Sunteți pe pagina 1din 11

International Journal of Mineral Processing 148 (2016) 48–58

Contents lists available at ScienceDirect

International Journal of Mineral Processing

journal homepage: www.elsevier.com/locate/ijminpro

Review of the flotation of molybdenite. Part I: Surface properties


and floatability
S. Castro a,⁎, A. Lopez-Valdivieso b, J.S. Laskowski c
a
Department of Metallurgical Engineering, University of Concepción, Chile
b
Instituto de Metalurgia, Universidad Autónoma de San Luis de Potosí, Mexico
c
NB Keevil Institute of Mining Engineering, University of British Columbia, Canada

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

Article history: Molybdenite (MoS2) is a naturally hydrophobic mineral with anisotropic surface properties. Its floatability is in-
Received 12 September 2014 fluenced by a number of factors such as, particle size and shape, face/edge ratio, degree of crystallization, face het-
Received in revised form 5 January 2016 erogeneity, pH, etc. Molybdenite is floated by using oily collectors, and its recovery is strongly affected by slime
Accepted 11 January 2016
coating phenomena. A number of hydrolyzable cations, such as Ca2+ and Mg2+, depress molybdenite in alkaline
Available online 14 January 2016
solutions. Depression of molybdenite in seawater and saline waters is mainly induced by the precipitation of col-
Keywords:
loidal magnesium hydroxide when pH is raised to depress pyrite. Other metal cations present in recycled process
Molybdenite water can also reduce the flotation recovery of molybdenite (e.g., Al3+, Fe2+, Cu2+ ions). The native floatability of
Native floatability molybdenite is highly depressed by natural polymers (starch, dextrin, guar gum, humic acids, etc.) and synthetic
Molybdenite depression polymers (e.g., flocculants of the polyacrylamide type). The advances in understanding the surface chemistry of
Cu–Mo sulfide ores these systems are reviewed and discussed in this paper.
© 2016 Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2. Crystalline structure and natural floatability of molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.1. Hydrophobic minerals with anisotropic surface properties — molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.2. The effect of particle size and morphology on the floatability of molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.3. Floatability of molybdenite and degree of crystallization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.4. Surface composition of molybdenite and its inherent hydrophobicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.5. Atomic force microscopy of basal planes of molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3. Molybdenite wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4. Zeta potential of molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.1. Slime coating and hetero-coagulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5. Molybdenite and flotation reagents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.1. Interactions with thiol collectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.2. Interaction with non-polar oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.3. Interactions with methyl isobutyl carbinol (MIBC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6. Depression of molybdenite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.1. Depression by polysaccharides and other natural polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.2. Depression by flocculants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7. Depression by hydrolyzable metal cations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.1. Depression by calcium ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.2. Depression in seawater — effect of magnesium ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.3. Depression by other metal cations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

⁎ Corresponding author.
E-mail address: scastro@udec.cl (S. Castro).

http://dx.doi.org/10.1016/j.minpro.2016.01.003
0301-7516/© 2016 Published by Elsevier B.V.
S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58 49

8. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

1. Introduction hydrophilic surfaces, faces and edges (Johnson et al., 2000; Laskowski,
2012). Molybdenite particles exhibit two types of surfaces: (1) those
Common sulfide mineral collectors are known to adsorb onto mo- that form by the break of the S–S bonds (non-polar faces) and
lybdenite but these reagents are not necessarily effective for molybde- (2) those that are generated by rupture of the strong covalent Mo–S
nite flotation. Flotation of molybdenite is carried out with the use of bonds (polar edges). As a result, the faces are hydrophobic while edges
hydrocarbons (e.g., diesel oil, kerosene, etc.). From this one concludes are hydrophilic. It is then reasonable to assume that crystal faces are re-
that molybdenite must differ from other sulfides in that its surface is al- sponsible for the natural floatability of molybdenite.
ready hydrophobic (Chander and Fuerstenau, 1972). Molybdenite, as all
other inherently hydrophobic minerals, cleaves by rupture of weak re-
2.2. The effect of particle size and morphology on the floatability of
sidual bonds (Gaudin et al., 1957) revealing the surfaces that are hydro-
molybdenite
phobic since they interact with water only via dispersion forces
(Laskowski and Kitchener, 1969). Because of large differences between
Particle size is a very important factor affecting flotation properties
the surface properties of molybdenite and other sulfides, molybdenite
of molybdenite. In porphyry copper–molybdenite ores the flotation
can be recovered from Cu–Mo ores as a by-product with copper at levels
conditions (grinding, reagents, bubble size, etc.) all are optimized with
as low as 0.01% Mo. Approximately half of the world's molybdenum
respect to Cu and not Mo. Fig. 2 shows that the optimum flotation par-
production comes from porphyry Cu–Mo ores (Bulatovic, 2007;
ticle size for molybdenite is around 20 μm, and the size fraction 6.8 μm
Guang-yi et al., 2012).
exhibits the lowest recovery and flotation rate (Castro and Mayta,
One of the most important publications in this area, edited by
1994). However, the optimum particle size for copper sulfide minerals
Sutulov, appeared in 1979 (International Molybdenum Encyclopaedia).
is around 15–100 μm. Therefore, usually it is not economically attractive
This volume also included two reviews on molybdenite flotation
to grind a copper ore to the optimum liberation size of molybdenite.
(Castro, 1979; Shirley, 1979). We feel that considerable advances in un-
The natural floatability of molybdenite is also governed by textural
derstanding of the surface properties and floatability of molybdenite
features (flatness, roundness, relative width, elongation ratio and surface
have recently been made and this calls for a new review.
roughness). Because of the preferential cleavage along the weak S–S
bonds during grinding, platelet shaped fragments exfoliating from larger
2. Crystalline structure and natural floatability of molybdenite particles, are produced (Fig. 4). In the case of flat and elongated particles,
low particle-bubble collision efficiency explains its poor flotation recov-
The crystalline structure of molybdenite (MoS2) was first deter- ery. Hence, the flotation behavior of molybdenite is governed by a com-
mined by Dikinson and Pauling (1923). Molybdenite crystallizes in the bination of: (a) the native floatability (Chander and Fuerstenau, 1972;
hexagonal system as the common polytype 2H and also in the trigonal Chander et al., 1975; Arbiter et al., 1975); and (b) the particle morphol-
system as the 3R polytype (less frequent). Molybdenite (2H) has a hex- ogy (shape and size) (Ametov et al., 2008; Nakhaei and Irannajad, 2014).
agonal layer structure with a complete basal cleavage, showing layers of Molybdenite particles with a high aspect ratio (major axis over minor
trigonal prismatic coordination polyhedra, where each Mo atom is axis) have higher probability of reporting to the concentrate, and coarse
surrounded by a trigonal prism of S atoms (Fig. 1). The sulfur atoms particles with high perimeter to area ratio tend to report to the tailings
form the upper and lower surfaces with the smaller Mo atoms (Triffett and Bradshaw, 2008; Triffett et al., 2008).
sandwiched between these layers. Sulfur–sulfur distances are 2.98 Å Similarly, the shape characteristics of talc particles produced by ball,
within the layers and 3.95 Å between the layers. rod and autogenous mills were investigated by Yekeler et al. (2004)
using Scanning Electron Microscope (SEM). Their results have shown
2.1. Hydrophobic minerals with anisotropic surface properties — that elongation and smoothness helped to increase the hydrophobicity,
molybdenite while roundness and roughness caused a decrease in hydrophobicity
and floatability of talc. The grinding of talc powder decreased crystallin-
A mineral with anisotropic surface properties is characterized by dif- ity due to selective destruction of the (00l) plane and its hydrophobicity
ferent surface properties on different sides of the crystal. This is the case decreases by the grinding effect (Terada and Yonemochi, 2004).
of the sheet structure of inherently hydrophobic minerals, such as mo- For molybdenite particles the faces-to-edges ratio is a function of
lybdenite, talc, graphite, etc. (Chander et al., 1975). In general, minerals particle size, it decreases for fine particles which are therefore less hy-
with anisotropic surface properties do not have to be hydrophobic, as in drophobic than coarse ones (Castro and Correa, 1995; Yang et al.,
the case of clays, which are well recognized to have two types of 2014). Accordingly, small size particles show low floatability (Fig. 3).

Fig. 1. Crystalline structure of molybdenite.


50 S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58

Fig. 2. The effect of particle size and of MIBC concentration on cumulative molybdenite
recovery after 12 min of flotation at pH 11 (NaOH) and IsopX = 28 g/ton (Castro and
Mayta, 1994).

In addition, fine particles show slow flotation kinetics (Hsu, 1982;


Castro and Mayta, 1994). The flotation rate also depends on pH and is
claimed to be faster at pH 7.5 than at pH 11 (Hsu, 1982). The floatability
of the overground molybdenite is significantly reduced (Hsu, 1982).

2.3. Floatability of molybdenite and degree of crystallization

The degree of crystallization is one of the relevant factors affecting


molybdenite flotation (Hernlund, 1961; Shirley, 1981; Podobnik and
Shirley, 1982; Zanin et al., 2009). Well crystallized molybdenite is con-
sidered fast floating, while almost amorphous variety is either slow
floating or non-floating (Hernlund, 1961).
Triffett et al. (2008) described two types of molybdenite: vein con-
trolled and disseminated. The vein controlled molybdenite is the most Fig. 4. Photomicrographs of molybdenite particles separated from a flotation concentrate.

common and its occurrence varies depending on the size of the vein
in which it forms. Wider – quartz filled – veins tend to form larger, from well crystallized veins, lack of liberation takes place with dispersed
cleaner crystals of molybdenite while narrow veins result in a coating in quartz veins and veinlets, and losses by sensitivity toward copper de-
of much finer molybdenite on fractured surfaces which can take on a pressant is observed in sooty or smear molybdenite types.
smeared appearance due to in situ movement within these fractures.
The main types of molybdenite in copper porphyry ores are: 2.4. Surface composition of molybdenite and its inherent hydrophobicity
(1) coarse, clean molybdenite with perfect basal cleavage which occurs
in quartz veins, (2) fine-grained scaly molybdenite which appears as Basal planes of molybdenite crystals are formed by rupture of the S–
fine coatings along the joints and fractures of breccia, and (3) amor- S week van der Waals bonds, and are considered hydrophobic due to
phous and dull variety (Sutulov, 1971; Hsu, 1982). Sutulov claimed the surface content of S atoms. According to Triffett et al. (2008), slower
that problems in the recovery of molybdenite arise mainly from the floating molybdenite shows decreasing S content. Therefore, particles
amorphous and finely disseminated molybdenite. According to Shirley reporting to the tailings have a lower concentration of S than those
(1979), molybdenite is commonly lost in copper flotation circuits, as a recovered with concentrates. Surface analysis of molybdenite particles
result of the following problems: (a) lack of liberation; (b) large flakes; by the Time of Flight Laser Ionization Mass Spectroscopy (TOFLIMS)
and (c) oxide coatings. This author observed that large flakes are formed has indeed demonstrated that surface S is depleted on slower floating
particles, while surface oxidation (MoO3) and calcium coating is
increased.

2.5. Atomic force microscopy of basal planes of molybdenite

Traditionally, the basal planes of molybdenite have been considered


homogeneously hydrophobic. However, this approach does not describe
well various phenomena. Recent atomic force microscopy (AFM) studies
have revealed that the faces of molybdenite particles are heterogeneous
in nature and terraces and clusters of micro-crystals of molybdenite hav-
ing nano-edges and nano-faces (Fig. 5) are observed on such faces. In ad-
dition, these studies have shown that the faces possess a high level of
roughness. Such heterogeneity of the faces leads to both hydrophilic
and hydrophobic areas on the basal planes (Lopez-Valdivieso et al.,
2012). These results are in good agreement with Komiyama et al.'s
findings (2004) (in this work the atomic structure of the basal planes
Fig. 3. Native floatability of molybdenite of various particles size fractions as a function of was examined by Ultra-High Vacuum Scanning Tunneling Microscopy,
pH (Lopez-Valdivieso et al., 2006). UHV-STM, and crater structures with diameters ranging from 5 to
S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58 51

Fig. 5. AFM images of different areas of MoS2 faces.

9 nm were observed). These findings are important because they offer angle depends on the method of preparation of the face. For a face
further understanding for the adsorption of some inorganic ions on cleaved under water, they obtained a contact angle of 90°, while for a
faces, such as Ca and Mg ions, and of organic polymers such as dextrin high luster polished face the contact angle was 70°. The surface rough-
and starch, which strongly reduce molybdenite hydrophobicity. They ness and the anisotropic nature of the face then are the main factors
showed that contact angle measurements on crystal faces of molybde- that affect the contact angle values.
nite are performed on rough and anisotropic surfaces composed of Contact angle is also a function of the degree of surface oxidation.
both hydrophobic and hydrophilic areas. Roasting molybdenite decreases contact angle from 80° to 30° depend-
ing on temperature (Chander and Fuerstenau, 1972). Depressing re-
3. Molybdenite wettability agents are able to reduce molybdenite hydrophobicity, e.g., calcium
ions decrease contact angle measured on faces (see Fig. 6).
Contact angle studies performed on molybdenite have shown that Wie and Fuerstenau (1974) also measured the contact angle be-
faces are hydrophobic while edges are hydrophilic. Natural crystals of tween a free iso-octane droplet in water (across water phase) on mo-
molybdenite have been used to obtain a contact angle around 80° for lybdenite surfaces. At pH 4, the contact angle was determined to be
faces (Chander and Fuerstenau, 1972; Kelebek, 1988; Yang et al., about 150°, in alkaline solutions it decreased to 100°, and so these mea-
2014). However, Lopez-Valdivieso et al. (2006) found a contact angle surements turned out to be much more sensitive to pH (to electrical
of 61° at pH 5, which decreases slightly as pH increases (Fig. 6). Fig. 6 charge at solid/liquid interface). In further analysis of these data,
also shows that calcium ions are able to reduce contact angle measured Chander et al. (2007) demonstrated an excellent correlation between
on the crystal faces. Arbiter et al. (1975) have reported that contact molybdenite oil flotation recovery and oil/water contact angle of

Fig. 7. The effect of pH on the oil flotation of fine molybdenite without the addition of a
Fig. 6. Contact angle on edges and faces of molybdenite particles as a function of pH, in the surfactant. Also plotted is the effect of pH on the oil/water contact angle of molybdenite,
absence and presence of calcium ions (Lopez-Valdivieso et al., 2006). expressed in terms of the flotation dewetting relation (1 − cosθ) (Chander et al., 2007).
52 S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58

molybdenite shown in Fig. 7 (in oil flotation an oil phase is substituted


for the gaseous phase).
In the thermodynamic criterion of flotation, ΔGflot = γLV(cosθ − 1),
but for correlation purposes, the parameter was needed that increases
with increasing flotation recovery, and that is why in Fig. 7 the results
were plotted in terms of the quantity (1 − cos θ). Fig. 7 shows a very
good correlation between oil flotation of molybdenite and the wettabil-
ity of molybdenite surface by oil. Since these contact angle measure-
ments were carried out on freshly cleaved faces of a molybdenite
crystal, these data seem to indicate that the zero-point of charge of mo-
lybdenite faces is situated around pH 4.

4. Zeta potential of molybdenite

Since in the case of the minerals with anisotropic surface properties


the faces and edges are formed by rupture of different bonds these
faces and edges carry different electrical charges. Electrical charge of
solid particles suspended in water is commonly estimated from the elec- Fig. 8. Effect of pH on zeta potential of molybdenite (+, after Lopez-Valdivieso et al., 2012)
trokinetic measurements and the value of the zeta potential. Micro- and of paraffin wax (●, after Parreira and Schulman, 1961).
electrophoretic equipment is utilized in such measurements and
Smoluchowski equation is used to calculate the zeta potential. However, formation and specific adsorption sites at the solid/solution interface of
this equation was derived for spherical isotropic particles and it is not HMoO− 4 /MoO4
2−
ions, formed by oxidation of Mo(IV) to Mo(VI), and by
clear what is calculated if the particles used in the experiment are non- the subsequent hydrolysis reactions (Chander and Fuerstenau, 1972).
spherical and carry different electrical charges on different sides of the
particles (Laskowski, 2012). Such experimental problems can result in 4.1. Slime coating and hetero-coagulation
misleading conclusions as was demonstrated in the measurements with
kaolinite (Johnson et al., 2000) and talc (Burdukova et al., 2007; Yan The generic term “slime coating” refers to the process in which
et al., 2013). “slimy” hydrophilic gangue particles coat the surfaces of valuable min-
In this paper we are dealing with the case of molybdenite, also an an- erals and make them hydrophilic (slime coating can also take place be-
isotropic mineral, which has very hydrophobic faces. This results in a tween “slimy” molybdenite and coarse hydrophilic gangue). The slime
high natural floatability of this mineral. However, since it is anisotropic, coating phenomenon has been identified as one of the factors affecting
finding correlations between the measured hydrophobicity and molybdenite recovery. Low hydrophobicity and poor floatability of mo-
floatability, or zeta potential and floatability, is not an easy task. lybdenite can result from slime coating on molybdenite. It is likely that
Healy and Fuerstenau (2007) for zero field strength materials, fine particles of non-sulfide gangue (quartz, feldspar, clays, etc.) adhere
i.e., nonpolar or hydrophobic materials, propose that the pzc will be in electrostatically to the molybdenite surface (Hsu, 1982; Triffett et al.,
the low acidic pH 1–3 region, depending on the value assigned to 2008). Lopez-Valdivieso (2004) has found that molybdenite fines are at-
water molecules or clusters at the interface; but for molybdenite the ex- tached to the surface of coarse quartz and pyrite in the feed of the rough-
perimental evidences presented show fluctuating values between pH 2 er flotation of the porphyry copper ore at Mexicana de Cobre (see Fig. 9).
and 4. As summarized by Beattie (2007), almost all of the experimental Hernlund (1961) claims that slime coating prevent the adsorption of col-
evidence is consistent with the idea that water at hydrophobic surfaces lector on the surface of molybdenite and it deteriorates its floatability.
acquires a negative charge above pH 3–4 from the preferential adsorp- The mechanism of slime coating is widely believed to be a coagulation
tion of hydroxide ions. Parreira and Schulman (1961) in the electroki- of the oppositely charged mineral particles on the surface of molydenite.
netic experiments with very pure paraffin wax found that the zeta As it is known, calcium ions are able to reduce or to reverse the zeta po-
potential of this solid exhibits i.e.p. value around pH 4 and that it is neg- tential of quartz from negative to positive values. Raghavan and Hsu
atively charged in distilled water. Zeta potential measurements for mo- (1984) reported that hetero-coagulation of molybdenite and quartz
lybdenite give very negative values over the whole pH range (Chander takes place mainly in the aqueous dispersions containing lime. They
et al., 1975; Lopez-Valdivieso, 1980). This is shown by the bottom also found that coarse molybdenite particles (120–200 mesh) float ex-
curve after Lopez-Valdivieso et al. (2012) in Fig.8. It can be expected tremely rapidly and completely without being influenced by hetero-
that for anisotropic minerals both the charge of the faces and that of coagulation. But the flotation of fine molybdenite particles (− 200
the edges contribute to the overall zeta potential of molybdenite parti- mesh) was highly sensitive to hetero-coagulation.
cles. As Fig. 7 indicates, the point-of-zero charge of molybdenite faces Dispersing agents, such as sodium silicate and polyphosphates are
seems to be situated around pH 4. If we assume that the zeta potential recommended to prevent the slime coating (Park and Jeon, 2010;
of the hydrophobic faces can be described by the Parreira and Huynh et al., 2000). Fig. 10 shows that hexametaphosphate improves
Schulman's data (1961) obtained for a hydrophobic paraffin wax, then the recovery of molybdenite in the rougher flotation of porphyry copper
the obtained picture (Fig. 8) can be used to estimate the electrical ore with 0.039% Mo. 7 g/ton thionocarbamate collector for copper and
charge of the edges. 5 g/ton non-polar collector for molybdenite were used in the flotation
Fig. 8 indicates that at pH N 4 the zeta potential of molybdenite is tests (Lopez-Valdivieso, 2004).
very negative since over this pH range both the faces and the edges
are charged negatively. Below pH 4, that is, below the i.e.p. for molybde- 5. Molybdenite and flotation reagents
nite faces, the charge on faces becomes positive while it must still be
very negative for the edges since the overall zeta potential data are 5.1. Interactions with thiol collectors
still very negative.
The origin of the electrical charge at the edges is very different from Relatively few papers have been published on the interaction of mo-
the way the charge is generated at the faces. The edges are chemically ac- lybdenite with thiol collectors (xanthates, dithiphosphates, dithiocarba-
tive and react with water or oxygen to form oxidized surface species. The mates, etc.). Xanthates are weak collectors for molybdenite and
negative electric charge at edges has been interpreted as a result of the molybdenite is commonly floated with the addition of a non-polar oily
S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58 53

Fig. 9. Slime coating of molybdenite on pyrite and quartz in the feed of rougher flotation of porphyry copper ore at Mexicana de Cobre, pH 10.5 (Lopez-Valdivieso, 2004).

collector. Fine particles of molybdenite (6.8 μm) practically do not re- depressant activity on molybdenite. However, this depression effect
spond to an increase in xanthate dose, but coarse particles (51.7 μm) flo- was overcome by small additions of fuel oil (Simpson et al., 1983).
tation recovery and rate are significantly enhanced (Castro and Mayta, Chander and Fuerstenau (1974a,b) studied the adsorption of potas-
1994). These results suggest that the effectiveness of xanthate as collec- sium diethyldithiophosphate (KDTP) on the surface of molybdenite as a
tor increases with particle size due to its higher degree of inherent hy- function of pH. They attributed the slight increase in contact angle to the
drophobicity given by the larger faces/edges ratio. physical adsorption of the oxidation product of KDTP, i.e. (DTP)2. The
In the presence of dissolved oxygen semi-conducting minerals (e.g. adsorption of KDTP on the molybdenite surface may proceeds through
sulfides) are able to electro-catalytically oxidize xanthate ions to form three mechanisms: (a) a reaction of ion exchange between the DTP−
non-polar dimmer molecules (e.g. dixanthogen, i.e., a dithiolate), which ions and the HMoO− 2−
4 or MoO4 adsorbed ions; (b) a catalytic oxidation
adsorb onto the faces sites, e.g. for graphite and molybdenite (Afenya, of the collector with formation of (DTP)2; (c) a cooperative adsorption
1982; Allison and Finkelstein, 1971; Allison et al., 1972). Pieces of evi- of DTP− ions favored by (DTP)2. The path (c) proceeds by electrochem-
dence of xanthate adsorption on molybdenite have been reported by ical mechanisms.
Russian researchers, as it was summarized by Castro (1979). However,
no insoluble metal xanthate has been detected on molybdenite. Allison 5.2. Interaction with non-polar oils
and Finkelstein (1971), and Allison et al. (1972) confirmed by a spectro-
photometric technique the formation of dixanthogen and another not- Commonly oily collectors are used to promote molybdenite flotation
identified product on molybdenite with different alkyl xanthates. The (and all other inherently hydrophobic minerals). For primary molybde-
molybdenite rest potential was +0.160 V (for 6.25 × 10−4 M KEX and nite ores a combination of sodium silicate (Na2SiO3) and lime (CaO) is
pH 7) and dixanthogen was the main product identified. Atomic force mi- used at pH 9–11. Sodium silicate (water glass) is employed as disper-
croscopy (AFM) was applied to study in situ the adsorption of potassium sant of the slime coating formed on the mineral particles, and as depres-
ethyl xanthate (KEX) on molybdenite at pH 11. AFM images show a very sant of gangue minerals (Park and Jeon, 2010). Another important factor
sparse adsorption on the basal planes. Adsorbed species are much less on is the use of an emulsifying agent for the non-polar oil collector. It is well
molybdenite compared with chalcopyrite under similar conditions. Then, established that the addition of the emulsifying agent Syntex-L im-
the obtained AFM images show a low adsorption of KEX on the molybde- proved molybdenite flotation at Climax mill, Colorado (Hoover and
nite basal planes, explaining why KEX is not an efficient collector for mo- Malhotra, 1976). Nishkov et al. (1994) studied the use emulsifying
lybdenite (Zhang and Zhang, 2010). Potassium trithiocarbonate (KTTC), a agents for diesel oil in the flotation of El Teniente Cu–Mo ore, reporting
degradation product of potassium ethylxanthate, exhibits a certain a significant improvement in molybdenite recovery.
Hydrocarbon oils are the most common collectors for molybdenite, in-
cluding kerosene, diesel oil, transformer oil, etc. Petroleum hydrocarbons
differed in their effectiveness for enhancing the natural floatability of mo-
lybdenite. Hydrocarbon oils of different aliphatic and aromatic content
have been tested on various types of molybdenite ores (Smith and
Bhappu, 1971). They found that for an unaltered ore containing very little
or no clays, highly aromatic oils proved the most effective for all types of
molybdenite occurrence, crystalline, finely divided, and sooty. However,
for altered ores, highly saturated oils proved more effective than the high-
ly aromatic oils (Hsu, 1982). Smit and Bhasin (1985) investigated in the
AMAX plant a series of commercial oils which were blended with
diesel fuel to form different molybdenite non-polar collectors. A two-
component blend of a high and low molecular-weight diluent oils dem-
onstrated to be better molybdenite collector than single-component oil.
Furthermore, collector blended with naphthenic oils gave superior flota-
tion results to those blended with aromatic oils.
Kerosene is widely used as the conventional non-polar collector for
molybdenite. However, it is difficult to ensure its stable composition,
which leads to a harmful effect on the molybdenum production, partic-
ularly with difficult ores, i.e., lower-grade and fine-disseminated ores
(He et al., 2011). As a substitute, these authors studied the influence
of diesel oil from different manufacturers on the flotation of molybde-
Fig. 10. Flotation recovery of Mo in copper rougher flotation of Mexicana porphyry copper
ore (0.039% Mo) without and with 10 g/ton sodium hexametaphosphate in the ore
nite and the influence of pulp temperature on the dispersibility of diesel
grinding. These are batch flotation tests with 7 g/ton thionocarbamate collector and 5 g/ oil. In flotation pulp temperatures ranging from 10° to 30 °C, the flota-
ton non-polar oil, pH 10.5 (Lopez-Valdivieso, 2004). tion recovery of molybdenite increases with increasing high-boiling
54 S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58

component in diesel oil. When pulp temperature is below 10 °C, the flo- weight (Beaussart et al., 2012). Ansari and Pawlik (2007a,b) studied
tation recovery of molybdenite is related to the dispersibility of diesel the adsorption and molybdenite depression by lignosulfonate, and con-
oil, i.e., the proportion of high-boiling and low-boiling component in cluded that the high adsorption density of lignosulfonates on molybde-
diesel oil. Therefore, a molybdenum flotation plant should not blindly nite was primarily due to the adsorption of high molecular weight
apply diesel oil instead of kerosene as the collector for molybdenite, fractions. Additionally, other natural polymers such as fulvic and
but should select diesel oil that is suitable for the properties of its ore. humic acids – which are poorly defined anionic polymers with phenolic
Many studies (Crozier, 1979; Smit and Bhasin, 1985; Xia and Peng, and carboxylic groups (products of plant decay) – have also been shown
2007; Wang et al., 2008; Song et al., 1999) have shown that the to depress flotation recovery of molybdenite in acidic solutions
collecting ability of hydrocarbon oils enhance as its carbon chain length (Laskowski and Yu, 1994) similarly to other inherently hydrophobic
increases, which contributes to the enhancement of collecting ability for minerals, such as graphite (Wong and Laskowski, 1984); and coal
coarse molybdenite and intergrowths of molybdenite with gangue. (Pawlik et al., 1997).
However, too long a carbon chain brings down the dispersibility of hy-
drocarbon oil in pulp, causing the adverse effect and recovery decline
6.2. Depression by flocculants
of the valuable mineral. As summarized by Bulatovic (2007) low-
viscosity oils give lower recoveries than the high viscosity oils but at a
Most of the commonly used flocculants depress molybdenite flota-
somewhat higher concentrate grade. Pine oil and alkoxy paraffins give
tion. Therefore, flocculants should not be utilized on middling thick-
better MoS2 recovery than MIBC or polyglycol, and the use of emulsifier
eners in the copper circuits or in the molybdenite plant feed thickener
improves molybdenum recovery significantly.
unless absolutely necessary (Shirley, 1979).
The main function of flocculants used in mineral processing circuits
5.3. Interactions with methyl isobutyl carbinol (MIBC)
is to produce large and strong flocs. It is generally accepted that poly-
mers used as flocculants aggregate suspensions of fine particles by a
Methyl isobutyl carbinol (MIBC) (4-methyl-2-pentanol), a common
bridging mechanism. The bridging is considered to be a consequence
frother, is a partially soluble short-chain aliphatic alcohol commonly
of the adsorption of the segments of the flocculant macromolecules
employed as a week but selective frother in the flotation of Cu–Mo sul-
onto the surfaces of more than one particle. The optimum flocculation
fide ores. For molybdenite MIBC is considered to be an effective frother-
occurs at flocculant dosages corresponding to a particle coverage that
collector. The adsorption of MIBC on molybdenite has been studied by
is significantly less than complete. Incomplete surface coverage ensures
gas chromatography (Veliz and Molina, 1984). It was found that MIBC
that there is sufficient unoccupied surface available on each particle for
is strongly adsorbed on fine molybdenite (6–9 μm), and in pH ranging
the adsorption of flocculant segments during collision of the particles.
from 8 to 12 adsorption decreases as pH increases. Adsorption by hydro-
Low molecular weight polymers are often used as depressant for
phobic interactions was suggested. These results agree with those re-
naturally hydrophobic minerals; South African platinum industry uti-
ported by Castro et al. at pH 11 (1997). Castro and Mayta (1994)
lizes carboxymethyl cellulose, and guar gum, to depress talc
reported that molybdenite recovery is significantly increased by in-
(Steenberg and Harris, 1984).
creasing the dose of MIBC for particle sizes from 6.8 μm to 51.7 μm,
Polymer chains in solution can be irreversibly degraded by breaking
being higher for the 6.8 μm size fraction (the size with poorest flotation
of the polymer chains by shearing forces (shear degradation) (Al
recovery) (Fig. 11).
Hashmi et al., 2013). Then, polymers undergo chain scission and the
On the other hand, it has been found that MIBC is adsorbed on other
overall result is a loss in its flocculation efficiency; this, however, does
inherently hydrophobic minerals promoting their flotation. For exam-
not reduce depressing ability of these polymers (Castro and
ple, Fuerstenau and Pradip (1982) and Miller et al. (1983) have shown
Laskowski, 2013).
that MIBC is adsorbed onto coal. They found that oxidized coal is more
The floatability of molybdenite – similarly to other naturally hydro-
hydrophilic and the adsorption of the non-ionic MIBC molecules is re-
phobic minerals – is strongly depressed by polymers employed as floc-
duced after oxidation. The results indicate that adsorption occurs
culants for example by polyacrylamide flocculants and non-ionic
through hydrophobic interactions between the frother molecules and
flocculants like poly(ethylene oxide) (PEO) (Castro and Laskowski,
the coal surface. The adsorption of MIBC on coal is closely related to
1997, 2013, 2004, 2015). Low molecular-weight shear-degraded poly-
coal floatability and combustible recovery increases with the MIBC
acrylamides also depress molybdenite.
dose (Qu et al., 2013). Similarly, Kho and Sohn (1989) reported a good
flotation of talc with MIBC. In a more general view Hui et al. (2011) re-
ported that weekly hydrophilic compounds, like MIBC, are positively 7. Depression by hydrolyzable metal cations
adsorbed on low energy surfaces by hydrophobic attractive forces.
This behavior includes plastics, coal, graphite, talc, molybdenite, etc. Molybdenite floatability is highly affected by the presence of hydro-
lyzable metal cations commonly found in process water. Similar effect is
6. Depression of molybdenite reported for talc, which has identical anisotropic surface properties
(Fuerstenau et al., 1988). The natural floatability of talc is depressed
6.1. Depression by polysaccharides and other natural polymers within the pH range from the onset of the precipitation pH of metal hy-
droxide to the i.e.p. of the metal hydroxide. Adsorption of hydrolyzable
Polysaccharides such as starch and dextrin have been used as molyb- metal cations at the mineral–solution interface depends not only on the
denite depressants for a long time (Hernlund, 1961). The adsorption of ionic charge but also on these ions hydrolysis (James and Healy, 1972c).
dextrin occurs through physical interactions with the molybdenite sur- James and Healy (1972a,b,c) studied the adsorption of various metal
face, possibly due to hydrophobic bonding (Wie and Fuerstenau, 1974). cations as a function of pH on two oxides: SiO2 (low dielectric constant)
Dextrin was found to be a very effective molybdenite flotation depres- and TiO2 (high dielectric constant) as model systems. These authors
sant in the absence of a collector, whereas it does not effectively depress found that a qualitative relationship exists between the pH where an
molybdenite when oil is used as a collector. Recently, substituted dex- abrupt rise in adsorption is observed and the pH of hydrolysis of the
trin has been studied in more detail (Beaussart et al., 2012). The adsorp- metal cations. Hydration shells of the ions constitute a barrier to adsorp-
tion of three dextrins on molybdenite (Wheat dextrin, dextrin TY, tion of highly charged ions on the mineral surface; this makes direct ad-
carboxymethyl (CM) dextrin, and hydroxypropyl (HP) dextrin) were sorption of un-hydrolyzed metal cations very difficult (in acid pH) in
studied through Tapping Mode Atomic Force Microscopy (TMAFM) at spite of the strong electrostatic attraction forces between a cation and
plateau coverage. The isotherms revealed the importance of molecular negatively charged solid surface. Fig. 12 shows the zeta potential of
S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58 55

molybdenite in the absence and presence of the hydrolyzable metal cat-


ions Fe3+, Al3+ and Mg2+.

7.1. Depression by calcium ions

An excess of lime can depress molybdenite flotation in cleaner cir-


cuits due to the high pH needed to depress pyrite (pH 11–12). In
some cases, the sensitivity of molybdenite to lime appears to be reduced
by using an oily collector (Shirley, 1979).
To understand the role of lime on the floatability of molybdenite the
effect of calcium ions has been studied (Lenkovskaya and Stepanov,
1968; Chander and Fuerstenau, 1972; Castro and Paredes, 1979;
Hoover, 1980; Hsu, 1982; Raghavan and Hsu, 1984). Calcium ions may
act as both, a promoter or depressant of molybdenite flotation.
In acid media (pH 3.8 and 5.0), experimental evidence shows that the
zeta potential of molybdenite becomes less negative or reverses with in-
creasing calcium ion concentration. The adsorption mechanisms of calci- Fig. 11. Effect of MIBC concentration and of particle size on recovery of molybdenite;
um ions seems to be different to the physical adsorption described by flotation time 12 min at pH 11 (NaOH), and IsopX consumption of 28 g/ton (Castro and
James and Healy (1972a,b,c). Specific adsorption of calcium ions by Mayta, 1994).
chemical interactions with edges to form a surface coating of CaMoO4
has been suggested (Chander and Fuerstenau, 1972). The adsorbed particles, but also on the basal planes. These basal planes – as discussed
Ca2+ ions would neutralize an equal amount of MoO2− 4 ions, decreasing in this paper – exhibit the pzc around pH of 4. If the i.e.p. for molybdenite
the negative surface charge and the value of zeta potential. Over this acid faces is pH 4 then better flotation of molybdenite at pH 3.8 than 5 (Fig. 13)
pH range the flotation recovery is increased by calcium ions becoming a could be explained by the lack of the electrical charge around pH 3.8
maximum at the concentration of Ca2+ ions required for reversal of the on the hydrophobic faces which are responsible for molybdenite
zeta potential (PZR). Therefore, under conditions where the zeta poten- flotation
tial is near zero, the flotation is found to be the best (Fig. 13). The subse-
quent adsorption of Ca2+ ions may explain the loss of floatability 7.2. Depression in seawater — effect of magnesium ions
observed at calcium concentration higher than the point of PZR.
Chander and Fuerstenau (1972) attributed this flotation depression in Molybdenite is strongly depressed around pH 9.5–10.0 when Cu–Mo
acid media to the competition between compression of the double ores are floated in seawater. The most important variable is pH and the
layer and specific adsorption of Ca2+ ions on the crystal faces. depression takes place approximately in the pH range where magne-
However, in the alkaline pH range a strong molybdenite depression sium hydroxide (Mg(OH)2(s)) starts precipitating (Castro, 2012; Castro
is usually observed at pH N 8 (Hoover, 1980). It is also reasonable to ex- et al., 2012, 2014; Laskowski and Castro, 2012; Laskowski et al., 2013).
pect a chemical adsorption of Ca(OH)+ species to form surface calcium In spite that MgMoO4 may be formed in aqueous solutions (Essington,
molybdate, followed by adsorption of Ca(OH)+ species which increases 1992), that does not seem to be relevant for molybdenite depression
the surface electrical charge. Then, adsorbed Ca(OH)+ species may nu- in seawater and alkaline media. Castro et al. (2012) demonstrated that
cleate and co-precipitate as calcium hydroxide on edges, and in a the native floatability of molybdenite is not affected by the presence of
small amount on faces. anions, such as, SO2−
4 and HCO− 3 , at concentrations similar to those oc-
The calcium speciation diagram shows that by increasing pH, Ca2+ curring in seawater (see Fig. 15). However, a strong depression of mo-
ions hydrolyze to form in a first step hydroxy species (CaOH+) with lybdenite at pH ranging from 9.5–10.0 is produced in seawater, and
subsequent precipitation of colloidal calcium this depressing effect of Mg2+ ions is significantly stronger than the ef-
hydroxide (Ca(OH)2(s)). The precipitation of calcium hydroxide de- fect of Ca2+ ions.
pends on calcium concentration and pH. Lopez-Valdivieso et al. Taking into account that the molar concentration of Mg in seawater
(2006) for a total Ca ions concentration of 10−3 M reported precipita- is about four times that of Ca (around 1300 mg/L for Mg and 400 mg/L
tion at a pH of 11.5. Other authors identified the precipitation zone of for Ca), and the lower Ksp (magnesium hydroxide Ksp = 1.8 × 10−11,
calcium hydroxide at higher pH values (Fuerstenau and Palmer, calcium hydroxide Ksp = 6 × 10−6) and lower critical pH of precipita-
1976). Taking into account that usually molybdenite depression takes tion for Mg(OH)2(s), it is reasonable to expect that the depressing effect
place over the pH range from 8 to 12, it seems that the depression mech- of magnesium ions in seawater (and also in saline and hyper-saline wa-
anism not necessarily involves the precipitation of calcium hydroxide. ters) should be much higher than that of calcium ions. The effect of
Systematic studies revealed that Ca2+/Ca(OH)+ ions are strongly Mg2+ ions on molybdenite is related to its hydrolysis. As a first step
adsorbed on molybdenite (see Fig. 14) and that adsorption increases MgOH+ hydroxy-complexes are formed and adsorb on the surface of
with increasing pH (Hoover, 1980; Lenkovskaya and Stepanov, 1968; molybdenite; in a second step surface precipitation of hydrophilic col-
Hsu, 1982). The depression mechanism assumes that calcium species loidal magnesium hydroxide takes place around pH 10. These species,
are adsorbed only onto the negatively charged edge surface sites. Never- as shown in Fig. 12, can reverse the zeta potential of molybdenite. Ac-
theless, in order to understand molybdenite depression, the adsorption of cording to Li and Somasundaran (1991), the pH of precipitation of mag-
calcium species on the hydrophobic surface sites needs to be considered. nesium hydroxide varies from 9.2 to 11, for 10− 2 and 10− 5 M Mg
Contact angles of edges are approximately 0°, i.e., they are initially hydro- concentrations, respectively. The hetero-coagulation of these cationic
philic before the adsorption of Ca(OH)+ ions. Additionally, it has been colloidal species on the molybdenite surface forms a hydrophilic coating
reported that basal planes are very heterogeneous and not fully hydro- and renders it hydrophilic as it has been recently suggested (Castro
phobic due to the presence of nano-edges, terraces and craters, where et al., 2014).
Ca2+ or Ca(OH)+ ions can be adsorbed (Lopez-Valdivieso et al., 2012).
Experimental evidence shows that the adsorption of these hydroxy spe- 7.3. Depression by other metal cations
cies decreases the contact angle of molybdenite faces (Lopez-Valdivieso
et al., 2012). Therefore, it can be expected that in alkaline media Celik and Somasundaran (1986), and Somasundaran et al. (2000)
Ca(OH)+ ions can be adsorbed not only on the edges of molybdenite demonstrated that multivalent ions, such as, Ca2 +, Fe3 +, and Al3 +
56 S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58

Fig. 12. Zeta potential of molybdenite in the absence and presence of Fe3+, Al3+ and Mg2+
cations (Lopez-Valdivieso, 1980).

depressed coal in the pH region of precipitation of metal hydroxides.


Adsorption tests as a function of pH showed that coal adsorbed these
multivalent ions in a similar way to oxides. For example, the adsorption
of Ca2 + and Al3 + ions on coal is governed by the formation of the
CaOH+ and AlOH2+ species, which drastically reduce the hydrophobic-
ity of coal, with good correlation between its floatability and the precip- Fig. 14. Adsorption of calcium ions by radioactive tracer technique as a function of free
Ca(OH)2 content in slurry (mg/L) and pH (adjusted with H2SO4 and NaOH)
itation of the metal hydroxy complexes. Molybdenite is depressed by a (Lenkovskaya and Stepanov, 1968; Hoover, 1980).
number of hydrolyzable metal cations, such as, Al3+, Pb2+, Fe2+, Fe3+,
Cu2+, etc. (Lenkovskaya and Stepanov, 1968; Hoover, 1980). The effect
of different ions on molybdenite flotation at pH 7.5 is shown in Fig. 16 In this paper – based on molybdenite surface anisotropic properties –
for Fe3+, Al3+, Cu2+, and Cu+ ions. we introduced the concept of various zeta potential-pH curves for mo-
It is to be pointed out that molybdenite case is different from other lybdenite faces and edges. Since for all hydrophobic surfaces the iso-
solids. MoO2− 4 surface sites at the edges may form insoluble salts with a electric points are reported to be in the range of pH 3–4, it is suggested
number of metal cations. The natural floatability of molybdenite in acid that the zeta potential-pH curve of molybdenite hydrophobic faces is
media seems to be affected by this phenomenon. However, in alkaline similar to such a curve for hydrophobic paraffin wax. Comparison
media molybdenite behaves like an oxide (James and Healy, 1972a,b,c), with the measured zeta potential values for molybdenite particles
the adsorption of the hydroxy species and the subsequent precipitation gives that the zeta potential of molybdenite hydrophilic edges is nega-
of the colloidal metal hydroxide seems to explain the depression tive in the whole pH range. This explains effect of pH on molybdenite
mechanism. floatability. AFM reveals that the molybdenite faces are not homoge-
neous and contain craters and nano-size structures. Such surfaces are
8. Summary then very heterogeneous not only because of the topography but also

Most of the factors affecting molybdenite floatability depend on


faces and edges surface sites which carry different electrical charges.

Fig. 13. Flotation recoveries of molybdenite as a function of calcium ion concentration at Fig. 15. Effect of pH (lime) on the natural flotation (+200 mesh pure molybdenite) in NaCl
an ionic strength of 10−2 M KCl (without flotation reagents) (Chander and Fuerstenau, solutions (0.6 M) containing different ions (SO2− 2−
4 ; CO3 ; Mg
2+
; and Ca2+ ions), without
1972). any collector and frother (Castro et al., 2012).
S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58 57

Ametov, I., Grano, S.R., Zanin, M., Gredelj, S., Magnuson, R., Bolles, T., Triffett, B., 2008. Cop-
per and molybdenite recovery in plant and batch laboratory cells in porphyry copper
rougher flotation. In: Wang, D.Z., Sun, C.Y., Wang, F.L., Han, L. (Eds.), Proceedings of
XXIV International Mineral Processing Congress. Science Press, Beijing,
pp. 1129–1137.
Ansari, A., Pawlik, M., 2007a. Floatability of chalcopyrite and molybdenite in the presence
of lignosulfonates. Part I. Adsorption studies. Miner. Eng. 20, 600–608.
Ansari, A., Pawlik, M., 2007b. Floatability of chalcopyrite and molybdenite in the presence
of lignosulfonates. Part II. Hallimond tube flotation. Miner. Eng. 20, 609–616.
Arbiter, N., Fujii, Y., Hansen, B., Raja, A., 1975. Surface properties of hydrophobic solids. In:
Somasundaran, P., Grieves, R.B. (Eds.), Advances in Interfacial Phenomena of Particu-
late/Solution/Gas Systems; Application to Flotation ResearchAIChE Symp. Series n°
150 vol. 71. American Institute of Chemical Engineering, New York, pp. 176–182.
Beattie, J.K., 2007. The intrinsic charge at the hydrophobe/water interface. In: Tadros, T.F.
(Ed.), Colloid Stability — The Role of Surface Forces, Part II. Wiley VCH Verlag,
pp. 153–164.
Beaussart, A., Parkinson, L., Miercsynska-Vasilev, A., Beattie, D.A., 2012. Adsorption of
modified dextrins on molybdenite: AFM imaging, contact angle and flotation studies.
J. Colloid Interface Sci. 368, 608–615.
Bulatovic, S.M., 2007. Handbook of Flotation Reagents: Chemistry, Theory and Practice.
Flotation of sulfide ores vol. 1. Elsevier B.V., Amsterdam, pp. 235–293.
Burdukova, E., Becker, M., Bradshaw, D.J., Laskowski, J.S., 2007. Presence of negative
charge on the basal planes of New York talc. J. Colloid Interface Sci. 315, 337–342.
Castro, S., 1979. In: A., S. (Ed.)Flotation of Molybdenite. In: International Molybdenum En-
cyclopaedia vol. II. Intermet Publications, Santiago-Chile, pp. 164–179.
Castro, S., Paredes, S., 1979. The depressant effect of lime on molybdenite flotation. Proc.
of the First Latin American Congress on Flotation. University of Concepcion,
pp. 290–306 (Spanish text).
Castro, S.H., Mayta, E., 1994. A kinetics approach to the effect of particle size on the flota-
tion of molybdenite. In: Castro, S., Alvarez, J. (Eds.), Flotation. A. Sutulov Memorial
Volume, pp. 331–344 (Concepción-Chile).
Castro, S., Correa, A., 1995. The effect of particle size on the surface energy and wettability
of molybdenite. Vancouver: 1st UBC-McGill International Symposium on Processing
of Hydrophobic Minerals and Fine Coal. CIM MET SOC, pp. 43–57.
Castro, S., Laskowski, J.S., 1997. The effect of hydrophilic and hydrophobic polymers on
molybdenite flotation. Proc. 5th Southern Hemisphere Meeting on Mineral Technol-
ogy. Intemin-Segemar, Buenos Aires, Argentine, pp. 117–120.
Fig. 16. Effect of soluble salts on floatability of molybdenite by xanthates (25 mg/L) at Castro, S., Stocker, R., Laskowski, J.S., 1997. The effect of hydrophobic agglomerants on the
pH 7.5 (Lenkovskaya and Stepanov, 1968; Hoover, 1980). flotation of fine molybdenite particles. Proceedings 20th IMPC. Aachen vol. 3,
pp. 559–572.
Castro, S., Laskowski, J.S., 2004. Molybdenite depression by shear degraded
because of difference in the electrical charge on these surface sites. This polyacrilamide solutions. In: Laskowski, J.S. (Ed.), Particle Size Enlargement in Miner-
might explain a strong effect of slime coatings, polycations, polysaccha- al Processing — Proc. 5th UBC-McGill Int. Symp.CIM Met Soc., Hamilton, pp. 169–178
Castro, S., 2012. Challenges in flotation of Cu–Mo sulfide ores in sea water. Water in Min-
rides and flocculants on molybdenite floatability. Likewise, this face het- eral Processing — Proceedings of the 1st International Symposium. SME, pp. 29–40.
erogeneity seems to be an important factor in the interaction of bubbles Castro, S., Rioseco, P., Laskowski, J.S., 2012. Depression of molybdenite in sea water. 26th
and oil droplets on molybdenite faces affecting the floatability of molyb- International Mineral Processing Congress, New Delhi, 2012, pp. 737–752.
Castro, S., Laskowski, J.S., 2013. The effect of flocculants and their degradation products on
denite particles. molybdenite flotation. Proc. Copper 2013 Conference, Santiago-Chile.
It is interesting to add that in the just published paper the zeta po- Castro, S., Uribe, L., Laskowski, J.S., 2014. Depression of inherently hydrophobic minerals
tential vs. pH curves for molybdenite faces and edges, back calculated by hydrolyzable metal cations: molybdenite depression in seawater. Proc. XVII
IMPC 2014, Santiago-Chile, pp. 737–752.
from the AFM measurements, were provided by Lu et al. (2015).
Castro, S., Laskowski, J.S., 2015. Depressing effect of flocculants on molybdenite flotation.
These experimentally determined values are different from what we es- Miner. Eng. 74, 13–19.
timated (Fig. 8). The zeta potential values for the edges given by Lu et al. Chander, S., Fuerstenau, D.W., 1972. On the natural floatability of molybdenite. Trans.
AIME 252, 62–69.
(2015) rather resemble our estimated curve for the faces. These discrep-
Chander, S., Fuerstenau, D.W., 1974a. Electrochemical study of the molybdenite–potassi-
ancies seem to point out that the traditional picture of the molybdenite um diethyldithiophosphate system. Trans. SME 265, 193–197.
face as a homogeneously hydrophobic surface is not necessarily correct Chander, S., Fuerstenau, D.W., 1974b. The effect of potassium diethyldithiophosphate on
(Komiyama et al., 2004; Lopez-Valdivieso et al., 2012). the interfacial properties of molybdenite. Trans. IMM 83, C180–C185.
Chander, S., Wie, J.M., Fuerstenau, D.W., 1975. On the native floatability and surface prop-
erties of naturally hydrophobic solids. In: Somasundaran, P., Grieves, R.B. (Eds.), Ad-
vances in Interfacial Phenomena of Particulate/Solution/Gas Systems; Applications to
Acknowledgments
Flotation Research. AIChE Symposium Series, 150 vol. 71, pp. 183–188.
Chander, S., Hogg, R., Fuerstenau, D.W., 2007. Characterization of the wetting and
The authors gratefully acknowledge the Water Research Center for dewetting behavior of powders. Kona 25, 56–75.
Agriculture and Mining (CRHIAM), University of Concepcion; and the Celik, M.S., Somasundaran, P., 1986. The effect of multivalent ions on the flotation of coal.
Sep. Sci. Technol. 21 (4), 393–402.
Chilean Research Council (CONICYT) for financial support through the Dikinson, R.G., Pauling, L., 1923. The crystal structure of molybdenite. J. Am. Chem. Soc. 45,
FONDAP—15130015 grant. 1466–1471.
Crozier, R.D., 1979. Flotation reagent practice in primary and by-product molybdenite re-
covery. Min. Mag. 140, 174–178.
References Essington, M.E., 1992. Formation of calcium and magnesium molybdate complexes in di-
luted aqueous solutions. Soil Sci. Soc. Am. J. 56 (4), 1124–1127.
Afenya, P.M., 1982. Adsorption of xanthate and starch on synthetic graphite. Int. J. Miner. Fuerstenau, M.C., Palmer, B.R., 1976. Anionic flotation of oxides and silicates. In:
Process. 9, 303–319. Fuerstenau, M.C. (Ed.), Flotation. A.M. Gaudin Memorial Volume vol. 1. AIME, New
Al Hashmi, A.R., Al Maamari, R.S., Al Shabibi, I.S., Mansoor, A.M., Zaitoun, A., Al Sharji, H.H., York, pp. 148–196.
2013. Rheology and mechanical degradation of high-molecular-weight partially hy- Fuerstenau, D.W., Pradip, 1982. Adsorption of frothers at coal/water interfaces. Colloids
drolyzed polyacrylamide during flow through capillaries. J. Pet. Sci. Eng. 105, Surf. 4, 213–227.
100–106. Fuerstenau, M.C., Lopez-Valdivieso, A., Fuerstenau, D.W., 1988. Role of hydrolyzed cations
Allison, S.A., Finkelstein, N.P., 1971. Study of the products of reactions between galena and in the natural hydrophobicity of talc. Int. J. Miner. Process. 23, 161–170.
aqueous xanthate solutions. Trans. Inst. Min. Metall. Sect. C 80, C235–C239. Gaudin, A.M., Miaw, H.L., Spedden, H.R., 1957. Native floatability and crystal structure.
Allison, S.A., Goold, L.A., Nicol, M.J., Granville, A.D., 1972. A determination of the products Proc. 2nd Int. Congress of Surface Activity vol. 3. Butterworths, London, pp. 202–219.
of reaction between various sulfide minerals and aqueous xanthate solution, and a Guang-yi, L., Yi-ping, L., Hong, Z., Zhan-fang, C., Zheng-he, X., 2012. A novel approach for
correlation of the products with electrode rest potentials. Metall. Trans. 3, preferential flotation recovery of molybdenite from a porphyry copper–molybdenum
2613–2618. ore. Miner. Eng. 36–38, 37–44.
58 S. Castro et al. / International Journal of Mineral Processing 148 (2016) 48–58

Healy, T.W., Fuerstenau, D.W., 2007. The isoelectric point/point-of zero-charge of inter- Nakhaei, F., Irannajad, M., 2014. Investigation of effective parameters for molybdenite re-
faces formed by aqueous solutions and nonpolar solids, liquids, and gases. J. Colloid covery from porphyry copper ores in industrial flotation circuit. Physicochem. Probl.
Interface Sci. 309, 183–188. Miner. Process. 50 (2), 477–491.
He, T., Wan, H., Song, N., Guo, L., 2011. The influence of composition of nonpolar oil on flo- Nishkov, I., Lazarov, D., Martinov, M., Beas, E., Henriquez, C., 1994. Surfactant-
tation of molybdenite. Miner. Eng. 24, 1513–1516. hydrocarbon oil emulsions for molybdenite flotation. In: Castro, S., Alvarez, J. (Eds.),
Hernlund, R.W., 1961. Extraction of molybdenite from copper flotation products. Q. Flotation. A Volume in memory of Alexander Sutulov. University of Concepcion,
J. Colorado Sch. Min. 56, 177–196. pp. 319–329.
Hoover, M.R., Malhotra, D., 1976. Emulsion flotation of molybdenite. In: Fuerstenau, M.C. Park, C., Jeon, H., 2010. The effect of sodium silicate as pH modifier and depressant in the
(Ed.)Flotation A. M. Gaudin Memorial Volume vol. 1. AIME, pp. 485–505. froth flotation of molybdenite ores. Mater. Trans. 51 (7), 1367–1369.
Hoover, M.R., 1980. Water chemistry effects in the flotation of sulphide ores — a review Parreira, H.C., Schulman, J.H., 1961. Streaming potential measurements on paraffin wax.
and discussion for molybdenite. In: Jones, M.J. (Ed.), Complex Sulphide Ores. IMM, In: Gould, R.G. (Ed.)Advances in Chemistry Series vol. 33. American Chem. Society,
London, pp. 100–112. pp. 160–171.
Hsu, L.K.L., 1982. Surface Chemistry of Molybdenite (MoS2) with Special Reference to its Pawlik, M., Laskowski, J.S., Liu, H., 1997. Effect of humic acid and coal surface properties on
Floatability from Porphyry Copper Ores (Master of Science Thesis) The University rheology of coal–water slurries. Coal Prep. 18, 129–150.
of Arizona. Podobnik, D.M., Shirley, J.F., Oct. 1982. Molybdenite recovery at Cuajone. Min, Eng.
Hui, W., Chao, G., Jiangang, F., Zhangxing, H., Wei, L., Xiaolei, C., Caihong, Z., 2011. Adsorp- 1473–1476.
tion behavior of week hydrophilic substances on low energy surface in aqueous Qu, X., Wang, L., Nguyen, A., 2013. Correlation of air recovery with froth stability and sep-
media. Appl. Surf. Sci. 257, 7959–7967. aration efficiency in coal flotation. Miner. Eng. 41, 25–30.
Huynh, L., Feiler, A., Michelmore, A., Ralston, J., Jenkins, P., 2000. Control of slime coating Raghavan, S., Hsu, L.L., 1984. Factors affecting the flotation recovery of molybdenite from
by the use of anionic phosphates: fundamental study. Miner. Eng. 13 (10–11), porphyry copper ores. Int. J. Miner. Process. 12 (1–3), 145–162.
1059–1069. Simpson, W.W., Nichols, I.L., Huiatt, J.L., 1983. Effect of potassium ethylxanthate degrada-
James, R.O., Healy, T.W.J., 1972a. Adsorption of hydrolysable metal ions at the oxide– tion on flotation of chalcopyrite and molybdenite. Bur. Mines Rep. Investig. 8787.
water interface, part I. J. Colloid Interface Sci. 40, 42–52. Steenberg, E., Harris, P.J., 1984. Adsorption of carboxymethylcellulose, guar gum and
James, R.O., Healy, T.W.J., 1972b. Adsorption of hydrolysable metal ions at the oxide– starch onto talc, sulphides, oxides and salt-type minerals. S. Afr. J. Chem. 37, 85–102.
water interface, part II. J. Colloid Interface Sci. 40, 53–63. Shirley, J.F., 1979. By-product molybdenum recovery. In: A., S. (Ed.)International Molyb-
James, R.O., Healy, T.W.J., 1972c. Adsorption of hydrolysable metal ions at the oxide– denum Encyclopaedia vol. II. Intermet Publications, Santiago-Chile, pp. 37–56.
water interface, part III. J. Colloid Interface Sci. 40, 65–81. Shirley, J.F., March 1981. Byproduct molybdenite plant design. Can. Min. J. 27–28.
Johnson, S.B., Franks, G.V., Scales, P.J., Boger, D.V., Healy, T.W., 2000. Surface chemistry — Smit, F.J., Bhasin, A.K., 1985. Relationship of petroleum hydrocarbon characteristics and
rheology relationships in concentrated minerals suspensions. Int. J. Miner. Process. molybdenite flotation. Int. J. Miner. Process. 15, 19–40.
58, 267–304. Smith, J.H., Bhappu, R.S., 1971. Paper Presented at the AIME Pacific Southwest Mineral In-
Kelebek, S., 1988. Critical surface tension of wetting and of floatability of molybdenite and dustry Conference, Reno, Nevada, May 5–7.
sulphur. J. Colloid Interface Sci. 124, 504–514. Song, S.X., Lopez-Valdivieso, A., Ding, Y.Q., 1999. Effects of nonpolar oil on hydrophobic
Kho, C.J., Sohn, H.J., 1989. Column flotation of talc. Int. J. Miner. Process. 27, 157–167. flocculation of hematite and rhodochrosite fines. Powder Technol. 101, 73–80.
Komiyama, M., Koyohara, K., Fujikawa, T., Ebihara, T., Kubota, T., Okamoto, Y., 2004. Crater Somasundaran, P., Zhang, L., Fuerstenau, D.W., 2000. The effect of environment, oxidation
structure on a molybdenite basal plane observed by ultrahigh vacuum tunneling mi- and dissolved metal species on the chemistry of coal flotation. Int. J. Miner. Process.
croscopy and its implication to hydrotreating. J. Mol. Catal. A Chem. 215, 143–147. 58, 85–97.
Laskowski, J.S., Kitchener, J.A., 1969. The hydrophobic–hydrophilic transition on silica. Sutulov, A., 1971. Recovery of Molybdenum and Rhenium from Porphyry Copper Ores.
J. Colloid Interface Sci. 29, 670–679. University of Conception, Chile.
Laskowski, J.S., Yu, Z., 1994. The effect of humic acids on the emulsion flotation of inher- Sutulov, A., 1979. International Molybdenum Encyclopaedia vol. II. Intermet Publications,
ently hydrophobic minerals. In: Castro, S., Alvarez, J. (Eds.), Flotation — Proc. of the Santiago-Chile.
4th Meeting of the Southern Hemisphere vol. 2. University of Concepcion, Terada, K., Yonemochi, E., 2004. Physicochemical properties and surface free energy of
pp. 397–411. ground talc. Solid State Ionics 172, 459–462.
Laskowski, J.S., 2012. Anisotropic minerals in flotation circuits. Canadian mineral process- Triffett, B., Bradshaw, D., 2008. The role of morphology and host rock lithology on the flo-
ing. CIM J. 3 (4), 203–213. tation behaviour of molybdenite at Kennecott Utah Copper. AusIMM Publication Se-
Laskowski, J.S., Castro, S., 2012. Hydrolyzing ions in flotation circuits: sea water flotation. ries, 9th International Congress for Applied Mineralogy, ICAM 2008 — Proceedings,
Proc. 13th International Mineral Processing Symposium (IMPS 2012), Bodrum- pp. 465–473.
Turkey, October 10–12, pp. 219–228. Triffett, B., Veloo, C., Adair, B.J.I., Bradshaw, D., 2008. An investigation of the factors affect-
Laskowski, J.S., Castro, S., Ramos, O., 2013. Effect of seawater main components on ing the recovery of molybdenite in the Kennecott Utah copper bulk flotation circuit.
frothability in the flotation of Cu–Mo sulfide ore. Physicochem. Probl. Miner. Process. Miner. Eng. 21 (12–14), 832–840.
50 (1), 17–29. Veliz, N., Molina, E., 1984. Study of the adsorption of frothers on molybdenite (Thesis De-
Lenkovskaya, G.L., Stepanov, B.A., 1968. Effect of the Ion Composition of the Pulp on Mo- partment of Chemical and Metallurgical Engineering) Universidad del Norte-Chile
lybdenite Flotation. Obogash Rud 13, n°1 pp. 6–9 (Russian text). (Spanish text).
Li, C., Somasundaran, P., 1991. Reversal of bubble charge in multivalent inorganic salt Wang, H., Gu, G.H., Fu, J.G., Hao, Y., 2008. Study of the interfacial interactions in the mo-
solutions—effect of magnesium. J. Colloid Interface Sci. 146, 215–218. lybdenite floatation system. J. China Univ. Min. Technol. 18, 82–87.
Lopez-Valdivieso, A., 1980. A Study of the Electrokinetics and Flotation Properties of Talc Wie, J.M., Fuerstenau, D.W., 1974. Effect of dextrin on surface properties and the flotation
and Molybdenite (Master of Science) South Dakota School of Mines and Technology. of molybdenite. Int. J. Miner. Process. 1, 17–32.
Lopez-Valdivieso, A., 2004. Design of a chemical scheme for the flotation of porphyry cop- Wong, K., Laskowski, J.S., 1984. Effect of humic acid on the properties of graphite aqueous
per ore with high contents of biotite. Internal Technical Report, Instituto de suspensions. Colloids Surf. 12, 317–332.
Metalurgia-Mexicana de Cobre, 2004. Xia, Y.K., Peng, F.F., 2007. Selection of frothers from residual organic reagents for copper–
Lopez-Valdivieso, A., Madrid-Ortega, I., Reyes-Bahena, J.L., Sánchez-López, A.A., Song, S., molybdenite sulfide flotation. Int. J. Miner. Process. 83, 68–75.
2006. Propiedades de la interface molibdenita/solución acuosa y su relación con la Yang, B., Song, S., Lopez-Valdivieso, A., 2014. Effect of particle size on the contact angle of
flotabilidad del mineral. Proc. XVI Congreso Internacional de Metalurgia Extractiva, molybdenite powders. Miner. Process. Extr. Metall. Rev. 35, 208–215.
Saltillo-México, pp. 226–235 (Spanish text). Yekeler, M., Ulusoy, U., Hiçyılmaz, C., 2004. Effect of particle shape and roughness of talc
Lopez-Valdivieso, A., Madrid-Ortega, I., Valdez-Pérez, D., Yang, B., Song, S., 2012. The het- mineral ground by different mills on the wettability and floatability. Powder Technol.
erogeneity of the basal plane of molybdenite: its effect on molybdenite floatability 140 (2004), 68–78.
and calcium ion adsorption. Santiago. : 9th International Mineral Processing Confer- Zanin, M., Ametov, I., Grano, S., Zhou, L., Skinner, W., 2009. A study of mechanisms affect-
ence. PROCEMIN, pp. 21–23. ing molybdenite recovery in a bulk copper molybdenum flotation circuit. Int. J. Miner.
Lu, Z., Qingxia, L., Xu, Z., Hongbo, Z., 2015. Probing anisotropic surface properties of mo- Process. 93, 256–266.
lybdenite by direct force measurements. Langmuir 11409–11418. Zhang, J., Zhang, W., 2010. Applying an atomic force microscopy in the study of mineral
Yan, L., Masliyah, J.H., Xu, Z., 2013. Understading suspension rheology of anisotropically- flotation. In: Méndez-Vilas, A., Diaz, J. (Eds.), Microscopy: Science, Technology, Appli-
charged platy minerals from direct interaction force measurement using AFM, cur- cations and Education, pp. 2028–2034.
rent opinion in coll. Interf. Sci. 18, 149–156.
Miller, J.D., Lin, C.L., Chang, S.S., 1983. MIBC adsorption at the coal/water interface. Col-
loids Surf. 7, 351–355.

S-ar putea să vă placă și