Você está na página 1de 14

ELSEVIER Int. J. Miner. Process.

55 (1999) 157–170

Effect of sodium sulfide additions on the pulp potential and


amyl xanthate flotation of cerussite and galena
R. Herrera-Urbina a,1 , F.J. Sotillo b , D.W. Fuerstenau a,*
a University of California at Berkeley, Department of Materials Science and Mineral Engineering,
Hearst Mining Building, Berkeley, CA 94720, USA
b Per Usa Environment Inc., Lakeland, FL 33813, USA

Received 15 November 1996; revised version received 5 April 1998; accepted 28 June 1998

Abstract

The flotation response of cerussite and galena treated with sodium sulfide solutions open
to the atmosphere was assessed at pH 9.5 using amyl xanthate as the collector. Since both
the hydrosulfide and xanthate anions are electrochemically active, the redox potential of the
suspensions was measured with a smooth-platinum=saturated calomel electrode system. Mineral
flotation depends on the sulfide dosage and is strongly affected by the potential. Cerussite
flotation is poor under sulfide-deficient and sulfide-rich conditions. Low sulfide dosages are
consumed by aqueous lead dissolved from the mineral and precipitate as lead sulfide. These
colloidal particles seem to deposit onto the cerussite surface and react with the collector. As
a result, a peak in the flotation curve is noted. High sulfide dosages oversulfidize the mineral
surface, shift the potential to reducing conditions, and depress the mineral. At two different
collector additions, optimum flotation of sulfide-treated cerussite occurs in the potential range
from 100 mV to about 150 mV. Galena flotation ceases above a certain sulfide dosage that
lowers the potential to reducing values thus inhibiting the chemisorption of xanthate onto the
mineral surface. The results of this research work indicate that the xanthate flotation of sulfidized,
mixed oxide–sulfide lead ores may be controlled by monitoring the pulp potential.  1999
Elsevier Science B.V. All rights reserved.

Keywords: cerussite; galena; sulfidization; pulp potential; amyl xanthate flotation

1 Present address: Universidad de Sonora, Departamento de Ingenierı́a Quı́mica y Metalurgia, Apartado

Postal 106, Hermosillo, Sonora, 83000, Mexico.


Ł Corresponding author. Tel.: C1 510 642 3826; Fax: C1 510 642 6623; E-mail:

dwfuerst@socrates.berkeley.edu

0301-7516/99/$ – see front matter  1999 Elsevier Science B.V. All rights reserved.
PII: S 0 3 0 1 - 7 5 1 6 ( 9 8 ) 0 0 0 2 9 - 5
158 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

1. Introduction

The effective concentration by flotation of mixed sulfide–oxide lead ores remains


a challenge to mineral engineers and scientists. Poor flotation response of these ores,
even at high levels of thio compounds used as collectors, is associated to the relatively
high solubility of the oxide-type lead minerals and oxidized–carbonated galena. High
collector consumption results from chemical reactions taking place both in the bulk and
at the mineral surface. The bulk reaction, however, has a dramatic impact in the flotation
performance of the system because it consumes a lot of collector. For the xanthate
flotation of cerussite, for example, collector requirements are about 20 times greater
than those for the flotation of galena (Gaudin, 1957).
Besides the high solubility and large specific area of the micro-crystalline cerussite,
the main reason for the high collector consumption appears to be the metathetical
surface reaction between xanthate and carbonate ions, forming solid lead xanthate and
releasing carbonate ions to the solution. This exchange reaction will continue with
the formation of multilayers of lead xanthate, proceeding until solution equilibrium
is reached. Fleming (1952) clearly showed that the flotability edge for cerussite is
controlled by the equilibrium ratio of carbonate and xanthate ions in solution. Recently,
an improvement in the xanthate flotation of cerussite has been reported to take place
when the mineral is conditioned in pure water and in lead solutions before the addition
of the collector (Popov and Vucinic, 1992a; Popov and Vucinic, 1992b).
Under usual laboratory=plant conditions galena readily oxidizes to lead sulfate and
lead thiosulfate (M.C. Fuerstenau et al., 1985; Pugh and Bergstrom, 1986). Furthermore,
in alkaline galena suspensions open to the atmosphere, the surface of oxidized galena
will carbonate at the expense of sulfate and thiosulfate. Lead carbonate has been
identified by FTIR spectroscopy as a surface product on wet ground galena (De Donato
et al., 1990). Surface oxidation and carbonation appear to be a necessary condition
for galena flotation with xanthate collectors (Gaudin, 1957; Page and Hazell, 1989;
De Donato et al., 1990). Eadington and Prosser (1969), and Toperi and Tolun (1969)
have explained the adsorption of xanthate on galena in terms of an electrochemical
mechanism. The chemical interaction between galena and xanthates has been related
to its rest potential (Allison et al., 1972). Lead xanthate was identified as the reaction
product, and the rest potential of the mineral was found to lie below the corresponding
reversible potential for the oxidation of xanthate to its disulfide. On the basis of modern
electrochemical studies a mixed-potential model has been proposed to rigorously explain
the electrochemical nature of the galena–xanthate interaction (Woods, 1984; Woods and
Richardson, 1986). The mixed-potential model states that this interaction involves
simultaneous electrochemical reactions: the cathodic step is usually the reduction of
oxygen, while the anodic step involves the oxidation of the xanthate. It is now widely
accepted that xanthate (X) chemisorbs on metal sites at the galena surface forming
a surface PbX chelate as the first layer (Buckley and Woods, 1994; Woods and
Richardson, 1986), over which layers of PbX2 build up (Woods and Richardson, 1986).
Buckley and Woods (1994) have concluded that there is no valid basis for the hypothesis
that adsorption of molecular lead xanthate is responsible for the xanthate flotation of
galena.
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 159

Sulfide compounds have long been used in oxide type base-metals mineral flotation
practice for decreasing the level of metal species in solution and for modifying the
surface of the mineral while still having a surface that will respond to flotation. In
sulfide minerals flotation environments made reducing by the addition of sulfide salts,
the sulfide ion acts as depressant. The surface characteristics of cerussite and galena in
sodium sulfide solutions have been investigated in detail by Herrera-Urbina et al. (1998).
Sulfidization of the mineral surface significantly improves the flotation of cerussite with
amyl xanthate as collector (M.C. Fuerstenau et al., 1985, 1987; Sotillo and Fuerstenau,
1988). The xanthate flotation of both sulfidized cerussite and oxidized galena, however,
is strongly affected by the sulfide dosage. Once the sulfide ion reports to the solution
at the end of the conditioning period, mineral flotation ceases. Depression of sulfidized
cerussite is analogous to the depression of galena by the addition of sulfide salts.
Several phenomena may participate in this process. First of all, lead sulfide is more
insoluble than lead xanthate. Furthermore, if the xanthate ion interacts with lead sulfide
by adsorption on a lead-surface site or by ion-exchange with oxidized–carbonated
products, the addition of sulfide could inhibit adsorption and ion-exchange by making
the surface strongly negatively charged or by removing the oxidation products. The
sulfide ion is a strong reductant, and at a certain concentration it can lower the potential
below the value required for the formation of lead xanthate (Guy and Trahar, 1985).
In recent years, considerable attention has been given to the direct correlation be-
tween sulfide mineral flotation and the pulp potential (Trahar, 1984; Labonté and Finch,
1990; Rao et al., 1991; Ralston, 1991). Rand and Woods (1984) have demonstrated
that under certain conditions measurements of the pulp potential with smooth noble
metal electrodes approach the value of the potential at the sulfide mineral=solution
interface. This methodology provides a useful tool for investigating changes that oc-
cur at the mineral surface upon the addition of electrochemically active reagents to
mineral suspensions. Smooth Pt electrodes have been found to respond to changes in
the hydrosulfide (Soto and Laskowski, 1973; Jones and Woodcock, 1978, 1979; Zhou
and Chander, 1991) and xanthate concentration in solution (Rand and Woods, 1984),
and to be sensitive to the concentration of oxygen in solution (Gebhardt and Shedd,
1988). The implications of pulp potential in sulfide mineral flotation have been recently
reviewed (Ralston, 1991; Cheng and Iwasaki, 1992). Sulfidized oxide-type minerals of
the base metals also exhibit a flotation behavior dependent on the pulp potential (Soto
and Laskowski, 1973; Castro et al., 1974; Jones and Woodcock, 1979; López Valdivieso
and Fuerstenau, 1988).
The aim of this publication is to elucidate the flotation behavior of natural and
sulfide-treated cerussite and galena conditioned at pH 9.5 with potassium amyl xanthate
as collector in the presence of air. Since both hydrosulfide and xanthate anions are
electrochemically active, the redox potential of these mineral suspensions was measured
using a smooth-platinum=saturated calomel electrode system. Residual aqueous sulfide,
sulfide uptake by cerussite and the number of lead sulfide monolayers were also
determined. The xanthate flotation of sulfidized cerussite and galena was correlated with
the pulp potential, residual sulfide and number of sulfide monolayers.
160 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

2. Experimental

2.1. Materials

Samples of cerussite from Tsumeb, South West Africa, and galena from Missouri
were obtained from Wards Natural Science Establishment, New York. Chemical anal-
yses of these materials showed that the galena contained 84.8% Pb and 13.1% S, and
the cerussite 77.4% Pb, 4.5% C and 17.9% O. The mineral samples were ground in a
porcelain mortar to produce 65 ð 200 mesh particles. The galena samples were stored
under argon immediately after being ground to avoid surface oxidation.
All reagents used in this investigation were analytical grade chemicals. Anhydrous
sodium sulfide was used as the sulfidizing agent. Potassium nitrate was added to
maintain the ionic strength of the solution constant at 5 ð 10 3 mol dm 3 . Nitric
acid and potassium hydroxide were added for pH adjustments. The collector used
was potassium amyl xanthate (Aero-xanthate 350 obtained from American Cyanamid
Co.), which was purified by repeated dissolution–recrystallization using acetone as
the solvent and petroleum ether as the precipitating agent. Stock aqueous solutions of
sodium sulfide and amyl xanthate were prepared daily. Triply-distilled water was used
throughout this research work.

2.2. Methods

Mineral suspensions used for flotation tests were prepared at 21ºC in the presence of
air as follows. After adding the mineral sample to a 5 ð 10 3 mol dm 3 KNO3 aqueous
solution, the suspension was first agitated for one minute using a magnetic stirrer and
then dispersed in a Megeason ultrasonic cleaner for an additional minute. The sulfidizing
reagent or the collector were added before adjusting the pH to 9.5, if necessary, and
then the suspension was conditioned for 15 min. Unless otherwise specified, when both
sodium sulfide and potassium amyl xanthate were added to the mineral suspension, the
mineral was first conditioned for 10 min at pH 9.5 in the presence of the sulfide salt,
then the xanthate was added, and the suspension conditioned for another 5 min. The
final volume was kept at 100 ml. Both pH and pulp potential (the actual measurements
were made with a smooth platinum electrode relative to a saturated calomel electrode
as reference) were recorded continuously during the conditioning time for all the ex-
periments. The platinum-saturated calomel electrode system was calibrated against a
reference ferro=ferric cyanide solution. Since sulfide ions are known to poison Pt elec-
trodes causing a sluggish response and often erroneous readings (Natarajan and Iwasaki,
1970), surface products were removed from the platinum electrode by polishing.
Flotation tests were conducted at 1.0% solids by weight with a Hallimond tube using
nitrogen as the flotation gas at a flow rate of 50 ml=min. All results are presented on the
basis of one liter of solution.
The sulfide ion activity in solution was measured with an Orion Ag=Ag2 S specific
ion electrode (Model 94-16) following the method given in the manual. This method
involves raising the pH by the addition of a sulfide antioxidant buffer solution to convert
any HS and H2 S(aq) species to the S2 form and to avoid oxidation of the sulfide
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 161

ion. After treatment of cerussite with a sulfide solution for 15 min, the solids were
separated by filtration and the potential of the solution was measured. Purified argon
was bubbled through the solution during filtration and measurement of the potential. A
saturated calomel electrode was used as the reference electrode. A calibration curve was
constructed using the electrode potential measurements of standard solutions at different
sulfide concentrations and at 22.5ºC. The exact concentration of sulfide ion in these
solutions was determined from potentiometric titration data. Standard sulfide solutions
were titrated with a Fisher Scientific Aided Tritimeter against a 0.1 mol dm 3 lead
perchlorate standard solution.

3. Results and discussion

Fig. 1a presents the flotation response of cerussite and galena at pH 9.5 as a function
of added potassium amyl xanthate. Cerussite flotation sets in only after enough xanthate
is added to precipitate all aqueous lead species as lead xanthate and to react at the
cerussite surface to form lead xanthate by an ion-exchange mechanism. IR spectra
of xanthate-treated cerussite indeed indicate that the product of this surface reaction
is stoichiometric lead xanthate (Marabini et al., 1984; Popov and Vucinic, 1992b).
Zeta potential results also suggest the formation of solid lead xanthate onto cerussite
treated with amyl xanthate solutions (D.W. Fuerstenau et al., 1985). Flotation results
show that in this system lead xanthate precipitated from the bulk does not render the
mineral hydrophobic. Fig. 1a also shows the effect of conditioning without collector
on the flotation response of cerussite. With 145 minutes of conditioning at pH 9.5
and 5 minutes of conditioning in the presence of the xanthate, the onset of mineral
flotation is displaced to a higher xanthate dosage, namely 5 ð 10 4 mol dm 3 . Raising
the pH of the cerussite suspension up to 9.5 before adding the xanthate causes the
precipitation of aqueous lead species as PbCO3 . The total surface area of the system
increases significantly with the presence of these colloidal precipitates, which consume
more collector. In addition, fine cerussite particles are likely to be formed by abrasion
during prolonged conditioning with a magnetic stirrer, thus increasing further the total
surface area and the consumption of collector. Precipitation of aqueous lead as lead
carbonate before collector addition does not improve significantly the flotation response
of cerussite. Small flotation recoveries at low xanthate dosages are perhaps due to
mechanical carry-over of fine particles. Contrary to our findings, the xanthate flotation
of cerussite has been reported to be enhanced by relatively short periods of conditioning
(up to 30 min) (Popov and Vucinic, 1992b).
Galena flotation is complete at collector concentrations greater than 10 5 mol
dm 3 . These flotation results agree with the concept of strong xanthate chemisorption
onto the galena surface inducing mineral flotability (Buckley and Woods, 1994), and
correlate with electrokinetic measurements that also suggest xanthate chemisorption
onto galena (D.W. Fuerstenau et al., 1985). It is now well established that chemisorption
of xanthate onto galena involves a charge transfer mechanism whereby the anodic
reaction is the oxidation of the mineral or the collector and the cathodic reaction
is the reduction of oxygen (Woods, 1984; Woods and Richardson, 1986). In sulfide
162 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

Fig. 1. (a). The flotation recovery of cerussite and galena at pH 9.5 as a function of added potassium amyl
xanthate in 5 ð 10 3 mol dm 3 KNO3 open to the atmosphere, and (b) The E h of the system under the
same conditions.

mineral dispersions, electrochemical potentials are commonly set through the addition
of reducing or oxidizing compounds. The potential measured in these systems with an
indicator electrode coupled to a reference electrode is known as the pulp potential, and
its value lies between that of the solution potential and of the galvanic potential of the
mineral (Rao et al., 1991).
The corresponding pulp potential measurements for the xanthate flotation of cerussite
and galena are plotted in Fig. 1b as E h . In the case of cerussite, the flotation response
of this mineral is closely related to the potential sensed by a smooth platinum electrode,
which responds to changes in the xanthate concentration in solution (Rand and Woods,
1984). At low xanthate dosages, the potential remains constant because all the collector
added precipitates from solution as lead xanthate and the final composition of the
solution is independent of the xanthate added. Under these conditions, mineral flotation
is poor. Once sufficient xanthate is added to precipitate aqueous lead dissolved from
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 163

the mineral, to react at the mineral surface and to report to the solution, the potential
begins to drop and flotation sets in. Since the surface reaction between cerussite and
xanthate is not electrochemical, the potential measured must be the mixed-potential at
the Pt=solution interface, where xanthate is oxidized and oxygen is reduced. As the
xanthate added is further increased, the potential continues to drop as a consequence
of excess free xanthate in solution, and complete flotation is observed. The correlation
between redox potential and the flotation behavior of galena is not as direct as in the case
of cerussite. Even though the potential also remains constant at low xanthate dosages
and begins to drop at a certain xanthate added, the flotation of galena is complete at
all xanthate dosages investigated. The pulp potential measured in this system must arise
from the oxidation of xanthate and the reduction of oxygen at both the galena=solution
interface and the electrode=solution interface.
The flotation behavior of cerussite and galena treated with different dosages of
sodium sulfide and 5 ð 10 5 mol dm 3 amyl xanthate as the collector at pH 9.5 is
shown in Fig. 2a. Complete galena flotation was obtained with sulfide additions smaller
than about 2 ð 10 4 mol dm 3 . At this sulfide dosage, however, mineral flotation
falls drastically. Optimum galena flotation seems to indicate that the main role of the
hydrosulfide is to react with aqueous lead dissolved from the mineral and with oxidized
surface sites (Gebhardt and Kotlyar, 1991). The edge of depression coincides with the
drop in pulp potential (Fig. 2b). Lack of flotation when the sulfide added is greater than
2 ð 10 4 mol dm 3 seems to be directly related to a cathodic shift in potential below
the value required for lead xanthate formation. Also, under reducing conditions the
mineral surface is free of exchangeable oxide species and highly negatively charged as
a result of a high adsorption density of hydrosulfide ions (Herrera-Urbina et al., 1998).
A combination of these phenomena may account for the exclusion of the xanthate
from the mineral=solution interface. As opposed to the flotation behavior of galena at
low sulfide additions, the flotation of cerussite is poor under these conditions because
dissolved lead consumes the sulfidizing agent in solution. At about 10 4 mol dm 3
added sulfide, however, flotation recovery reaches a maximum of 50%, but then drops
at about 2 ð 10 4 mol dm 3 sulfide. This flotation peak appears to be primarily related
to the interaction between xanthate and precipitated PbS deposited at the surface of the
cerussite particles. Colloidal lead sulfide was identified onto the surface of cerussite
by means of scanning and transmission electron microscopy studies (Sotillo, 1985).
Precipitated PbS may oxidize and carbonate before the addition of the collector, which
is then ion-exchanged with the surface oxide and carbonate species. This mechanism
agrees with the surface oxidation–carbonation concept proposed to explain the decrease
in the negative zeta potentials of cerussite and galena that occurs at about 10 4 mol
dm 3 added sulfide (Herrera-Urbina et al., 1998). Fine, collector-coated PbS particles
deposited onto cerussite may cause the mineral to respond to flotation. As the sodium
sulfide addition is increased, more sulfide adsorbs onto precipitated PbS and its degree of
oxidation–carbonation decreases. As a result, xanthate adsorption onto the PbS particles
decreases, and flotation falls. Under these conditions the cerussite particles are still very
light colored, which indicates that they are only partially sulfidized. At higher sulfide
additions, the surface of cerussite itself is sulfidized to lead sulfide, which adsorbs the
collector and flotation ensues. Sulfide requirements for the best flotation of cerussite,
164 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

Fig. 2. (a). The flotation recovery of cerussite and galena at pH 9.5 as a function of added sodium sulfide in
5 ð 10 3 mol dm 3 KNO3 open to the atmosphere and 5 ð 10 5 mol dm 3 added potassium amyl xanthate,
and (b) The E h of the system under the same conditions.

however, are very critical because excess sulfide causes complete depression of the
mineral. The optimum sulfide dosage was about 5 ð 10 4 mol dm 3 . Further additions
of sodium sulfide bring about the depression of the mineral by the mechanism of galena
depression by sulfide. The collectorless flotation of cerussite and galena sulfidized with
different levels of added sodium sulfide (from 10 5 mol dm 3 to 5 ð 10 3 mol dm 3 ) at
pH 9.5 was also assessed. Even though the addition of sodium sulfide has been reported
to enhance the flotation of galena without collector in a certain potential range (Guy
and Trahar, 1985), neither cerussite nor galena were found to respond significantly to
flotation after sulfidization in the absence of collector. The corresponding final pulp
potential ranges were from C244 mV to 81 mV for galena and from C33 mV to
473 mV for cerussite. In the case of a lead=zinc ore concentrated by flotation without
collector, the recovery of galena is considerable at E h values greater than 100 mV
(Grano et al., 1990).
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 165

Platinum electrodes respond to changes in the hydrosulfide ion concentration in


solution (Zhou and Chander, 1991). As Fig. 2b shows, the potential sensed by a smooth
platinum electrode immersed in sulfide-treated cerussite and galena dispersions is
independent of sulfide added at low sulfide dosages, but drops at higher concentrations
of added sulfide. At low sulfide dosages, hydrosulfide ions react with aqueous lead
dissolved from the mineral and do not affect the pulp potential. Oxidation of aqueous
sulfide by dissolved oxygen also keeps unaffected the potential. Once all lead has
been precipitated as lead sulfide, residual sulfide appears in solution and causes a drop
in potential. This potential drop coincides with the fall in mineral flotation shown
in Fig. 2a. These results clearly show a close correlation between pulp potential and
xanthate flotation of both sulfidized cerussite and galena. Mineral depression occurs
when the addition of sulfide is sufficient to shift the final pulp potential cathodically
(Gebhardt and Kotlyar, 1991). The fall in mineral flotation as the E h of the pulp
decreases corroborates the electrochemical nature of xanthate chemisorption at the
sulfidized mineral surface.
Fig. 3 presents the flotation recovery of cerussite as a function of added sodium
sulfide with 5 ð 10 5 mol dm 3 and 1 ð 10 4 mol dm 3 added potassium amyl xanthate.
The binodal flotation behavior is also found for the high collector addition. With an
increase in the concentration of collector a lower concentration of sulfide is required for
the onset of flotation. Regardless of xanthate added, however, mineral depression sets in
at the same sulfide dosage. The dependence of cerussite flotation on the E h , chemically
modified by the addition of sodium sulfide, is depicted by Fig. 4. Optimum mineral
flotation occurs in the potential range from 100 mV to about 150 mV, and a sharp drop
takes place above and below these potential values. The E h range of optimum flotation
was found to be independent of the xanthate added. Under oxidizing conditions, that is
at low additions of sulfide, cerussite responds poorly to flotation because the mineral
surface is not sulfidized and the collector is consumed by lead species in solution. Below

3
Fig. 3. The flotation recovery of cerussite at pH 9.5 as a function of added sodium sulfide in 5 ð 10 mol
dm 3 KNO3 open to the atmosphere with two different additions of potassium amyl xanthate.
166 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

Fig. 4. The flotation recovery of sulfide-treated cerussite at pH 9.5 as a function of E h , chemically modified
by the addition of sodium sulfide, in 5 ð 10 3 mol dm 3 KNO3 open to the atmosphere with two different
additions of potassium amyl xanthate.

100 mV the reducing conditions inhibit xanthate adsorption onto the oversulfidized
cerussite surface, and mineral flotation ceases.
The concentration of sulfide ion remaining in solution after the conditioning of
cerussite with different dosages of sodium sulfide, and the number of sulfide monolayers
are given in Fig. 5 together with the flotation response of sulfide-treated cerussite with
5 ð 10 5 mol dm 3 added amyl xanthate. Residual sulfide was not detected in solution
at the end of conditioning when the sodium sulfide dosage is smaller than 5 ð 10 4 mol
dm 3 . Sulfide uptake by cerussite was first determined to calculate the number of sulfide
monolayers. For calculating the sulfide uptake by the mineral, the amount of sulfide

Fig. 5. The flotation recovery of cerussite at pH 9.5 with 5 ð 10 5 mol dm 3 added potassium amyl
xanthate, the residual aqueous sulfide and the number of sulfide monolayers as a function of added sodium
sulfide in 5 ð 10 3 mol dm 3 KNO3 open to the atmosphere.
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 167

required to precipitate aqueous lead dissolved from cerussite .1:4 ð 10 4 mol dm 3 in


these experiments) and the amount of sulfide remaining in solution had to be subtracted
from the sulfide added. To convert the sulfide uptake by the mineral to surface coverage
in terms of number of monolayers we considered the following. The specific surface
area of the 65 ð 200 mesh size fraction is 0.92 m2 g 1 ; 90% of the surface planes are
assumed to be (110) planes with a lead density of 3:4 ð 108 atoms m 2 and 10% (021)
planes lacking lead atoms (Hurlbut and Klein, 1977; Glembovsky and Anfimova, 1964);
each HS ion will react with a lead atom on the cerussite surface giving an average
adsorption density of 5:1 ð 10 6 mol m 2 for the monolayer. As Fig. 5 shows, the
concentration of sulfide ion in solution becomes significant (reaches a value of 5 ð 10 5
mol dm 3 ) at about 6 ð 10 4 mol dm 3 added sulfide. It is at this same sulfide dosage
that mineral flotation ceases. These results clearly show that the depression of sulfidized
cerussite is closely related to the final concentration of aqueous sulfide, and occurs once
residual sulfide is detected in solution.
Fig. 5 also shows that the number of sulfide monolayers onto cerussite begins to build
up above about 10 4 mol dm 3 added sulfide. That is, only after all aqueous dissolved
lead precipitates as lead sulfide from solution. At 1:5 ð 10 4 mol dm 3 added sulfide,
the uptake of sulfide by the mineral is only 10 5 mol dm 3 , the adsorption density
is equivalent to 0.21 monolayers, and the flotation of sulfidized cerussite sets in. The
number of sulfide monolayers increases as the sodium sulfide added is further increased
and reaches a value of nearly 8 monolayers when the flotation response is maximum. At
this sulfide film thickness the mineral surface is still able to take up more sulfide but at a
lower rate so that aqueous sulfide is now at the outer surface. Hence, depression begins
due to the negatively charged surface, lack of oxygen-containing surface products, and
reducing potentials, all of which inhibit collector adsorption. Oversulfidized cerussite,
with nearly 15 sulfide monolayers, does not respond to flotation with amyl xanthate.
Marabini et al. (1984) have found that the thickness of the lead sulfide film onto
sulfidized cerussite is of the order of ten monolayers.
It is now possible to identify three distinct steps that occur during a sulfidization
process, as illustrated schematically in Fig. 6. Initially, all sulfide added is consumed by
aqueous cations dissolved form the mineral, and precipitated as metal sulfide (Stage 1).
Upon adding more sulfide, the mineral surface is then converted to metal sulfide and the
sulfidization process sets in (Stage 2). At the beginning of this surface transformation
a porous metal sulfide, which does not present too much resistance to the hydrosulfide
and mineral anion transport, must be formed at the surface. This porous structure may
promote a high initial reaction rate enabling further growth of the sulfidized layer to take
place. As sulfidization proceeds, the sulfidized outer layers attain a uniform coating,
sulfide uptake by the mineral slows down, and sulfide starts reporting to the solution
at the end of the conditioning period (Stage 3). The distribution of sulfide species
as a function of sulfide added in a sulfidization system is schematically given by the
diagram presented as Fig. 6, which also shows the three sulfidization steps. A direct
correlation exists between these steps in sulfidization and cerussite flotation. Although
the precipitation of PbS (Stage 1) seems to induce some mineral flotability, it is not until
the mineral itself is sulfidized (Stage 2) that complete flotation is attained. Once the
concentration of aqueous sulfide reaches significant values (Stage 3), flotation ceases.
168 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

Fig. 6. Schematic representation of the distribution of sulfide species in a sulfidization system, and its
relation to mineral flotation, as a function of added sodium sulfide.

4. Conclusions

The xanthate flotation of sulfide-treated cerussite and galena has been assessed at
different levels of added sodium sulfide, and the E h of the pulp determined from
the redox potential sensed by a smooth platinum electrode immersed in the mineral
suspension. Galena floats well with amyl xanthate as the collector at sodium sulfide
additions lower than 10 4 mol dm 3 . Cerussite responds best to amyl xanthate flotation
after enough sodium sulfide is added to precipitate aqueous lead dissolved from the
mineral as lead sulfide and to sulfidize the mineral surface. Oversulfidized cerussite,
however, does not float with amyl xanthate. The flotation response of sulfide-treated
cerussite and galena falls sharply at the onset of pulp potential drop. Measurements
of sulfide ion activity in solution show that mineral depression and the drop in pulp
potential occur when the concentration of aqueous residual sulfide becomes significant.
The results of this investigation have clearly demonstrated that there is a maximum
sulfide dosage that limits the xanthate adsorption onto sulfidized cerussite and galena,
and therefore the flotation of these minerals. Since mineral depression seems to be
directly related to excess sulfide reporting to the solution and, consequently, to a
significant decrease of the pulp potential, the concentration of mixed sulfide–oxide lead
ores by sulfidization–xanthate flotation may be controlled by monitoring the potential
sensed by a smooth platinum electrode immersed in the pulp.

References
Allison, S.A., Goold, L.A., Nicol, M.J., Granville, A., 1972. A determination of the products of reaction
between various sulphide minerals and aqueous xanthate solutions, and a correlation of the products with
electrode rest potentials. Metall. Trans. 3, 2613–2618.
Buckley, A.N., Woods, R., 1994. Xanthate chemisorption on lead sulfide. Colloids Surf. 89, 71–76.
Castro, S., Soto, H., Goldfarb, J., Laskowski, J., 1974. Sulphidizing reactions in the flotation of oxidized
R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170 169

copper minerals, II. Role of the adsorption and oxidation of sodium sulphide in the flotation of chrysocolla
and malachite. Int. J. Miner. Process. 1, 151–161.
Cheng, X., Iwasaki, I., 1992. Pulp potentials and its implications to sulfide flotation. Miner. Process. Ext.
Met. Rev. 11, 187–210.
De Donato, P., Cases, J.M., Kongolo, M., Michot, L., Burneau, A., 1990. Infrared investigation of
amylxanthate adsorption on galena: influence of oxidation, pH and grinding. Colloids Surf. 44, 207–228.
Eadington, P., Prosser, A.P., 1969. Trans. IMM, Sec. C 78, 74–82.
Fleming, M.G., 1952. Trans. AIME 193, 231–236.
Fuerstenau, D.W., Sotillo, F., Valdivieso, A., 1985. Sulfidization and flotation behavior of anglesite, cerussite
and galena. In: Proceedings, XV Int. Miner. Process. Congress, Cannes, France, pp. 74–86.
Fuerstenau, M.C., Miller, J.D., Kuhn, M.C., 1985. Chemistry of Flotation. Society of Mining
Engineers=AIME, Littleton, CO, 177 pp.
Fuerstenau, M.C., Olivas, S.A., Herrera-Urbina, R., Han, K.H., 1987. The surface characteristics and
flotation behavior of anglesite and cerussite. Int. J. Miner. Process. 20, 73–85.
Gaudin, A.M., 1957. Flotation. McGraw-Hill, New York, 573 pp.
Gebhardt, J.E., Kotlyar, D.G., 1991. Hydrosulphide depression of copper-sulphide minerals floated by xan-
thate and thionocarbamate collectors. In: Dobby, G.S., Argyropoulos, S.A., Rao, S.R. (Eds.), Proceedings
of the Copper 91-Cobre 91 Int. Symposium, vol. II. Pergamon, Oxford, pp. 201–215.
Gebhardt, J.E., Shedd, K.B., 1988. Effect of solution composition on redox potentials of platinum and
sulfide mineral electrodes. In: Richardson, P.E., Woods, R. (Eds.), Proceedings, Electrochemistry in
Mineral and Metal Processing II. The Electrochemcial Society, pp. 84–100.
Glembovsky, V.A., Anfimova, E.A., 1964. Specific crystallochemical and structural features of oxidized
minerals of lead and their influence on the choice of reagents for the flotation of these minerals. In:
Arbiter, N. (Ed.), VII Int. Miner. Process. Congress, AIME, vol. 1, pp. 329–336.
Grano, S., Ralston, J., Smart, R.S.C., 1990. Influence of electrochemical environment on the flotation
behaviour of Mt. Isa copper and lead–zinc ore. Int. J. Miner. Process. 30, 69–97.
Guy, P.J., Trahar, W.J., 1985. The effects of oxidation and mineral interaction on sulphide flotation. In:
Forssberg, K.S.E. (Ed.), Flotation of Sulphide Minerals. Elsevier, Amsterdam, pp. 91–109.
Herrera-Urbina, R., Sotillo, F.J., Fuerstenau, D.W., 1998. Amyl xanthate uptake by natural and sulfide-
treated cerussite and galena. Int. J. Miner. Process. 55, 113–128.
Hurlbut Jr., C.S., Klein, C., 1977. Dana James Dwight Manual of Mineralogy, 19th ed. Wiley, New York,
532 pp.
Jones, M.H., Woodcock, J.T., 1978. Evaluation of ion-selective electrode for control of sodium sulphide
additions during laboratory flotation of oxidized ores. Trans. IMM 87, C99–C105.
Jones, M.H., Woodcock, J.T., 1979. Control of laboratory sulphidization with a sulphide ion-selective
electrode before flotation of oxidized lead–zinc–silver dump material. Int. J. Miner. Process. 6, 17–30.
Labonté, G., Finch, J.A., 1990. Behavior of redox electrodes during flotation and relationship to mineral
floatabilities. Miner. Metall. Process. 7 (2), 106–109.
López Valdivieso, A., Fuerstenau, D.W., 1988. Control de la sulfidizacion y flotación de minerales oxidados
de plomo por mediciones de potencial de oxidación. Geomimet 152, 64–70.
Marabini, A.A., Alesse, V., Garbassi, F., 1984. Role of sodium sulphide, xanthate and amine in flotation of
lead–zinc oxidized ores. In: Jones, M.J., Oblatt, R. (Eds.), Reagents in the Mineral Industry. The IMM,
London, pp. 125–136.
Natarajan, K.A., Iwasaki, I., 1970. Behavior of platinum electrodes as redox potential indicators in some
systems of metallurgical interest. Trans. AIME 247, 317–324.
Page, P.W., Hazell, L.B., 1989. X-ray photoelectron spectroscopy (XPS) studies of potassium amyl xanthate
(KAX) adsorption on precipitated PbS related to galena flotation. Int. J. Miner. Process. 25, 87–100.
Popov, S.R., Vucinic, D.R., 1992a. Adsorption characteristics and floatability of cerussite with ethylxanthate
in the presence of dissolved lead ion. Int. J. Miner. Process. 34, 307–319.
Popov, S.R., Vucinic, D.R., 1992b. The effect of prolonged agitation in lead ion solution on ethylxanthate
adsorption and surface characteristics of cerussite. Int. J. Miner. Process. 35, 85–100.
Pugh, R.J., Bergstrom, L., 1986. Surface and solution chemistry studies on galena suspensions. Colloids
Surf. 19, 1–20.
Ralston, J., 1991. Eh and its consequences in sulphide mineral flotation. Miner. Eng. 4 (7891011), 859–878.
170 R. Herrera-Urbina et al. / Int. J. Miner. Process. 55 (1999) 157–170

Rand, D.A.J., Woods, R., 1984. Eh measurements in sulphide mineral slurries. Int. J. Miner. Process. 13,
29–42.
Rao, S.R., Labonte, G., Finch, J.A., 1991. Electrochemistry in the plant. In: Dobby, G.S., (Organizer),
Froth Flotation Systems. Professional Enhancement Short Course. 30th Annual CIM Conference of
Metallurgists, Ottawa, Canada, 43 pp.
Sotillo, F.J., 1985. The sulfidization and flotation of cerussite and galena. M.S. Thesis. University of
California at Berkeley, 74 pp.
Sotillo, F.J., Fuerstenau, D.W., 1988. The sulfidization and flotation of cerussite and galena. In: Castro
Flores, S.H., Alvarez Moisan, J., (Eds.), Froth Flotation. Elsevier, Amsterdam, pp. 271–288.
Soto, H., Laskowski, J., 1973. Redox conditions in the flotation of malachite with sulphidizing agent. Trans.
IMM 82, C153–C157.
Toperi, D., Tolun, R., 1969. Electrochemical study and thermodynamic equilibria of the galena–oxygen–
xanthate flotation system. Trans. IMM 78, C191–C197.
Trahar, W.J., 1984. The influence of pulp potential in sulphide flotation. In: Jones, M.H., Woodcock, J.T.
(Eds.), Principles of Mineral Flotation. The Wark Symposium. The Australasian Inst. of Mining and
Metallurgy, Parkville, Australia, pp. 117–136.
Woods, R., 1984. Electrochemistry of sulfide flotation. In: Jones, M.H., Woodcock, J.T. (Eds.), Principles
of Mineral Flotation. The Wark Symposium. The Australasian Inst. of Mining and Metallurgy, Parkville,
Australia, pp. 91–115.
Woods, R., Richardson, P.E., 1986. The flotation of sulfide minerals — electrochemistry aspects. In:
Somasundaran, P. (Ed.), Advances in Mineral Processing. Society of Mining Engineers, Littleton, CO, pp.
154–170.
Zhou, R., Chander, S., 1991. Comparison of gold, platinum and sulfide ion selective electrodes as sensors
for Eh measurements in sulfide solutions. Miner. Metall. Process. 8 (2), 91–96.

Você também pode gostar