
advance of collector development, modifying the molecular structure of collectors to improve their
activity and selectivity has been an on-going subject. Water chemistry needs to be closely monitored
and controlled. In general, the presence of multivalent cations and anions in high concentration can
impose a detrimental effect on quartz flotation by adsorbing on the quartz surface and resisting the
adsorption of collectors.
The mixture of surfactants has been attracting increasing attention. Anionic surfactant ions can
co-adsorb specifically with cations as ion pairs at the non-charged sites of gangue mineral through
van der Waals interaction; the presence of nonionic surfactant also increases the adsorption of cationic
collector at the solid surface by reducing the electrostatic repulsion between ionic head groups. Both of
these mechanisms can lead to increased adsorption density at the mineral surface and enhanced
hydrophobicity of the mineral surface.
There are several other non-quartz gangues commonly present in iron ores. The present work
provides a review of the interactions between collectors and the non-quartz gangues. In general, there is
no universal collector regime and solution chemistry for the removal of various non-quartz gangues.
The selection of collector types should be tailored based on the surface characteristics of the gangues.
To process iron ores in the presence of multiple gangues, using a more selective depressant would be
necessary for an efficient separation.

Author Contributions: Conceptualization, C.L. and Y.C.; investigation, G.F.; writing—original draft
preparation, G.F.; writing—review and editing, L.W.; funding acquisition, C.L. All authors have read and
agreed to the published version of the manuscript.
Funding: This work was funded by National Natural Science Foundation of China, grant number 51704263,
and China Postdoctoral Science Foundation, grant number 2019M652580.
Acknowledgments: The authors also gratefully acknowledge Yuhua Wang for his fruitful advice on this work.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Filippov, L.; Severov, V.; Filippova, I. An overview of the beneficiation of iron ores via reverse cationic flotation.
Int. J. Miner. Process. 2014, 127, 62–69. [CrossRef]
2. Bruckard, W.; Smith, L.; Heyes, G. Developments in the physiochemical separation of iron ore. In Iron Ore
Processing and Environmental Sustainability; Elsevier Science: Burlington, ON, Canada, 2015; pp. 339–356.
3. Standardization Administration of China. Iron Ore Concentrate, Chinese Standard Full-Text Database;
Standardization Administration of China: Beijing, China, 2018.
4. Wang, J.; Lu, S.; Rong, L.; Li, D. Effect of silicon contents on the microstructures and mechanical properties of
heat affected zones for 9Cr2WVTa steels. J. Nucl. Mater. 2016, 470, 1–12. [CrossRef]
5. Schrama, F.N.H.; Beunder, E.M.; Berg, B.V.D.; Yang, Y.; Boom, R. Sulphur removal in ironmaking and oxygen
steelmaking. Ironmak. Steelmak. 2017, 44, 333–343. [CrossRef]
6. Nunes, A.P.L.; Peres, A.E.C.; De Araujo, A.C.; Valadão, G.E.S. Electrokinetic properties of wavellite and
its floatability with cationic and anionic collectors. J. Colloid Interface Sci. 2011, 361, 632–638. [CrossRef]
[PubMed]
7. Naoi, H.; Ohgami, M.; Liu, X.; Fujita, T. Effects of aluminum content on the mechanical properties of a
9Cr-0.5Mo-1.8W steel. Met. Mater. Trans. A 1997, 28, 1195–1203. [CrossRef]
8. Suhasini, R.; Mallick, A.K.; Vasumathi, N.; Kumar, T.; Rao, S.; Prabhakar, S.; Raju, G.B.; Kumar, S. Evaluation
of Flotation Collectors in Developing Zero Waste Technology for Processing Iron Ore Tailings. Int. J. Eng. Res.
2015, 4, 604–608. [CrossRef]
Minerals 2020, 10, 681 18 of 22

9. Araújo, A.; Viana, P.R.; Peres, A.E.C. Reagents in iron ores flotation. Miner. Eng. 2005, 18, 219–224. [CrossRef]
10. Rao, K.H.; Forssberg, K. Mixed collector systems in flotation. Int. J. Miner. Process. 1997, 51, 67–79. [CrossRef]
11. Vidyadhar, A.; Rao, K.H.; Ari, V. Adsorption mechanism of mixed cationic/anionic collectors in feldspar-quartz
flotation system. J. Colloid Interface Sci. 2007, 306, 195–204. [CrossRef]
12. Wang, L.; Sun, W.; Hu, Y.H.; Xu, L.H. Adsorption mechanism of mixed anionic/cationic collectors in
Muscovite-Quartz flotation system. Miner. Eng. 2014, 64, 44–50. [CrossRef]
13. Filippov, L.; Filippova, I.; Severov, V. The use of collectors mixture in the reverse cationic flotation of magnetite
ore: The role of Fe-bearing silicates. Miner. Eng. 2010, 23, 91–98. [CrossRef]
14. Quast, K. Literature review on the use of natural products in the flotation of iron oxide ores. Miner. Eng.
2017, 108, 12–24. [CrossRef]
15. Liu, X.H.; Yu, Y.F.; Chen, W.; Yan, X.H. Effect of selective flocculation desliming on flotation of fine grained
yuanjiacun iron ore. In Proceedings of the XXVI International Mineral Processing Congress, Delhi, India,
24–28 September 2012; pp. 2980–2992.
16. Luo, X.; Wang, Y.; Wen, S.; Ma, M.; Sun, C.; Yin, W.; Ma, Y. Effect of carbonate minerals on quartz flotation
behavior under conditions of reverse anionic flotation of iron ores. Int. J. Miner. Process. 2016, 152, 1–6.
[CrossRef]
17. Ma, X.; Marques, M.; Gontijo, C. Comparative studies of reverse cationic/anionic flotation of Vale iron ore.
Int. J. Miner. Process. 2011, 100, 179–183. [CrossRef]
18. Nakhaei, F.; Irannajad, M. Reagents types in flotation of iron oxide minerals: A review. Miner. Process. Extr.
Met. Rev. 2017, 39, 89–124. [CrossRef]
19. Hu, Y.; Wei, S.; Hao, J.; Miller, J.D.; Fa, K. The anomalous behavior of kaolinite flotation with dodecyl
amine collector as explained from crystal structure considerations. Int. J. Miner. Process. 2005, 76, 163–172.
[CrossRef]
20. Filippov, L.; Severov, V.; Filippova, I. Mechanism of starch adsorption on Fe-Mg-Al-bearing amphiboles.
Int. J. Miner. Process. 2013, 123, 120–128. [CrossRef]
21. Ma, X.; Bruckard, W.; Holmes, R. Effect of collector, pH and ionic strength on the cationic flotation of kaolinite.
Int. J. Miner. Process. 2009, 93, 54–58. [CrossRef]
22. Moudgil, B.M.; Ince, D.E. Flotation separation of apatite from dolomite using dodecylamine and
sodium chloride. In Particle Technology and Surface Phenomena in Minerals and Petroleum; Sharma, M.K.,
Sharma, G.D., Eds.; Springer: Boston, MA, USA, 1991; pp. 191–197.
23. Severov, V.; Filippov, L.; Filippova, I. Relationship between cation distribution with electrochemical and
flotation properties of calcic amphiboles. Int. J. Miner. Process. 2016, 147, 18–27. [CrossRef]
24. Ma, M. Froth Flotation of Iron Ores. Int. J. Min. Eng. Miner. Process. 2012, 1, 56–61. [CrossRef]
25. Rao, K.H.; Forssberg, K.S.E. Chemistry of Iron Oxide Flotation. In Froth Flotation—A Century of Innovation;
Fuerstenau, M.C., Jameson, G.J., Yoon, R.H., Eds.; Society for Mining, Metallurgy, and Exploration, Inc.:
Littleton, CO, USA, 2007; pp. 498–513.
26. Kou, J.; Xu, S.; Sun, T.; Sun, C.; Guo, Y.; Wang, C. A study of sodium oleate adsorption on Ca2+ activated
quartz surface using quartz crystal microbalance with dissipation. Int. J. Miner. Process. 2016, 154, 24–34.
[CrossRef]
27. Quast, K. The use of zeta potential to investigate the interaction of oleate on hematite. Miner. Eng. 2016, 85,
130–137. [CrossRef]
28. Fuerstenau, D.W.; Cummins, W.F.J. The role of basic aqueous complexes in anionic flotation of quartz.
Trans. Am. Inst. Min. Metall. Pet. Eng. 1967, 238, 196–200.
29. Jung, R.F.; James, R.O.; Healy, T.W. Adsorption, precipitation, and electrokinetic processes in the iron oxide
(Goethite)-oleic acid-oleate system. J. Colloid Interface Sci. 1987, 118, 463–472. [CrossRef]
30. Liu, A.; Fan, J.-C.; Fan, M.-Q. Quantum chemical calculations and molecular dynamics simulations of amine
collector adsorption on quartz (0 0 1) surface in the aqueous solution. Int. J. Miner. Process. 2015, 134, 1–10.
[CrossRef]
31. Ozkan, A.; Ucbeyiay, H.; Duzyol, S. Comparison of stages in oil agglomeration process of quartz with sodium
oleate in the presence of Ca(II) and Mg(II) ions. J. Colloid Interface Sci. 2009, 329, 81–88. [CrossRef]
32. Liu, N.; Wang, Z.; Xiao, J.; Wang, H.; Deng, B.; Zhang, Y.; Chen, C. Novel Selective Depressant of Titanaugite
and Implication for Ilmenite Flotation. Minerals 2019, 9, 703. [CrossRef]
Minerals 2020, 10, 681 19 of 22

33. Wills, B.A.; Finch, J.A. Wills’ Mineral Processing Technology—An Introduction to the Practical Aspects of Ore
Treatment and Mineral Recovery, 8th ed.; Elsevier: Amsterdam, The Netherlands, 2016.
34. Somasundaran, P.; Nagarai, D.R. Chemistry and applications of chelating agents in flotation and flocculation.
In Reagents in the Minerals Industry; IMM: London, UK, 1984; pp. 209–219.
35. Quast, K. Flotation of hematite using C6-C18 saturated fatty acids. Miner. Eng. 2006, 19, 582–597. [CrossRef]
36. Fuerstenau, D.W.; Jia, R.H. The role of molecular structure of surfactants on the interfacial and flotation
behavior of oxide minerals particularly quartz. In Proceedings of the XXIV International Mineral Processing
Congress, Beijing, China, 24–28 September 2008.
37. Ogata, Y.; Sugimoto, T.; Inaishi, M. α-Chlorination of Long-chain Aliphatic Acids. Bull. Chem. Soc. Jpn. 1979,
52, 255–256. [CrossRef]
38. Zhu, Y.; Luo, B.; Sun, C.; Li, Y.; Han, Y. Influence of bromine modification on collecting property of lauric
acid. Miner. Eng. 2015, 79, 24–30. [CrossRef]
39. Luo, B.; Zhu, Y.; Sun, C.; Li, Y.; Han, Y. Flotation and adsorption of a new collector α-Bromodecanoic acid on
quartz surface. Miner. Eng. 2015, 77, 86–92. [CrossRef]
40. Yang, H.; Tang, Q.; Wang, C.; Zhang, J. Flocculation and flotation response of Rhodococcus erythropolis to
pure minerals in hematite ores. Miner. Eng. 2013, 45, 67–72. [CrossRef]
41. Pattanaik, A.; Venugopal, R. Investigation of Adsorption Mechanism of Reagents (Surfactants) System and
its Applicability in Iron Ore Flotation—An Overview. Colloid Interface Sci. Comm. 2018, 25, 41–65. [CrossRef]
42. Ofor, O. Effect of Inorganic Ions on Oleate Adsorption at a Nigerian Hematite-Water Interface. J. Colloid
Interface Sci. 1996, 180, 323–328. [CrossRef]
43. Tang, M.; Wen, S.M. Effects of Cations/Anions in Recycled Tailing Water on Cationic Reverse Flotation of
Iron Oxides. Minerals 2019, 9, 161. [CrossRef]
44. Vidyadhar, A.; Kumari, N.; Bhagat, R.P.; Ari, V. Adsorption Mechanism of Mixed Cationic/Anionic Collectors
in Quartz–Hematite Flotation System. Miner. Process. Extr. Met. Rev. 2013, 35, 117–125. [CrossRef]
45. Wang, J.; Somasundaran, P. Adsorption and conformation of carboxymethyl cellulose at solid-liquid interfaces
using spectroscopic, AFM and allied techniques. J. Colloid Interface Sci. 2005, 291, 75–83. [CrossRef] [PubMed]
46. Ignatow, A. Cationic Flotation of Siderite. Master’s Thesis, Department of Metallurgical Engineering,
Mcgill University, Montreal, QC, Canada, November 1975.
47. Shrimali, K.; Jin, J.; Hassas, B.V.; Wang, X.; Miller, J.D. The surface state of hematite and its wetting
characteristics. J. Colloid Interface Sci. 2016, 477, 16–24. [CrossRef] [PubMed]
48. Suman, S.K.; Kumar, S. Reverse flotation studies on iron ore slime by the synergistic effect of cationic
collectors. Sep. Sci. Technol. 2019, 55, 1–13. [CrossRef]
49. Abaka-Wood, G.B.; Addai-Mensah, J.; Skinner, W. A study of flotation characteristics of monazite, hematite,
and quartz using anionic collectors. Int. J. Miner. Process. 2017, 158, 55–62. [CrossRef]
50. Somasundaran, P.; Wang, D.Z. Solution Chemistry: Minerals and Reagents; Elsevier Science: Amsterdam,
The Netherlands, 2006.
51. Somasundaran, P.; Fuerstenau, D.W. Mechanisms of Alkyl Sulfonate Adsorption at the Alumina-Water
Interface1. J. Phys. Chem. 1966, 70, 90–96. [CrossRef]
52. Zhang, R.; Somasundaran, P. Advances in adsorption of surfactants and their mixtures at solid/solution
interfaces. Adv. Colloid Interface Sci. 2006, 123, 213–229. [CrossRef] [PubMed]
53. Huang, Z.; Zhong, H.; Wang, S.; Xia, L.; Zou, W.; Liu, G. Investigations on reverse cationic flotation of iron
ore by using a Gemini surfactant: Ethane-1,2-bis(dimethyl-dodecyl-ammonium bromide). Chem. Eng. J.
2014, 257, 218–228. [CrossRef]
54. Fuerstenau, D.W.; Jia, R. The adsorption of alkylpyridinium chlorides and their effect on the interfacial
behavior of quartz. Colloids Surf. A Physicochem. Eng. Asp. 2004, 250, 223–231. [CrossRef]
55. Gao, Y.; Du, J.; Gu, T. Hemimicelle formation of cationic surfactants at the silica gel-water interface. J. Chem.
Soc. Faraday Trans. 1 Phys. Chem. Condens. Ph. 1987, 83, 2671. [CrossRef]
56. Metzer, A.; Lin, I.J. Effect of dissolved paraffinic gases on the surface tension and critical micelle concentration
(cmc) of aqueous solutions of dodecylamine hydrochloride (DACl). J. Phys. Chem. 1971, 75, 3000–3004.
[CrossRef]
57. Hoyer, H.W.; Greenfield, A. The critical micelle concentrations of decyl-, dodecyl- and tetradecylamine
hydrochloride. J. Phys. Chem. A 1957, 61, 818–819. [CrossRef]
Minerals 2020, 10, 681 20 of 22

58. Wang, Y.H.; Ren, J.W. The flotation of quartz from iron minerals with a combined quaternary ammonium
salt. Int. J. Miner. Process. 2005, 77, 116–122.
59. Jiang, H.; Liu, G.; Hu, Y.; Xu, L.; Yu, Y.; Xie, Z.; Chen, H. Flotation and adsorption of quaternary ammonium
salts collectors on kaolinite of different particle size. Int. J. Min. Sci. Technol. 2013, 23, 249–253. [CrossRef]
60. Montes, S.; Atenas, G.M. Hematite floatability mechanism utilizing tetradecylammonium chloride collector.
Miner. Eng. 2005, 18, 1032–1036. [CrossRef]
61. Fuerstenau, D.W.; Modi, H.J. Streaming Potentials of Corundum in Aqueous Organic Electrolyte Solutions.
J. Electrochem. Soc. 1959, 106, 336. [CrossRef]
62. Liu, W.; Liu, W.-G.; Zhao, Q.; Peng, X.; Wang, B.; Zhou, S. Investigating the performance of a novel polyamine
derivative for separation of quartz and hematite based on theoretical prediction and experiment. Sep. Purif.
Technol. 2020, 237, 116370. [CrossRef]
63. Houot, R. Beneficiation of iron ore by flotation—Review of industrial and potential applications. Int. J. Miner.
Process. 1983, 10, 183–204. [CrossRef]
64. Papini, R.M.; Brandão, P.R.G.; Peres, A.E.C. Cationic flotation of iron ores: Amine characterization and
performance. Min. Met. Explor. 2001, 18, 5–9. [CrossRef]
65. Zana, R. Alkanediyl-α,ω-bis(dimethylalkylammonium bromide) Surfactants: II. Krafft Temperature and
Melting Temperature. J. Colloid Interface Sci. 2002, 252, 259–261. [CrossRef] [PubMed]
66. Weng, X.; Mei, G.; Zhao, T.; Zhu, Y. Utilization of novel ester-containing quaternary ammonium surfactant as
cationic collector for iron ore flotation. Sep. Purif. Technol. 2013, 103, 187–194. [CrossRef]
67. Sahoo, H.; Rath, S.S.; Das, B. Use of the ionic liquid-tricaprylmethyl ammonium salicylate (TOMAS) as a
flotation collector of quartz. Sep. Purif. Technol. 2014, 136, 66–73. [CrossRef]
68. Sahoo, H.; Rath, S.S.; Jena, S.; Mishra, B.K.; Das, D. Aliquat-336 as a novel collector for quartz flotation. Adv.
Powder Technol. 2015, 26, 511–518. [CrossRef]
69. Sahoo, H.; Sinha, N.; Rath, S.S.; Das, B. Ionic liquids as novel quartz collectors: Insights from experiments
and theory. Chem. Eng. J. 2015, 273, 46–54. [CrossRef]
70. Krishnan, S.; Iwasaki, I. Pulp dispersion in selective desliming of iron ores. Int. J. Miner. Process. 1984, 12,
1–13. [CrossRef]
71. Somasundaran, P. Cationic depression of amine flotation of quartz. Trans. Metall. Soc. AIME 1974, 256, 64–68.
72. Ahmed, S.M.; van Cleave, A.B. Adsorption and flotation studies with quartz: Part, I. Adsorption of calcium,
hydrogen and hydroxyl ions on quartz. Can. J. Chem. Eng. 1965, 43, 23–26. [CrossRef]
73. Tang, M.; Tong, X. The relationship between anion distribution in process water and flotation properties of
iron oxides. Miner. Eng. 2020, 154, 106378–106387. [CrossRef]
74. Rama Murthy, R.K.; Mallikajunan, R. Flotation of quartz with soaps of bombax malabarica oil and shark
liver oil as collectors. J. Indian Inst. Sci. 1959, 41, 30–35.
75. Lin, X.H.; Lu, P.; Chen, R.H.; Chen, J.; Ma, X.; Lin, B. Preparation and application of a new type of efficient
collector—RA-315. Min. Metall. Eng. 1993, 13, 31–35.
76. Wei, Y.H.; Wei, J.X.; Guo, W.D.; Zhou, G.Y. Reverse flotation of iron ore using purified cotton seed fatty acid
at ambient temperature. J. Wuhan Univ. Technol. 2015, 37, 36–40.
77. Huang, Z.; Yan, Z.; Gu, T. Mixed adsorption of cationic and anionic surfactants from aqueous solution on
silica gel. Colloids Surf. 1989, 36, 353–358. [CrossRef]
78. Filippov, L.O.; Duverger, A.; Filippova, I.V.; Kasaini, H.; Thiry, J. Selective flotation of silicates and Ca-bearing
minerals: The role of non-ionic reagent on cationic flotation. Miner. Eng. 2012, 36–38, 314–323. [CrossRef]
79. Vidyadhar, A.; Rao, K.; Chernyshova, I.V.; Pradip; Forssberg, K. Mechanisms of Amine-Quartz Interaction in
the Absence and Presence of Alcohols Studied by Spectroscopic Methods. J. Colloid Interface Sci. 2002, 256,
59–72. [CrossRef]
80. Vidyadhar, A.; Kumari, N.; Bhagat, R.; Ari, V. Adsorption mechanism of mixed collector systems on hematite
flotation. Miner. Eng. 2012, 26, 102–104. [CrossRef]
81. Leja, J.; Schulman, J.H. Flotation theory: Molecular interaction between frothers and collectors at
solid-liquid-air interfaces. Trans. Am. Inst. Min. Metall. Pet. Eng. 1954, 199, 221–228.
82. Sahoo, H.; Rath, S.S.; Rao, D.S.; Mishra, B.K.; Das, D. Role of silica and alumina content in the flotation of
iron ores. Int. J. Miner. Process. 2016, 148, 83–91. [CrossRef]
83. Ravishankar Pradip, S.; Khosla, N. Selective flocculation of iron oxide from its synthetic mixtures with clays:
A comparison of polyacrylic acid and starch polymers. Int. J. Miner. Process. 1995, 43, 235–247. [CrossRef]
Minerals 2020, 10, 681 21 of 22

84. Thella, J.S.; Mukherjee, A.K.; Srikakulapu, N.G. Processing of high alumina iron ore slimes using classification
and flotation. Powder Technol. 2012, 217, 418–426. [CrossRef]
85. Kumar, D.; Jain, V.; Rai, B. Can carboxymethyl cellulose be used as a selective flocculant for beneficiating
alumina-rich iron ore slimes? A density functional theory and experimental study. Miner. Eng. 2018, 121,
47–54. [CrossRef]
86. Ma, X.; Bruckard, W. The effect of pH and ionic strength on starch–kaolinite interactions. Int. J. Miner. Process.
2010, 94, 111–114. [CrossRef]
87. Ndlovu, B.; Forbes, E.; Farrokhpay, S.; Becker, M.; Bradshaw, D.; Deglon, D. A preliminary rheological
classification of phyllosilicate group minerals. Miner. Eng. 2014, 55, 190–200. [CrossRef]
88. Rodrigues, O.M.S.; Peres, A.E.C.; Martins, A.H.; Pereira, C.A. Kaolinite and hematite flotation separation
using etheramine and ammonium quaternary salts. Miner. Eng. 2013, 40, 12–15. [CrossRef]
89. Xu, L.; Hu, Y.; Dong, F.; Gao, Z.; Wu, H.; Wang, Z. Anisotropic adsorption of oleate on diaspore and kaolinite
crystals: Implications for their flotation separation. Appl. Surf. Sci. 2014, 321, 331–338. [CrossRef]
90. Li, H.; Franks, G.V. Role of Temperature-Sensitive Polymers in Hydrophobic Aggregation/Flotation of Silicate
Minerals. In Proceedings of the XXIV International Mineral Processing Congress, Beijing, China, 24–28
September 2008; pp. 1261–1269.
91. Pereira, A.C.; Papini, R.M. Processes for phosphorus removal from iron ore—A review. Rem Rev. Esc. Minas
2015, 68, 331–335. [CrossRef]
92. Subramanian, S.; Rao, K.H.; Forssberg, K.S.E. Dispersion characteristics of apatite and magnetite fines in
the presence of inorganic and organic reagents and its influence on the dephosphorization of magnetite
ore. In Beneficiation of Phosphates III—Fundamentals and Technology; Zhang, P., El-Shall, H., Somasundaran, P.,
Stana, R., Eds.; SME: Littleton, CO, USA, 2002; pp. 21–31.
93. Nunes, A.P.L.; Pinto, C.L.L.; Valadão, G.E.S.; Viana, P.R.D.M. Floatability studies of wavellite and preliminary
results on phosphorus removal from a Brazilian iron ore by froth flotation. Miner. Eng. 2012, 39, 206–212.
[CrossRef]
94. Miller, J.D.; Wang, X.M.; Li, M.H. Selective Flotation of Phosphate Minerals with Hydroxamate Collectors.
U.S. Patent 6341697 B1, 29 January 2002.
95. Su, F.; Hanumantha Rao, K.; Forssberg, K.S.E.; Samskog, P.O. Dephosphorization of magnetite fines: Part
2: Influence of chemical variables on flotation kinetics. Trans. Inst. Min. Metall. Sec. C Miner. Process.
Extr. Metall. 1998, 107, c103–c110.
96. Kou, J.; Tao, D.; Xu, G. Fatty acid collectors for phosphate flotation and their adsorption behavior using
QCM-D. Int. J. Miner. Process. 2010, 95, 1–9. [CrossRef]
97. Cao, Q.; Cheng, J.; Wen, S.-M.; Li, C.; Bai, S.; Liu, D. A mixed collector system for phosphate flotation.
Miner. Eng. 2015, 78, 114–121. [CrossRef]
98. Horta, D.; de Mello Monte, M.B.; de Salles Leal Filho, E.L. The effect of dissolution kinetics on flotation
response of apatite with sodium oleate. Int. J. Miner. Process. 2016, 146, 97–104. [CrossRef]
99. Finkelstein, N.P. Review of interactions in flotation of sparingly soluble calcium minerals with anionic
collectors. Trans. Inst. Min. Metall. Sect. C 1989, 988, 157–178.
100. Sis, H.; Chander, S. Adsorption and contact angle of single and binary mixtures of surfactants on apatite.
Miner. Eng. 2003, 16, 839–848. [CrossRef]
101. Soto, H.; Iwasaki, I. Flotation of apatite from calcareous ores with primary amines. Mining Met. Explor. 1985,
2, 160–166. [CrossRef]
102. Li, L.X.; Yin, W.Z.; Wang, Y.B.; Tao, S.J. Effect of Siderite on Flotation Separation of Martite and Quartz.
J. Northeast. Univ. 2012, 33, 431–434.
103. Yin, W.-Z.; Li, D.; Luo, X.-M.; Yao, J.; Sun, Q.-Y. Effect and mechanism of siderite on reverse flotation of
hematite. Int. J. Miner. Met. Mater. 2016, 23, 373–379. [CrossRef]
104. Gu, X.T.; Zhu, Y.M.; Li, Y.J.; Han, Y.X. Selective flotation of siderite and quartz from a carbonate-containing
refractory iron ore using a novel amino-acid-based collector. Physicochem. Probl. Miner. Process. 2018, 54,
803–813.
105. Chen, W.; Yu, Y.F.; Feng, Z.L.; Lu, X.S.; Zhao, Q.; Liu, X.Y. Six hundred thousand t/a refractory siderite flash
magnetizing roasting complete sets technique and equipment. Met. Mine 2017, 3, 54–58.
106. Laskowski, J.; Nyamekye, G. Colloid chemistry of weak electrolyte collectors: The effect of conditioning on
flotation with fatty acids. Int. J. Miner. Process. 1994, 40, 245–256. [CrossRef]
Minerals 2020, 10, 681 22 of 22

107. Feng, B.; Luo, X.-P. The solution chemistry of carbonate and implications for pyrite flotation. Miner. Eng.
2013, 53, 181–183. [CrossRef]
108. Forssberg, K.E.; Subrahmanyam, T.; Nilsson, L.K. Influence of grinding method on complex sulphide ore
flotation: A pilot plant study. Int. J. Miner. Process. 1993, 38, 157–175. [CrossRef]
109. Shi, Q.; Zhang, G.; Feng, Q.; Deng, H. Effect of solution chemistry on the flotation system of smithsonite and
calcite. Int. J. Miner. Process. 2013, 119, 34–39. [CrossRef]
110. Pinto, C.; Peres, A.E.C.; De Araujo, A. The effect of starch, amylose and amylopectin on the depression of
oxi-minerals. Miner. Eng. 1992, 5, 469–478. [CrossRef]
111. Sis, H.; Chander, S. Reagents used in the flotation of phosphate ores: A critical review. Miner. Eng. 2003, 16,
577–585. [CrossRef]
112. Quast, K. Use of conditioning time to investigate the mechanisms of interactions of selected fatty acids on
hematite. Part 1: Literature survey. Miner. Eng. 2015, 79, 295–300. [CrossRef]
113. Zhang, M. Study on the Floatation Behavior for Donganshan Carbonate-Containing Iron Ore. Ph.D. Thesis,
Northeast University, Shenyang, China, May 2009.
114. Abido, A.M. Contribution to concentration of Tin ores by flotation. J. Inst. Eng. India Ser. D 1973, 53, 66–70.
115. Manser, R.M. Handbook of Silicate Flotation; Warren Spring Laboratory: Stevenage, UK, 1975.
116. Mei, G.J.; Rao, P.; Yu, Y.F. Flotation separation of hematite and iron-containing silicate using ammonium
hexafluorosilicate depressant. In Proceedings of the XXIV International Mineral Processing Congress,
Beijing, China, 24–28 September 2008; pp. 1255–1260.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like