You are on page 1of 10

Minerals Engineering 26 (2012) 70–79

Contents lists available at SciVerse ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Benchmarking the flotation performance of ores


S. Muganda a, M. Zanin a, S.R. Grano b,⇑
a
Ian Wark Research Institute, University of South Australia, The ARC Special Research Centre for Particle and Material Interfaces, Mawson Lakes Campus, Adelaide,
South Australia 5095, Australia
b
Institute for Mineral and Energy Resources, University of Adelaide, South Australia 5005, Australia

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

Article history: A porphyry copper ore containing chalcopyrite as the principal copper bearing mineral, and pyrite as the
Received 1 July 2011 only other sulphide mineral, was treated in batch flotation tests under well defined physical conditions.
Accepted 5 November 2011 The size-by-size flotation response was benchmarked against established calibration curves to infer an
Available online 10 December 2011
operational contact angle of the sulphide minerals as a function of particle size. The inferred operational
contact angle values of the sulphide minerals were validated by independent measurements of contact
Keywords: angle on the concentrates and, in the case of chalcopyrite, by an indirect approach using Time of Flight
Flotation
Secondary Ion Mass Spectrometry (ToF-SIMS).
Contact angel
Floatability
Recovery, flotation rate, and inferred operational contact angle increased with collector addition across
all size fractions, with the intermediate and coarse size fractions benefitting the most from increased col-
lector addition. The directly measured and inferred operational contact angles were in reasonable agree-
ment, with an R2 value of 0.7 across all size fractions. There was good agreement between the advancing
contact angle values determined using ToF-SIMS and those calculated from direct contact angle measure-
ment on the 53–75 lm size fraction for the case of chalcopyrite. A method for benchmarking flotation
response has been developed, which may lead to better flotation process diagnostics and modelling.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction size. In other studies, hydrophobicity was shown to depend on par-


ticle size and surface roughness (Drelich and Miller, 1992). In these
The flotation behaviour of model chalcopyrite particles has been studies, a pseudo-line tension was introduced to describe more
characterised as a function of particle size fraction and advancing realistically the effect of surface roughness on the line tension,
contact angle range under well defined physical conditions which controls the critical contact angle of flotation, and therefore
(Muganda et al., 2011a). The model chalcopyrite particles were lib- is a measure for particle hydrophobicity.
erated and of high purity such that the flotation response could be Previous investigations on the relationship between the particle
attributed to the effects of particle size and advancing contact an- size, contact angle and flotation behaviour of chalcopyrite (Muganda
gle. In the case of a natural ore, the principal scope of this current et al., 2011a) demonstrated that the advancing contact angle varies
study, the degree of liberation has been identified as an important with particle size due to differences in surface species even for nar-
parameter determining floatability (Runge et al., 2003). Sutherland row particle size fractions subjected to the same treatment. This
(1989) observed that copper bearing value minerals within all size finding necessitates the determination of contact angle and flotation
fractions and in all liberation classes including those fully liberated, behaviour on a size-by-size basis, at least. For coarse particle size
displayed two-component flotation behaviour. In the study by fractions, incomplete liberation may reduce the particle contact an-
Sutherland (1989) it was not possible to link one floatability com- gle, while at fine size fractions (less than 10 lm) oxidation and metal
ponent to the degree of liberation or composition of the composite ion hydrolysis may play key roles. Heterogeneity in surface oxida-
particle. Sutherland (1989) assumed that the hydrophobicity of the tion of sulphides results in the variation of hydrophobicity, even
chalcopyrite surface was independent of the particle size and de- within the same particle size fraction. Furthermore, the develop-
gree of liberation. Rather, differences in floatability were attributed ment of the contact angle with increasing collector addition is also
to the amount of chalcopyrite in a particle, and not to differences in apparently particle size dependent (Trahar, 1981). A distribution
chalcopyrite hydrophobicity, implying that all exposed chalcopy- of particle contact angles within individual size fractions of a feed
rite surfaces had the same contact angle, irrespective of the particle sample was demonstrated by direct contact angle measurements
on flotation concentrates and tailings (Muganda et al., 2011b). In this
⇑ Corresponding author. Tel.: +61 08 83130626. current paper, it is assumed that multi-component flotation behav-
E-mail address: stephen.grano@adelaide.edu.au (S.R. Grano). iour within a size fraction, and within any liberation class, may be

0892-6875/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2011.11.002
S. Muganda et al. / Minerals Engineering 26 (2012) 70–79 71

explained in terms of the distribution of particle advancing contact phobic copper bearing mineral in the ore precluded the direct mea-
angle. Thus floatability, and more specifically, the flotation rate surement of contact angle using the Washburn technique, as
and floatability components, is a function of particle size and contact discussed in this paper.
angle, for a given hydrodynamic condition.
While the contact angle is evidently one of the most important
2. Experimental
parameters determining particle floatability, its measurement on
heterogeneous particle mixtures in which the bulk is hydrophilic
2.1. Sample preparation (natural ore)
gangue is very problematic. As a result, researchers have resorted
to characterising flotation response of natural ores in terms of parti-
A porphyry copper ore with chalcopyrite as the principal copper
cle size, liberation class, and collector surface coverage (Sutherland,
bearing mineral, and pyrite as the only other sulphide mineral,
1989; Runge et al., 2003; Vianna, 2004). The surface coverage of col-
supplied by Kennecott Utah Copperton Concentrator, USA, was
lector molecules largely determines the hydrophobicity of the parti-
used in the current test work. Crushed ore (<2.4 mm) was blended
cle. The particle contact angle controls the rate of bubble–particle
and riffled into 2 kg samples. A grind size calibration was carried
attachment (Holuszko et al., 2008), and hence the flotation rate con-
out with 2 kg feed samples to obtain a d80 of 220 lm (similar to
stant. It may be possible to infer the contact angle of particles of a
the plant operation). As a standard procedure, a 2 kg ore sample
natural ore by benchmarking the flotation rate constant as demon-
was ground in a stainless steel rod mill with about 1.8 g of lime
strated in our previous paper which focussed on model, single min-
and 1 dm3 of water for 12 min. The pH of the mill discharge was
eral chalcopyrite and pyrite particles (Muganda et al., 2011b).
in the range 9.4–9.8. The size distribution after grinding, as well
The flotation rate constant has been considered to be the best
as the copper and sulphur assays per size fraction, are shown in Ta-
descriptor of floatability (Imaizumi and Inoue, 1963). The chal-
ble 1. BMA analysis by QEM-Scan confirmed that the sulphur bear-
lenge now, is to be able to predict flotation behaviour through a
ing minerals were chalcopyrite and pyrite only. The sulphur
description of the rate constants for a given particle size fraction
recovery is used to represent the combined recovery of both chal-
and contact angle range, but under strictly controlled and specific
copyrite and pyrite.
hydrodynamic conditions (Prestidge and Ralston, 1996). An at-
tempt is made here to characterise the flotation response of chal-
copyrite under well-defined physical conditions, and to use the 2.2. Flotation tests
floatability of single minerals to infer the effective operational con-
tact angles of particles in an ore. The characterisation of the flota- A 5 dm3 bottom driven flotation cell was used in the flotation
tion response of chalcopyrite, as a single mineral, has led to the tests. The ground ore was transferred to the flotation cell where
development of calibration curves of the undistributed rate con- the pulp volume was increased to 5 dm3 with water to give a solid
stant, k, and collection efficiency, Ecoll, against particle size fraction content of 30% (w/w). The pH was adjusted to 10 where necessary,
for different contact angle ranges (Muganda et al., 2011a). These by adding lime, and the pulp subsequently conditioned with col-
curves have been used to benchmark the flotation responses of lector and frother. Sodium dicresyl dithiophosphate, DTP (S-
chalcopyrite with a different type of collector (Muganda et al., 8989) was used as collector while methyl isobutyl carbinol (MIBC)
2011b). It was shown that the flotation response of chalcopyrite was used as frother, the latter at a constant addition of 37.5 g/t
is the same when the advancing contact angle, measured by the (60 ll) of ore in all tests. Four flotation tests with collector
Washburn method on particle ensembles of a given size fraction, additions of 2, 15, 30 and 40 g/t were carried out. These collector
is the same, within experimental error. It was also shown that additions were chosen to give discernible differences in particle
the calibration curves were valid at a higher pulp density (30% sol- hydrophobicity. The conditioning time for collector and frother
ids) of a mixture of chalcopyrite and non-interacting gangue min- was 1 min each in turn. The superficial gas velocity in flotation
eral, paving the way for benchmarking the flotation response of a was 0.3 cm s1 (volumetric flow rate of air per sec divided by the
natural ore against the established calibration. cross-sectional area of the flotation cell), and the impeller rota-
This paper tackles the determination of the operational advanc- tional speed was 1200 rpm, while concentrates were collected at
ing contact angle of both chalcopyrite and combined sulphide min- cumulative times of 1, 3, 5, and 8 min, with a scraping rate of 1
erals (i.e., chalcopyrite and pyrite combined) in an ore feed using every 10 s. Each test was repeated four times to generate sufficient
their flotation behaviour. A porphyry copper ore is floated at differ- mass for sizing, advancing contact angle measurements, and assay.
ent collector additions and the flotation behaviour of chalcopyrite
in each size fraction is benchmarked against the calibration curves. 2.3. Chemical assay and mineralogical analysis
An attempt is made to verify the inferred contact angle values by
directly measuring the advancing contact angles on size fractions The feed, concentrates and tail samples were wet sieved into
of the concentrates and tailings from the flotation tests, and back the same size fractions shown in Table 1. The concentrates were fil-
calculating the contact angle of the feed. The approach is some- tered and dried in a desiccator under vacuum. Samples were col-
what difficult because the bulk of the tailing is composed of hydro- lected for chemical assay. The chemical assays of the feed (Table
philic gangue and the Washburn method, as demonstrated in this 1) and flotation product size fractions allowed the recovery to be
paper, is insensitive to low concentrations of hydrophobic particles calculated as a function of particle size fraction. The samples were
in a bulk hydrophilic matrix. In the same vein, direct contact angle wet-sieved to ensure that the surfaces of particles were cleaned of
measurements on the feed size fractions were not successful be- slimes. This was particularly important for concentrate samples
cause the hydrophobic component was also too low. Furthermore, destined for advancing contact angle measurements and surface
there is the potential issue of surface oxidation which may take analysis. The remainder of the concentrates for each size fraction
place during sample preparation of sulphide minerals for the con- were combined and used in direct contact angle measurements
tact angle measurement. However, the techniques developed for and surface analysis by ToF-SIMS (for the 53–75 lm size fraction
sample preparation seemed to preserve the surface species only) as explained further below.
(Muganda et al., 2011b). Indeed, these difficulties were precisely Mineralogical analysis of all feed size fractions and selected
the primary motivation for developing the method of inferring concentrate and tailing size fractions was carried out by means
the advancing contact angle by benchmarking flotation response of QEM-Scan analysis. Samples of the 150–210, 210–300, and
against a well-defined standard. The low concentration of hydro- +300 lm size fractions of the concentrate for the test at 15 g/t
72 S. Muganda et al. / Minerals Engineering 26 (2012) 70–79

Table 1
Size distribution for 2 kg of KUCC ore ground for 12 min (10 rods, 9.3 kg).

Size fraction, lm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 +300 All sizes
Size distribution, wt% 15.8 6.9 5.5 7.7 10.3 11.5 14.7 12.8 14.8 100
Cu, wt% in sampled feed 0.47 0.61 0.80 0.59 0.54 0.40 0.36 0.30 0.22 0.43
Typical recalculated Cu, wt% in feed 0.62 0.62 0.81 0.58 0.54 0.39 0.36 0.30 0.22 0.46
S, wt% in sampled feed 0.74 1.17 1.77 1.45 1.68 1.56 1.75 1.49 1.09 1.47
Typical recalculated S, wt% in feed 0.92 1.34 1.54 1.57 1.67 1.52 1.84 1.94 1.13 1.47

collector were analysed to gain an understanding of the liberation Table 2


characteristics, and the type of gangue minerals in the concentrate. Bulk mineralogical assay (wt%) for KUCC ore by QEM-Scan.

The chalcopyrite liberation profile of the feed size fractions is Mineral Wt%
shown in Fig. 1, which shows that for size fractions up to 105– Chalcopyrite 1.5
150 lm, 80% of the chalcopyrite was at least in the 80% liberation Pyrite 1.8
class or greater. Chalcopyrite liberation decreased above the 105– Quartz 30.2
150 lm size fraction, with the +300 lm fraction still having 50% of Feldspars 36.7
Micas 24.4
the chalcopyrite in at least the 80% liberation class or greater. This
Chlorite 1.1
liberation profile has significance to the assignment of contact an- Other sulphides 0.02
gle values for the floatability components discussed further below.
The bulk mineralogical composition of the ore feed is shown in
The sensitivity of the Washburn method was tested using mix-
Table 2. The only significant copper bearing mineral was chalcopy-
tures of chalcopyrite and quartz in the 75–105 lm size fraction.
rite. There was significant pyrite, relative to chalcopyrite, which
The quartz had an advancing contact angle of 0°. The proportion
complicated the direct determination of the contact angle on chal-
of chalcopyrite, previously conditioned with sodium dicresyl dithio-
copyrite using the Washburn method. The other gangue minerals
phosphate (DTP) (with an advancing contact angle of 90°), was var-
were quartz, feldspars and micas.
ied from 0% to 100% volume fraction of the mixture. The results of
The presence of pyrite, in addition to chalcopyrite, as minerals
these measurements provides a guide to the reliability of the
for which the contact angle can change with collector addition,
advancing contact angles measured directly on concentrates, and
complicated the direct measurement of contact angle, necessitat-
their potential use in the calculation of the contact angle of chalco-
ing also the consideration of sulphur recovery representing the
pyrite in the ore concentrate. Advancing contact angles were mea-
combined recovery of pyrite and chalcopyrite. Since there were
sured on all combined concentrate size fractions from the three
no other significant sulphur bearing minerals, sulphur recovery
flotation tests at 2, 15, and 30 g/t of DTP. Contact angle values for
can be equated to the recovery of sulphide minerals, both chalco-
the concentrates of the test at 40 g/t were not measured because
pyrite and pyrite.
the recovery and rate constant for each size fraction was the same,
within experimental error, as the 30 g/t test. Attempts were also
made to measure the advancing contact angle on the size fractions
2.4. Contact angle measurement on heterogeneous mixtures of
of the feed and tailing samples. However, this approach gave a con-
particles
tact angle of zero under all conditions, effectively ruling out the pos-
sibility of measuring the contact angle directly on the feed sample.
Advancing contact angle measurements using the Washburn
technique (Washburn, 1921) were carried out on heterogeneous
2.5. Surface analysis by Time of Flight Secondary Ion Mass
mixtures of particles. This method was chosen for its simplicity
Spectrometry (ToF-SIMS)
and the fact that several measurements can be done in a short time
(Chau, 2009) with good reproducibility (Muganda et al., 2008). The
An independent surface analysis study via ToF-SIMS (Brito E
Washburn method makes use of capillary pressure to drive a liquid
Abreu et al., 2010) showed a distribution in surface chemistry of
at an observable rate through a packed bed of particles in a capil-
chalcopyrite particles within a sample treated with collector. This
lary tube to measure the advancing contact angle. The description
heterogeneity in surface chemistry was shown to correlate with
of the method used is given elsewhere (Stevens, 2005; Chau, 2009;
a distribution of contact angle. The study, which used the same
Muganda et al., 2011a).
chalcopyrite particles as in our previous paper (Muganda et al.,
100
2011b), resulted in an empirical equation to correlate the advanc-
Cumulative Liberation Yield (%)

ing contact angle and surface species, a correlation which was


independent of particle size fraction, and with a range of validity
80
in contact angle up to a value of 90° (Brito E Abreu, 2011). In this
current paper, the methodology was trialled on the ore concen-
60
trate, in which it was possible to detect a statistically significant
number of chalcopyrite particles by ToF-SIMS. The empirical rela-
40 tionship (Eq. (1)), which is specific to the chalcopyrite-DTP system,
was developed for the calculation of the contact angle of chalcopy-
20 20-38 38-53 53-75 rite using ToF-SIMS intensities (I). Important species considered
75-105 105-150 150-210 are oxygen (O), sulphur (S) and collector (coll), in this case DTP
210-300 +300
0 (Brito E Abreu et al., 2010).
0 20 40 60 80 100
Contact Angle ¼ 45:74  1:208IO þ 3:065IS þ 15:82Icoll ð1Þ
Liberation (%)
This methodology was used to ‘‘measure’’ the contact angle of
Fig. 1. Cumulative liberation yield as a function of percent liberation for chalco- chalcopyrite particles in the 53–75 lm size fraction of the concen-
pyrite in KUCC ore feed size fractions. trate for the flotation test at 15 g/t collector. This approach
S. Muganda et al. / Minerals Engineering 26 (2012) 70–79 73

provided a way to indirectly determine the contact angle of chalco-


pyrite particles in the concentrate, which was compared with di-
rectly measured value on the heterogeneous mixture (using the
Washburn technique). The result provided confidence in the
remainder of the measured advancing contact angles of the size
fractions of the concentrate.

2.6. Data analysis

2.6.1. Recovery by entrainment


Recovery by entrainment in a single batch flotation test was
estimated using the method by Ross (1991), and was subtracted
from the size-by-size recovery data to give recovery by bubble–
particle attachment. Details of the calculations for entrainment
subtraction were given in our previous paper (Muganda et al.,
2011a). Fig. 2. Maximum recovery (Rmax) of chalcopyrite as a function of particle size at
different collector additions for KUCC ore. Collector (DTP) added at 2 (e), 15 (h), 30
2.6.2. Recovery and rate constant (D), and 40 g/t (o). Frother 37.5 g/t MIBC, 1200 rpm, 30% solids (w/w).
The rate constant, after entrainment subtraction, was calculated
assuming that there exists in each size fraction a single floatable
fraction and a non-floating fraction (Muganda et al., 2011a,b). In the consequent increase in the contact angle. Fine and intermedi-
general, this gave a very good fit to the experimental data. ate particles apparently respond in flotation at lower collector
For batch flotation, the recovery, R, at time t is related to the additions, an observation similar to that made by Trahar (1981).
maximum recovery, Rmax at infinite time, for a floatable component Increased collector addition, and hence surface coverage of col-
with recovery less than 100%, by (Ralston, 1992; Rahal et al., 2000), lector, causes more particles to be hydrophobised to contact angles
above the critical value for stable bubble–particle attachment.
R ¼ Rmax ð1  ekt Þ ð2Þ Apparently, the flotation behaviour of particle size fractions below
where Rmax is the maximum recovery at infinite time or maximum 53 lm (Fig. 2) was nearly the same in the four tests.
recovery possible, k is the flotation rate constant of the floatable The fine to intermediate size fractions appear to preferentially
component. A modified flotation rate constant which takes into ac- adsorb collector, possibly because they have a higher surface area
count both maximum recovery and rate constant has been pro- to solution volume ratio than the coarse particles. Another possi-
posed (Agar et al., 1986; Sripriya et al., 2003). In this paper, the bility is that the degree of liberation is also higher for the fine to
modified flotation rate constant is called the ‘undistributed’ rate intermediate size fractions, thus more surface is available to adsorb
constant, k, and is calculated using: collector. As the collector addition is increased, coarse particles ad-
sorb more collector and become more hydrophobic, hence there is
 kRmax
k ¼ ð3Þ an increase in recovery with an increase in collector addition. The
100 maximum recovery was highest for the 38–53 lm size fraction in
From a hydrodynamic point of view, the flotation rate constant all three tests. This is the optimum size fraction for flotation under
(k) can be related to gas dispersion parameters, such as the bubble the test conditions.
surface area flux (Sb), and to the probability of particle collection by The decrease in recovery with an increase in size above 53 lm is
an air bubble, i.e., the collection efficiency (Ecoll), (Jameson et al., attributed to the increase in the critical contact angle with particle
1997) by: size (Muganda et al., 2011a). A smaller mass of particles have a
contact angle above the critical value when collector addition is
1
k¼ Sb Ecoll ð4Þ low, hence recovery is lower. The degree of liberation of the parti-
4
cles also affects the contact angle achieved by a mineral aggregate,
especially when the mineral particle exists as a composite in asso-
3. Results ciation with hydrophilic gangue. As the particle size increases, the
stability of the bubble particle aggregate in agitated pulps de-
3.1. Natural ore flotation and inference of an operational contact angle creases due to an increase in mass, thus a higher contact angle is
required for bubble–particle aggregate stability.
3.1.1. Chalcopyrite flotation response A fit of the first order rate equation to the experimental recov-
The flotation behaviour of chalcopyrite obtained in laboratory ery with time, after subtraction of entrainment, was very good
tests on the natural ore is typical of plant practice in terms of the across all size fractions. The recovery with time reached an asymp-
recovery dependence on particle size (Fig. 2). The recovery-time totic value (Rmax) across all size fractions, giving justification to the
profile is similar to that obtained with single minerals (Muganda two-floatability components model (i.e., a floating and a non-float-
et al., 2011a,b). The final (unsized) recovery increased with collec- ing fraction).
tor addition, reaching 71%, 80%, 85%, and 87% for 2, 15, 30, and
40 g/t of DTP, respectively. The test at 40 g/t DTP produced the 3.1.1.1. Benchmarking chalcopyrite flotation behaviour. The objective
same results, within experimental error, as the 30 g/t test, showing of this study is to infer the contact angle of the hydrophobised
that collector was in excess at this high addition. The plant gener- chalcopyrite particles in the ore. The flotation behaviour of chalco-
ally uses approximately 20 g/t of DTP for this ore (Triffett et al., pyrite is benchmarked against calibration curves previously devel-
2008). It is evident (Fig. 2) that coarse particles benefit more from oped for the chalcopyrite-amyl xanthate system (Muganda et al.,
increased collector addition, similar to that observed in other min- 2011a). Further, the difference in pulp density between the single
eral systems (Polat and Chander, 2000; Vianna, 2004). The increase mineral work and ore work necessitated taking into account inev-
in recovery with collector addition may be explained in terms of an itable changes in bubble size distribution which accompanies
increase in surface coverage of the coarse particles by collector and changes in pulp density. In the single mineral work with 2% solids,
74 S. Muganda et al. / Minerals Engineering 26 (2012) 70–79

the Sauter mean bubble diameter was 0.48 mm while that with copper and may become more hydrophobic in the presence of col-
30% solids (chalcopyrite mixed with clean quartz) the Sauter mean lector. The presence of pyrite in the ore complicates an already dif-
bubble diameter was 0.57 mm (Muganda et al., 2011b). The Sauter ficult task in terms of determining the hydrophobicity of
mean bubble size measured was also 0.57 mm throughout flota- chalcopyrite in the ore. It is difficult enough to deal with one float-
tion in the current ore test work. The method used in measuring able species due to the existence of a distribution of both contact
bubble size and calculating the Sauter mean bubble diameter is gi- angle and floatability components, but when there are two differ-
ven in previous publications (Muganda et al., 2011a,b). ent minerals floating at the same time, the task becomes onerous.
The collection efficiency, Ecoll, (Eq. (4)) was used to benchmark It was decided, therefore, to consider sulphur recovery, represent-
the flotation behaviour of chalcopyrite in an ore. This conversion ing chalcopyrite plus pyrite recovery, and is referred to below as
from rate constant to collection efficiency eliminates the bubble sulphide recovery. QEM-Scan analysis justified this assignment,
surface area flux as a variable arising from differences in bubble as the only significant sulphur bearing minerals are chalcopyrite
size distribution. Empirical equations (Table 3) relating the contact and pyrite.
angle to the collection efficiency, based on the undistributed rate The maximum recovery of sulphides (chalcopyrite and pyrite)
constant (Muganda et al., 2011a), were used, so that for a given va- as a function of particle size at 2, 15, and 30 g/t collector addition
lue of the collection efficiency in the ore, an inferred operational is shown in Fig. 4. At 2 g/t collector addition, chalcopyrite recovery
contact angle value can be calculated. It is assumed that mineral (Fig. 2) is higher than sulphide recovery across all size fractions.
particles of different sizes in the ore float independently of each Although the recovery of chalcopyrite for the fine to intermediate
other, as observed in single mineral experiments described in pre- size fractions is high at this collector addition, it appears that pyrite
vious publications (Muganda et al., 2011a,b). The results of the flotation remains suppressed, possibly because there is insufficient
empirical relationships are shown in Table 4. collector to hydrophobise it to a contact angle above the critical va-
The concept of the benchmarking procedure is illustrated in lue. At higher collector additions, chalcopyrite and sulphide recov-
Fig. 3, where the collection efficiencies in Table 4 are plotted ery are the same, within experimental error, up to 53 lm. Above
against the calibration curves to infer an operational contact angle. 53 lm, the sulphide recovery is higher than chalcopyrite recovery,
This illustration, however, does not give a single contact angle va- which may be explained by increased recovery of pyrite as it be-
lue as obtained with the empirical equations, but a narrow range of comes sufficiently hydrophobic to float at higher collector addi-
contact angle, and the situation can be difficult if there is overlap of tions. These results seem to suggest that collector preferentially
error bars as in the coarse end of the calibration. adsorbs onto the chalcopyrite rather than pyrite, or that collector
An increase in the collector addition above 2 g/t did not increase adsorption may be the same, but that pyrite requires more collec-
the collection efficiency significantly for particle size fractions less tor to attain sufficient hydrophobicity to float, since, as discussed
than 53 lm. Generally, the collection efficiency increased with col- in our previous paper (Muganda et al., 2011b), pyrite has a higher
lector addition for size fractions above 53 lm only, although the critical contact angle for stable bubble–particle attachment.
increase was small for the very coarse particles. As the particle size Assays of the concentrates and the calculated weight percent
increases, the collection efficiency increases and reaches a maxi- and volume fractions of the chalcopyrite, pyrite and non-sulphide
mum for the 38–53 lm size fraction, and decreases with an in- gangue minerals are considered in Table 5, where it becomes evi-
crease size above this fraction. The maximum collection dent that pyrite recovery remained low at 2 g/t collector addition
efficiency is obtained with the 53–75 lm size fraction at the high- and was higher across all size fractions greater than 20 lm for
est collector addition. The inferred operational contact angle values higher collector additions. A multiple regression on the three com-
increase with an increase in collector addition for each size frac- ponents of the concentrates (chalcopyrite, pyrite, non-sulphide
tion. Assuming that these results are a true indication of the re- gangue) shows that the amount of chalcopyrite is negatively corre-
sponse of particles to conditioning with collector, the collection lated to the amount of non-sulphide gangue up to about 210 lm,
efficiency of copper bearing particles increases with an increase while pyrite becomes significant in its association with gangue
in the degree of hydrophobicity of individual particles within the minerals above that size. The recovery of sulphides is significantly
same size fraction. It is important to note that the inferred opera- higher than that of chalcopyrite alone at the coarse end of the size
tional contact angle is the contact angle on the feed of the valuable distribution for higher collector additions, and this is borne out by
mineral, as both the ore and the benchmark (single mineral chalco- the calculated concentrate compositions shown in Table 5.
pyrite-amyl xanthate system from Muganda et al., (2011a)) collec-
tion efficiencies are calculated from the undistributed rate
constant, taking into account both the floating and non-floating 3.1.2.1. Benchmarking sulphide flotation response. The flotation re-
components in the feed. sponse of the sulphides (chalcopyrite and pyrite) was bench-
The flotation behaviour of chalcopyrite in a natural ore may be marked against the calibration previously developed for the
affected by other minerals in the ore such as pyrite. In the follow- chalcopyrite-amyl xanthate system using the collection efficiency
ing section, the flotation behaviour of sulphides (i.e., chalcopyrite as before (Table 6). The inferred operational contact angles gener-
and pyrite) is considered. ally increase with collector addition and decrease with particle size
above 38 lm. Hydrophobic particles in the size fractions between
20 and 300 lm appear to have the same collection efficiency
3.1.2. Sulphide flotation response (and thus flotation rate) for the 15 g/t collector addition, the same
The natural ore has a considerable amount of pyrite from the flotation behaviour is noticed for the 30 g/t collector addition for
bulk mineralogical analysis (Table 2). Pyrite may be activated by the same range of sizes. Also significant is the difference between

Table 3
Empirical equations from single mineral work relating the contact angle and the collection efficiency, Ecoll. Contact angle (±2.5°) = c(Ecoll)d (Muganda et al., 2011a).

Particle size, lm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 +300
c 111 164 140 104 97 101 123 108 120
d 0.166 0.429 0.410 0.293 0.251 0.222 0.267 0.169 0.173
R2 0.95 0.99 0.95 0.97 0.97 0.96 0.96 0.96 0.95
S. Muganda et al. / Minerals Engineering 26 (2012) 70–79 75

Table 4
Collection efficiency, Ecoll, and inferred operational contact angles of chalcopyrite in KUCC ore feed after different collector additions.

Particle size fraction, lm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 300+
2 g/t Ecoll, ±0.02 0.18 0.24 0.32 0.22 0.14 0.05 0.02 0.01 0.002
Inferred contact angle, ±2.5 (°) >90 90 88 66 59 52 43 50 41
15 g/t Ecoll,±0.02 0.20 0.25 0.34 0.37 0.27 0.17 0.12 0.07 0.03
Inferred contact angle, ±2.5 (°) >90 90 90 77 70 68 70 69 65
30 g/t Ecoll, ±0.02 0.20 0.25 0.34 0.41 0.36 0.24 0.14 0.08 0.04
Inferred contact angle, ±2.5 (°) >90 90 90 80 75 73 73 71 69

at 15 g/t, there is agreement between the inferred contact angle


values of chalcopyrite and sulphides, showing that more pyrite
was hydrophobised to higher inferred contact angles, thus floating
at the same rate as the chalcopyrite particles (Fig. 5b). At 30 g/t col-
lector addition, the sulphides (combined) attain inferred contact
angles equal to, or greater than that of chalcopyrite (Fig. 5c).
An attempt is made below to validate some of the contact angles
inferred using the benchmarking procedure from calibration curves
developed for the chalcopyrite-amyl xanthate system (Muganda
et al., 2011a). The inferred operational contact angles are based
on the feed contact angle for each size fraction before particle sep-
aration by flotation. The approaches for validation included an indi-
rect method using surface analysis by ToF-SIMS specifically
addressing chalcopyrite, and direct contact angle measurements
on the ore feed, concentrate, and tailings addressing the sulphide
minerals (chalcopyrite and pyrite combined). These approaches
Fig. 3. Collection efficiencies of particle size fractions for 2, 15, 30 g/t ore flotation are discussed further below.
tests benchmarked against calibration curves of the collection efficiency, Ecoll,
against particle size for different contact angle ranges (Muganda et al., 2011a).
3.2. Validation of inferred operational contact angles

3.2.1. Surface analysis technique for contact angle measurement


A contact angle/species intensity (ToF-SIMS) correlation (Brito E
Abreu et al., 2010) was developed for the chalcopyrite-DTP system
using the same chalcopyrite sample used in our previous paper
(Muganda et al., 2011b). The developed methodology was trialled
on the 53–75 lm size fraction of the concentrate for the test at
15 g/t on the ore in the current study. Only mineral particles of
chalcopyrite were considered in the analysis. The Washburn
advancing contact angle measurement method incorporates the ef-
fects of particle surface roughness and shape, so that the correla-
tion between the advancing contact angle and ToF-SIMS species
intensity may be valid.
The predicted average contact angle of chalcopyrite particles in
the concentrate is 74° over 33 different chalcopyrite particles (Bri-
to E Abreu, 2011). The standard error of prediction is 14°. The dis-
tribution of contact angles of this group of particles is displayed in
Fig. 4. Maximum recovery (Rmax) of sulphide minerals (chalcopyrite and pyrite) as Fig. 6. The median is 69° and the mode is around 70°. The standard
a function of particle size at different collector additions 2 (e), 15 (h), and 30 (D) g/ deviation is 21°. According to the ToF-SIMS analysis the chalcopy-
t DTP for KUCC ore. rite particles in the concentrate have a range of abundances of the
surface species which would give rise to a range of contact angle
the flotation rate of chalcopyrite and sulphides for the coarse size values (Brito E Abreu et al., 2010).
fractions at 30 g/t collector addition. It appears that pyrite is The Washburn technique can only measure advancing contact
hydrophobised to high contact angles and its recovery is higher angles up to 90°. Some chalcopyrite particles in the ore appear to
than that of chalcopyrite. Pyrite and other non-sulphide gangue have adsorbed more collector than that observed in the calibration
minerals (NSG) appear in abundance in the concentrate as shown of the advancing contact angle and ToF-SIMS intensity correlation
in Table 5. Assuming that coarse non-sulphide gangue particles using single mineral chalcopyrite, resulting in contact angle values
that are liberated are hydrophilic and do not float on their own, greater than 90°. All particle contact angles above 90° are in the
all the non-sulphide gangue minerals recovered into the concen- extrapolated range and it is not yet clear how accurate these values
trate should then be in the form of composites with pyrite and are. As discussed previously, it is unlikely that the chalcopyrite par-
chalcopyrite. ticle advancing contact angle would be much greater than 100° for
A comparison of the inferred operational contact angles of chal- this mineral-collector system. Further, the ToF-SIMS correlation
copyrite and sulphide minerals is made in Fig. 5. For the test at 2 g/ holds for oxygen, sulphur and collector species present on chalco-
t collector addition, chalcopyrite generally has higher inferred con- pyrite in the ore system being the same as on the single mineral
tact angle values than the combined sulphides (Fig. 5a). For the test system from which it is derived. If for example, oxygen is derived
76 S. Muganda et al. / Minerals Engineering 26 (2012) 70–79

Table 5
Cu and S assays of KUCC ore concentrates from flotation tests with calculated weight % and volume fractions of chalcopyrite (chp), pyrite (py) and non-sulphide gangue (NSG).

Collector addition Size fraction, lm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 +300
2 g/t Cu, wt% 5.1 13.9 18.5 23.2 22.7 16.4 10.1 4.1 2.1
S, wt% 6.0 16.3 24.0 28.1 30.1 31.5 26.9 9.4 3.5
Wt% chp 14.8 40.2 53.4 66.9 65.7 47.5 29.2 11.7 6.1
Vol. chp 0.10 0.30 0.44 0.59 0.59 0.43 0.24 0.08 0.04
Wt% py 1.6 4.2 10.0 8.9 13.3 27.8 31.2 10.0 2.6
Vol. py 0.01 0.03 0.07 0.07 0.10 0.21 0.22 0.06 0.01
Wt% NSG 83.6 55.6 36.6 24.2 21.0 24.7 39.6 78.3 91.3
Vol. NSG 0.89 0.67 0.49 0.34 0.31 0.36 0.54 0.86 0.95
15 g/t Cu 4.6 11.4 14.1 15.6 12.1 8.1 4.8 2.7 2.0
S 6.92 20.78 28.18 34.36 37.71 39.19 25.68 22.72 23.76
Wt% chp 13.2 32.9 40.8 45.0 34.9 23.4 13.8 7.8 5.7
Vol. chp 0.09 0.26 0.35 0.42 0.34 0.23 0.11 0.06 0.05
Wt% py 4.3 17.4 26.1 34.9 47.7 58.1 39.0 37.4 40.7
Vol. py 0.02 0.11 0.19 0.27 0.39 0.48 0.27 0.25 0.27
Wt% NSG 82.5 49.7 33.1 20.1 17.4 18.6 47.2 54.8 53.6
Vol. NSG 0.89 0.63 0.46 0.30 0.27 0.29 0.62 0.69 0.68
30 g/t Cu 3.6 7.4 11.2 11.2 11.8 8.2 4.7 2.7 2.1
S 5.62 15.99 21.35 29.81 36.59 33.91 27.60 24.02 22.79
Wt% chp 10.4 21.4 32.3 32.4 34.0 23.8 13.4 7.9 6.2
Vol. chp 0.07 0.16 0.25 0.28 0.32 0.22 0.11 0.06 0.05
Wt% py 3.7 16.0 18.8 34.6 46.2 47.9 42.9 39.8 38.6
Vol. py 0.02 0.10 0.12 0.25 0.37 0.37 0.30 0.27 0.25
Wt% NSG 85.9 62.7 48.9 33.0 19.8 28.3 43.7 52.3 55.2
Vol. NSG 0.91 0.74 0.62 0.46 0.30 0.42 0.59 0.67 0.70

Table 6
Collection efficiency, Ecoll, and inferred operational contact angles of sulphides (chalcopyrite and pyrite) in KUCC ore feed after different collector additions.

Particle size fraction, lm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 +300
2 g/t Ecoll, ±0.02 0.14 0.22 0.22 0.08 0.05 0.03 0.01 0.01 0.001
Inferred contact angle, ±2.5 (°) >90 86 75 49 46 44 39 50 36
15 g/t Ecoll, ±0.02 0.16 0.26 0.23 0.23 0.22 0.22 0.21 0.2 0.11
Inferred contact angle, ±2.5 (°) >90 >90 77 67 66 72 81 83 55
30 g/t Ecoll, ±0.02 0.19 0.34 0.33 0.35 0.39 0.33 0.33 0.33 0.17
Inferred contact angle, ±2.5 (°) >90 >90 89 76 77 79 91 90 88

from silicates for which the oxygen yield is different from oxygen ores have less than 1% chalcopyrite in them, the bulk being hydro-
bearing species on chalcopyrite (e.g., iron oxide) in the calibration philic gangue minerals, so that direct contact angle measurements
stage, discrepancies may occur. The measured average contact an- do not yield meaningful results. In the current study, direct mea-
gle by ToF-SIMS appears reasonable when compared with the in- surements on the feed and tailing size fractions yielded zero con-
ferred contact angle of 77° (Table 4, 53–75 lm, 15 g/t), but the tact angle values. However, this result is very encouraging in that
inferred contact angle is for the feed, not the concentrate only. for concentrations of the hydrophobic mineral greater than 2%, it
The concentrate value is compared with the calculated value from should be possible to determine the contact angle on the hydro-
direct contact angle measurements described further below. phobic entity through use of the Cassie equation and armed with
The contact angle calculated using surface analysis is for the assumptions on the contact angle of the gangue component. This
floatable component only, while the tail component of the valuable methodology is described further below.
mineral could not be determined due to the very low concentration It is possible that the detection limit for the hydrophobic com-
of the chalcopyrite particles that did not float. Another approach ponent in the chalcopyrite–quartz mixture may be particle size
involving direct contact angle measurements on concentrates from dependent, and may also be influenced by the particle size in the
all flotation tests was considered, and is discussed further below. mixture, and the contact angle of the hydrophobic component.
These possibilities require further study.
3.2.2. Advancing contact angle measurements on heterogeneous
mixtures of particles 3.2.2.2. Ore system. QEM-Scan showed that the bulk of the chalco-
3.2.2.1. Model system. The sensitivity of the Washburn method was pyrite in the ore is more than 80% liberated (Fig. 1). The chalcopy-
evaluated using single mineral chalcopyrite–quartz mixtures of rite in the natural ore is associated with other minerals as noted
varying proportions and known contact angles. The Cassie equa- previously. Particle maps show three principal components in the
tion was used to calculate a theoretical contact angle value for each ore feed, i.e., silicates, pyrite and chalcopyrite (Fig. 8), consistent
mixture. There is good agreement between the measured and the with the assay data in Table 5. It is evident from the particle maps
calculated contact angle for chalcopyrite concentrations above that pyrite which is recovered into the concentrate occurs mainly
approximately 1% by volume (Fig. 7). It is evident that the Wash- in the fully liberated form, and less likely in association with chal-
burn technique is not sensitive enough to measure the contact an- copyrite and/or silicates. This result shows that any contact angle
gle on particle beds in which there is 1% by volume or less of the measured on the concentrate is an average of at least three major
hydrophobic component (the measured contact angle for such components, hydrophobic chalcopyrite and pyrite, and hydrophilic
beds was always zero, as depicted in Fig. 7). Most natural copper quartz.
S. Muganda et al. / Minerals Engineering 26 (2012) 70–79 77

90
Measured contact angle Cassie Equation
80

Advancing contact angle, °


70

60

50

40

30

20

10

0
0.001 0.01 0.1 1
Volume fraction of chalcopyrite

Fig. 7. Advancing contact angle as a function of the volume fraction of hydrophobic


chalcopyrite in a mixture of clean quartz (zero contact angle) and chalcopyrite. 75–
105 lm size fraction for both minerals.

The contact angles measured on the concentrate (Table 7)


showed small variations, and could be said to be the same, within
experimental error, for each size fraction over the range of collec-
tor additions. This is because of the presence of hydrophobic pyrite
as part of the gangue component, and it is interpreted as showing
that pyrite attains the same contact angle value as chalcopyrite
after collector addition (which also explains the very high flotation
recovery).
The contact angle measured on the concentrates includes
contributions from both the hydrophobic minerals and the non-
sulphide gangue, which is assumed to be hydrophilic. In order to
determine the contact angle of chalcopyrite in the concentrate, it
was assumed that chalcopyrite and pyrite had the same contact
angle within a size fraction and for a given collector addition (indi-
cated as sulphide minerals CA in Table 7). The non-sulphide gan-
gue component was assumed to have a contact angle of zero
(hydrophilic quartz). The Cassie equation was used to calculate
the mean contact angle (Priest et al., 2008). For a mixture of chal-
copyrite, pyrite and non-sulphide gangue, the average contact an-
gle is given by:

Fig. 5. A comparison of the inferred operational contact angle of chalcopyrite and cos haveragevalue ¼ x cos hchalcopyrite þ y cos hpyrite þ z cos hgangue ð5Þ
sulphide (chalcopyrite and pyrite) for different size fractions at collector additions
of (a) 2 g/t, (b) 15 g/t, and (c) 30 g/t DTP. where h is the mean contact angle of a given phase. The coefficients
x, y, and z, are the respective volume (surface area) fractions of the
three minerals. The contact angle of pyrite and chalcopyrite are
equal in Eq. (5), and the contact angle of the gangue component is
assumed to be zero.
Based on knowledge of the composition of each size fraction
(Table 5), the volume fraction of chalcopyrite, pyrite, and non-sulphide
gangue was calculated and, by means of Eq. (5), the contact angle
values of the sulphide minerals (Table 7) as also calculated.
A contact angle of 67° for chalcopyrite particles in the concen-
trate for the 53–75 lm size fraction is in good agreement with
the value determined indirectly using ToF-SIMS (74 ± 14°) at
15 g/t DTP. The agreement between the two methods is an encour-
aging development set to lead to new approaches of determining
particle wettability, down to the individual particle level. The re-
sults confirm the reliability of the Washburn technique for contact
angle measurement of powder materials. The advancing contact
angle values of chalcopyrite in the concentrates for the remainder
of the tests are shown in Table 7.
It must be borne in mind that the contact angle calculated using
Fig. 6. Distribution of the advancing contact angles measured using ToF-SIMS of
Eq. (5) and the procedure explained above gives the contact angle
chalcopyrite particles in the concentrate of the 53–75 lm size fraction of the test at of chalcopyrite particles in the concentrates, while the calibration
15 g/t of collector. curves developed with single mineral chalcopyrite (Muganda
78 S. Muganda et al. / Minerals Engineering 26 (2012) 70–79

Fig. 8. Particle maps for the ore (a) feed, (b) concentrate, and (c) tailing, for the 150–210 lm size fraction determined by QEM-Scan. Flotation test at 15 g/t collector addition.
NSG is non-sulphide gangue excluding silicates. Cu sulphides is chalcopyrite. Fe sulphides is pyrite.

Table 7 angle of zero, and reports to the tailings even at very high col-
Contact angles measured on the concentrates, the calculated sulphide contact angle
values within concentrate, and the back calculated feed contact angles. The back
lector additions. The liberation data for chalcopyrite (Fig. 1)
calculated feed contact angles use the Rmax values. ND – not determined. seem to justify this approach.
 The floatable fraction Rmax was assigned the contact angle value
Size Collector Sulphide Measured Calculated Back
fraction, addition, Rmax, % CA of CA of calculated
hc measured on the concentrates by the Washburn method.
lm g/t concentrate sulphide CA of  The non-floatable component is the fraction of the sulphides
sample, by minerals in sulphide that are lost to the tailings at 30 g/t collector addition and no
Washburn concentrate, minerals in further collector addition can recover (as evidenced by the sim-
technique by Eq. (5) (°) feed, by Eq.
ilar recovery at 40 g/t collector addition – Fig. 2). This fraction
(°) (8) (°)
was calculated by subtracting the maximum recovery at 30 g/t
20 2 77 – ND ND
collector from 100%.
15 86 – ND ND
30 89 – ND ND
 The difference between the maximum recovery at 30 g/t DTP
and the maximum recovery at either 2 or 15 g/t for each size
20–38 2 71 55 107 105
15 97 51 90 89
fraction represents the component of the chalcopyrite in the
30 98 49 108 107 tailing that is recoverable by increasing collector addition. This
38–53 2 82 57 84 83
component requires greater collector to attain a contact angle
15 97 53 75 75 above the critical value for stable bubble–particle attachment,
30 98 53 92 91 and was assigned a contact angle equal to the critical value
53–75 2 35 54 68 66 for that size fraction, hcrit, calculated in previous work (Muganda
15 87 55 67 66 et al., 2011a,b).
30 96 53 74 72
75–105 2 31 49 60 57 The contact angle of sulphide particles in the feed (last column
15 88 50 59 57 in Table 7) was also calculated using the Cassie equation, as the
30 95 56 68 66
weighted average of the calculated contact angle of the sulphides
105–150 2 26 47 60 53 in the concentrate, the component at the critical contact angle hcrit,
15 89 49 59 54
which may report to the tailing at low collector addition, and the
30 93 53 72 66
150–210 2 11 46 70 53 component that is locked within gangue, assigned a zero contact
15 80 48 82 69 angle for all collector additions. The feed contact angles were cal-
30 94 52 86 76 culated using the following equations at 2, 15 and 30 g/t collector
210–300 2 9 – ND ND additions, respectively, and for each size fraction:
15 83 49 96 72
30 91 51 97 76 cos hfeed;2 ¼ cos hc  Rmax2 þ cos hcrit  ðRmax30  Rmax2 Þ
+300 2 2 – ND ND þ ð1  Rmax30 Þ  cos 0 ð6Þ
15 65 50 97 51
30 74 50 101 55
cos hfeed;15 ¼ cos hc  Rmax2 þ cos hcrit  ðRmax30  Rmax15 Þ
þ ð1  Rmax30 Þ  cos 0 ð7Þ

et al., 2011a,b) were based on the feed contact angle. The inferred cos hfeed;30 ¼ cos hc  Rmax30 þ ð1  Rmax30 Þ  cos 0 ð8Þ
operational contact angles (Section 3.1) are based on the feed, and
thus the advancing contact angle of the sulphide particles in the where Rmax2, Rmax15 and Rmax30 are the fractional maximum recov-
feed, conditioned with collector, needed to be calculated. eries at 2, 15 and 30 g/t, respectively, for each size fraction.
A back calculation was carried out with the following There was a good agreement between the calculated and in-
assumptions: ferred operational contact angle (Fig. 9), with an R2 value of 0.7
across most of the size fractions. Given the complexity of a natural
 The sulphides in the feed are in two components, a floatable ore with two hydrophobic minerals, the agreement between the
(Rmax) and a non-floatable (100  Rmax) component. The latter inferred operational and measured feed contact angles is accept-
may be locked within gangue and may be assigned a contact able at this stage in the development of the methodology.
S. Muganda et al. / Minerals Engineering 26 (2012) 70–79 79

Brito E Abreu, S., 2011. Correlation of Surface Analysis with Particle Hydrophobicity
and Flotation. PhD Thesis, Ian Wark Research Institute. Adelaide, University of
South Australia.
Brito E Abreu, S., Brien, C., Skinner, W., 2010. ToF-SIMS as a new method to
determine the contact angle of mineral surfaces. Langmuir 26 (11), 8122–
8130.
Chau, T.T., 2009. A review of techniques for measurement of contact angles and
their applicability on mineral surfaces. Minerals Engineering 22, 213–219.
Drelich, J., Miller, J.D., 1992. The effect of surface heterogeneity on pseudo-line
tension and the flotation limit of fine particles. Colloids and Surfaces 69, 35–43.
Holuszko, M.E., Franzidis, J.P., Manlapig, E.V., Hampton, M.A., Donose, B.C., Nguyen,
A.V., 2008. The effect of surface treatment and slime coatings on ZnS
hydrophobicity. Minerals Engineering 21 (12–14), 958–966.
Imaizumi, T., Inoue, T., 1963. Kinetic consideration of froth flotation. In: Roberts, A.
(Ed.), Mineral Processing: Proceedings of the Sixth International Congress, May
26–June 2, 1965 ed. Pergamon Press, Cannes.
Jameson, G.J., Nam, S., Young, M., 1997. Physical factors affecting recovery rates in
flotation. Mineral Science Engineering 9 (3), 103–118.
Muganda, S., Zanin, M., Grano, S., 2008. Flotation behaviour of sulphide mineral size
fractions with controlled contact angle. In: Chemeca Conference Proceedings:
Towards a sustainable Australasia, Newcastle, New South Wales, Australia,
Engineers Australia.
Muganda, S., Zanin, M., Grano, S.R., 2011a. Influence of particle size and contact
Fig. 9. Comparison of the inferred (operational) and measured contact angle of
angle on the flotation of chalcopyrite in a laboratory batch flotation cell.
sulphide minerals in a natural ore. R2 = 0.7.
International Journal of Mineral Processing 98 (3–4), 150–162.
Muganda, S., Zanin, M., Grano, S.R., 2011b. Benchmarking flotation performance –
4. Conclusions single minerals. International Journal of Mineral Processing 98 (3–4), 182–194.
Polat, M., Chander, S., 2000. First-order flotation kinetics models and methods for
The flotation behaviour of particles is closely related to the par- estimation of the true distribution of flotation rate constants. International
Journal of Mineral Processing 58, 145–166.
ticle size and advancing contact angle under constant hydrody- Prestidge, C.A., Ralston, J., 1996. Contact angle studies of particulate sulphide
namic conditions. A method to determine an operational contact minerals. Minerals Engineering 9, 85–102.
angle of the sulphide mineral particles was developed and uses Priest, C., Stevens, N., Sedev, R., Skinner, W., Ralston, J., 2008. Inferring wettability of
heterogeneous surfaces by ToF-SIMS. Journal of Colloid and Interface Science
the undistributed flotation rate constant and collection efficiency 320, 563–568.
to infer the operational contact angle by benchmarking against a Rahal, K.R., Franzidis, J.-P., Manlapig, E.V., 2000. The application of the floatability
calibration developed with single mineral chalcopyrite. The ap- characterisation test rig (FCTR) in flotation research. In: Seventh Mill Operators
Conference, 12–14 October, Kalgoorlie, WA.
proach takes into account variations of the contact angle with par- Ralston, J., 1992. The influence of particle size and contact angle in flotation. In:
ticle size and infers a contact angle of the feed. This approach has Laskowski, J.S., Ralston, J. (Eds.), Developments in Mineral Processing. Colloid
been validated by direct contact angle measurements on combined Chemistry in Mineral Processing. Elsevier.
Ross, V.E., 1991. Comparison of methods for evaluation of true flotation and
concentrates on most of the size fractions, and indirectly for the entrainment. Transactions of the Institution of Mining and Metallurgy Section C
53–75 lm size fraction using a correlation involving ToF-SIMS – Mineral Processing and Extractive Metallurgy. The Institution of Mining and
intensities, for a natural porphyry sulphide ore. The calibration Metallurgy 100, C121–C126.
Runge, K.C., Franzidis, J.P., Manlapig, E.V., 2003. A study of the flotation
curves that were generated using single minerals can be used to in-
characteristics of different mineralogical classes in different streams of an
fer the operational contact angles of copper minerals in an ore, industrial circuit. In: Proceedings: XXII International Mineral Processing
irrespective of the degree of liberation. It appears that the degree Congress. Cape Town, South Africa, 29 September–3 October, 2003.
of liberation, at least for classes above 10% liberation, is less impor- Sripriya, R., Rao, P.V.T., Roy Choudhury, B., 2003. Optimisation of operating variables
of fine coal flotation using a combination of modified flotation parameters
tant than the contact angle of the chalcopyrite itself in determining and statistical techniques. International Journal of Mineral Processing 68, 109–
the flotation response of particles for the ore type investigated. 127.
Stevens, N.I., 2005. Contact Angle Measurements on Particulate Systems. Ian Wark
Research Institute, Adelaide, PhD Thesis. University of South Australia.
Acknowledgements Sutherland, D.N., 1989. Batch flotation behaviour of composite particles. Minerals
Engineering 2, 351–367.
Trahar, W.J., 1981. A rational interpretation of the role of particle size in flotation.
Funding from AMIRA P260E project is gratefully acknowledged,
International Journal of Mineral Processing 8, 289–327.
and the ToF-SIMS analysis carried out by Susana Brito E Abreu is Triffett, B., Veloo, C., Adair, B.J.I., Bradshaw, D., 2008. An investigation of the factors
appreciated. affecting the recovery of molybdenite in the kennecott utah copper bulk
flotation circuit. Minerals Engineering 21, 832–840.
Vianna, S., 2004. The Effect of Particle Size Collector Coverage and Liberation on the
References Floatability of Galena Particles in an Ore, PhD Thesis. Julius Kruttschnitt Mineral
Research Centre, The University of Queensland.
Agar, G.E., Stralton-Crawly, R., Bruce, T.J., 1986. Optimising the design of flotation Washburn, E.W., 1921. The dynamics of capillary flow. The Physical Review XVII (3),
circuits. CIM Bulletin 73, 173–181. 273–283.

You might also like