You are on page 1of 16

International Journal o f Mineral Processing, 6 (1979) 1--16 1

© Elsevier Scientific Publishing Company, Amsterdam -- Printed in The Netherlands

AN ELECTROCHEMICAL INVESTIGATION OF THE NATURAL FLOT-


ABILITY OF CHALCOPYRITE

J.R. GARDNER and R. WOODS


CSIRO Division of Mineral Chemistry, Port Melbourne, Vic. 3207 (Australia)
(Received July 27, 1978; revised and accepted December 12, 1978)

ABSTRACT

Gardner, J.R. and Woods, R., 1979. An electrochemical investigation of the natural
flotability of chalcopyrite. Int. J. Miner. Process., 6: 1--16.

The flotation of chalcopyrite particles was investigated using a modified Hallimond


tube cell, in which the potential at the mineral-- solution interface was controlled poten-
tiostatically. Previous reports that this mineral displays natural flotability at oxidizing
potentials and non-flotability at reducing potentials have been confirmed. Anodic oxida-
tion of the mineral surface is responsible for the change from a hydrophilic to a hydro-
phobic condition. Linear potential sweep voltammetry has identified the products of the
anodic reaction as CuS, Fe(OH)3 and S. The presence of sulphur on the mineral surface
is considered to be the critical factor in rendering chalcopyrite flotable.

INTRODUCTION

A l t h o u g h the generally a c c e p t e d view has b e e n t h a t sulphide minerals,


including c h a l c o p y r i t e , are h y d r o p h i l i c ( S u t h e r l a n d a n d Wark, 1955), t h e r e
have b e e n r e p o r t s in t h e l i t e r a t u r e (e.g. L e p e t i c , 1 9 7 4 ; F i n k e l s t e i n et al.,
1975) t h a t c h a l c o p y r i t e c o u l d f l o a t in the a b s e n c e o f collectors in certain
c i r c u m s t a n c e s . In a r e c e n t p u b l i c a t i o n , H e y e s and T r a h a r ( 1 9 7 7 ) p r e s e n t e d
investigations w h i c h e l u c i d a t e d t h e n a t u r a l f l o t a t i o n characteristics o f chal-
c o p y r i t e and e x p l a i n e d the a p p a r e n t c o n t r a d i c t i o n s b e t w e e n p r e v i o u s find-
ings.
H e y e s and T r a h a r d e m o n s t r a t e d t h a t c h a l c o p y r i t e w o u l d f l o a t in the
absence of collector under some experimental conditions but not under
others, the f l o t a t i o n r e s p o n s e o f the m i n e r a l being d e p e n d e n t on the con-
s t i t u t i o n o f the grinding m e d i u m , the gas used f o r f l o t a t i o n and t h e p r e s e n c e
o f r e d u c i n g or oxidizing agents in the f l o t a t i o n p u l p . T h e y w e r e able to
c o r r e l a t e f l o t a t i o n e f f i c i e n c y w i t h the p o t e n t i a l o f the p u l p ( t h e Eh) in the
various c h e m i c a l e n v i r o n m e n t s investigated. C h a l c o p y r i t e was f o u n d to
f l o a t n a t u r a l l y in t h e m o r e oxidizing e n v i r o n m e n t s , i.e. at high E h , b u t n o t
t o f l o a t at l o w Eh.
T h e e l e c t r o c h e m i c a l investigations p r e s e n t e d h e r e w e r e u n d e r t a k e n t o
verify the conclusion that the critical factor in the natural flotability of
chalcopyrite is the value of the potential of the mineral particles and not
the chemical system used to control that potential. Investigations to identify
the surface reactions on chalcopyrite are also discussed.

EXPERIMENTAL

Flotation

A Hallimond tube flotation cell was adapted to enable the potential of


the mineral particles to be controlled electrochemically. The cell, incor-
porating the necessary electrodes, is shown in Fig. 1. The working electrode

N~ A

9as
Fig. 1. H a l l i m o n d t u b e f l o t a t i o n cell a d a p t e d f o r e l e c t r o c h e m i c a l c o n t r o l o f t h e mineral
p o t e n t i a l . A = p l a t i n u m grid w o r k i n g e l e c t r o d e ; B = r e f e r e n c e e l e c t r o d e ; C = c o u n t e r
electrode; D = collection tube.

A is a grid of platinum wire formed into a cylinder. Its potential is measured


against a saturated calomel reference electrode B {S.C.E.), which is connect-
ed to the flotation compartment through a Luggin probe capillary. Under
potentiostatic control, the current is passed between the working electrode
and a platinum counter electrode housed in compartment C, separated from
the main cell by a sintered glass disc. The potential was controlled with a
68 TS1 Wenking potentiostat programmed with a sweep generator construct-
ed in these laboratories. All potentials reported in this paper have been
converted to the standard hydrogen electrode (S.H.E.) scale, assuming the
S.C.E. has a potential of 0.245 V (Bates, 1964) on this scale.
The principle on which the cell operates is that the slurry of mineral
particles in the cell will take up the potential of the working electrode
through c o n t a c t with the platinum grid during agitation (Gardner and Woods,
1973; Chander and Fuerstenau, 1975). More rapid potential control of the
mineral particles was achieved with a platinum foil working electrode, but
its presence was f o u n d to interfere with the fluid dynamics necessary for
efficient flotation. Care was taken in designing the cell so that mineral
particles would n o t accumulate in the side-arms.
The chalcopyrite sample was the same as that used by Heyes and Thahar
(1977) and originated at Mt. Lyell, Tasmania. The mineral specimens were
ground by hand in a glass m o r t a r and pestle under distilled water in a ni-
trogen atmosphere in a glove bag. The mineral was ground until it passed
through a stainless-steel basket constructed from Tyler 48 mesh (0.295
mm opening). Fines were removed by three decantations from a beaker,
and a glass spoon used to deliver a 1 g (-+ 0.1 g) sample to a deaerated pH
11 buffer solution (Table I) in the " e l e c t r o h a l l i m o n d " tube. The liquid
level in the cell was kept low at this stage so that no floated material would
carry over into the collection tube.

TABLE I

Composition of buffer solutions employed

4.6 0.5 M CH3COOH + 0.5 M CH3COONa


6.82 0.194MH,B40~ +0.0015MNa2B407 +0.049MNa2SO 4
6.97 0.025 M KH:PO4+ 0.025 M Na:HPO 4
7.63 0.17 M H:B407 + 0.0075 M Na2B407 + 0.047 M Na2SO4
9.2 0.05 M Na2B40T
11.0 0.025 M NaHCO3 + 0.023 M NaOH
13.0 0.1 M NaOH

The sample was conditioned at the chosen potential for 30 min with
nitrogen passing through the cell and the Teflon-coated magnetic stirrer in
operation. The mineral would retain the set potential on open circuit when
it was conditioned for this length of time. After conditioning, the gas flow
was stopped but stirring was continued, and this resulted in any floating
mineral returning to the base of the cell. The cell was then filled with buffer
solution and three flotation cuts made by passing nitrogen gas at 60 cm 3/min
for 5 min, then further periods of 10 and 15 min. The floats, collected in
tube D (Fig. 1), and the tails were separated from the solution on sintered-
glass crucibles, dried and weighed to determine flotation recovery.
Flotation was also carried out in the cell using additions of a reductant,
sodium dithionite, or an oxidant, sodium hypochlorite, to control the po-
tential o f the mineral, the potential being m o n i t o r e d through the platinum
grid. Runs were also pe r f or m e d on tails from experiments perform ed at
potentials where little flotation occurred, the potential being adjusted for
the second run w i t h o u t removing the mineral particles from the cell.

Voltammetry

Linear potential sweep voltammetry was performed on chalcopyrite


electrodes using the contact-angle cell described previously (Gardner and
Woods, 1974). This is a three-compartment electrochemical cell fitted with
an optical window for contact-angle determination.
Voltammograms were recorded for chalcopyrite electrodes in a range of
buffer solutions. The compositions of these buffers are given in Table I.
The electrode was prepared from a massive specimen of chalcopyrite set
in mounting epoxy as described previously for galena and pyrite (Gardner
and Woods, 1977). A new surface was produced before each run by wet
grinding on 600-paper and the electrode immediately transferred to the cell.
For one experiment, the electrode surface was polished to the 1 pm diamond
stage. The voltammograms were recorded at 4 mV s TM in quiescent solutions
and in solutions stirred by the passage of nitrogen gas. In the figures pre-
sented, anodic currents are positive and cathodic currents negative.

Contact angles

The contact angle was measured as a function of electrode potential


using the technique described previously (Gardner and Woods, 1974, 1977).
Low angles are difficult to determine accurately by photographing the bub-
ble profile. In order to ascertain whether such angles were zero or had a
small but finite value, the bubble was slowly withdrawn from the mineral
surface. A zero angle is characterized by a lack of coalescence between the
bubble and the surface, "stickiness" denotes a finite angle.

RESULTS AND DISCUSSION

Flotation

The flotation response of chalcopyrite in pH 11 buffer using nitrogen gas


is shown in Fig. 2. Recovery is plotted as a function of time for different
conditioning potentials. The flotation recovery after 30 min is presented as
a function of potential in Fig. 3. This figure also contains the results of
Heyes and Trahar (1977) for flotation in a modified Denver cell under
chemical control of the redox potential. The two sets of results show similar
behaviour, flotation occurring when the potential is above about --0.1 V, but
not if it is below this value. However, the performance of the potentiosta-
tically controlled runs is lower. Lower performance was also obtained
using chemical oxidation and reduction in the "electrohallimond" tube. With
dithionite additions, the open circuit potential was stable during flotation;
100

80 Set potential
(V vs SHE)

t0 ~+0"~
*0'4

a 4o
fX_

Io ZO 30 4~
Flotahon hme (rnin)
Fig. 2. R a t e o f f l o t a t i o n of c h a l c o p y r i t e c o n d i t i o n e d at d i f f e r e n t p o t e n t i a l s .

100

f
/
80- /

v g0- I
I
40--
I
/ /
20 -

o × dJ-.-~.-..-~ l I
-0"4 0 04 0$
Potential (V vs. SHE)
Fig. 3. R e c o v e r y o f c h a l c o p y r i t e a f t e r 30 rain f l o t a t i o n as a f u n c t i o n o f c o n d i t i o n i n g
p o t e n t i a l . O = p o t e n t i o s t a t i c c o n t r o l ; X = c h e m i c a l c o n t r o l . Dashed curve is t h e relation-
ship f o u n d b y Heyes a n d T r a h a r ( 1 9 7 7 ) for f l o t a t i o n in a m o d i f i e d D e n v e r cell.

the results obtained in the presence of this reagent are shown as the crosses
in Fig. 3. With h y p o c h l o r i t e the potential was n o t constant, but the results
obtained were o f the same order as those observed for potentiostatic control
in the same potential region. Therefore, the difference between the two
sets o f results in Fig. 3 is due to the different fluid dynamics established in
the Hallimond tube and Denver cells.
Heyes and Trahar (1977) found that chalcopyrite could be made to float
and sink alternately by cycling the pulp between oxidizing and reducing
conditions. Similar reversibility is observed with potentiostatic control.
These results confirm the conclusion that the value of the electrode
potential of chalcopyrite determines whether or not the mineral will float.

Identification of surface reactions

The chalcopyrite--aqueous solution interface was investigated using linear


potential sweep voltammetry in order to identify the processes which in-
duce flotation, i.e. cause a change in the mineral surface from a hydrophilic
to a hydrophobic condition.
Voltammograms for a freshly ground chalcopyrite surface in oxygen-free
solution of pH 11 are shown in Fig. 4. The potential sweeps were com-

L __ I L __ L i
-0~ -0'4 -0'2 0 0'Z
Potential (Y vS SHE)
Fig. 4. V o l t a m m o g r a m s for c h a l c o p y r i t e in p H 11 b u f f e r s o l u t i o n . L i n e a r p o t e n t i a l
sweeps at 4 m V s TM are t a k e n f r o m t h e rest p o t e n t i a l t o o n e limit f o l l o w e d b y 11/2 cycles.
Initial scan positive-going in A a n d negative-going in B.

menced from the rest potential of the mineral in this solution, in initial
scan being taken to more positive potentials in Fig. 4A and to more negative
values in Fig. 4B. It can be seen that an anodic current flows at potentials
above about --0.1 V and that this gives rise to a peak at about 0.15 V. Chal-
copyrite becomes naturally flotable at about the same potential as the com-
mencement of the anodic process (cf. Figs. 3 and 4). This correspondence
leads to the conclusion that the products of the electrochemical reaction
are responsible for the change in flotation response.
There is a cathodic peak on the negative-going scan in Fig. 4A which one
can assume is due to the reduction of the products of the anodic process.
This conclusion is confirmed by triangular potential sweeps of varying
amplitude, commencing from either of the two potential limits of Fig. 4.
From -0.65 V, a cathodic peak is observed on the return scan only when
the sweep is extended into the region where anodic currents flow, the charge
associated with the peak increasing as the positive potential limit is in-
creased. The analogous situation is observed on sweeps taken from 0.3 V.
The anodic peak on the second positive-going scan in Fig. 4A is larger than
that on the initial scan from the rest potential. Some oxidation of the
mineral surface occurring during the grinding procedure and transfer to the
electrochemical cell could account for this behaviour. If this explanation
is correct, then a cathodic peak corresponding to the reduction of the pro-
ducts of air-ox:Ldation should be observed when the initial scan is taken to
negative potentials. This is the case, for a cathodic peak is apparent in Fig.
4B. The magnitude of the charge associated with this peak is 1.6 mC c m - 2 ,
which is a b o u t 1/3 that after taking the potential to 0.3 V. The difference
between the anodic charges in Fig. 4A, which should also be a measure of
the extent of air-oxidation, is approximately the same value.
Exposure of chalcopyrite to oxygen-saturated solution prior to insertion
into the cell results in an increase in the charge associated with the cathodic
peak on the initial negative-going scan taken from the rest potential. The
magnitude of this peak becomes constant after about 1 hour exposure, at
which time it is approximately equal to that on the second negative-going
scan in Fig. 4B. This observation supports the conclusion that the products
of air-oxidation of chalcopyrite are identical to those on a potential sweep.
The anodic and cathodic charges are equal for a complete cycle between
- 0 . 6 5 and 0.3 V (Fig. 4), and the shape of the voltammogram does not
change greatly between consecutive triangular sweeps. This indicates that
the oxidized surface is reduced back to chalcopyrite by the time the sweep
returns to --0.65 V.
The reversibility of the system would suggest that only surface species
are formed under the conditions of Fig. 4. It has been established (Wads-
worth, 1972; Linge, 1976) that iron is dissolved preferentially from the
surface layers of chalcopyrite in acid solutions during oxidative leaching.
The analogous process in alkaline solution would result in the formation of
a hydrated iron oxide or hydroxide. It has been demonstrated by X-ray
emission spectroscopy (Michell and Woods, unpublished work) and Auger
electron spectroscopy (Eadington, 1977) that oxygen is incorporated into
the surface of chalcopyrite when the mineral is exposed to oxygen-contain-
ing solutions. In the light of these considerations, we suggest that the oxida-
tion reaction in alkaline solution can be represented by:

CuFeS2 + 3H20-~ CuS + Fe(OH)3 + S + 3H+ + 3e (i)


The standard potential for this reaction is 0.547 V [calculated from the
free-energy data of Latimer (1952) and the value for chalcopyrite of --45.5
kcal gives by Young {1967)]. In acid solutions, the reaction is expect ed to
be:

CuFeS2 -* CuS + Fe 2. + S + 2e (E ° = 0.293 V) (2)

It should be not e d that t h e r m o d y n a m i c considerations favour the forma-


tion o f sulphate rather than sulphur. However, oxidation of sulphides to
sulphate generally exhibits a considerable degree of irreversibility {Peters,
1977) and sulphur can exist as a metastable phase.
In order to confirm that reactions (1) and (2) account for the electro-
chemical behaviour, voltammograms were recorded in quiescent and stirred
solutions of various pH values. This approach, which was e m p l o y e d by Ri-
chardson and Maust (1976) to examine galena, allows the pH dependence of
the reactions taking place to be determined and a distinction to be made
between reactions which produce surface species and those which yield
soluble products. The results for chalcopyrite are shown in Fig. 5. If reac-
P01enhal (V vs. SHE)
-0~ -0'4 -0'2 0 0'2 0"4
r r i l i i i I i ~ i

pH 4 . 0

" "'._./ pH$'BZ

50~A crn-z[ ~ pH$gr


pH?.03

pHSZ

, pH II

pH f3

Fig. 5. VoitammoBTams for chalcopyrite in a range of buffer solutions. Potential stepped


to low potential limit before recording a complete potential cycle at 4 mV s-'. - -
quiescent solution; .............. stirred solution. Arrows indicate the calculated reversible
potentials for reaction (2) with 10 -1 M Fe 2÷ (pH 4.6 and 6.82) and for reaction (1)
(pH 6.97 to 13).

tions (1) and (2) occurred reversibly, the anodic wave would c o m m e n c e
at the relevant equilibrium potential, although it is possible that currents
would flow below this value because the initial products of reaction may
n o t exhibit the t h e r m o d y n a m i c properties of the bulk phases. The theoretic-
al reversible potentials for reaction (1), and for reaction (2) with 1 0 -6 M
Fe 2÷, are presented on Fig. 5. It can be seen that the anodic wave appears
at potentials close to those expected for the two processes.
The presence of iron oxide on pyrite was identified (Janetski et al., 1977;
Michell and Woods, 1978) from voltammetric peaks which result from
oxidation and reduction between iron (II) and iron (III) valence states. This
is not observed for chalcopyrite in alkaline solution, because the iron (III)
hydroxide reacts to form sulphide by the reverse of reaction (1) at poten-
tials above those at which it would be reduced to an iron (II) hydroxide.
All the oxidized iron must reform the sulphide, since if any iron (II) hydrox-
ide were formed on the cathodic scan its oxidation to the iron (III) state
would give rise to a peak on a subsequent anodic sweep:

Fe(OH)2 + H20 -~ Fe(OH)3 + H ÷ + e (E ° = 0.271 V) (3)

Since this is not observed, no sulphur can be lost to the solution by anodic
oxidation to sulphur--oxygen ions under the present experimental condi-
tions.
In near-neutral solutions, it should be possible to detect the reduction of
Fe(OH)3 to Fe 2. because this reaction has a much greater pH dependence
than reaction (1) and will occur at more anodic potentials (Fig. 6). In fact,
two cathodic peaks are observed on the voltammograms for solutions of pH
6.82, 6.97 and 7.63. The first peak appears at the same potential as that
for iron oxide deposited on a gold electrode (Janetski et al., 1977) in each
solution. The cathodic curves at these pH values are stirring-dependent {Fig.
5), as is to be expected if the first peak results in the formation of ferrous
ions. In quiescent solution, the Fe 2÷ will remain in the vicinity of the mineral
surface and react with CuS and S in the reverse of reaction (2). In stirred
solution, the Fe 2÷ will be dispersed into the bulk solution and, as the po-
tential is decreased, sulphur will be reduced to sulphide ions at a potential
dependent on pH (Fig. 6):

S + 2H ÷ + 2e-~ H2S (E ° = 0.141 V) (4)

This reaction gives rise to the second cathodic peak. If stirring is terminated
after the iron has been dispersed, the sulphide ions subsequently produced
by reaction (4) will remain at the electrode surface. Their presence can be
detected by reversing the potential sweep. An additional anodic peak is then
observed (Fig. 7) due to the oxidation process which is the reverse of re-
action (4).
In proposing that the mechanism can be represented by reactions (1) and
(2), we have assumed that the iron in the surface zone is completely oxi-
dized before any copper-oxygen species are formed, i.e. the product of the
initial reaction is best represented by CuS rather than a copper--iron sul-
phide. This assumption is supported by potential sweeps (Fig. 8) in which
the positive potential limit is greater than for the voltammograms presented
I0

0-4

02

O0

Eb

-0.2

-0,4

0 2 4 G S fO 12 14
pH
Fig. 6. R e v e r s i b l e p o t e n t i a l s f o r t h e F e 2 + / F e ( O H ) 2 / F e ( O H ) 3 ( . . . . . ), S/H2 S / H S - l
( - - ) a n d C u F e S 2 / C u S , F e ( O H ) 3 , S (. . . . . . . ) s y s t e m s as a f u n c t i o n o f pH.

i i i ~ r i

40 i/x',
I \
I \\
2O

-zo

~ -40

[ I I [ I I I
-0"3 -O'Z -0'1 0 OI 02 0'3 0'4
I°otentLaL (V vs. ,SHE)

Fig, 7. V o l t a m m o g r a m s f o r c h a l c o p y r i t e in p H 6 . 8 2 b u f f e r s o l u t i o n . P o t e n t i a l s w e e p s
c o m m e n c e d at t h e l o w p o t e n t i a l l i m i t a n d r u n f o r 11/2 c y c l e s at 4 m V s -1 . -....... s t i r r e d ,
- - unstirred.
11

I)H ~;g7

g 5o

~ -50
[ I I_ I
-02 0 02 0'4
Potential (V vs ,SHE)

Fig. 8. Voltammograms for chalcopyrite in pH 6.97 buffer solution. Linear potential


sweep at 4 mV s-' commenced at the rest potential and reversed at 0.47 V.

in previous figures. There is an increase in current at potentials above the


region where the anodic peak occurs and an additional cathodic peak appears
at the start of the negative-going scan. The potentials at which these reac-
tions take place are consistent with the reactions:

CuS-~ Cu 2÷ + S + 2e (E ° = 0.590 V) (5)


CuS + H20 -+ Cu(OH): + S + 2H÷ + 2e (E° = 0.863 V) (6)

The pH determines whether (5) or (6) is appropriate for a particular solu-


tion.
The relative charges passed during the various processes occurring during
the potential sweeps at neutral pH provide further evidence in support of
the proposed mechanism. For example, the anodic charge on the voltammo-
gram (Fig. 5) for a stirred solution of pH 6.97 was 2.51 mC cm -2 , and the
charge associated with the first cathodic peak (integrated to --0.14 V) was
0.71 mC cm -2 . Since the former corresponds to a 3e process [reaction (1)]
and the latter to a l e [reaction (3}], the ratio of these charges is expected
to be 3. The experimental ratio of 3.5 is close to this value. The fact that it
is somewhat greater could be due to some Fe 2~ being formed on the anodic
scan by reaction (2). Integration of the sulphur reduction peak is difficult,
because the current does not decrease to zero when the sweep is extended
to more negative potentials because the associated CuS begins to be reduced:

2CuS + 2H+ + 2e -~ Cu2S + H2S (E° = 0.081 V) (7)

An estimate of the charge associated with the reduction of sulphur was made
by assuming that the peak for this process alone would be symmetrical.
Thus integration from --0.14 V to the peak potential of -0.3 V should re-
present half the total charge. This procedure have a value of 1.80 mC cm -2
for the sulphur reduction charge. The reaction involves 2e and hence the
12

charge is expected to be 2/3 that for the anodic process. The experimental
ratio of 0.72 is close to this value, and the difference could also be explained
b y reaction (2) accounting for a fraction of the anodic charge.
At pH 4.6, reductive dissolution of iron (III) species is n o t evident (Fig.
5) because at this pH the oxidation reaction proceeds only to Fe 2÷. In stirred
solution, the ferrous ions produced anodically will be dispersed and hence
reaction (2) cannot be reversed. The cathodic charge on the return sweep
must then be due to reactions (4) and (7) in which the other oxidation
products are reduced. If both these processes proceed to completion on the
negative-going scan, the cathodic charge should be 3/2 times the anodic
charge. The experimental ratio for stirred solution was found to be 1.4,
in reasonable agreement with the calculated value.
Estimation of the thickness of the chalcopyrite layer taking part in the
initial oxidation process requires a knowledge of the surface roughness in
addition to the charge passed. Since the roughness factor of electrodes
ground on 600 paper was not known, voltammograms were recorded for a
polished chalcopyrite surface in a solution of pH 11. An anodic wave com-
mencing at --0.4 V was observed in addition to the peak at positive potentials
evident in Fig. 4. We assign this anodic wave to oxidation of iron (II) hy-
droxide by reaction (3). The presence of iron hydroxide implies that sulphur
was lost from the surface during the polishing procedure. This conclusion is
in agre.ement with the report of Peters {1977) that polished chalcopyrite was
nearly passive while rough saw-cut specimens were reactive.
In order to estimate the thickness of the layer involved in reaction (1),
we have assumed that the charge associated with the anodic peak corre-
sponds to this process. The charge integrated between --0.1 and 0.25 V
amounted to 0.9 mC cm -2 which, assuming a roughness factor of unity for
the polished surface, is equivalent to 1.4 nm of chalcopyrite. The structure
of this mineral involves cubic close-packed sulphur layers 0.26 nm apart in
which copper and iron occupy tetrahedral sites in an ordered configuration
(Hall and Stewart, 1973). Thus the oxidized layer extends over about 5
sulphur layers.
The oxidation process is remarkably reversible, considering the complex
nature of the chalcopyrite structure, though continuous cycling does ul-
timately lead to significant changes in the voltammogram (Fig. 9). The
anodic and cathodic peaks corresponding to reaction (1) diminish in size and
additional peaks appear. This behaviour can be explained by dissolution of
a small fraction of the iron on each cycle. The additional peaks at the lower
potential limit can be assigned to reactions of the sulphur originally associat-
ed with the iron that has dissolved [reaction (4)] and the increase in anodic
current and the additional cathodic peak at the upper limit to reactions of
the resulting excess of CuS [reaction (6}].
13

i T F [
pH 92

~ 50 I
~g

~., 50
o

~ -5o
k
j ~
I__ I I
o!4 -o z o 02 --04
Poter~hal (V V5 5HE)

Fig. 9. Voltammograms for chalcopyrite in pH 9.2 buffer solution, C o n t i n u o u s triangular


potential s w e e p s a t 4 m V s TM ,

Contact angles

A zero contact angle with no apparent tendency for "cling" was observed
at pH 11 over the potential range --0.6 to 0.8 V, i.e. the range covered by
the flotation experiments in Fig. 3. The contact angle was also found to be
zero at oxidizing as well as reducing potentials in the other solutions em-
ployed in the electrochemical investigations. Thus, we could not detect a
finite angle under the conditions in which efficient flotation was observed.
Heyes and Trahar (1977) report similar findings for chemical control of the
potential of chalcopyrite. These results contrast with previous investigations
of the interaction of surfaces with xanthate collectors (Gardner and Woods,
1974, 1977) in which good correlations were obtained between flotation
and contact-angle measurements under potential control.
Sutherland and Wark (1955) report investigations on the behaviour of
galena in the presence of paraffin-chain collectors such as sodium hexadecyl
sulphate in which conditions were defined where (1) the mineral floated in
a pneumatic cell but not in a cylinder and the contact angle was zero, (2)
the mineral floated in both a pneumatic cell and a cylinder but no contact
could be induced, and (3) flotation occurred in both experimental systems
and there were finite contact angles. Regime (3) was characteristic of low
collector concentrations. As the concentration was increased, (2) and then
(1) were observed. These findings were explained in terms of adsorption
of the collector at the gas/solution interface resulting in a long induction
period for bubble/mineral contact. Systems which create fresh bubbles,
which can contact the mineral before collector species have time to adsorb
at the gas/solution interface, have short induction times and efficient flo-
tation. Aged bubbles on which collector has adsorbed can be ineffective
for flotation.
The present results cannot be accounted for by a mechanism involving
14

kinetics of adsorption on the bubble since no collector species was present.


However, it is possible that an energy barrier exists which must be overcome
before a particle will attach to a gas bubble. This would explain why a
stationary bubble fails to coalesce with a chalcopyrite surface, the mineral
will float, b u t not very efficiently, in a Hallimond tube while flotation is
rapid in a conventional flotation cell. No reason for such a barrier can be
presented at this time and the explanation must be considered as speculative.
However, it is clear that differences do exist between the results of the dif-
ferent techniques. We conclude from these observations that the modified
Denver cell designed by Heyes and Trahar is particularly useful for the study
of flotation as a function of pulp potential, because it involves the same
pulp density and particle--bubble contact as in a practical flotation system.

Surface composition and flotability

The electrochemical investigations demonstrate that the onset of natural


flotability of chalcopyrite coincides with the appearance of an anodic
reaction. We therefore conclude that the oxidation of the mineral surface
is responsible for the mineral becoming naturally flotable. However, it must
be pointed out that flotation response is the same at pH 8 and 11 (Trahar,
private communication) while the oxidation reaction is pH dependent.
The initial oxidation products in alkaline solution are CuS, Fe(OH)3 and
S. These products behave as definite chemical entities, and the reaction
can not be considered as simply adsorbing oxygen or incorporating oxygen
into the chalcopyrite lattice. Flotation persists at high potentials (Fig. 3),
conditions under which further oxidation of the CuS to sulphur by reaction
(6) will take place. Iron oxide minerals and hydrated iron oxide layers on
pyrite are quite hydrophilic (Gardner and Woods, 1977). There is no evi-
dence to suggest that CuS would display hydrophobic properties and flota-
tion still occurs when this material is itself oxidized. We must, therefore,
identify flotability with the presence of sulphur on the mineral surface.
Sulphur (Sutherland and Wark, 1955) and sulphides with layer structures,
which cleave readily along sulphur layers, such as molybdenite, (Gaudin et
al., 1957), are known to be naturally flotable. Finkelstein et al. (1975)
have shown that sulphides can be rendered flotable by the formation of
sulphur layers on their surfaces but no correlation could be established
between flotation efficiency and the quantity of sulphur present. Indeed,
instances of high flotability at low surface coverages and low flotability in
the presence of multilayers of sulphur were observed. Only when the amount
of sulphur at the surface was high (much greater than 6 layers) was good
flotation obtained consistently. They determined the thickness of the
sulphur layers by extraction with acetone. It was pointed out by the authors
that this technique could not distinguish between sulphur being concentrat-
ed in clusters on the surface and being evenly distributed. Differences in
the manner of distribution of the surface sulphur could explain some of the
differences observed.
15

The investigations o f Finkelstein et al. ( 1 9 7 5 ) suggest t h a t f l o t a b i l i t y can-


n o t be a s s u m e d s i m p l y because s u l p h u r is p r e s e n t o n the mineral surface.
The s t r u c t u r e o f the surface layers including the s u l p h u r c o u l d be i m p o r t a n t
in d e t e r m i n i n g the influence on interfacial p r o p e r t i e s and this c o u l d be p H
d e p e n d e n t . It is a p p a r e n t t h a t f u r t h e r s t u d y on these aspects is required in
o r d e r to elucidate the role o f sulphur in d e t e r m i n i n g natural flotability.

CONCLUSIONS

(1) C h a l c o p y r i t e is naturally flotable if the p o t e n t i a l o f the mineral -


s o l u t i o n interface is above a critical value b u t n o t if it is b e l o w this value.
(2) F l o t a b i l i t y is associated with a n o d i c o x i d a t i o n o f the surface chalco-
pyrite.
(3) T h e p r o d u c t s o f the o x i d a t i o n r e a c t i o n are CuS, Fe(OH)3 and S.
(4) The presence o f s u l p h u r on the mineral surface is the critical f a c t o r
in r e n d e r i n g c h a l c o p y r i t e flotable.

ACKNOWLEDGEMENT

The a u t h o r s are i n d e b t e d to Mr. W.H. T r a h a r f o r suggesting t h a t t h e y un-


d e r t a k e this s t u d y and for useful discussions during the course o f the w o r k .

REFERENCES

Bates, R.G., 1964. Determination of pH. Wiley, New York, N.Y., pp. 458--483.
Chander, S. and Fuerstenau, D.W., 1975. Effect of potential on the flotation and wet-
ting behaviour of chalcocite and copper. Trans. A.I.M.E., 258: 284--285.
Eadington, P., 1977. Study of oxidation layers on surfaces of chalcopyrite by use of
Auger electron spectroscopy. Trans. Inst. Min. Metal., 86: C186--189.
Finkelstein, N.P., Allison, S.A., Lovell, V.M. and Stewart, B.V., 1975. Natural and in-
duced hydrophobicity in sulphide mineral systems. In: P. Somasundaran and R.G.
Grieves (Editors), Advances in Interfacial Phenomena of Particulate/Solution/Gas
Systems. Applications to Flotation Research, Am. Inst. Chem. Eng., Syrup. Ser. No.
150, 71: 165--175.
Gardner, J.R. and Woods, R., 1973. The use of a particulate bed electrode for the electro-
chemical investigation of metal and sulphide flotation. Aust. J. Chem., 26: 1635--
1644.
Gardner, J.R. and Woods, R., 1974. An electrochemical investigation of contact angle
and of flotation in the presence of alkylxanthates. I. Platinum and gold surfaces. Aust.
J. Chem., 27: 2139--2148.
Gardner, J.R. and Woods, R., 1977. An electrochemical investigation of contact angle
and of flotation in the presence of alkylxanthates. II. Galena and pyrite surfaces.
Aust. J. Chem., 30: 981--991.
Gaudin, A.M., Miaw, H.L. and Spedden, H.R., 1957. Native flotability and crystal struc-
ture. Proc. 2nd Int. Congr. of Surface Activity, 3. Butterworths, London, pp. 202--
219.
Hall, S.R. and Stewart, J.M., 1973. The crystal structure refinement of chalcopyrite,
CuFeS:. Acta Cryst., B29: 579--585.
Heyes, G.W. and Trahar, W.J., 1977. The natural flotability of chalcopyrite. Int. J.
Miner. Process., 4: 317--344.
16

Janetski, N.D., Woodburn, S.I. and Woods, R., 1977. A n electrochemical investigation
of pyrite flotation and depression. Int. J. Miner. Process., 4: 227--239.
Latimer, W.M., 1952. Oxidation Potentials.Prentice-Hall,N e w York, N.Y., 2nd ed.,
409 pp.
Lepetic, V.M., 1974. Flotation of chalcopyrite without collector after dry autogenous
grinding. C.I.M. Bull.,June, pp. 71--77.
Linge, H.G., 1976. A study of chalcopyrite dissolution in acidic ferricnitrate by poten-
tiometric titration.Hydrometallurgy, 2 : 51--64.
Michell, D. and Woods, R., 1978. Analysis of oxidized layers on pyrite surfaces by X-
ray emission spectroscopy and cyclic voltammetry. Aust. J. Chem., 31: 27--34.
Peters, E., 1977. Electrochemistry of sulphide minerals. In: J.O'M. Bockris, D.A.J.
Rand and B.J. Welch (Editors), Trends in Electro-Chemistry. Plenum, N e w York,
N.Y., pp. 167--290.
Richardson, P.E. and Maust, E.E., Jr., 1976. Surface stoichiometry of galena in aqueous
electrolytes and its effect on xanthate interactions.In: M.C. Fuerstenau (Editor),
Flotation: A.M. Gaudin Memorial Volume, 1. A.I.M.E., N e w York, N.Y., pp. 364--
392.
Sutherland, K.L. and Wark, I.W., 1955. Principles of Flotation. Australas. Inst. Min.
Metall., Melbourne, Vic., 489 pp.
Wadsworth, M.E., 1972. Advances in the leaching of sulphide minerals. Min. Sci. Eng.
4(4): 36--47.
Young, P.A., 1967. The stability of copper--iron sulphides. Amdel Bull., No. 3, pp. 1--
19.

You might also like