You are on page 1of 42

Journal Pre-proof

Recent progress on research of molybdenite flotation: A review

Gaosong Yi, Eloy Macha, Jeff Van Dyke, Rafael Macha, Tim
McKay, Michael L. Free

PII: S0001-8686(21)00107-X
DOI: https://doi.org/10.1016/j.cis.2021.102466
Reference: CIS 102466

To appear in: Advances in Colloid and Interface Science

Revised date: 12 June 2021

Please cite this article as: G. Yi, E. Macha, J. Van Dyke, et al., Recent progress on research
of molybdenite flotation: A review, Advances in Colloid and Interface Science (2021),
https://doi.org/10.1016/j.cis.2021.102466

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2021 Elsevier B.V. All rights reserved.


Journal Pre-proof

Recent Progress on Research of Molybdenite Flotation: A Review


Gaosong Yi1, Eloy Macha1, Jeff Van Dyke1, Rafael Macha1, Tim McKay1, Michael L Free2
1
Reliable Controls Corporation, Salt Lake City, UT 84117, USA
2
Department of Materials Science and Engineering, University of Utah, Salt Lake City, UT 84112, USA

Abstract
Molybdenum is an important alloy element for metallurgical industry because of its high temperature
stability. As the major mineral reserve for molybdenum, molybdenite (MoS2) is commonly found in
porphyry copper deposits. Molybdenite is naturally floatable and can be separated from copper sulfide

of
mineral using froth flotation. Properties of molybdenite such as mineralogy, microstructure, surface
wettability, zeta potential, etc. can have a great effect on its floatability. Organic and inorganic

ro
depressants and surface pre-treatment methods are applied to improve the recovery of molybdenite.
Electrochemical potential measurements using different electrodes are used to monitor process conditions

-p
and enable processing parameter adjustments to improve flotation circuit performance and reduce
operating costs. Cations like Ca2+ and Mg2+ are reported to have negative effects on the flotation of
re
molybdenite in alkaline solution, and dispersants and oil collectors need to be added to restore the
lP

flotation of molybdenite. In addition, effects of gangue minerals, particle size, and oil collectors and
surfactants on molybdenite recovery are also discussed in this manuscript.
na

Keywords: Molybdenite flotation, depressant, electrochemical potential, cationic depressants, oil


collectors, surfactants
ur

1. Introduction

Molybdenum is widely utilized in high temperature alloys, electrical and electronic devices, thermal spray
Jo

coatings, medical equipment as well as aerospace and defense components because of its strength and
high temperature stability.[1,2] Through the addition of molybdenum, unique combinations of properties
such as thermal and electrical conductivity, high temperature strength and creep resistance, environmental
stability, and resistance to abrasion and wear, can be achieved for alloys for special applications.[1]
According to the data from the International Molybdenum Association (IMOA) [3] over 80% of
molybdenum products are used in metallurgy, and around 14% of molybdenum products are used in the
chemical industry. Molybdenite is the principal ore mineral for molybdenum, and almost half of the
world’s molybdenum reserves are in porphyry copper deposits.[4,5] When the MoS2 grade is 0.01% -
0.07% (or higher), the recovery of molybdenite as a byproduct from copper concentrates is economic.[6]

1
Journal Pre-proof

In a copper-molybdenum processing plant, copper sulfide concentrate (such as chalcopyrite) is floated


together with molybdenite in the first stage.[7] The separation of molybdenite and copper sulfide is
achieved in the second stage through flotation. [8] In most scenarios, copper sulfides, such as chalcopyrite,
are depressed using a depressant (e.g. NaHS, NaCN, or Nokes reagent (P2S5 + NaOH))[9] and
molybdenite is recovered as a flotation concentrate. Afterwards, Mo oxides (MoO3) are produced from
molybdenite concentrate by roasting or acid pressure-oxidation processes.[10,11] The recovery of
molybdenite during flotation is affected by many factors. For instance, mineral type, liberation, and
particle size are reported to have a critical impact on molybdenite recovery and concentrate grade.[12,13]
In addition, process related parameters such as feed solids concentration, pH, collector type, collector to
frother ratio, and depressants can also affect the flotation circuit performance.[14,15] Other factors like

of
organic and inorganic matter/ion concentrations in the water, pulp electrochemical potential, and flotation

ro
gas utilization should also be considered and adjusted as needed for optimal molybdenite recovery.[16–19]
Extensive research has been conducted on molybdenite flotation in past decades to improve the

-p
understanding and optimizing its flotation. In this manuscript, recent progress on molybdenite flotation
and the corresponding mechanisms are discussed.
re
2. Molybdenite mineralogy and surface properties
lP

a) Mineralogy & microstructure

Molybdenite structure and morphology can be different form mine to mine, or even within the same mine
na

deposit. Fig. 1 (a) and (b) show molybdenite mineral specimens from the Zinngrube Mine of Germany
and the Shinkolobwe Mine of the Democratic Republic of Congo, illustrating totally different
ur

morphologies.[20] At the Bingham Canyon mine of Kennecott Utah Corporation, vein-controlled and
disseminated-molybdenite are observed, as illustrated in Fig. 1 (c) and (d).[13] The size of the vein-
Jo

controlled molybdenite is decided by that of the vein, while the disseminated molybdenite is very fine and
associated with copper sulfides. Molybdenite and graphite are similar in appearance and feel because both
of them exhibit a laminar crystalline structure as demonstrated in Fig. 2(a).[6,21] The high-resolution
scanning transmission electron microscopy (HR STEM) image of the edge surface is displayed in Fig. 2(b)
[22], and an obvious laminar atomic structure can be observed. Molybdenite consists of hexagonal
crystals where each molybdenum atom is surrounded by and covalently bonded to six sulfur atoms. Fig. 2
(c) reveals the HR STEM image of the plane view of monolayer MoS2. Molybdenum and sulfur atoms are
differentiated from each other because of the high Z-contrast as highlighted by dots of different colors in
Fig. 2(c).[23]

2
Journal Pre-proof

of
ro
Fig. 1. Optical microscopy image of molybdenite minerals from (a) Zinngrube Mine of Germany and (b)
the Shinkolobwe Mine of Democratic Republic of Congo[20], and image of (c) quartz vein and (d) a thin

-p
layer of smeared molybdenite from Bingham Canyon Mine[13].
re
Different S-Mo-S layers are held together by weak Van de Waals forces and can be easily pulled apart
under stress or shear forces in the grind circuit.[24,25] The newly generated surface (basal surface) is
lP

non-polarized and composed of weak unsaturated S-S bonds, thus it is naturally hydrophobic. When in-
plane S-Mo covalent bonds are broken, hydrophilic edge surfaces are formed.[6] Consequently,
na

molybdenite mineral particles exhibit anisotropic surface properties and are naturally hydrophobic due to
the large aspect ratio (basal to edge). As a matter of fact, the basal surface is not flat and homogeneously
ur

hydrophobic everywhere. Extensive studies[26–28] using Atomic Force Microscopy (AFM) and
Scanning Electron Microscopy (SEM) reveal that terraces and micro-edges, as shown in Fig. 2(d), are
Jo

randomly distributed on the basal surface, and the presence of these micro-edges leads to the local
hydrophilicity of molybdenite basal surface.

3
Journal Pre-proof

of
ro
-p
Fig. 2. Diagram of crystal structure of molybdenite (2H polytype) (a) side view and top view[6], (b) High
angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) image of MoS2
re
view along [1-210][22], (c) HAADF-STEM image of the plane view of monolayer 2H-MoS2, and (d)
micro-edges on basal plane of molybdenite[23].
lP

b) Surface wettability
na

Surface wettability is a key factor for mineral floatability, and it can be estimated by surface contact angle.
As an anisotropic mineral, the wettability of basal and edge surface for molybdenite is different, as shown
ur

in Fig. 3(a)[29]. The contact angle for the hydrophobic basal plane is reported to be around 70°.[30] In
contrast, the hydrophilic edge surfaces have small contact angle ranging from 0 to 48° in most
Jo

cases.[24,29,31,32] Due to different surface preparation methods and associated edge surface roughness
and cleanliness, the corresponding contact angles vary a lot.[6] For instance, the contact angle obtained
from molecular dynamic simulations for armchair edge surface is 54º and for zigzag edge surfaces is
24.3º.[33] In addition, surface oxidation and the adsorption of organic matter (e.g. dextrin and humic acid)
[34–36]or inorganic ions (e.g. calcium ions)[37,38] onto surfaces of molybdenite can also affect its
contact angle. Fig. 3(b) illustrated the contact angle of basal and edge surfaces of molybdenite at different
pHs, and the contact angle for basal surface decreases slightly in alkaline solution.[26,39–41] This has
been attributed to the formation of molybdate ( ) [38] as an oxidation product of MoS2 at micro-
edges of basal plane. Fig. 3(c) shows the measured contact angle of molybdenite minerals of different size
classes, and it decreases as particle size decreases.[42] This could be explained by the fact that smaller

4
Journal Pre-proof

molybdenite particles have a lower aspect ratio, which leads to a smaller basal area relative to the edge
area. Therefore, the overall contact angle of the mineral decreases as size decreases.

of
Fig. 3. (a) Image of water contact angle of molybdenite basal and edge plane[29], (b) contact angle of

ro
basal and edge plane from previous studies, (c) contact angle of molybdenite of different particle size[42].

-p
re
lP
na

Fig. 4 (a) surface potentials of face and edge plane of molybdenite in 10 mM NaCl solution and zeta
ur

potentials from literature sources [39] (zeta potential in the literature * was in 2 mM KCl. Zeta potential
in the literature** was in 5 mM KNO3.), (b) Initial OCP (mV) of MoS2 basal and edge electrodes in 0.1 M
Jo

NaCl at pH 9 and pH 11[31], (c) results of potentiostatic experiments performed on molybdenite


electrodes at 1.4 V for 800s, in 10-3 M KCl (pH=9) solution[43].

c) Surface potential and open circuit potential

Surface potential is another crucial factor for molybdenite flotation and has been extensively studied.
Surface charge is expected to be different for basal and edge surfaces because of the isotropic structure of
molybdenite. Lu et al.[39] reported the surface potential of basal and edge surface of molybdenite by
fitting the force profile measured using AFM with DLVO (Derjaguin−Landau−Verwey−Overbeek)
theory. Both basal and edge surfaces are negatively charged in the pH range of 3-11, as shown in Fig. 4(a).
The point of zero charge (PZC) for the edge surface is detected around pH 3, and PZC for the basal
surface is expected to be lower than pH 3.0. Moreover, surface potentials of both surfaces decrease with

5
Journal Pre-proof

pH, and the value for the edge surface decreases faster than that of the basal surface. Previous research
suggests that the surface potential of the basal surface is negative as a result of adsorption of hydroxide
ions,[38] similar to the charging mechanisms of hydrophobic oil droplets and air bubble surfaces. [44,45]
While the potential of the edge surface is negative because of the formation of / ions after
the oxidation of Mo.[39]

Wang et. al.[31] characterized the surface reactivity of molybdenite by examining the open circuit
potential (OCP). Fig. 4(b) shows the initial OCP of basal and edge plane for molybdenite at pH 9 and 11
in 0.1 M NaCl solution. The initial OCP for the basal surface is 41 mV (Ag/AgCl) at pH 9, and it is
higher than that of the edge surface (-90 mV). When pH is increased to 11, the initial OCP decreases to -

of
17 and -121 mV respectively. The lower initial OCP indicates the edge surface is easier to lose electrons
than the basal surface during the redox reaction[40,46] where the cathodic reaction is:

ro
(1)
And the anodic reaction is:
-p (2)
re
Similarly, the oxidation of both basal and edge surfaces is easier at pH 11. Potential polarization research
lP

performed by Miki et al.[43] demonstrated that the edge surface was more conductive and easier to
oxidize than the basal surface during potential scanning, as displayed in Fig.4(c).
na

Therefore, basal surface oxidization, especially at micro-edges, could be another reason for the negative
charge of basal surface.[26,39,47] Moreover, the high conductivity and the reactivity of edge surface
could lead to the high sensitivity of the surface potential to solution pH as shown in Fig. 4(a)[31,43].
ur
Jo

Fig. 5. (a) Zeta potentials for molybdenite of different particles size[42], and (b) Surface potential of the
MoS2 face and edge in 10 mM MgCl2 and 10 mM NaCl by DLVO fitting; zeta potential of MoS2 particle
in 10 mM MgCl2 by electrophoresis measurement[48].

6
Journal Pre-proof

d) Zeta potential

Fig. 4(a) reveals zeta potential of molybdenite at different pH levels from previous research. The zeta
potential of molybdenite is negative at pH above 2 and becomes more negative at higher pH values[38].
As an average value of both basal and edge surface potential, the relationship between zeta potential and
pH is close to that of the basal surface and pH as shown in Fig. 4(a), indicating the majority of
molybdenite surface is basal.[39] However, for smaller molybdenite particles, the basal-to-edge plane
ratio is low. Consequently, a more negative zeta potential can be expected, as indicated in Fig. 5(a)[42].
Other factors, such as surface oxides[49] and adsorption of organic or inorganic matters[48,50,51] on

of
molybdenite could also affect zeta potential of molybdenite. For example, Fig. 5(b) showed the zeta
potential change of molybdenite in MgCl2 and CaCl2 solution at different pH.[48] The increase of zeta

ro
potential is caused by the adsorption of positively charged Mg2+, Mg(OH)+ and Ca2+ and Ca(OH)+.[51]
However, when pH is higher than 9.5, Ca(OH)2 (s) precipitates onto molybdenite, resulting in a decrease
-p
of zeta potential. It is worth noting that zeta potential is measured under the assumption that a
re
molybdenite particle is spherical and isotropic.[52] Therefore, zeta potential of molybdenite should be
treated as apparent and comparative rather than as absolute due to its non-spherical, plate-like shape.
lP

3. Depressants for molybdenite flotation


na

Copper sulfides, such as chalcopyrite and bornite, are hydrophilic before treatment and therefore can be
easily separated from molybdenite using froth flotation.[9] However, most processing plants recover
chalcopyrite and molybdenite together in the rougher circuit.[4,8,11] Chalcopyrite becomes floatable after
ur

the adsorption of collectors. Therefore, extra treatment or selective depressants are required to separate
Jo

them. Two types of depressants[9], i.e., inorganic and organic depressants, are commonly applied to
separate molybdenite and chalcopyrite. In addition, pretreatment methods such as microwave
treatment[53,54], high temperature aging[55], and electrolysis[40,43] can be used to separate them.

a) Inorganic depressants for Cu sulfide

Sulfur-containing compounds,[56] such as Na2S, NaHS, sodium thiocarbonate (Na2CS3), P-Nokes


(P2S5+NaOH), As-Nokes (As2O3+Na2S), Na2SO3, and Na2S2O3, are widely implemented to depress
chalcopyrite in a Cu-Mo process plant. Nitrogen-containing compound[11,56] is another type of Cu
depressant, which include NaCN, Na4Fe(CN)6, Na3Fe(CN)6, KCN, Zn(CN)2, Ca(CN)2 , etc. The flotation
of chalcopyrite is caused by the formation of the Cu(I)-xanthate complex (denoted by CuX) and
dixanthogen as shown in Eq(3) and (4)[57,58]:

7
Journal Pre-proof

(3)
(4)
With the addition of Cu depressant, for example NaHS, the desorption of xanthate adsorbed on
chalcopyrite surface occurs as shown in Eq.(5), (6), and (7)[9,59]:

(5)

(6)
(7)

of
As a result, chalcopyrite flotation was depressed, as shown in Fig. 6(a)[59]. Recent research by Peng et
al.[60] demonstrated that HS-/S2- preferred to adsorb on chalcopyrite, and xanthate disappeared from

ro
chalcopyrite after treated by sodium sulfide. In fact, the xanthate induced flotation of copper sulfides

-p
could be affected by pulp potential, because Eq.(3) and (4) are favored in a nonoxidizing environment.
Nagaraj et al.[61] investigated the effect of pulp potential on chalcopyrite-molybdenite separation, and
re
their results are shown in Fig. 6(b). The flotation of chalcopyrite was achievable only when the pulp
potential exceeded 0.1V (versus SHE) at pH of 10.5-11.0, which is consistent with the contact angle
lP

results reported by Chander[62]. Further study by Tolley et al.[63] reveals that chalcopyrite and
chalcocite (Cu2S) were totally depressed when potential is lower than -0.4V, as shown in Fig. 6(c).
na

Therefore, the addition of sulfur-containing and nitrogen-containing salts leads to a reductive pulp
potential, resulting in the desorption of xanthate and the depression of copper sulfide flotations.
ur

Furthermore, other actions, such as applying nitrogen[64] as a flotation gas and creating a reductive pulp
potential through an electrolysis[43] method were reported to be helpful for depressing copper sulfide
Jo

flotation in plant or lab scale tests.

Even though NaHS/Na2S is effective in separating molybdenite and chalcopyrite. Over-dosage of


NaHS/Na2S could also cause the depressing of molybdenite. Fig. 6(d) [65] illustrated the recovery of
molybdenite versus different NaHS concentrations. Molybdenite recovery increased greatly after the
addition of 3mM of NaHS, while further increase (from 3 to 12 mM) of NaHS decreases the recovery of
molybdenite. The edge surface of molybdenite is of high energy and excellent electrochemical activity
due to the broken of covalent S-Mo bond. At a low NaHS concentration, HS- could be oxidized through
the reaction as shown in Eq.(8):

( ) ( ) (8)

Further oxidation could occur on the surface of edge surface of molybdenite as show in Eq.(9):

8
Journal Pre-proof

(9)

where Sx represent elemental-sulfur or polysulfide attached to the edge surface. X-ray Photoelectron
Spectroscopy (XPS) results confirmed the formation of ⁄ on molybdenite surface. Polysulfide or
element sulfur on the edge surface makes molybdenite more hydrophobic and thus can enhance flotation
recovery. However, when NaHS concentration is relatively high (e.g. 12mM), it creates a reductive
environment that prevents the formation of polysulfide or element sulfur. XPS results reported by Chen et
al.[59] revealed that HS- and SO42- are identified at the surface of molybdenite when 36 mM NaHS was
added. Micro-edges of basal surfaces are believed to play a critical role during the adsorption of HS -
through the following reaction:

of
( ) (10)
As a result, the hydrophobicity and floatability of molybdenite decreases slightly as excess HS - is present

ro
as shown in Fig.6(d).

-p
re
lP
na
ur
Jo

Fig. 6. (a) chalcopyrite recovery in 10% seawater of different NaHS concentrations[59], (b) recovery of
Cu (chalcopyrite) and Mo (molybdenite) as a function of potential (SHE) [61], (c) flotation recovery of
common sulfides in nitrogen purged borate solution[63], (d) The effect of NaHS on the recovery of
molybdenite (53 µm)[65].

The application of inorganic depressants such as NaHS and NaCN can effectively depress copper sulfides
and separate Cu and Mo sulfides. However, they still have several disadvantages, such as potentially

9
Journal Pre-proof

releasing toxic gases (H2S and HCN) when pH is out of control, corrosion of storage tanks and pipes by
Cu depressants, inadequate molybdenite recovery, and a possible loss of gold and silver when NaCN is
used.[9,66,67] Therefore, extensive study has be conducted to develop alternative depressants. Among
them, organic depressants are promising candidates because they are eco-friendly and inexpensive.

b) Organic depressants for copper sulfide

Organic copper sulfide depressants are attractive because they are less toxic, more effective at low
dosages, eco-friendly, diverse, and easy to obtain. Thioglycollic acidis (TGA) is one of earlier organic
depressants used for copper sulfide minerals, and “favorable results” were obtained by the Utah Copper-
Company[68]. Thereafter, more organic depressants, such as disodium carboxymethyltrithiocarbonate

of
(DCMT)[67], disodium bis(carboxymethyl) trithiocarbonate (DBT)[69], 2,3-disulfanylbutanedioic acid
(DMSA)[70], acetic acid-[(hydrazinylthioxomethyl)thio]-sodium (AHS)[71] , chitosan[72,73], pseudo

ro
glycolythiourea (PGA)[74], 4-amino-3-thioxo-3,4-dihydro-1,2,4-triazin-5(2H)-one (ATDT)[75], 4-
amino-5-mercapto- 1,2,4-triazole (AMT)[76], thiocarbonohydrazide (TCH)[77] etc. are designed,
-p
synthesized, and applied to depress chalcopyrite flotation. Donor atoms such as S, N, and O in these
re
compounds coordinate with transition metal ions (i.e. Cu and Fe) on chalcopyrite in a neutral or
deprotonated form and lead to the formation of mono- or poly-nuclear complexes.[78] For instance, Guan
lP

et al.[77] reported greater than 70% separation of molybdenite and chalcopyrite at pH 4-8 by applying
TCH as a depressant. Further research using Fourier-transform infrared spectroscopy (FTIR), XPS and
na

Time-of-Flight Secondary Ion Mass Spectrometry (ToF-SIMS) reveal that the chemisorption of TCH was
probably induced by the reaction of thiol and primary amine functional groups of TCH with Cu atoms to
form Cu-S and Cu-N bonds on the chalcopyrite surface. In addition, L-cysteine, sodium alginate, and
ur

poly(acrylamide-allyl thiourea) (PAM-ATU) were reported to be good copper sulfide depressants.


Jo

L-cysteine is a type of naturally occurring proteinogenic amino acid, which is non-toxic and water soluble.
In addition, thiol (-SH), carboxyl (-COOH), and primary amine (-NH2) functional groups, as shown in Fig.
7(a), ensure a strong coordination ability between L-cysteine and metal ions.[79] L-cysteine was
originally used as a corrosion inhibitor for copper and copper alloys because thiol and amine function
groups could strongly adsorb onto copper.[80] Moreover, poly (dimethylsioxane)-gold electrode treated
by L-cysteine was identified to be a sensitive sensor for copper(II).[81] Considering the unsaturated
copper ions on chalcopyrite surfaces, the large aspect ratio of molybdenite, and the weak interaction force
between L-cysteine and molybdenite, L-cysteine was selected as a promising chalcopyrite depressant.

10
Journal Pre-proof

of
ro
Fig. 7. (a) Flotation recovery of molybdenite and chalcopyrite as function of pH with or without L-

-p
cysteine and the molecular formula of L-cysteine, (b) diagram of the adsorption model of L-cysteine on
chalcopyrite surface[79].
re
lP

Fig. 7(a) shows the recovery of molybdenite and chalcopyrite with or without L-cysteine at different pH
levels.[79] The recovery of molybdenite does not change too much during the testing pH range, while the
chalcopyrite recovery decreases from over 80% to lower than 20% after the addition of L-cysteine.
na

Furthermore, bench-scale tests using chalcopyrite-molybdenite concentrate from a processing plant


demonstrates the recovery of molybdenite is approximately 90%. The ratio of Mo recovery and Cu
ur

recovery in concentrate and tailing are 1.6 and 0.25 respectively after adding L-cysteine, which is better
than the performance of NaHS, Na2S, and sodium thioglycolate. Density-functional theory (DFT)
Jo

calculation reveals the adsorption reaction energy of L-cysteine is much lower than that of water at Cu or
Fe active sites, indicating L-cysteine can easily replace water molecules on chalcopyrite. Also, the
adsorption reactions of L-cysteine with Cu2+ and Fe2+ are thermodynamically favorable. An adsorption
model is proposed that a L-cysteine-Cu compound is formed through the interaction of Cu atoms and
thiol and primary amine groups, as shown in Fig. 7(b),[79] which is confirmed by FTIR, Raman, and
ToF-SIMS measurement results.

Macromolecular polymer such as polysaccharide[34,82], polyacrylamide[83] and their derivatives are


also reported to be promising chalcopyrite depressants. PAM-ATU [84] with an average molecular weight
of approximately 1.66 × 104 g/mol was reported to achieve a chalcopyrite to molybdenite mass separation
ratio of 3:2 with a 78.2% recovery and 58.9% concentrate grade of molybdenum. The chemisorption of

11
Journal Pre-proof

PAM-ATU onto chalcopyrite surface can be attributed to the interactions between S, N, and Cu sites as
revealed by FTIR, zeta potential, and XPS results. Sodium alginate (SA)[85], a natural polysaccharide,
contains many hydroxyl (-OH) and carboxyl (-COO-) groups in its molecular structure as shown in Fig.
8(a). Therefore, it was used as an inhibitor for molybdenite or galena during flotation, and it could also be
used as a depressant for chalcopyrite. Fig. 8(b) [86] shows the recovery of molybdenite and chalcopyrite
with or without SA at different pH levels. The recovery of chalcopyrite could be depressed significantly
by SA at pH 3-9, while the recovery of molybdenite changed slightly. FTIR results illustrated that the
chemical adsorption of SA onto chalcopyrite was achieved by chelation between Cu and hydroxyl and
carboxyl groups, and the interaction between these groups and molybdenite is negligible. However,
during the flotation test of mixed minerals (chalcopyrite and molybdenite), molybdenite recovery

of
decreases to 35.97%. FTIR, zeta potential, and adsorption tests reveal that Cu2+ and Cu(OH)+ ions from

ro
chalcopyrite could dissolve in the water and adsorb onto molybdenite surface. SA could adsorb onto
molybdenite surface through chelation or electrostatic attraction with Cu species as highlighted in Fig.

-p
8(c).[86] The flotation of molybdenite could be restored by pretreatment using kerosene before the
addition of SA.
re
lP
na
ur
Jo

Fig. 8. (a) molecular structure of sodium alginate (SA)[85], (b) flotation recovery of chalcopyrite and
molybdenite at different pH levels with and without SA (40 mg/L SA and 2 mg/L kerosene), (c) diagram
of interaction mechanisms between chalcopyrite and SA[86].

12
Journal Pre-proof

c) Depressants for molybdenite reverse flotation

Instead of depressing copper sulfide ores, several studies[87] have been conducted on depressants for
molybdenite aiming at possible reverse flotation in talc and molybdenite separation[88,89]. Organic
depressants for molybdenite such as dextrin[34,35,90], lignosulfonate[91,92], Carboxymethylcellulose
(CMC)[29,31], O-carboxymethyl chitosan (O-CMC)[73,93], humic acid (HA)[94,95], and xanthan
gum[96] are reported to be possible alternatives for inorganic depressants like NaHS and NaCN. Yuan et
al.[73] report that molybdenite recovery decreases sharply from 97% to 11% after the addition of 150
ppm O-CMC, while the recovery of chalcopyrite (>94%) does not change too much at pH 9. The

of
interaction mechanism between dextrin, lignosulfonates, HA, and O-CMC and molybdenite have been
studied and summarized in the work of Yuan[93] and Park[9]. The adsorption of O-CMC onto

ro
molybdenite is a result of attractive force arising from the hydrophobic interaction between O-CMC and
the basal surface of molybdenite and the repulsive force caused by electrostatic repulsion between
-p
them.[93] When the repulsive force is weaker than the attractive force, the adsorption of O-CMC as well
re
as the depression of molybdenite occur. Moreover, the adsorption of O-CMC onto the edge surface of
molybdenite is negligible due to the presence of HMoO4- and MoO42- , which cause a strong electrostatic
lP

repulsion between O-CMC and the edge surface of molybdenite. Hydrogen bonding is not considered as
an important adsorption mechanism between O-CMC and molybdenite, because adsorption through
na

hydrogen bonding is not thermodynamically favorable.

Xanthan gum (XG) is a water-soluble polymer polysaccharide with a chemical formula of (C35H49O29)n,
ur

and its molecular structure is shown in Fig.9(a).[97] XG is produced by fermentation as


an exopolysaccharide from the microorganism Xanthomonas campestris, and it is widely used in food, oil,
Jo

and mining industries. Functional groups of XG, such as hydroxyl groups, pyruvic acid and acetyl group,
and free carboxyl groups, make it water soluble and easy to interact with mineral surfaces. Fig. 9(b)
illustrated the recovery of chalcopyrite and molybdenite at different pH, and the recovery of molybdenite
decreased sharply to 10% after the addition of 100 ppm XG at pH>10.[96] While the recovery of
chalcopyrite was barely affected. After treated by XG, the -OH, -CH2, -COO− and C-O-C stretching peaks
were identified on molybdenite surface, and a decrease of zeta potential of molybdenite was also observed,
indicating considerable XG were adsorbed on molybdenite surface. Even though the interaction
mechanism between XG and molybdenite was not discussed in the same study, it is highly possible that
the interaction is dictated by the attractive hydrophobic force and repulsive electrostatic force as reported
in studies of other polysaccharide type depressants [93].

13
Journal Pre-proof

Fig. 9. (a) molecular structure of xanthan gum[97], and (b) recovery of molybdenite and chalcopyrite at

of
different pH with or without XG[96].

ro
-p
re
lP
na
ur

Fig. 10. molecular structures of xylose, mannose and glucose, and (b) effect of D-xylose, D-mannose, and
Jo

D-glucose, and kerosene on the recovery of molybdenite[98].

Castillo et al.[98] investigated effects of three types of hemicelluloses monosaccharides, D-xylose, D-


mannose and D-glucose on the flotation behavior of molybdenite. Hemicelluloses are reported to be able
to improve the flotation of chalcopyrite and depress the recovery of molybdenite in the presence of clays
and slimes. Monosaccharides, for example, xylose, mannose, and glucose, are the basic compounds of the
structure of hemicelluloses polysaccharides, and the molecular structures of them are shown in Fig. 10(a).
The recovery of molybdenite decreases from pH 7.0 to 10.0, as shown in Fig. 10(b). D-mannose and D-
glucose are stronger depressants for molybdenite than D-xylose, which could be attributed to more carbon
atoms and hydroxyl groups in their molecular structures. The interaction between these three
monosaccharides and molybdenite is believed to be caused by the interaction between the hydroxyl

14
Journal Pre-proof

groups of hemicelluloses and hydroxylated metallic basic/acidic sites on molybdenite surfaces. However,
the depressing effect of monosaccharides can be reduced by adding kerosene, as indicated in Fig.10(b).
Molybdenite surface covered by kerosene prevents the adsorption of monosaccharides.

d) Surface treatments

Surface oxidation treatments, such as addition of oxidant (H2O2 and ozone)[99,100], plasma treatment
[54], thermal pretreatment [55], electrolysis oxidation[43] can be applied to depress copper sulfides
during Cu-Mo separation. Hydrophilic oxidation products (e.g., CuO, Cu(OH)2, FeOOH, and Fe2(SO4)3)
formed on the surface of chalcopyrite after oxidation treatment.[55] These oxidation products have low
solubility in alkaline water and will precipitate on the surface of chalcopyrite and reduce its floatability.

of
However, the oxidation products of molybdenite (e.g., MoO2 and MoO3) are water soluble under alkaline
condition.[54] Therefore, the recovery of molybdenite changes slightly before and after oxidation

ro
treatment. Surface oxidation treatments also have their own limitations, for example, electrolysis
oxidation is difficult to apply on small mineral particles, plasma and thermal treatment requires dry
-p
conditions and extra equipment, and H2O2 and ozone treatment can be expensive. Consequently, new
re
oxidants have been tested for copper sulfide depression.
lP
na
ur
Jo

Fig. 11. Flotation recovery of copper and molybdenum in mineral mixtures in ferrate solution (3×10−4
mol/L) as a function of pH[101].

Liao et al.[101] examined the depressing effect of ferrate(VI) on chalcopyrite during the separation of
molybdenite and chalcopyrite. As shown in Fig. 11, chalcopyrite flotation was strongly depressed over a
wide pH range (pH 4-9) after the addition of ferrate (3 × 10−4 mol/L). XPS results reveals that ferrate, as a
strong oxidant, can react with metal ions on the surface of chalcopyrite and form hydrophilic Cu(OH)2
and Fe(OH)3 compounds. The deposition of these hydrophilic products on the surface of chalcopyrite

15
Journal Pre-proof

reduces its floatability. While ferrate does not react with the surface of molybdenite, thus the recovery of
molybdenite is not affected.

Microencapsulation is another possible method used to separate molybdenite and chalcopyrite.


Microencapsulation treated chalcopyrite using Fe2+ and PO43- at pH 4 becomes hydrophilic because of a
precipitated layer of FePO4 and/or FeO(OH) as shown in Fig. 12.[102] Because chalcopyrite is a good
conductor compared with molybdenite,[43] oxidation of Fe2+ prefers to occur on chalcopyrite rather than
on molybdenite, which is demonstrated by cyclic voltammetry results. Further investigation using SEM-
EDS and XPS[102] proved that FePO4 preferentially coated on chalcopyrite rather than on molybdenite.

of
ro
-p
re
lP
na
ur

Fig. 12. Scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-EDX) results of
chalcopyrite without and with microencapsulation treatment[102].
Jo

4. Effect of electrochemical potential measurements and control on molybdenite flotation

Mehrabani et al.[64] investigated the effect of pulp potential on the separation of molybdenite and
chalcopyrite by manipulating Na2S dosage and aeration gases. Fig. 13(a) illustrated the oxidation-
reduction potential (ORP) decrease from +228mV to -597 mV with the addition of Na2S. However, when
Na2S concentration is higher than 16kg/t, no further ORP drop was observed. Fig. 13(b) shows ORP
change versus time using different Na2S dosage and flotation gas. ORP decreases immediately each time
after the addition of Na2S, and ORP increases gradually versus time when air is applied as the flotation
gas. However, when N2 is applied as the flotation gas, ORP remained under -200 mV during the entire
test time with 6 kg/t of Na2S. A high recovery of molybdenite, 92.88%, is achieved using 6kg/t Na2S and

16
Journal Pre-proof

nitrogen, which is very similar to that when 16kg/t Na2S and air was applied. Therefore, ORP is a key
factor for molybdenite flotation, and N2 can greatly reduce the consumption of Na2S. A 42-month
industrial test conducted by Poorkani and Banisi [103]in the Sarcheshmeh processing plant reveals that N2,
as a byproduct from the oxygen plant, could reduce the consumption of Na2S by 19.7%. An earlier study
reported by Aravena[104] demonstrated that NaHS consumption was reduced by 48% at the
Chuquicamata mine in Chile by using N2, and a further decrease of 21% was achieved when NaHS was
controlled by potential measurement. Furthermore, both lab and plant scale test demonstrated that the
application of potential control and measurement could greatly reduce the consumption of depressant.

of
ro
-p
re
Fig. 13. Pulp potential after the addition of Na2S dosage with different flotation gas (air and nitrogen), (b)
lP

pulp ORP change with the addition of Na2S[64].


na

Table 1 The separation of chalcocite from molybdenite in an electrochemical flotation cell from a 50:50
percent mixture. Potential of precondition -1.2V[40].
ur

Potential of Molybdenite recovery Chalcocite recovery Chalcocite in


flotation(V) (%) (%) concentrate (%)
-1.2 62 14 18
Jo

-0.6 39 6 14
+0.3 (50 mV/s) 21 63 75
+0.3 (5 mV/s) 17 100 86

Pulp electrochemical potential control in lab scale testing was performed by Chander and Fuerstenau[40]
using a three-electrode system and a potentiostat. The test results are shown in Table 1. At low potential
(-1.2 V) chalcocite can be depressed because the collector attachment reaction onto chalcocite was
depressed. At high potentials, chalcocite flotation was restored while molybdenite flotation was depressed
due to oxidation at edge surface. The recovery of chalcocite is better if the sweep rate is slow (5mV/s). A
slower sweep rate results in a better recovery for chalcocite indicating oxidation rate at the surface is the
control step for chalcocite flotation. Electrochemical studies by Outokumpu Oy[105], and the Original

17
Journal Pre-proof

Potential Flotation (OPF) technology[106] based on potential monitoring are successfully used to develop
flotation strategies. In addition, the OK-PCF potential monitoring system developed by Outokumpu Oy
was applied in several process plants. For instance, lime and collector consumption decreased to one third
and profitability increased by 10-20% due to the installation of the OK-PCF system.[107]

Metal electrodes, mineral electrodes, and ion selective electrode are usually applied to measure
electrochemical potential.[108] As a noble metal, platinum is stable in different solutions, thus it is the
most widely used metal electrode. The Eh results in Fig. 6(b) were measured using platinum electrode,
and 0.1 V was observed to be the critical potential value to achieve molybdenite and chalcopyrite
separation.[61] However, coating could readily form on it in many cases, and pretreatment is usually

of
needed before using a Pt electrode. Mineral electrodes[108] measured the redox potential established at
the mineral/solution interface. Therefore, different mineral electrodes may obtain different Eh potentials

ro
in the sample pulp. For example, when NaCN was added into the pulp, a pyrite electrode[109] and a
galena electrode[110] exhibited totally different potentials, suggesting different reactions occurred on

-p
them. Further research reveals that electrode potential of these minerals is a function of the solution pH
rather than the concentration of the respective ions.[111] The potential response curves for all these
re
minerals had slopes close to 59 mv/pH. The reactions occurred at sulfide mineral surfaces were ascribed
lP

to the formation of oxides or hydroxides.


na
ur
Jo

Fig. 14. The change of pulp potential measured by Pt and chalcopyrite electrodes in chalcopyrite pulp in
the presence of K2Cr2O7[112].

Kocabag and Guler [112]compared the potential value measured using platinum and mineral electrodes.
They found that under reducing potentials and acidic pH values, the mineral and platinum electrodes
displayed similar potential value. However, they acted quite differently in oxidizing and alkaline media,
as shown in Fig. 14.[112] The potential measured with Pt-electrode could only represent the solution

18
Journal Pre-proof

potential, i.e. oxidation level. As a matter of fact, Teague et al.[113] proved that there is a linear
relationship between the potential of clear solution and that of slurry. But they are not identical to each
other. On the other hand, the potentials obtained using mineral electrodes could be the true potential of
mineral particles surfaces. Therefore, a mineral electrode is preferred in the practical Eh measurement
related to flotation, and proper pretreatment and short measurement times are usually necessary to obtain
consistent results.

5. Effect of cations on molybdenite flotation

Cations, such as Ca2+, Mg2+, Na+, K+, and Cu2+ etc, from seawater or reused water have a significant
effect on the recovery of molybdenite.[114–116] Ca2+ and Mg2+ are reported to have a negative effect on

of
molybdenite flotation when pH>8. For example, molybdenite recovery dropped by 30% when 10-2 M

ro
MgCl2 was added at pH 11, as shown in Fig. 15(a).[51] Hydroxides and precipitates, such as Ca(OH)+,
Mg(OH)+, and Mg(OH)2, adsorbed onto the surface of molybdenite can make it hydrophilic, leading to a

-p
large drop in molybdenite recovery.[48] A study[38,48] of the interaction mechanisms between
molybdenite surfaces and Ca2+ and Mg2+ reveals that, in acidic environment, the oxidation of molybdenite
re
edge surface produces HMoO4-. Ca2+ and Mg2+ could adsorb onto the edge surface and reverse the
negative surface potential (or even make it zero). Therefore, as the repulsive force between a molybdenite
lP

surface and an air bubble decreases, the molybdenite recovery increases.


na
ur
Jo

Fig. 15. (a) recovery of molybdenite in solutions of different Mg ion concentrations as a function of pH,
(b) AFM image of molybdenite surface before and after treatment in MgCl2 solutions[51].

19
Journal Pre-proof

In alkaline environment, metal hydroxide cations such as Ca(OH)+ and Mg(OH)+ can adsorb onto
molybdenite. A surface potential and ToF-SIMS study[48] reveals that the basal surface can adsorb more
metal hydroxide cations than the edge surface. The adsorption of Ca(OH)+ and Mg(OH)+ onto
molybdenite could occur through physical interaction as well as chemical interaction of Ca2+ and
Mg2+ with MoO42- on the edge surface.[48,117] The amount of adsorbed metal hydroxide cations
increases with pH, and when the saturated concentration of cation was reached, for example 10 mM
MgCl2 at pH 11, Mg(OH)2 precipitates start to form, as shown in Fig. 15(b)[51]. Characterization
techniques, such as ToF-SIMS and SEM EDS confirmed the precipitation of Mg(OH)2 and CaCO3 on
molybdenite at high pH levels. [48,118]

of
Na+ and K+ in water of high salinity could also have a negative effect on molybdenite flotation.[119,120]
However, the mechanism of depression for Na+ and K+ is different from that of Ca2+ and Mg2+. Na+ and

ro
K+ are believed to be helpful for oxidation and leaching of Mo on the edge or micro-edge surfaces. On the
one hand, the oxidation/leaching leads to more negative charges on the surface, and on the other hand the

-p
oxidation reaction makes molybdenite surface less hydrophobic. As a result, a repulsive force between
molybdenite and air bubbles prevents the collision and attachment of them. What’s more, Ca2+, Mg2+, Na+,
re
and K+ are also harmful for the kinetics of molybdenite flotation. For example, at pH 10, the flotation rate
lP

constant drop from 2.5 to 0.96 when Mg2+ concentration increases from 10-4 to 10-2M.[120]

Cu2+ and Fe3+ ions are also common cations in flotation pulp because of the dissolution of copper
na

minerals. Yang et al[121] reported that molybdenite recovery dropped from 42% to less than 12% when
Cu2+ increased to 50 mg/L at pH 4.2. They also found that the depressing effect of Cu2+ is stronger in the
pH range of 5-8.4 than that in the pH range of 2-5. Hydrophilic species such as CuS, Cu(OH)2, and Cu2+
ur

formed in acidic environment, and Cu(OH)2 precipitates formed in alkaline environment causes the
Jo

depression of molybdenite flotation. When Na2S was added into the pulp[122], copper hydroxides react
with HS- to form Cu(I)-S species (Cu2S). Excess HS- could adsorb onto molybdenite surface and form a
hydrophilic layer. The existence of these hydrophilic copper species depresses the flotation of
molybdenite. However, Cu2+ could improve the separation performance of chalcopyrite and molybdenite
when Na2S2O3 was applied as a depressant for chalcopyrite. [123] Cu2+ could react with S2O32− ions
adsorbed onto chalcopyrite surface and form hydrophilic species such as CuS2O3−, Cu(S2O3)23− and
Cu(S2O3)35−, which further decrease the floatability of chalcopyrite, while the flotation of molybdenite
was only slightly affected under these conditions.

Compared with Cu2+, Fe3+ is a stronger depressant for molybdenite.[121] In the pH range of 2-8, the
flotation of molybdenite was strongly depressed (Recovery<10%) after the addition of 20 mg/L Fe3+ ions.

20
Journal Pre-proof

The depression was mainly attributed to the precipitation of Fe(OH)3 as indicated by SEM EDS and XPS
results.

Pb2+ ion, as a common activator, has been used to improve the flotation of sulfide minerals.[124]
Moreover, Pb bearing minerals could also induce Pb2+ ion formation in the pulp because of dissolution of
surface lead oxides.[125] Pb2+ can greatly depress the flotation of molybdenite. For example, at pH 9, the
recovery of molybdenite can drop 80% when 6×10-4 M Pb2+ is added.[28] However, in the acidic
environment (pH=2-4), the depression effect of Pb2+ can be neglected. When Na2S and Pb2+ were added
simultaneously, molybdenite was strongly depressed in the pH range of 8-10.5. XPS results demonstrated
that the depression was caused by the formation of PbS on molybdenite in Pb2+ solution of weak

of
alkalinity. Strong depression through the addition of Na2S and Pb2+ in solution was attributed to the
precipitation of Pb(OH)2 on molybdenite. Other metal cations, such as Al3+, Fe2+, and Cu+, could also

ro
depress the flotation of molybdenite in a similar way as that of Cu2+ and Pb2+.[24]

-p
re
lP
na
ur

Fig. 16. (a)Effect of Mg ion and SHMP on the recovery of molybdenite[119], and (b) interaction energy
Jo

between MoS2 particles and Mg(OH)2[30].

6. Reducing the effect of Ca and Mg cations on molybdenite flotation

Sodium silicate (SS) and sodium hexametaphosphate (SHMP or SH) are widely used as dispersants of
hydrophilic substances from the surface of valuable minerals. Fig. 16(a)[119] illustrated that the recovery
of molybdenite increased from 20% to 79% in MgCl2 solution with the addition of 50 mg/L SHMP at pH
10. Further examination of results of contact angle and zeta potential measurements, FTIR data, and XPS
evaluations[30] confirmed that the adsorption of hydrophilic species of Mg and Ca was greatly reduced
after the addition of SHMP or SS. Interaction force study based on Derjaguin–Landau–Verwey–Overbeek
(EDLVO) theoretical calculations revealed that SHMP and SS were able to change the interaction force

21
Journal Pre-proof

between molybdenite and Mg(OH)2 colloids from attractive to repulsive, as shown in Fig. 16(b)[30,126]
due to its long chain structure. A mechanism for the formation of a soluble complex of SHMP with Ca2+
and Mg2+ [119] was proposed as another possible explanation for reducing the depressing effect of Ca and
Mg cations. However, considering the low concentration of SHMP, the first mechanism should play a
more important role. In addition, SS does not rely on competitive adsorption on molybdenite surface to be
functional.

Oil collectors like kerosene are also effective in reducing the depressing effect of Mg2+ and Ca2+ ions. The
recovery of molybdenite maintained at a high level in artificial seawater (48.7 mM Mg2+ and 10.2 mM
Ca2+) when emulsified kerosene was used at pH 10. Contact angle measurements[37] reveal the surface

of
wettability of molybdenite was greatly reduced because of the coating of a thin kerosene film. Bubble-
particle attachment [127] tests show an oil film can also coat on air bubbles and prevent the attachment of

ro
Mg hydroxide products on them. The thin kerosene film[128] can also improve the formation of three-
phase (water-air-molybdenite) contact. Moreover, Mg colloids precipitated[128] could form aggregates

-p
with emulsified kerosene. As a result, the depressing effect of Mg and Ca cations decreases significantly.
re
Polycyclic aromatic hydrocarbons (PAHs) were added into diesel oil to further improve the performance
of oil collectors.[118] Industrial scale tests demonstrated that the average recovery of molybdenite
lP

increases 1.8% after adding 4% PAHs into diesel oil.[129] Ca2+ helps the adsorption of PAHs onto the
edge surfaces of molybdenite because CaMoO4 precipitates form on edge surfaces can react with PAHs
na

and form PAHs-Ca2+-MoO42- with a structure of π-cation-π. This structure can increase the contact angle
of molybdenite and reduce the depressing effect of Ca2+. Since fine molybdenite particles have more edge
ur

surfaces, better recovery of molybdenite is obtained for smaller molybdenite particles.[129]

Oil collector emulsion is usually applied together with SHMP or SS[130] to enhance flotation of
Jo

molybdenite in alkaline solution with high Mg2+ and Ca2+ concentrations. Ramirez et al.[127] reported
that a high recovery of molybdenite is achieved in seawater at pH 10 with the addition of 100 ppm of
kerosene and SHMP. The recovery of molybdenite is similar to that achieved using fresh water (95-96%).
Water pre-treatment is also applied to reduce the depressing effect of Mg2+ and Ca2+. Jeldres et al.[131]
pre-treated sea water using a mixture of Na2CO3 and CaO (50%-50%), and a final concentration of 176
ppm and 190 ppm were achieved for Ca2+ and Mg2+ respectively. The recovery of molybdenite increase
accordingly from 12.7% to 73.1% at pH 10.5. However, extra cost induced by the high consumption (0.05
M each) of Na2CO3 and CaO is the main concern for the application of this method.

22
Journal Pre-proof

7. Effect of gangue minerals on molybdenite flotation

Limestone skarn ore (LSN) is reported to have a negative effect on molybdenite recovery in the Bingham
Canyon mine.[14] Laboratory and plant scale tests demonstrated that molybdenite recovery decreased
significantly when limestone skarn ore was blended with the quartzite ore (with high molybdenum
content).[15] Further study on gangue mineralogy was conducted using composite samples collected for a
period of 18 months from the Bingham Canyon mine.[13] Four gangue minerals, talc (as shown in
Fig.17(a)), andradite, calcite and amphibole associated with LSN are found to have a negative effect on
the recovery of molybdenite. Mineral surface characterization performed by Gerson et al.[14] reveals that
LSN ore could induce Mg and Ca-containing species formation in the pulp. The precipitation of Mg and

of
Ca hydroxides onto molybdenite bearing ores results in reduced selectivity, flotation kinetics, and
recovery of molybdenite.

ro
-p
re
lP

Fig. 17. (a) relationship between Talc percentage in feed and Molybdenum recovery[13], (b) recovery of
na

molybdenite with or without 15% kaolinite as a function of pH in fresh water and seawater[90], and (c)
SEM image of kaolinite coated on molybdenite surface[132].
ur

Talc, with a theoretical chemical formula of Mg3Si4O10(OH)2 [133], is a common gangue mineral for
Jo

sulfide ores. It is one of the possible sources for Mg2+, which could cause severe depression of
molybdenite flotation. Moreover, the surface of talc is isotropic, and the non-polar and hydrophobic basal
plane accounts for 90% of total surface[88]. Therefore, talc is easy to be brought into molybdenite
concentrate due to its excellent natural floatability. The presence of talc in molybdenite concentrate
causes extra cost for future smelting and refining. Addition of inorganic and organic depressants as well
as surface oxidation by roasting[6,89] have been applied to separate molybdenite and talc.

Kaolinite is another common clay mineral with a chemical formula of Al2Si2O5(OH)4.[133] It can cause
many negative effects such as loss in recovery (as shown in Fig. 17(b)), extra reagents consumption,
unstable froth zone, and high impurity levels in the concentrate.[90] The depressing effect of kaolinite can
be enhanced by Ca2+ and Mg2+ ions because of the heterocoagulation between kaolinite and molybdenite

23
Journal Pre-proof

(as indicated in Fig. 17(c)) induced by the hydroxides of these cations.[132] Dispersants like SS and
SHMP can restore the floatability of molybdenite and chalcopyrite in the presence of kaolinite in alkaline
water of high Ca2+ and Mg2+ concentrations.[134] Moreover, reducing pulp density has also been shown
to be an effective way to increase molybdenite recovery and concentrate grade in presence of
kaolinite[90].

8. Effect of particle size on molybdenite flotation

Molybdenite in porphyry ores is a byproduct of copper concentrate, therefore it is more reasonable to


establish both the grinding size and the flotation scheme for copper minerals.[8] Fig. 18(a) [13] displays
the average size by size recovery of molybdenum and copper for 18 months in the Bingham Canyon mine.

of
The highest recovery of molybdenum was achieved in the 38–75 μm size range. While the recovery of

ro
molybdenum drops significantly in the coarse and fine sizes. The poor flotation performance of coarse
particle[135] is attributed to the disruption of bubble-particle aggregates in turbulent zone and the

-p
decrease in buoyancy of bubble-particle aggregate. For fine molybdenite particles, on the one hand, the
basal/edge plane ratio was reduced, and thus the hydrophobicity of the overall mineral surfaces
re
decreased.[136] On the other hand, forth flotation poorly responds to fine mineral particles because of the
small mass of fine particles and thus low probability of collision and adhesion to air bubbles[32,137],
lP

which leads to a low flotation rate. Accordingly, the flotation recovery of molybdenite fines dropped
sharply. Henceforth, an appropriate range of particle size should be identified for the grinding circuit of
na

Cu and Mo minerals.
ur
Jo

24
Journal Pre-proof

of
ro
-p
re
Fig. 18. (a) Average size by size recovery of Mo and Cu for the 18 months from April 2005 until October
lP

2006 of Bingham Canyon mine[13], (b) effect of kerosene, diesel, and transformer oil on molybdenite
recovery[136], (c) SEM image of agglomerates of molybdenite fines in the presence of kerosene[138], (d)
effect of different mechanical agitation on molybdenite recovery[138].
na
ur

Neutral oil, such as kerosene and transformer oil can improve the recovery of fine molybdenite, as shown
in Fig. 18(b)[136]. Fine molybdenite particles aggregated around oil droplets, as indicated in Fig.
Jo

18(c)[138], because of hydrophobic attraction between oil droplets and molybdenite. Excellent
aggregation among molybdenite particles makes good flotation achievable. Transformer oil is the most
effective in improving the recovery of molybdenite among kerosene, diesel, transformer, and rapeseed
oils.[136] While rapeseed oil [139] shows little effect on the recovery fine molybdenite due to the fact
that it is an unsaturated fatty acid. The dosage of oil collector is critical. For example, a higher dosage of
transformer oil (from 2.0 to 13.8 kg/t) leads to a larger average size of fine molybdenite agglomerates
(from 0.15 to 0.68mm)[139]. Song et al.[138] demonstrated that the smaller the droplet size of kerosene,
the better the recovery of fine molybdenite. In addition, increased mechanical conditioning [140,141] can
lead to more collisions between oil droplets and molybdenite fines, and a better recovery of fine
molybdenite is achieved as displayed in Fig. 18(d)[138]. However, the conditioning time should be

25
Journal Pre-proof

relatively short, because long duration conditioning (over 2 hours) treatment can destroy the trapped
bubbles on molybdenite and lead to a large decrease in molybdenite recovery.[142]

The addition of polyethylene oxide (PEO) could further increase the flotation efficiency of fine
molybdenite. As reported by Alvarez et al. [143], the recovery of molybdenite fines (<10 μm) increased
from 50% to 85% with the addition of 10 ppm of PEO and 50 ppm of diesel at pH 9. Even in seawater, a
20% increase in fine molybdenite (<10 μm) recovery was achieved by using the same dosage of PEO and
diesel. Further examination[32] reveals that PEO molecules could easily attached to the (010) edge plane
of molybdenite. As a result, the hydrophobicity of molybdenite fine particles increases and the attachment
time between air bubbles and molybdenite particles is reduced. Moreover, AFM results show larger

of
attractive forces are detected between molybdenite particles over a longer separation distance when PEO
is added.

ro
9. Effect of oil collector and surfactants on molybdenite flotation

-p
Kerosene, diesel oil, and transformer oil have been widely used as collectors to improve the recovery of
re
molybdenite. Oil collectors can improve the fine molybdenite recovery by forming agglomerates of larger
sizes.[138] Collector oil can also form a thin film on air bubbles and molybdenite particles, which prevent
lP

the precipitation of hydrophilic Ca or Mg hydroxides.[127] Moreover, kerosene could improve the bubble
aspect ratio and reduce the bubble rise velocity in artificial seawater.[128,144] The draining of
intervening thin liquid film on the mineral surface and the formation of a three-phase (air-liquid-mineral)
na

contact could also be accelerated by kerosene.


ur

The adsorption of collector oil on molybdenite was firstly attributed to the attraction force between the
hydrogen atoms of the oil hydrocarbon chain and the open spaces at the centers of the sulfur triangles in
Jo

the molybdenite cleavage plane.[145] However, Wang et al.[146,147] found that the Van der Waals force
and the hydrophobic attractive force caused by the Lewis acid-base interaction at the molybdenite surface
lead to the adsorption of oil onto molybdenite surface. Feng et al.[148] investigated the interaction
mechanism between molybdenite and a dodecane oil droplet through AFM measurements and theoretical
calculation. The interaction force was composed of attractive van der Waals and hydrophobic forces, and
a repulsive electrical double-layer (EDL) force. At high pH and low salinity, a strong EDL force caused
by negatively charge oil droplet and molybdenite surfaces prevented attachment of collector, as shown in
Fig. 19[148]. On the contrary, at low pH and high salinity, the EDL force was reduced, and the
hydrophobic force became dominant, which allowed the adsorption of oil onto molybdenite surface.

26
Journal Pre-proof

of
ro
Fig. 19. Stability map of the interaction between dodecane and the molybdenite basal surface in an
electrolyte solution for varying salt concentrations, salt types, and pH values[148].

-p
Table 2 Physicochemical properties of the oily collectors[136]
re
Oil Type Density/(kg/ Kinematic Surface Fundamental Length of
m3, 20 °C) viscosity/(mm2/s, 40 °C) tension/(mN/m) components carbon chain
lP

Kerosene 800 1.6 24.0 n-Alkanes C12-C15


Diesel Oil 840 3.7 27.2 n-Alkanes C15-C18
Transformer Oil 895 11.5 29.5 Cycloalkanes C16-C23
na

The performance of oil collectors is influenced by many factors such as viscosity, chain structure, type of
ur

function groups, and temperature.[145,149] For instance, oil with a straight chain could create a dry and
Jo

collapsing froth, and naphthenic base oil could achieve better performance than the aromatic base
oils.[145] In addition, collector oils with high viscosity were able to achieve a higher molybdenum
recovery but at a lower concentrate grade[7]. Table 2[136] shows the properties of commonly used oil
collectors. Kerosene has the lowest viscosity and the smallest number of carbon atoms in its chain, while
transformer oil is the most viscous and of the longest carbon chain among the three. Flotation results
confirmed that transformer oil led to the best recovery of molybdenum. The interaction and flotation test
results indicated that the dispersion capability[136,149] of the oily collector played a more important role
in molybdenite recovery than the strength of interaction between collectors and molybdenite. When the
pulp temperature is relatively low (i.e. <10 °C), the flotation recovery of molybdenite is closely related to
the dispersibility of diesel oil.[150]

27
Journal Pre-proof

To stabilize the emulsified collector oil, surfactants were also added into the pulp.[151–153] Syntex, as a
surfactant, has been successfully applied in Climax mill, Colorado to improve the recovery of
molybdenite.[154] A small addition of surfactant could lower the interfacial energy of oil/water, modify
zeta potential of oil droplets, and prevent the coalescence of oil droplets during collision. Liu et al.[151]
reported that the addition of dodecyl amine hydrochloride into kerosene could change the zeta potential of
oil droplets from negative to positive and cause strong attachment on sphalerite. In contrast, potassium
dodecyl xanthate (KDX) made the zeta potential of kerosene droplet more negative, and the oil droplet
attaches strongly to galena and weakly to sphalerite. Surfactant mixtures were also applied to obtain a
better emulsion result. For example, Prapagen WKT with Emulsogen EL and Emulsogen A with
Emulsogen EL were reported to achieve the best metallurgical results among nine surfactants and their

of
combinations.[153]

ro
10. Summary

-p
Molybdenum is critical for metallurgical and chemical industries. Molybdenite is the most important
mineral resource for molybdenum, and it can be recovered as a byproduct from porphyry copper ores.
re
Froth flotation, which is the most cost-effective way to separate Mo and Cu in industry, has been studied
extensively. Previous research showed that the mineralogy of molybdenite could be different from mine
lP

to mine, and microstructure results revealed molybdenite was anisotropic with a hydrophobic basal plane
and a hydrophilic edge plane. Zeta/surface potential, open circuit potential, and surface force
na

measurements can be applied to interpret different reactions on molybdenite surfaces. Organic and
inorganic copper sulfide depressants are utilized to improve the recovery of molybdenite in Mo flotation
ur

circuits. Surface oxidation treatment, ferrate(VI), and microencapsulation are also reported to be effective
in depressing copper sulfide flotation in laboratory tests. Molybdenite depressants were also investigated
Jo

and applied to reverse flotation of molybdenite from other sulfide minerals. Electrochemical potential
methods are adopted in both laboratory and plant scale to improve the separation performance of
molybdenite and copper sulfides, and the consumption of depressant can be reduced substantially by
applying potential control.

Cations such as Ca2+, Mg2+, Na+, K+, and Cu2+ etc. can cause negative effects on the recovery of
molybdenite in alkaline environments. Dispersants such as sodium silicate (SS) and sodium
hexametaphosphate (SHMP) and oil collectors need to be added to restore the flotation of molybdenite in
such alkaline environments. Gangue minerals like talc and kaolinite can depress the flotation of
molybdenite because Mg and Ca ions dissolved from their surfaces can precipitate on molybdenite as
hydrophilic colloids. To achieve the best recovery of molybdenite, an appropriate range of particle size
needs to be utilized. Particles that are too large or too fine decrease molybdenite recovery. Oil collectors

28
Journal Pre-proof

and polyethylene oxide (PEO) are effective in improving the recovery of fine molybdenite. The addition
of oil collectors and surfactants can also increase hydrophobicity of molybdenite and facilitate the
attachment of molybdenite onto air bubbles. Hydrophobic forces are reported to be the major force for oil
adsorption onto molybdenite surface.

Even though some other factors such as pH, solids concentration, grinding time etc. are not discussed in
the current manuscript, they are of great importance for the recovery of molybdenite and have been
studied extensively.

of
ro
-p
re
lP
na
ur
Jo

29
Journal Pre-proof

Reference:
[1] Shields J. Applications of Molybdenum Metal and its Alloys. 2013.
[2] Wadsworth, J ; Wittenauer JP. The history of development of molybdenum alloys for structural
applications, United States: 1993.
[3] IMOA. Use of new Molybdenum 2018. https://www.imoa.info/molybdenum-uses/molybdenum-
uses.php.
[4] Abdollahi M, Bahrami A, Mirmohammadi MS, Kazemi F, Danesh A, Ghorbani Y. A process
mineralogy approach to optimize molybdenite flotation in copper – molybdenum processing plants.
Miner Eng 2020;157:106557. https://doi.org/10.1016/j.mineng.2020.106557.
[5] Bahrami A, Abdollahi M, Mirmohammadi M, Kazemi F, Danesh A, Shokrzadeh M. A process
mineralogy approach to study the efficiency of milling of molybdenite circuit processing. Sci Rep

of
2020;10:1–14. https://doi.org/10.1038/s41598-020-78337-8.
[6] Yuan D, Cadien K, Liu Q, Zeng H. Separation of talc and molybdenite: challenges and

ro
opportunities. Miner Eng 2019;143. https://doi.org/10.1016/j.mineng.2019.105923.
[7] Bulatovic SM. Flotation of Copper Sulfide Ores. Handb Flotat Reagents 2007:235–93.

[8]
-p
https://doi.org/10.1016/b978-044453029-5/50021-6.
Bulatovic SM, Wyslouzil DM, Kant C. Operating practices in the beneficiation of major porphyry
re
copper/molybdenum plants from Chile innovated technology and opportunities, a review. Miner
Eng 1998;11:313–31. https://doi.org/10.1016/s0892-6875(98)00011-9.
lP

[9] Park I, Hong S, Jeon S, Ito M, Hiroyoshi N. A review of recent advances in depression techniques
for flotation separation of cu–mo sulfides in porphyry copper deposits. Metals (Basel) 2020;10:1–
26. https://doi.org/10.3390/met10091269.
na

[10] Prasad P, Mankhand T, Prasad A. Molybdenum extraction process: An overview. NML Tech J
1997;39:39–58.
[11] Gupta CK. Extractive Metallurgy of Molybdenum. Taylor \& Francis; 1992.
ur

[12] Raghavan S, Hsu LL. Factors affecting the flotation recovery of molybdenite from porphyry
copper ores. Int J Miner Process 1984;12:145–62. https://doi.org/10.1016/0301-7516(84)90026-7.
Jo

[13] Triffett B, Veloo C, Adair BJI, Bradshaw D. An investigation of the factors affecting the recovery
of molybdenite in the Kennecott Utah Copper bulk flotation circuit. Miner Eng 2008;21:832–40.
https://doi.org/10.1016/j.mineng.2008.03.003.
[14] Gerson AR, Smart RSC, Li J, Kawashima N, Weedon D, Triffett B, et al. Diagnosis of the surface
chemical influences on flotation performance: Copper sulfides and molybdenite. Int J Miner
Process 2012;106–109:16–30. https://doi.org/10.1016/j.minpro.2012.01.004.
[15] Zanin M, Ametov I, Grano S, Zhou L, Skinner W. A study of mechanisms affecting molybdenite
recovery in a bulk copper/molybdenum flotation circuit. Int J Miner Process 2009;93:256–66.
https://doi.org/10.1016/j.minpro.2009.10.001.
[16] Rao SR, Finch JA. A review of water re-use in flotation. Miner Eng 1989;2:65–85.
https://doi.org/10.1016/0892-6875(89)90066-6.
[17] Liu W, Moran CJ, Vink S. A review of the effect of water quality on flotation. Miner Eng
2013;53:91–100. https://doi.org/10.1016/j.mineng.2013.07.011.

30
Journal Pre-proof

[18] Chimonyo W, Corin KC, Wiese JG, O’Connor CT. Redox potential control during flotation of a
sulphide mineral ore. Miner Eng 2017;110:57–64. https://doi.org/10.1016/j.mineng.2017.04.011.
[19] Peres AEC, Gonçalves KLC, Andrade VLL, Turrer HDG. Effect of pulp potential on the flotation
of a sulfide copper ore. Miner Metall Process 2005;22:168–72.
https://doi.org/10.1007/bf03403132.
[20] Pracejus B. Sulfides with Metal : Sulfur, Selenium, and Tellurium < 1 : 1, Tellurides with Copper,
Silver, and Gold. Ore Miner Under Microsc 2014:408–547. https://doi.org/10.1016/b978-0-444-
62725-4.50007-7.
[21] Dickinson RG, Pauling L. THE CRYSTAL STRUCTURE OF MOLYBDENITE. J Am Chem Soc
1923;45:1466–71. https://doi.org/10.1021/ja01659a020.
[22] Janish MT. In-situ TEM Lithiation of Alternative Battery Electrode Materials. Univ Connect Grad
Sch 2017.

of
[23] Lu N, Wang J, Oviedo JP, Lian G, Kim MJ. Atomic Resolution Scanning Transmission Electron
Microscopy of Two-Dimensional Layered Transition Metal Dichalcogenides. Appl Microsc

ro
2015;45:225–9. https://doi.org/10.9729/am.2015.45.4.225.
[24] Castro S, Lopez-Valdivieso A, Laskowski JS. Review of the flotation of molybdenite. Part I:

-p
Surface properties and floatability. Int J Miner Process 2016;148:48–58.
https://doi.org/10.1016/j.minpro.2016.01.003.
re
[25] Jeon J, Jang SK, Jeon SM, Yoo G, Jang YH, Park JH, et al. Layer-controlled CVD growth of
large-area two-dimensional MoS2 films. Nanoscale 2015;7:1688–95.
lP

https://doi.org/10.1039/c4nr04532g.
[26] Tabares JO, Ortega IM, Bahena JLR, López AAS, Pérez DV, Valdivieso AL. Surface properties
and floatability of molybdenite. Proc 2006 China--Mexico Work Miner Part Technol San Luis
na

Potosi, Mex 2006.


[27] Qiu Z, Liu G, Liu Q, Zhong H. Understanding the roles of high salinity in inhibiting the
molybdenite flotation. Colloids Surfaces A Physicochem Eng Asp 2016;509:123–9.
ur

https://doi.org/10.1016/j.colsurfa.2016.08.059.
[28] Zhang Q, Zhu H, Yang B, Jia F, Yan H, Zeng M, et al. EFFECT of Pb2+ on the flotation of
Jo

molybdenite in the presence of sulfide ion. Results Phys 2019;14.


https://doi.org/10.1016/j.rinp.2019.102361.
[29] Xie L, Wang J, Huang J, Cui X, Wang X, Liu Q, et al. Anisotropic polymer adsorption on
molybdenite basal and edge surfaces and interaction mechanism with air bubbles. Front Chem
2018;6:1–11. https://doi.org/10.3389/fchem.2018.00361.
[30] Li Y, Yang X, Fu J, Li W, Hu C. New insights into the beneficial roles of dispersants in reducing
negative influence of Mg2+on molybdenite flotation. RSC Adv 2020;10:27401–6.
https://doi.org/10.1039/d0ra05556e.
[31] Wang J, Xie L, Lu Q, Wang X, Wang J, Zeng H. Electrochemical investigation of the interactions
of organic and inorganic depressants on basal and edge planes of molybdenite. J Colloid Interface
Sci 2020;570:350–61. https://doi.org/10.1016/j.jcis.2020.03.007.
[32] Li S, Ma X, Wang J, Xing Y, Gui X, Cao Y. Effect of polyethylene oxide on flotation of
molybdenite fines. Miner Eng 2020;146:106146. https://doi.org/10.1016/j.mineng.2019.106146.

31
Journal Pre-proof

[33] Jin J, Miller JD, Dang LX. Molecular dynamics simulation and analysis of interfacial water at
selected sulfide mineral surfaces under anaerobic conditions. Int J Miner Process 2014;128:55–67.
https://doi.org/10.1016/j.minpro.2014.03.001.
[34] Braga PFA, Chaves AP, Luz AB, França SCA. The use of dextrin in purification by flotation of
molybdenite concentrates. Int J Miner Process 2014;127:23–7.
https://doi.org/10.1016/j.minpro.2013.12.007.
[35] Beaussart A, Parkinson L, Mierczynska-Vasilev A, Beattie DA. Adsorption of modified dextrins
on molybdenite: AFM imaging, contact angle, and flotation studies. J Colloid Interface Sci
2012;368:608–15. https://doi.org/10.1016/j.jcis.2011.10.075.
[36] Sinche-Gonzalez M, Fornasiero D, Zanin M. Flotation of chalcopyrite and molybdenite in the
presence of organics in water. Minerals 2016;6:1–13. https://doi.org/10.3390/min6040105.
[37] Suyantara GPW, Hirajima T, Miki H, Sasaki K. Floatability of molybdenite and chalcopyrite in

of
artificial seawater. Miner Eng 2018;115:117–30. https://doi.org/10.1016/j.mineng.2017.10.004.
[38] Li Y, Lartey C, Song S, Li Y, Gerson AR. The fundamental roles of monovalent and divalent

ro
cations with sulfates on molybdenite flotation in the absence of flotation reagents. RSC Adv
2018;8:23364–71. https://doi.org/10.1039/C8RA02690D.
[39]
-p
Lu Z, Liu Q, Xu Z, Zeng H. Probing Anisotropic Surface Properties of Molybdenite by Direct
Force Measurements. Langmuir 2015;31:11409–18. https://doi.org/10.1021/acs.langmuir.5b02678.
re
[40] Chander S, Fuerstenau DW. Electrochemical flotation separation of chalcocite from molybdenite.
Int J Miner Process 1983;10:89–94. https://doi.org/10.1016/0301-7516(83)90035-2.
lP

[41] Li Y, Zhu H, Li W, Zhu Y. A fundamental study of chalcopyrite flotation in sea water using
sodium silicate. Miner Eng 2019;139:105862. https://doi.org/10.1016/j.mineng.2019.105862.
[42] He T, Li H, Jin J, Peng Y, Wang Y, Wan H. Improving fine molybdenite flotation using a
na

combination of aliphatic hydrocarbon oil and polycyclic aromatic hydrocarbon. Results Phys
2019;12:1050–5. https://doi.org/10.1016/j.rinp.2018.12.010.
ur

[43] Miki H, Matsuoka H, Hirajima T, Suyantara GPW, Sasaki K. Electrolysis oxidation of


chalcopyrite and molybdenite for selective flotation. Mater Trans 2017;58:761–7.
https://doi.org/10.2320/matertrans.M-M2017807.
Jo

[44] Kudin KN, Car R. Why Are Water−Hydrophobic Interfaces Charged? J Am Chem Soc
2008;130:3915–9. https://doi.org/10.1021/ja077205t.
[45] Tian CS, Shen YR. Structure and charging of hydrophobic material/water interfaces studied by
phase-sensitive sum-frequency vibrational spectroscopy. Proc Natl Acad Sci U S A
2009;106:15148–53. https://doi.org/10.1073/pnas.0901480106.
[46] Saji VS, Lee CW. Molybdenum, molybdenum oxides, and their electrochemistry. ChemSusChem
2012;5:1146–61. https://doi.org/10.1002/cssc.201100660.
[47] Pan L, Yoon RH. Effects of elctrolytes on the stability of wetting films: Implications on seawater
flotation. Miner Eng 2018;122:1–9. https://doi.org/10.1016/j.mineng.2018.03.011.
[48] Lu Z, Ralston J, Liu Q. Face or Edge? Control of Molybdenite Surface Interactions with Divalent
Cations. J Phys Chem C 2019. https://doi.org/10.1021/acs.jpcc.9b07632.
[49] Kozbial A, Gong X, Liu H, Li L. Understanding the Intrinsic Water Wettability of Molybdenum
Disulfide (MoS2). Langmuir 2015;31:8429–35. https://doi.org/10.1021/acs.langmuir.5b02057.

32
Journal Pre-proof

[50] Xie L, Wang J, Yuan D, Shi C, Cui X, Zhang H, et al. Interaction Mechanisms between Air
Bubble and Molybdenite Surface: Impact of Solution Salinity and Polymer Adsorption. Langmuir
2017;33:2353–61. https://doi.org/10.1021/acs.langmuir.6b04611.
[51] Hirajima T, Suyantara GPW, Ichikawa O, Elmahdy AM, Miki H, Sasaki K. Effect of Mg2+ and
Ca2+ as divalent seawater cations on the floatability of molybdenite and chalcopyrite. Miner Eng
2016;96–97:83–93. https://doi.org/10.1016/j.mineng.2016.06.023.
[52] J. S. Laskowski NBK. Anisotropic minerals in flotation circuits. CIM Journal, Vol 3, No 4 2012.
https://store.cim.org/en/anisotropic-minerals-in-flotation-circuits (accessed April 1, 2021).
[53] Azghdi SMS, Barani K. Effect of microwave treatment on the surface properties of chalcopyrite.
Miner Metall Process 2018;35:141–7. https://doi.org/10.19150/mmp.8463.
[54] Hirajima T, Mori M, Ichikawa O, Sasaki K, Miki H, Farahat M, et al. Selective flotation of
chalcopyrite and molybdenite with plasma pre-treatment. Miner Eng 2014;66:102–11.

of
https://doi.org/10.1016/j.mineng.2014.07.011.
[55] Tang X, Chen Y, Liu K, Zeng G, Peng Q, Li Z. Selective flotation separation of molybdenite and

ro
chalcopyrite by thermal pretreatment under air atmosphere. Colloids Surfaces A Physicochem Eng
Asp 2019;583. https://doi.org/10.1016/j.colsurfa.2019.123958.
[56]
-p
DENG S-M. Current Situation of Flotation Process and Reagents of Cu-Mo Sulphide Ores.
DEStech Trans Environ Energy Earth Sci 2018:289–93.
re
https://doi.org/10.12783/dteees/epee2017/18140.
[57] Ralston J. Eh and its consequences in sulphide mineral flotation. Miner Eng 1991;4:859–78.
lP

https://doi.org/10.1016/0892-6875(91)90070-C.
[58] Allison SA, Goold LA, Nicol MJ, Granville A. Determination of the Products of Reaction
Between Various Sulfide Minerals and Aqueous Xanthate Solution, and a Correlation of the
na

Products With Electrode Rest Potentials. Met Trans 1972;3:2613–8.


https://doi.org/10.1007/bf02644237.
[59] Chen Y, Chen X, Peng Y. The effect of sodium hydrosulfide on molybdenite flotation in seawater
ur

and diluted seawater. Miner Eng 2020;158. https://doi.org/10.1016/j.mineng.2020.106589.


[60] Peng H, Wu D, Abdalla M, Luo W, Jiao W, Bie X. Study of the effect of sodium sulfide as a
Jo

selective depressor in the separation of chalcopyrite and molybdenite. Minerals 2017;7.


https://doi.org/10.3390/min7040051.
[61] Nagaraj DR, Wang SS, Avotins P V, Dowling E. Structure--Activity Relationship for Copper
Depressants. Trans Inst Min Met C 1986;95.
[62] Chander S. Electrochemistry of sulfide flotation: Growth characteristics of surface coatings and
their properties, with special reference to chalcopyrite and pyrite. Int J Miner Process
1991;33:121–34. https://doi.org/10.1016/0301-7516(91)90047-M.
[63] Tolley W, Kotlyar D, Van Wagoner R. Fundamental electrochemical studies of sulfide mineral
flotation. Miner Eng 1996;9:603–37. https://doi.org/10.1016/0892-6875(96)00051-9.
[64] Vazife J, Pourghahramani P, Asqarian H, Bagherian A. Effects of pH and pulp potential on
selective separation of Molybdenite from the Sungun Mine Cu-Mo concentrate 2017:10–3.
[65] Chen Y, Chen X, Peng Y. The effect of sodium hydrosulfide on molybdenite flotation as a
depressant of copper sulfides. Miner Eng 2020;148. https://doi.org/10.1016/j.mineng.2020.106203.

33
Journal Pre-proof

[66] Liu GY, Lu YP, Zhong H, Cao ZF, Xu ZH. A novel approach for preferential flotation recovery of
molybdenite from a porphyry copper-molybdenum ore. Miner Eng 2012;36–38:37–44.
https://doi.org/10.1016/j.mineng.2012.02.008.
[67] Yin Z, Sun W, Hu Y, Zhai J, Qingjun G. Evaluation of the replacement of NaCN with depressant
mixtures in the separation of copper–molybdenum sulphide ore by flotation. Sep Purif Technol
2017;173:9–16. https://doi.org/10.1016/j.seppur.2016.09.011.
[68] Gibbs HL. Differential froth flotation of sulfide ores. US2449984A, 1948.
[69] YIN Z gang, SUN W, HU Y hua, GUAN Q jun, ZHANG C hu, GAO Y sheng, et al. Depressing
behaviors and mechanism of disodium bis (carboxymethyl) trithiocarbonate on separation of
chalcopyrite and molybdenite. Trans Nonferrous Met Soc China (English Ed 2017;27:883–90.
https://doi.org/10.1016/S1003-6326(17)60100-6.
[70] Li MY, Wei DZ, Shen YB, Liu WG, Gao SL, Liang GQ. Selective depression effect in flotation

of
separation of copper-molybdenum sulfides using 2,3-disulfanylbutanedioic acid. Trans Nonferrous
Met Soc China (English Ed 2015;25:3126–32. https://doi.org/10.1016/S1003-6326(15)63942-5.

ro
[71] Yin Z, Sun W, Hu Y, Zhang C, Guan Q, Liu R, et al. Utilization of acetic acid-
[(hydrazinylthioxomethyl)thio]-sodium as a novel selective depressant for chalcopyrite in the

-p
flotation separation of molybdenite. Sep Purif Technol 2017;179:248–56.
https://doi.org/10.1016/j.seppur.2017.01.049.
re
[72] Li M, Wei D, Liu Q, Liu W, Zheng J, Sun H. Flotation separation of copper-molybdenum sulfides
using chitosan as a selective depressant. Miner Eng 2015;83:217–22.
https://doi.org/10.1016/j.mineng.2015.09.013.
lP

[73] Yuan D, Cadien K, Liu Q, Zeng H. Flotation separation of Cu-Mo sulfides by O-Carboxymethyl
chitosan. Miner Eng 2019;134:202–5. https://doi.org/10.1016/j.mineng.2019.02.007.
na

[74] Chen JH, Lan LH, Liao XJ. Depression effect of pseudo glycolythiourea acid in flotation
separation of copper-molybdenum. Trans Nonferrous Met Soc China (English Ed 2013;23:824–31.
https://doi.org/10.1016/S1003-6326(13)62535-2.
ur

[75] Yin Z, Sun W, Hu Y, Zhang C, Guan Q, Zhang C. Separation of molybdenite from chalcopyrite in
the presence of novel depressant 4-amino-3-thioxo-3,4-dihydro-1,2,4-triazin-5(2H)-one. Minerals
2017;7. https://doi.org/10.3390/min7080146.
Jo

[76] Yin Z, Hu Y, Sun W, Zhang C, He J, Xu Z, et al. Adsorption Mechanism of 4-Amino-5-mercapto-


1,2,4-triazole as Flotation Reagent on Chalcopyrite. Langmuir 2018;34:4071–83.
https://doi.org/10.1021/acs.langmuir.7b03975.
[77] Guan C, Yin Z, Khoso SA, Sun W, Hu Y. Performance analysis of thiocarbonohydrazide as a
novel selective depressant for chalcopyrite in molybdenite-chalcopyrite separation. Minerals
2018;8. https://doi.org/10.3390/min8040142.
[78] Netalkar PP, Netalkar SP, Revankar VK. Transition metal complexes of thiosemicarbazone:
Synthesis, structures and invitro antimicrobial studies. Polyhedron 2015;100:215–22.
https://doi.org/10.1016/j.poly.2015.07.075.
[79] Yin Z, Chen S, Xu Z, Zhang C, He J, Zou J, et al. Flotation separation of molybdenite from
chalcopyrite using an environmentally-efficient depressant L-cysteine and its adsoption
mechanism. Miner Eng 2020;156:106438. https://doi.org/10.1016/j.mineng.2020.106438.
[80] Kuruvilla M, John S, Joseph A. Electrochemical studies on the interaction of l-cysteine with

34
Journal Pre-proof

metallic copper in sulfuric acid. Res Chem Intermed 2013;39:3531–43.


https://doi.org/10.1007/s11164-012-0860-y.
[81] Fan D, Bi L, Tang F, Zheng H, Xu Q, Wang W. L-Cysteine modified flexible PDMS-gold
electrode for sensing ascorbic acid and copper. Sensors Actuators, B Chem 2012;161:1124–8.
https://doi.org/10.1016/j.snb.2011.11.001.
[82] Chen Z, Gu G, Li S, Wang C, Zhu R. The effect of seaweed glue in the separation of copper–
molybdenum sulphide ore by flotation. Minerals 2018;8:1–15.
https://doi.org/10.3390/min8020041.
[83] Wang K, Wang L, Cao M, Liu Q. Xanthation-modified polyacrylamide and spectroscopic
investigation of its adsorption onto mineral surfaces. Miner Eng 2012;39:1–8.
https://doi.org/10.1016/j.mineng.2012.06.005.
[84] Zhang X, Lu L, Cao Y, Yang J, Che W, Liu J. The flotation separation of molybdenite from

of
chalcopyrite using a polymer depressant and insights to its adsorption mechanism. Chem Eng J
2020;395:125137. https://doi.org/10.1016/j.cej.2020.125137.

ro
[85] Aktar B, Erdal MS, Sagirli O, Güngör S, Özsoy Y. Optimization of biopolymer based transdermal
films of metoclopramide as an alternative delivery approach. Polymers (Basel) 2014;6:1350–65.

[86]
https://doi.org/10.3390/polym6051350.
-p
Zeng G, Ou L, Zhang W, Zhu Y. Effects of Sodium Alginate on the Flotation Separation of
re
Molybdenite From Chalcopyrite Using Kerosene as Collector. Front Chem 2020;8:1–9.
https://doi.org/10.3389/fchem.2020.00242.
lP

[87] Castro S, Laskowski JS. Depressing effect of flocculants on molybdenite flotation. Miner Eng
2015;74:13–9. https://doi.org/10.1016/j.mineng.2014.12.027.
[88] Tang X, Chen Y, Liu K, Peng Q, Zeng G, Ao M, et al. Reverse flotation separation of talc from
na

molybdenite without addition of depressant: Effect of surface oxidation by thermal pre-treatment.


Colloids Surfaces A Physicochem Eng Asp 2020;594:1–8.
https://doi.org/10.1016/j.colsurfa.2020.124671.
ur

[89] Yuan D, Xie L, Shi X, Yi L, Zhang G, Zhang H, et al. Selective flotation separation of
molybdenite and talc by humic substances. Miner Eng 2018;117:34–41.
https://doi.org/10.1016/j.mineng.2017.12.005.
Jo

[90] Jorjani E, Barkhordari HR, Tayebi Khorami M, Fazeli A. Effects of aluminosilicate minerals on
copper-molybdenum flotation from Sarcheshmeh porphyry ores. Miner Eng 2011;24:754–9.
https://doi.org/10.1016/j.mineng.2011.01.005.
[91] Gutierrez L, Uribe L, Hernandez V, Vidal C, Texeira Mendonça R. Assessment of the use of
lignosulfonates to separate chalcopyrite and molybdenite by flotation. Powder Technol
2020;359:216–25. https://doi.org/10.1016/j.powtec.2019.10.015.
[92] Ansari A, Pawlik M. Floatability of chalcopyrite and molybdenite in the presence of
lignosulfonates. Part I. Adsorption studies. Miner Eng 2007;20:600–8.
https://doi.org/10.1016/j.mineng.2006.12.007.
[93] Yuan D, Cadien K, Liu Q, Zeng H. Adsorption characteristics and mechanisms of O-
Carboxymethyl chitosan on chalcopyrite and molybdenite. J Colloid Interface Sci 2019;552:659–
70. https://doi.org/10.1016/j.jcis.2019.05.023.
[94] Lai RWM, Stone LC, Rimmasch BE. Effect of humus organics on the flotation recovery of

35
Journal Pre-proof

molybdenite. Int J Miner Process 1984;12:163–72. https://doi.org/10.1016/0301-7516(84)90027-9.


[95] Yuan D, Cadien K, Liu Q, Zeng H. Selective separation of copper-molybdenum sulfides using
humic acids. Miner Eng 2019;133:43–6. https://doi.org/10.1016/j.mineng.2019.01.005.
[96] Yan H, Yang B, Zeng M, Huang P, Teng A. Selective flotation of Cu-Mo sulfides using xanthan
gum as a novel depressant. Miner Eng 2020;156:106486.
https://doi.org/10.1016/j.mineng.2020.106486.
[97] Jindal N, Singh Khattar J. Microbial Polysaccharides in Food Industry. Biopolym. Food Des.,
Elsevier Inc.; 2018, p. 95–123. https://doi.org/10.1016/B978-0-12-811449-0.00004-9.
[98] Castillo I, Gutierrez L, Hernandez V, Diaz E, Ramirez A. Hemicelluloses monosaccharides and
their effect on molybdenite flotation. Powder Technol 2020;373:758–64.
https://doi.org/10.1016/j.powtec.2020.07.032.

of
[99] Hirajima T, Miki H, Suyantara GPW, Matsuoka H, Elmahdy AM, Sasaki K, et al. Selective
flotation of chalcopyrite and molybdenite with H2O2 oxidation. Miner Eng 2017;100:83–92.
https://doi.org/10.1016/j.mineng.2016.10.007.

ro
[100] Suyantara GPW, Hirajima T, Miki H, Sasaki K, Yamane M, Takida E, et al. Selective flotation of
chalcopyrite and molybdenite using H2O2 oxidation method with the addition of ferrous sulfate.

-p
Miner Eng 2018;122:312–26. https://doi.org/10.1016/j.mineng.2018.02.005.
[101] Liao R, Feng Q, Wen S, Liu J. Flotation separation of molybdenite from chalcopyrite using
re
ferrate(VI) as selective depressant in the absence of a collector. Miner Eng 2020;152.
https://doi.org/10.1016/j.mineng.2020.106369.
lP

[102] Park I, Hong S, Jeon S, Ito M, Hiroyoshi N. Flotation separation of chalcopyrite and molybdenite
assisted by microencapsulation using ferrous and phosphate ions: Part i. selective coating
formation. Metals (Basel) 2020;10:1–11. https://doi.org/10.3390/met10121667.
na

[103] Poorkani M, Banisi S. Industrial use of nitrogen in flotation of molybdenite at the Sarcheshmeh
copper complex. Miner Eng 2005;18:735–8. https://doi.org/10.1016/j.mineng.2004.10.013.
[104] Aravena JJ. Column flotation applications at Chuquicamata’s molybdenite flotation plant. A Mular,
ur

G Gonz~ Ilez C Barahona (Editors), Copp 1988;87:155–69.


Jo

[105] Hintikka V V, Leppinen JO. Potential control in the flotation of sulphide minerals and precious
metals. Miner Eng 1995;8:1151–8.
[106] Gu G, Hu Y, Wang H, Qiu G, Wang D. Original potential flotation of galena and its industrial
application. J Cent South Univ Technol 2002;9:91–4.
[107] Woods R. Electrochemical potential controlling flotation. Int J Miner Process 2003;72:151–62.
https://doi.org/10.1016/S0301-7516(03)00095-4.
[108] Chander S. A brief review of pulp potentials in sulfide flotation. Int J Miner Process 2003;72:141–
50. https://doi.org/10.1016/S0301-7516(03)00094-2.
[109] Abramov AA. Effect of ph on the condition of the pyrite surface. Sov J NONFERROUS METL
Sov J NONFERROUS Met DEC 1965, P 33-36 1965.
[110] Richardson PE, Maust Jr EE. Surface stoichiometry of galena in aqueous electrolytes and its effect
on xanthate interactions. Flot M Gaudin Meml 1976.
[111] Sato M. Oxidation of sulfide ore bodies; II, Oxidation mechanisms of sulfide minerals at 25

36
Journal Pre-proof

degrees C. Econ Geol 1960;55:1202–31.


[112] Kocabaǧ D, Güler T. A comparative evaluation of the response of platinum and mineral electrodes
in sulfide mineral pulps. Int J Miner Process 2008;87:51–9.
https://doi.org/10.1016/j.minpro.2008.01.005.
[113] Teague AJ, Van Deventer JSJ, Swaminathan C. The effect of galvanic interaction on the behaviour
of free and refractory gold during froth flotation. Int J Miner Process 1999;57:243–63.
[114] Castro S. Physico-chemical factors in flotation of Cu-Mo-Fe ores with seawater: A critical review.
Physicochem Probl Miner Process 2018;54:1223–36. https://doi.org/10.5277/ppmp18162.
[115] Jeldres RI, Forbes L, Cisternas LA. Effect of Seawater on Sulfide Ore Flotation: A Review. Miner
Process Extr Metall Rev 2016;37:369–84. https://doi.org/10.1080/08827508.2016.1218871.
[116] Lin S, Liu R, Wu M, Hu Y, Sun W, Shi Z, et al. Minimizing beneficiation wastewater through

of
internal reuse of process water in flotation circuit. J Clean Prod 2020;245:118898.
https://doi.org/10.1016/j.jclepro.2019.118898.

ro
[117] Laskowski JS, Castro S, Gutierrez L. Flotation in Seawater. Mining, Metall Explor 2019;36:89–98.
https://doi.org/10.1007/s42461-018-0018-6.

-p
[118] Wan H, Yang W, He T, Yang J, Guo L, Peng Y. The influence of Ca2+ and pH on the interaction
between PAHs and molybdenite edges. Minerals 2017;7:1–12.
https://doi.org/10.3390/min7060104.
re
[119] Li W, Li Y, Wei Z, Xiao Q, Song S. Fundamental studies of shmp in reducing negative effects of
divalent ions on molybdenite flotation. Minerals 2018;8:1–17. https://doi.org/10.3390/min8090404.
lP

[120] Zhu H, Li Y, Lartey C, Li W, Qian G. Flotation kinetics of molybdenite in common sulfate salt
solutions. Miner Eng 2020;148:106182. https://doi.org/10.1016/j.mineng.2020.106182.
na

[121] Yang B, Wang D, Wang T, Zhang H, Jia F, Song S. Effect of Cu2+ and Fe3+ on the depression of
molybdenite in flotation. Miner Eng 2019;130:101–9.
https://doi.org/10.1016/j.mineng.2018.10.012.
ur

[122] Zhao Q, Liu W, Wei D, Wang W, Cui B, Liu W. Effect of copper ions on the flotation separation
of chalcopyrite and molybdenite using sodium sulfide as a depressant. Miner Eng 2018;115:44–52.
Jo

https://doi.org/10.1016/j.mineng.2017.10.008.
[123] Yang B, Yan H, Zeng M, Huang P, Jia F, Teng A. A novel copper depressant for selective
flotation of chalcopyrite and molybdenite. Miner Eng 2020;151:106309.
https://doi.org/10.1016/j.mineng.2020.106309.
[124] Peng Y, Grano S. Effect of grinding media on the activation of pyrite flotation. Miner Eng
2010;23:600–5.
[125] Liu J, Wang Y, Luo D, Chen L, Deng J. Comparative study on the copper activation and xanthate
adsorption on sphalerite and marmatite surfaces. Appl Surf Sci 2018;439:263–71.
[126] Peng Y, Li Y, Li W, Fang X, Liu C, Fan R. Elimination of adverse effects of seawater on
molybdenite flotation using sodium silicate. Miner Eng 2020;146:106108.
https://doi.org/10.1016/j.mineng.2019.106108.
[127] Ramirez A, Gutierrez L, Laskowski JS. Use of “oily bubbles” and dispersants in flotation of
molybdenite in fresh and seawater. Miner Eng 2020;148:106197.
https://doi.org/10.1016/j.mineng.2020.106197.

37
Journal Pre-proof

[128] Suyantara GPW, Hirajima T, Elmahdy AM, Miki H, Sasaki K. Effect of kerosene emulsion in
MgCl2 solution on the kinetics of bubble interactions with molybdenite and chalcopyrite. Colloids
Surfaces A Physicochem Eng Asp 2016;501:98–113.
https://doi.org/10.1016/j.colsurfa.2016.04.039.
[129] Wan H, Qu J, He T, Bu X, Yang W, Li H. A new concept on high-calcium flotation wastewater
reuse. Minerals 2018;8. https://doi.org/10.3390/min8110496.
[130] Park CIH, Jeon HS. The effect of sodium silicate as pH modifier and depressant in the froth
flotation of molybdenite ores. Mater Trans 2010;51:1367–9.
https://doi.org/10.2320/matertrans.M2009397.
[131] Jeldres RI, Arancibia-Bravo MP, Reyes A, Aguirre CE, Cortes L, Cisternas LA. The impact of
seawater with calcium and magnesium removal for the flotation of copper-molybdenum sulphide
ores. Miner Eng 2017;109:10–3. https://doi.org/10.1016/j.mineng.2017.02.003.

of
[132] Ramirez A, Gutierrez L, Vega-Garcia D, Reyes-Bozo L. The depressing effect of kaolinite on
molybdenite flotation in seawater. Minerals 2020;10:1–14. https://doi.org/10.3390/min10060578.

ro
[133] Pracejus B. Sulfides with Metal : Sulfur, Selenium, and Tellurium < 1 : 1, Tellurides with Copper,
Silver, and Gold. Ore Miner. Under Microsc., Elsevier; 2014, p. 408–547.

-p
https://doi.org/10.1016/b978-0-444-62725-4.50007-7.
[134] Ramirez A, Rojas A, Gutierrez L, Laskowski JS. Sodium hexametaphosphate and sodium silicate
re
as dispersants to reduce the negative effect of kaolinite on the flotation of chalcopyrite in seawater.
Miner Eng 2018;125:10–4. https://doi.org/10.1016/j.mineng.2018.05.008.
lP

[135] Soto H, Barbery G. Flotation of coarse particles in a counter-current column cell. Mining, Metall
Explor 1991;8:16–21. https://doi.org/10.1007/BF03402925.
[136] Lin Q quan, Gu G hua, Wang H, Liu Y cai, Fu J gang, Wang C qing. Flotation mechanisms of
na

molybdenite fines by neutral oils. Int J Miner Metall Mater 2018;25.


https://doi.org/10.1007/s12613-018-1540-8.
[137] Laskowski JS. Oil assisted fine particle processing. Dev Miner Process 1992;12:361–94.
ur

https://doi.org/10.1016/B978-0-444-88284-4.50017-2.
[138] Song S, Zhang X, Yang B, Lopez-Mendoza A. Flotation of molybdenite fines as hydrophobic
Jo

agglomerates. Sep Purif Technol 2012;98:451–5. https://doi.org/10.1016/j.seppur.2012.06.016.


[139] Jiangang F, Kaida C, Hui W, Chao G, Wei L. Recovering molybdenite from ultrafine waste
tailings by oil agglomerate flotation. Miner Eng 2012;39:133–9.
https://doi.org/10.1016/j.mineng.2012.07.006.
[140] Estrada D, Echeverry L, Ramirez A, Gutierrez L. Molybdenite flotation in the presence of a
polyacrylamide of low anionicity subjected to different conditions of mechanical shearing.
Minerals 2020;10:1–12. https://doi.org/10.3390/min10100895.
[141] Tabosa E, Rubio J. Flotation of copper sulphides assisted by high intensity conditioning (HIC) and
concentrate recirculation. Miner Eng 2010;23:1198–206.
https://doi.org/10.1016/j.mineng.2010.08.004.
[142] Chen Y, Bu X, Truong VNT, Peng Y, Xie G. Study on the effects of pre-conditioning time on the
floatability of molybdenite from the perspective of cavitation threshold. Miner Eng 2019;141.
https://doi.org/10.1016/j.mineng.2019.105845.

38
Journal Pre-proof

[143] Alvarez A, Gutierrez L, Laskowski JS. Use of polyethylene oxide to improve flotation of fine
molybdenite. Miner Eng 2018;127:232–7. https://doi.org/10.1016/j.mineng.2018.08.018.
[144] Suyantara GPW, Hirajima T, Miki H, Sasaki K. Bubble interactions with chalcopyrite and
molybdenite surfaces in seawater. Miner Eng 2020;157:106536.
https://doi.org/10.1016/j.mineng.2020.106536.
[145] Smit FJ, Bhasin AK. Relationship of petroleum hydrocarbon characteristics and molybdenite
flotation. Int J Miner Process 1985;15:19–40. https://doi.org/10.1016/0301-7516(85)90021-3.
[146] Wang H, Guo C, Fu J, He Z, Liang W, Chen X, et al. Adsorption behavior of weak hydrophilic
substances on low-energy surface in aqueous medium. Appl Surf Sci 2011;257:7959–67.
https://doi.org/10.1016/j.apsusc.2011.03.158.
[147] Wang H, GU G hua, FU J gang, CHEN L, HAO Y. Study of the interfacial interactions in the
molybdenite floatation system. J China Univ Min Technol 2008;18:82–7.

of
https://doi.org/10.1016/S1006-1266(08)60018-8.
[148] Feng L, Manica R, Grundy JS, Liu Q. Unraveling Interaction Mechanisms between Molybdenite

ro
and a Dodecane Oil Droplet Using Atomic Force Microscopy. Langmuir 2019;35:6024–31.
https://doi.org/10.1021/acs.langmuir.9b00203.

-p
[149] Wan H, Qu J, Li H, He T, Bu X, Yang W. A novel method for improving low-temperature
flotation performance of nonpolar oil in the molybdenite flotation. Minerals 2019;9.
re
https://doi.org/10.3390/min9100609.
[150] He T, Wan H, Song N, Guo L. The influence of composition of nonpolar oil on flotation of
lP

molybdenite. Miner Eng 2011;24:1513–6. https://doi.org/10.1016/j.mineng.2011.07.003.


[151] Liu J, Mak T, Zhou Z, Xu Z. Fundamental study of reactive oily-bubble flotation. Miner Eng
2002;15:667–76. https://doi.org/10.1016/S0892-6875(02)00158-9.
na

[152] Cao L, Chen X, Peng Y. The Formation and Stabilization of Oily Collector Emulsions–A Critical
Review. Miner Process Extr Metall Rev 2020;00:1–18.
https://doi.org/10.1080/08827508.2020.1776279.
ur

[153] Nishkov I, D.Lazarov, M.Marinov, E.Beas, C.Henriquez. Surfactant-Hydrocarbon Oil Emulsions


for Molybdenite Flotation. 1994.
Jo

[154] Hoower and Malhotra. Emulsion Flotation of Mlybdenite. AIME 1976;1:485–506.

39
Journal Pre-proof

Highlights:
 Commercial production and process flowsheets for molybdenite flotation are reviewed.
 Mineralogy, microstructure, and surface properties of molybdenite are discussed.
 Effects of factors such as depressants, electrochemical potential measurement and control, and
cations, etc. on molybdenite flotation are described.
 Interaction mechanisms correlated with these factors are illustrated.

of
ro
-p
re
lP
na
ur
Jo

40
Journal Pre-proof

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

41

You might also like