Sie sind auf Seite 1von 13

International Journal of Mineral Processing 98 (2011) 182–194

Contents lists available at ScienceDirect

International Journal of Mineral Processing


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / i j m i n p r o

Benchmarking flotation performance: Single minerals


S. Muganda a, M. Zanin a, S.R. Grano b,⁎
a
Ian Wark Research Institute, University of South 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: Chalcopyrite, conditioned with sodium dicresyl dithiophosphate (DTP), was floated under standard and well-
Received 26 January 2010 defined hydrodynamic conditions. The advancing contact angle values of the flotation feed and products were
Received in revised form 24 November 2010 measured and the flotation response benchmarked against a calibration previously established for the
Accepted 5 December 2010
chalcopyrite-amyl xanthate (KAX) system. Furthermore, the flotation response of pyrite, separately
Available online 10 December 2010
conditioned with KAX and DTP, was also evaluated under the same hydrodynamic conditions. When the
Keywords:
advancing contact angle of chalcopyrite conditioned with DTP was the same (within 5°) as that of chalcopyrite
Advancing contact angle conditioned with KAX, the flotation response was, within experimental error, the same. For both chalcopyrite
Floatability component and pyrite, heterogeneity of the advancing contact angle within the feed size fractions was demonstrated by
Critical contact angle significant differences in contact angle values measured on the flotation concentrate and tailings size
fractions. The mean contact angle of the chalcopyrite and pyrite particles remained constant, within
experimental error, through both flotation and sample preparation under the test conditions. The flotation
response of chalcopyrite at 2% solids (w/w) was the same, within experimental error, as that at 30% solids
(w/w) in the presence of silicate gangue, suggesting non-interaction of this gangue mineral with chalcopyrite
under the test conditions. The operational advancing contact angles inferred for pyrite using the calibration
established for the chalcopyrite-KAX system were, however, lower than the measured feed advancing contact
angles, while the maximum recovery of pyrite was also lower than for chalcopyrite for the same feed
advancing contact angle values, in the contact angle range less than 80°. The difference in flotation response
between these two minerals for the same feed contact angle values was interpreted in terms of a difference in
critical contact angle value for stable bubble–particle attachment. The critical advancing contact angle values
for the pyrite size fractions were higher than values for chalcopyrite for size fractions above 20 μm. This
difference in critical advancing contact angle was attributed to the difference in mineral specific gravity
between chalcopyrite and pyrite.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction cells, the froth effect is assumed negligible when the flotation process
is carried out at shallow froth depths (Vera et al., 1999).
The key steps in the collection of particles by bubbles in the pulp Hydrodynamic interactions are dominant in the collision efficiency
involve bubble–particle collision, attachment, and detachment (Dai while interfacial forces (e.g., particle hydrophobicity) play a major role
et al., 1998; Bloom and Heindel, 2003; Duan et al., 2003). The collection in the attachment efficiency. The stability of the bubble–particle
of particles is represented by the collection efficiency, Ecoll, which is a aggregate, a function of both hydrodynamics and interfacial forces
product of the efficiencies of the three sub-processes: against disruptive forces in the pulp, determines the stability efficiency
(Duan et al., 2003). An increase in the pulp density decreases the
turbulence and results in an increase in bubble size for solids content
Ecoll = Ec :Ea :Es ð1Þ
above 20% (Grau, 2006). A change in the bubble size and turbulence may
result in changes in the flotation rate constant. In situations where the
where Ec, Ea, and Es are the collision, attachment, and stability hydrodynamics is constant, interfacial forces determine the collection
efficiencies, respectively, of the bubbles and particles in the pulp efficiency of particles by bubbles.
(Derjaguin and Dukhin, 1961; Ralston et al., 1999b). A froth recovery Mineral surfaces are chemically heterogeneous so that the overall
factor, Ef, may also be incorporated into the above equation if the froth hydrophobicity and flotation response is controlled by the abundance
height is significant (Harris et al., 2002). In laboratory batch flotation of hydrophilic and hydrophobic surface patches (Fairthorne et al.,
1997; Smart et al., 1998; Hart et al., 2006). Mineral surfaces
⁎ Corresponding author. Tel.: +61 8 83130626. apparently have different collector adsorption densities leading to
E-mail address: stephen.grano@adelaide.edu.au (S.R. Grano). variations in surface hydrophobicity (Song et al., 2001). In flotation,

0301-7516/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.minpro.2010.12.001
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 183

this leads to different flotation rates (Drelich, 2001) and the same advancing contact angle (within a range no greater than 5°), the
occurrence of floatability components, even for particles within the flotation response of the chalcopyrite particles, both in the absence and
same size fraction and with the same density. Piantadosi (2001) found presence of amyl xanthate collector, was, within the determined
a large variation in adsorbed collector amount on particle surfaces experimental error of flotation, the same (Muganda et al., 2011). This
within the same size fraction confirming observations by Song et al. study led to the generation of master curves (i.e., calibration curves) of
(2001). Possible explanations given were a non-uniform distribution rate constant and collection efficiency against particle size fraction for
of active sites and their degree of activity on mineral particles different contact angle ranges, based on the chalcopyrite-amyl xanthate
(Piantadosi, 2001). The variation of adsorbed collector amount from system (Muganda et al., 2011). Empirical correlations between the
particle to particle may lead to different advancing contact angles on advancing contact angle and both the undistributed rate constant and
the surfaces of individual particles and a distribution of advancing collection efficiency for each size fraction were derived. These
contact angles in a powder sample (Brito et al., 2010). In this current correlations are used in this paper to benchmark flotation performance
study, distribution of contact angle values within a feed sample is to infer an operational contact angle for particles.
demonstrated by direct contact angle measurements on the flotation In this paper, the validity of the calibration curves is evaluated using
products (i.e., concentrate and tailings). chalcopyrite conditioned with a different collector (i.e., potassium amyl
The mean advancing contact angle of a chemically heterogeneous, xanthate, KAX) at 2% and 30% solids (w/w), and also pyrite conditioned
but smooth, surface may be approximated by the Cassie equation separately with KAX and DTP. The evaluation is carried out by
(Cassie and Baxter, 1944). In the Cassie equation, the functional groups benchmarking the flotation responses of chalcopyrite-DTP, pyrite-
or species on the mineral surface are not considered in the calculation, KAX, and pyrite-DTP systems against the chalcopyrite-KAX calibration.
but the contact angle of the domains is important in the determination The 30% solids pulp density is a typical plant pulp content and was used
of the overall hydrophobicity. It may therefore be expected that smooth, as a first step to establishing whether the calibration curves may be used
liberated particles with the same advancing contact angle have the same to benchmark the flotation response of particles in a natural ore. The
wettability and flotation behaviour, irrespective of the surface species. specific task of benchmarking a natural ore against the calibration is
When a particle collides with a bubble in the pulp zone, the liquid tackled in a separate paper (Muganda et al., submitted for publication).
separating the two must be displaced if attachment is to occur. The The overall objective of the study is to develop a method to infer an
induction time is the minimum time necessary for the attachment of an operational contact angle of particles based on the flotation process
air bubble to the mineral surface (Gu et al., 2003, 2004). The water film itself.
between a mineral surface and air bubble must thin to a critical value
(drainage) and rupture, forming the three-phase wetting perimeter, 2. Experimental
which then recedes before successful attachment can take place (Ralston
et al., 1999a). The rate of movement of the three-phase contact (TPC) line 2.1. Sample preparation
depends on the surface chemistry and roughness of the particle
(Krasowska et al., 2006). As noted before, surfaces of mineral particles Two sulphide minerals were used in this work, chalcopyrite and
are known to be heterogeneous in terms of chemistry (Piantadosi et al., pyrite, which have densities of 4.2 and 5.0 g/cm3, respectively. Both
2000), shape, and roughness (Sedev et al., 2004; Krasowska et al., 2009). minerals were supplied in lump form and were crushed to less than
These heterogeneities may result in differences in wetting behaviour 2.4 mm and separately blended. The chalcopyrite, as supplied, was
(Miller et al., 1996). Surface roughness and high hydrophobicity shorten significantly contaminated with quartz (8.5% SiO2 un-sized assay),
the induction time and attachment may take place in the first collision while the pyrite was relatively pure, with less than 1% SiO2
(Krasowska and Malysa, 2007; Najafi et al., 2008). Thus, surface contamination (Table 1). The single minerals (2 kg charges) were
roughness may assist in the thinning and rupturing of the intervening ground for 15 min in a stainless steel rod mill (10 stainless steel rods
liquid film, decreasing the attachment time and increasing the with a total weight of 9.3 kg), and 1 dm3 water at natural pH.
attachment efficiency. These findings demonstrate the role of surface The maximum size of the ground product targeted was 420 μm.
roughness in bubble–particle attachment, and why contact angles The ground product was screened with a 420 μm sieve to remove
measured on smooth surfaces do not necessarily correlate well with oversize material for regrinding. In the case of chalcopyrite, the
flotation recovery (Subrahmanyam et al., 1999). Particles of the same grinding product was immediately floated without collector in a pre-
mineral and size fraction prepared under the same conditions of grinding cleaning stage to reduce the amount of impurities, notably silica, to
may be assumed to have very similar physical properties in terms of very low levels. Polypropylene glycol (PPG425) was used as the
shape and roughness (i.e., a narrow spread in shape and roughness frother. A 5% aqueous solution of the PPG425 was prepared. One
heterogeneity). The expectation is that the induction time is the same for molar analytical grade sodium hydroxide was used to adjust the pH to
the same mean advancing contact angle values. 10. The rejection of silica at this stage relied on the natural
In a previous paper (Muganda et al., 2011), the effect of particle size hydrophobicity of the freshly prepared chalcopyrite particles. This
and contact angle on the flotation response of chalcopyrite under well- was important to ensure that multi-component flotation behaviour
defined and constant hydrodynamic conditions and at 2% solids (w/w) was not due to hydrophilic impurities such as quartz. Concentrates
was investigated. In that work, it was demonstrated that when from this stage were screened into nine size fractions, filtered, dried,
chalcopyrite particles of the same narrow size fraction exhibit the stored in sealed bottles, and used as the feed to subsequent

Table 1
Chemical assays (%) of pyrite size fractions.

Size fraction, μm −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Fe 47.7 46.9 45.5 45.8 46.3 47.0 47.1 45.5 46.0


Cu 0.1 0.11 0.23 0.18 0.11 0.12 0.1 0.12 0.09
S 48.6 50.1 52.3 52.8 51.9 51.1 51.3 53.0 52.6
Al2O3 0.9 0.9 0.3 0.1 0.1 b 0.09 b0.09 b 0.09 b0.09
SiO2 1.6 1.5 1.3 b 0.4 1.1 b 0.42 b0.42 b 0.4 b0.4
184 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

Table 2
Contact angle manipulation for chalcopyrite size fractions.

Size fraction, μm −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Test 1 a b b a a a a a b
2 c c c c c c c c c
3 c c c c c c a a a
4 a a a a a a c c c

a—Storage (6 months), then conditioned with 2 g/t DTP at pH 10 for 1 h; b—thermally oxidised (2 h), then conditioned with 2 g/t DTP at pH10 for 1 h; c—storage (6 months), then
conditioned with 20 g/t DTP at pH 10 for 1 h.

experiments. The chemical composition of the chalcopyrite size 2.3. Advancing contact angle measurement
fractions have been reported previously (Muganda et al., 2011).
In the case of pyrite, no cleaning pre-flotation stage was required, 2.3.1. Washburn technique
so the ground sample was wet sieved to separate the −20, 20–38, and The method chosen for the measurement of advancing contact
38–53 μm sizes. The +53 μm size sample was dried in an oven at angle on the powdered sample was the Washburn technique
50 °C, dry-sieved, and stored in containers as separate size fractions. (Washburn, 1921). The Washburn method makes use of capillary
The chemical assays of the pyrite size fractions suggested high purity pressure to drive a liquid at an observable rate through a packed bed
(N95%) of the pyrite sample. of particles in a capillary tube. The wetting velocity is then related to
the advancing contact angle. Details of the method used here are
given elsewhere (Muganda et al., 2008).
2.2. Advancing contact angle manipulation In this current paper, the advancing contact angles of both the
flotation feed and product size fractions were separately measured.
The first stage in the preparation of a flotation feed sample was the Chalcopyrite and pyrite size fractions were conditioned with collector
manipulation of the feed advancing contact angle and its measure- as explained previously to different advancing contact angles and
ment on each size fraction independently. Freshly prepared size floated at 2% solids (w/w), according to the procedure described
fractions of chalcopyrite were hydrophobic and needed to be oxidised further below. Flotation products were collected separately into three
to decrease the advancing contact angle to lower values that could samples to measure the advancing contact angle: the first concentrate
eventually be increased to desired values by collector addition. (collected in the first minute), the second concentrate (combined
Selected size fractions were thermally oxidised at 150 °C to achieve concentrates from 1–8 min of cumulative flotation time), and the
surface alterations. Other selected samples were used directly from tailing in the flotation cell after 8 min of flotation. The flotation
storage, the storage time being noted. After surface oxidation or products were wet sieved followed by drying of the size fractions at
storage, the advancing contact angle of each size fraction was ambient temperature, and the advancing contact angles were
increased by collector conditioning with dicresyl dithiophosphate measured. The advancing contact angles measured on the flotation
(DTP). A mass of 50 g of each size fraction was conditioned with products were used to recalculate the feed advancing contact angle
collector at pH10 for 1 h. The sample was then filtered and dried at using the mass recovery into each flotation product. Evidence that the
ambient temperature, and the advancing contact angles were sample preparation method and flotation procedure did not appre-
measured on 10 g of each size fraction sample. The remainder of the ciably alter the advancing contact angle values on the flotation
size fraction was then used in the flotation test, conducted products is provided by comparison of the original measured feed and
immediately after measurement of the feed contact angle. Details of recalculated feed advancing contact angle values.
the experimental conditions to achieve the contact angle manipula-
tion for each size fraction of chalcopyrite are given in Table 2. 2.3.2. Equilibrium capillary pressure technique
In the case of pyrite, each size fraction was activated separately in The Washburn technique was the principal method used in the
the presence of 500 g/t CuSO4.5H2O at pH10, followed by the addition measurement of advancing contact angles in this work. However, there
of collector to achieve different advancing contact angle values. A was a need to compare the advancing and receding contact angles of
mass of 50 g of each size fraction was conditioned with collector at chalcopyrite and pyrite at high and low advancing contact angle regimes.
pH10 for 1 h. The sample was then filtered and dried at ambient This determination provides information on the contact angle hysteresis
temperature, and the advancing contact angle was measured on 10 g for the two minerals. Both advancing and receding contact angle
of each size fraction sample. The remainder of the size fraction was measurements were carried out using the Equilibrium Capillary Pressure
then used in the flotation test, conducted immediately after (ECP) technique. Details of the method are given elsewhere (Dunstan and
measurement of the feed contact angle. Details of the experimental White, 1986; Diggins and Ralston, 1993; Stevens, 2005).
conditions to achieve the contact angle manipulation for each size In the case of pyrite, the 75–105 μm fraction was copper-activated in
fraction of pyrite are given in Table 3. Potassium amyl xanthate (KAX) the presence of 500 g/t CuSO4.5H2O at pH 10. For each mineral, two
and sodium dicresyl dithiophosphate (DTP) collectors were used to samples (100 g each) were prepared, one with a high advancing contact
increase the contact angle as indicated in Table 3. angle and the other with a lower advancing contact angle. This was

Table 3
Contact angle manipulation for copper activated pyrite using KAX and DTP.

Size fraction, −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300
μm

Test 5 a a a a a a a a a
6 b b b b b b b b b
7 c c c c c c c c c

a—Conditioned with 50 g/t KAX at pH10 for 1 h, b—conditioned with 300 g/t KAX at pH10 for 1 h, c—conditioned with 30 g/t DTP at pH10 for 1 h.
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 185

Table 4
Size distribution of flotation feed (mass, %), and recalculated distribution by addition of the flotation products.

Particle size, μm −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Weight%, ±0.5 15 10 15 10 10 10 15 10 5
Typical* recalculated feed, weight%, ± 0.5 14.6 10.1 15.3 10 10 10.3 14.8 10.1 4.8
Nominal size range Fine Intermediate Coarse Very coarse

*Calculated from flotation test products.

achieved by addition of 200 g/t and 80 g/t KAX at pH10 to the 75–105 μm 2.5. Bubble size measurement
size fraction of each mineral separately for 1 h, respectively. After collector
conditioning, the samples were filtered, and dried in air overnight at The bubble size in flotation was measured using the photographic
ambient temperatures. The advancing and receding contact angle method (Chen et al., 2001; Hernandez-Aguilar et al., 2005). A CCD camera
measurements were carried out using the 75–105 μm size fraction of was used to capture images of bubbles in a viewing chamber as they rose
the feed for each mineral after drying at ambient temperature. from the pulp zone through a tube. Bubble size measurements with MIBC
were carried out in a pulp made of cleaned, hydrophilic quartz, at 30%
2.4. Flotation tests solids (w/w). Images captured on the digital camera were processed
using Image J software. The Sauter mean bubble diameter was calculated
After the measurement of advancing contact angles on all feed size (Gorain et al., 1997).
fractions, a flotation feed sample (100 g) was constituted (Table 4) and
the sample was tested in flotation within two h. This particle size 2.6. Data analysis
distribution (Table 4) was chosen as a typical size distribution in sulphide
copper flotation. 2.6.1. Recovery by entrainment
Each test condition was repeated at least twice, and the corresponding Particles are recovered via two primary mechanisms (1) by true
concentrates at each time interval were combined for particle sizing. The flotation in which hydrophobic particles are recovered by attachment to
conditioning time at pH 10 was 1 min before and after addition of frother, air bubbles and (2) by entrainment, which involves the non-selective
giving a total of 2 min before flotation air was introduced. Polypropylene recovery of particles in water recovered to the concentrate irrespective of
glycol (PPG425) was used as frother. Concentrates were collected at 1, 3, whether they are hydrophobic or hydrophilic (Harris et al., 2002).
5, and 8 min cumulative flotation time with a scraping rate of one every Recovery by entrainment is subtracted from the total recovery to obtain
10 s. The following conditions were constant in all tests, unless otherwise recovery by bubble–particle attachment (Savassi et al., 1998).
stated: (1) superficial gas velocity, Jg 0.3 cm/s, or gas flow rate of 7.2 l/min, Recovery by entrainment in a single batch flotation test was estimated
(2) impeller rotational speed 1200 rpm, (3) solids concentration 2%, (4) using the method by Ross (1991). The rate of recovery (g/min) of
frother (PPG425) addition 10 ppm at the beginning of the flotation entrained particles in the ith size fraction, Ei(t), at time t during flotation is
sequence and another 10 ppm addition after collection of the second (Ross, 1991):
concentrate. All concentrates and tailing samples were separated into size
fractions (the same as Table 4) to extract size-by-size recovery data. A W ðt Þ⋅Cmi ðt Þ
Ei ðt Þ = Xi ðt Þ⋅ ð2Þ
bottom driven 5 dm3 flotation cell (Runge et al., 2003) was used to carry Cw ðt Þ
out the flotation tests. The cell impeller diameter was 10.5 cm and the
flotation cell had dimensions of 20×20×13.5 cm for the length, width, where W(t) is the rate of recovery of water (g/min), Cmi(t) and Cw(t) are
and height, respectively. the concentrate mass of solids (g/l) and water at time, t. Xi(t) is a transfer
factor, which is particle size-dependent, and corresponds to the degree of
entrainment (Lynch et al., 1981; Warren, 1985). The method takes into
2.4.1. Chalcopyrite flotation at 30% solids
account that the pulp in batch flotation is continuously depleted of
Chalcopyrite size fractions were conditioned individually with DTP at
particles due to entrainment and true flotation. Recovery by entrainment
pH10 (Tests 3 and 4 in Table 2). Advancing contact angles of the
was subtracted from the size-by-size recovery data to determine the
chalcopyrite size fractions were measured and a flotation feed sample of
recovery by true flotation.
100 g was constituted as before. The feed sample (100 g) with the same
size distribution as in the single mineral test work (Table 4) was mixed
2.6.2. Recovery, rate constant, and collection efficiency
with hydrophilic quartz (1900 g) to obtain the same feed size distribution
For each particle size fraction, the rate of particle recovery, after
(Table 5) as a natural ore used in other tests (Muganda et al., submitted for
subtraction of entrainment, was calculated assuming a single floatable
publication).
component and a non-floating component. The size-by-size recovery
The frother used was MIBC (Methyl Iso-Butyl Carbinol) at 38 g/t
data, after subtraction of entrainment, was fitted to the first-order rate
(60 μl). PPG425 was used in tests at 2% solids (i.e., in the presence of
equation:
chalcopyrite only) to achieve a stable froth, while MIBC was used in
tests at 30% solids (w/w). Sodium hydroxide was used, instead of lime,  
−kt
to raise the pulp pH to 10, and no collector was added to the pulp Rðt Þ = R max 1−e ð3Þ
because the chalcopyrite size fractions had already been manipulated
to known advancing contact angles prior to the constitution of the using a least squares method, where R(t) and Rmax are the recovery at
feed and its flotation. Each test was repeated three times to generate time t and maximum recovery at infinite time, respectively, for the
enough concentrate mass for sizing and assay. size fraction, and k is the distributed rate constant, which is the rate

Table 5
Feed size distribution for 2 kg of chalcopyrite/quartz mixture.

Particle size, μm −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Weight%, ±0.5 15.8 6.9 5.5 7.7 10.3 11.5 14.7 12.8 14.8
186 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

constant of the single floatable species. A modified flotation rate 3.2.1. Heterogeneity and conservation of the advancing contact angle
constant that takes into account both the maximum recovery and rate The approach adopted in this study is to assume two floatability
constant has been proposed (Agar et al., 1986; Sripriya et al., 2003) components in the feed for each size fraction, a floating, and a non-
and is the rate constant for all floatability components in the size floating component. The advancing contact angles of flotation
fraction. In this paper, the modified flotation rate constant is called the products were measured to ascertain the basis of the separation and
undistributed rate constant, k*, and is calculated using: establish if there was conservation of advancing contact angle values.
Results of the advancing contact angle measurements carried out on
kR max the original feed and the flotation products demonstrate that the
k⁎ = ð4Þ
100 advancing contact angle of particles in the feed remains largely
unchanged during both flotation and sample preparation under the
The collection efficiency, Ecoll, was calculated using the equation experimental conditions (Table 6). The recalculated feed advancing
relating the rate constant (k or k*), the bubble surface area flux (Sb) contact angle is a weighted average of the advancing contact angle of
and collection efficiency as follows (Jameson et al., 1997): the flotation products using the respective mass fractions (Fig. 1). The
recalculated and measured feed advancing contact angle values are,
3Jg 1 within the experimental error of the contact angle measurement
k= E = Sb Ecoll ð5Þ of ± 2.5° (Muganda et al., 2011), the same for virtually all size
2d32 coll 4

The bubble surface area flux is a measure of the bubble surface area
rising up through the cell per unit cross-sectional area and is given by
(Finch and Dobby, 1990):

6Jg
Sb = ð6Þ
d32

where Sb = bubble surface area flux, Jg = superficial gas velocity,


d32 = Sauter mean bubble diameter.

3. Results

3.1. Introduction

In previous work, flotation characterisation of the chalcopyrite-


amyl xanthate system resulted in the development of calibration
curves of the undistributed rate constant and collection efficiency
versus particle size fraction for different advancing contact angle
ranges (Muganda et al., 2011). Modelling of the maximum recovery
was carried out, which resulted in the calculation of the critical
advancing contact angle for stable bubble–particle attachment for
each size fraction. The applicability of these calibration curves is now
evaluated against the chalcopyrite-dicresyl dithiophosphate (DTP)
system at 2% and 30% (w/w) solids in the pulp, and the pyrite-
xanthate and pyrite-DTP systems at 2% (w/w) solids.

3.2. Flotation response of the chalcopyrite-DTP system

In this section, the flotation response of the chalcopyrite-DTP


system is compared with the calibration curves developed previously
(Muganda et al., 2011). For each size fraction, the flotation response of
high and low advancing contact angle regimes was probed, coupled
with advancing contact angle measurements on the flotation
products. The recovery–time profiles (Fig. 1) for chalcopyrite size
fractions conditioned with DTP to different contact angles (Table 6)
show that the maximum recovery is a function of both particle size
and advancing contact angle. In the two tests, low and high contact
angle regimes for each size fraction were probed. Apparently, the
recovery plateaus at longer flotation times for almost all size fractions
and feed contact angles such that a two-component model of floatable
and non-floatable particles is reasonably assumed. While only a single
(mean) contact angle is measured in the feed, there exist under most
conditions at least two floatability components (i.e., Rmax significantly
less than 100%). The non-floatable component, which reports to the
tailing, has a significantly lower advancing contact angle than the
floatable component in the concentrates, based on direct contact Fig. 1. Cumulative recovery with time for chalcopyrite size fractions conditioned with
angle measurements of the products (Table 6). Implications of this DTP, (a) Test 1 (b) Test 2. The continuous lines are for Eq. (3), after subtraction of
finding are discussed further below. recovery by entrainment.
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 187

Table 6
Contact angle values of flotation products and the measured and recalculated advancing contact angles (°) for DTP conditioned chalcopyrite before and after flotation, and the
corresponding maximum recovery and undistributed rate constants. Low (Test 1) and high (Test 2) advancing contact angle regimes.

Size fraction, μm −20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Test 1 First concentrate (°) 77 62 75 78 68 68 56 77 79


Second concentrate (°) 70 59 73 70 61 64 54 71 75
Tail (°) 67 21 26 39 46 50 48 61 42
Recalculated feed (°) 72 42 67 70 59 62 51 69 55
Measured feed (°) 72 46 66 70 58 66 50 70 56
Rmax (%) 58 53 84 81 60 70 48 58 37
k* (min−1) 0.86 0.49 1.44 2.68 1.13 1.86 0.22 0.61 0.16
Test 2 First concentrate (°) 77 86 80 83 88 89 81 86 90
Second concentrate (°) 74 82 68 80 85 86 76 82 89
Tail (°) 66 46 45 44 44 51 57 63 76
Recalculated feed (°) 73 83 77 81 87 86 73 77 86
Measured feed (°) 75 90 76 84 86 86 70 72 86
Rmax (%) 65 92 94 96 98 92 69 62 71
k* (min−1) 0.90 2.30 1.96 4.49 4.99 4.12 1.26 0.92 1.10

fractions and for both low and high advancing contact angle regimes apparently different surface functional groups exhibit the same
(Test 1 and Test 2 in Table 6). flotation behaviour, within experimental error. This result implies
Furthermore, the concentrates collected in the first minute (first that the efficiency of the sub-processes of attachment and detachment
concentrate) have consistently higher advancing contact angle values for chalcopyrite are the same when the advancing contact angle falls
than the second concentrate. The tailing, which is solely the non- within the same range (±2.5°).
floating component (Fig. 1), has lower advancing contact angle than The critical advancing contact angle values by size fraction
both concentrates, showing that the separation of particles is contact determined for the chalcopyrite-amyl xanthate system (Muganda
angle-dependent. This, of course, is expected given the basis of et al., 2011) are now compared with the measured advancing contact
separation in flotation. The results of the advancing contact angle angles of the flotation products of the chalcopyrite-DTP system
measurements on the flotation products further confirm that (Fig. 3). The advancing contact angle values of the concentrates are
individual particles within individual size fractions of the feed achieve above the critical advancing contact angle for both high and low
different advancing contact angles when conditioned with collector advancing contact angle regimes, while the tailing advancing contact
under the same conditions. The heterogeneity of the advancing angle values are at, or below, the critical value determined by
contact angle is probably due to differences in the abundance of active Muganda et al. (2011). The tailing advancing contact angle value in
sites on the mineral surface (Piantadosi et al., 2000), at least. the intermediate (20–53 μm) and very coarse (+ 300 μm) particle size
Differences in contact angle of particles separated by flotation have
been demonstrated for the first time in this study, and it is anticipated
that future studies relating particle size, flotation rate constants, and
contact angle may lead to the determination of contact angle
distribution within a given size fraction depending on flotation
behaviour. This may represent a leap forward in the development of
diagnostic tools for process control in mineral processing.

3.2.2. Benchmarking flotation response


The undistributed rate constant for each size fraction in Tests 1 and
2 (Table 6) is clearly dependent on the advancing contact angle
measured on the feed. To benchmark the flotation response of the
chalcopyrite-DTP system against the chalcopyrite-amyl xanthate
system, empirical equations relating the advancing contact angle to
the undistributed rate constant determined for the chalcopyrite-amyl
xanthate system were used to calculate an inferred operational
advancing contact angle value. The inferred operational advancing
contact angle is given by:

 b
θadv = a k ð7Þ

where a and b are constants and dependent on particle size fraction


(Muganda et al., 2011).
The measured feed and inferred operational advancing contact
angles, based on the undistributed rate constant, are in good
agreement for both the low and high advancing contact angle
regimes, with R2 values of 0.94 and 0.95, respectively (Fig. 2). The
result further confirms that the flotation behaviour of DTP condi-
tioned chalcopyrite is the same, within experimental error, as amyl
xanthate conditioned chalcopyrite, provided that they have the same
advancing contact angle (within 5° of each other). Liberated Fig. 2. Inferred operational advancing contact angle compared with the measured feed
chalcopyrite particles exhibiting the same advancing contact angle, advancing contact angle of DTP conditioned chalcopyrite size fractions with (a) low,
within experimental error of the Washburn method, but with R2 = 0.94 (Test 1), (b) high advancing contact angle (Test 2) regimes, R2 = 0.95.
188 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

Fig. 3. Advancing contact angles of concentrates and tails for DTP conditioned
chalcopyrite with (a) low (2 g/t DTP, Test 1) and (b) high (20 g/t DTP, Test 2)
advancing contact angles compared with the critical advancing contact angle curve
(chalcopyrite-amyl xanthate system) (Muganda et al., 2010).

fractions for the low advancing contact angle regime (Fig. 3a) have Fig. 4. Operational advancing contact angle compared with measured feed advancing
contact angle for chalcopyrite feed size fractions floated at 30% solids (w/w), (a) Test 3,
significantly lower advancing contact angles than the critical value, (b) Test 4. R2 = 0.98 in both tests.
due to the fact that these size fractions were intentionally thermally
oxidised (Table 2) to achieve low contact angles in the feed (Table 6).
For the high advancing contact angle regime (20 g/t DTP in Test 2) specific gravity, as discussed further below. It could be concluded that
the advancing contact angles of the tailing size fractions are, within the critical advancing contact angle and the flotation response may be
experimental error, the same as the critical advancing contact angle independent of the hydrophobisation route for a specific mineral.
values determined for the chalcopyrite-amyl xanthate system Given a measured feed mean advancing contact angle and assuming a
(Muganda et al., 2011). The higher collector addition in Test 2 is distribution of advancing contact angle values about the mean
likely to give a narrower distribution of advancing contact angle (Brito (Muganda et al., 2011), then it may be possible to predict the
et al., 2010), so that the particles that do not float are at, or close to, the maximum recovery (i.e., floatable components) of the mineral.
critical value required for stable bubble–particle attachment. There is
remarkable agreement between the chalcopyrite-amyl xanthate and 3.3. Tests at higher solids percent
chalcopyrite-DTP system. Flotation clearly separates on the basis of
advancing contact angle; particles with advancing contact angle lower 3.3.1. Flotation response at high pulp density
than the critical value report to the tailing stream as the non-floatable Flotation of ores is normally carried out at higher pulp density than
component. that used in the study of the chalcopyrite-amyl xanthate (Muganda
It appears that the critical advancing contact angles calculated et al., 2011) and chalcopyrite-DTP single mineral systems discussed thus
using the chalcopyrite-amyl xanthate system (Muganda et al., 2011) far. This disparity of pulp density (i.e., %solids w/w) warranted further
are valid for the chalcopyrite-DTP system, possibly implying that the examination. The ultimate aim is to provide a useful benchmark against
critical advancing contact angle may be independent of the hydro- which ore flotation behaviour can be compared. The effect of pulp
phobising species at the molecular scale. The critical advancing density on the flotation response was probed using chalcopyrite
contact angle is a function of the particle size, and possibly mineral particles preconditioned with DTP (Table 2, Tests 3 and 4) for which

Table 7
Flotation response of chalcopyrite with quartz at 30% solids (w/w). Chalcopyrite conditioned with DTP (2 and 20 g/t) prior to flotation. Size fractions conditioned individually and
floated together.

Particle size fraction, μm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 + 300

Test 3 Feed advancing contact angle,° 84 82 79 89 86 85 60 64 66


Rmax, % 84 91 97 96 89 85 59 41 35
Ecoll 0.14 0.21 0.32 0.55 0.55 0.42 0.08 0.04 0.04
Test 4 Feed advancing contact angle,° 65 58 65 63 55 59 83 80 90
Rmax, % 51 59 67 72 63 52 90 86 78
Ecoll 0.10 0.10 0.17 0.17 0.12 0.14 0.26 0.18 0.18
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 189

the individual contact angle values were measured, and then mixed that, within the experimental error of flotation, there is very good
with clean, hydrophilic quartz. Two tests were carried out so that the agreement between the measured feed and inferred operational
flotation behaviour of individual size fractions, at both low and high advancing contact angles (Fig. 4), with an R2 value of 0.98 for both
advancing contact angle regime, was probed (Table 7). As in previous Tests 3 and 4. Changes in energy dissipation (lower turbulence) at
tests, there is clearly a floating and non-floating component in each size higher solids percent may benefit the stability efficiency, while also
fraction. The higher the contact angle measured on the feed, the lower reducing the collision frequency due to reduced energy dissipation
the non-floating component (100–Rmax). The feed advancing contact and increased bubble size. The net effect of these subtle and opposing
angle values measured on each of the size fractions, before mixing with influences seems to be minimal in any case. The conclusion that the
quartz, in the two tests are shown in Table 7. percent solids (up to 30% w/w) do not affect the calibration may be
valid in the specific case of a chemically inert and non-interacting
3.3.2. Benchmarking against the calibration gangue component such as quartz at pH 10. Chalcopyrite and pyrite
Changes in pulp density may affect bubble size (Grau, 2006) and, particles may not interact with quartz in the circumstances because
therefore, the bubble–particle collision efficiency. The level of both are negatively charged under alkaline conditions and changes in
turbulent energy dissipation is also lower at higher solids percent, pulp viscosity may be minimal over this range. The conclusion has also
and both the bubble–particle attachment and detachment processes been tested in another paper in which a natural ore with chalcopyrite
may be affected (Pyke et al., 2003) to an extent. as the main copper mineral and poly-component gangue mineralogy
The Sauter mean bubble diameter was determined to be 0.57 mm was examined (Muganda et al., submitted for publication).
at high (30% w/w) solids percent and was constant through flotation.
At low (2% w/w) solids concentration, the Sauter mean bubble
diameter was 0.48 mm and was also constant through flotation. This 3.4. Flotation of pyrite
difference in bubble size was incorporated into the calibration by
considering the collection efficiency (Ecoll), rather than simply the The calibration curves developed previously on the chalcopyrite-
undistributed rate constant (k*). To benchmark the flotation response amyl xanthate system (Muganda et al., 2011) were used to
of the chalcopyrite at high solids percent, empirical equations relating benchmark the flotation response of pyrite under the same hydrody-
the advancing contact angle to the collection efficiency (based on the namic conditions. The pyrite-amyl xanthate and pyrite-DTP systems
undistributed rate constant) at 2% solids were used to calculate an were investigated to probe the effect of collector species and mineral
inferred operational advancing contact angle in the feed. The inferred type (e.g., specific gravity) on flotation response. For each size
operational advancing contact angle was calculated by Muganda et al. fraction, high and low advancing contact angle regimes with copper
(2011): activated pyrite conditioned with potassium amyl xanthate (KAX)
were probed (Table 8, Tests 5 and 6), while the flotation response of
d
pyrite conditioned with dicresyl dithiophosphate (DTP) was tested at
θadv = cðEcoll Þ ð8Þ a low advancing contact angle regime only (Table 8, Test 7).
Apparently, only low advancing contact angle values could be
where c and d are constants and dependent on particle size fraction. obtained with the pyrite-DTP system, which would be a result of
An inferred operational advancing contact angle was then differences in the adsorption mechanism between copper sulphide
calculated for each size fraction in Tests 3 and 4. The results show and pyrite when conditioned with DTP (Fuerstenau et al., 1971). It is

Table 8
Advancing contact angles of pyrite flotation feed and products, and corresponding maximum recovery and undistributed rate constants. Corresponding Rmax values for chalcopyrite
size fractions within the same advancing contact angle range as in Tests 5–7, obtained from previous work (Muganda et al., 2010).

Particle size, μm 20 20–38 38–53 53–75 75–105 105–150 150–210 210–300 300+

Test 5 (KAX) First concentrate (°), ±2 80 66 69 73 63 70 82 75 86


Second concentrate (°), ±2 75 61 64 63 60 66 69 70 80
Tail (°), ± 2 48 41 26 38 35 42 46 43 58
Recalculated feed (°), ± 2 advancing contact angle 50 51 34 53 39 47 58 45 66
Measured feed (°), ±2 51 54 38 56 43 50 60 49 69
Rmax, %, ± 3 2 39 20 47 16 19 35 9 33
k*, min−1, ± 0.2 0.03 0.6 0.17 0.51 0.09 0.23 0.49 0.05 0.2
Test 6 (KAX) First concentrate (°), ±2 81 81 77 79 75 83 90 90 90
Second concentrate (°), ±2 77 73 68 71 66 73 81 81 84
Tail (°), ± 2 66 50 47 48 48 54 60 60 65
Recalculated feed (°), ± 2 advancing contact angle 76 75 72 73 67 77 87 84 85
Measured feed (°), ±2 76 78 73 74 70 78 88 85 90
Rmax, %, ± 3 71 85 89 81 74 81 91 84 85
k*, min−1, ± 0.2 0.6 1.51 1.27 2.61 1.77 2.29 2.2 1.96 0.95
Test 7 (DTP) First concentrate (°), ±2 80 60 54 62 66 68 72 76 86
Second concentrate (°), ±2 75 56 52 59 58 64 66 71 79
Tail (°), ± 2 49 45 42 24 22 46 41 46 54
Recalculated feed (°), ± 2 advancing contact angle 53 52 46 34 29 49 46 49 58
Measured feed (°), ±2 58 56 50 38 34 54 50 53 61
Rmax, %, ± 3 11 52 36 28 18 15 19 9 15
k*, min−1, ±0.2 0.11 0.26 0.15 0.13 0.07 0.35 0.10 0.06 0.09

Rmax for chalcopyrite size fractions with the same corresponding contact angle range as in Tests 5–7, respectively
Feed contact angle range (°) 51–55 51–55 36–40 56–60 41–45 46–50 56–60 46–50 66–70
Rmax, %, ± 3 18 64 30 64 45 48 55 25 46
Feed contact angle range (°) 76–80 76–80 71–75 71–75 66–70 76–80 86–90 81–85 86–90
Rmax, %, ± 3 83 86 85 98 84 93 95 88 74
Feed contact angle range (°) 56–60 56–60 46–50 36–40 b 36 51–55 46–50 51–55 61–65
Rmax, %, ± 3 20 65 65 40 30 51 45 37 41
190 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

the advancing contact angle measured on the feed. This finding


further supports the conclusion that the advancing contact angle of
particles present in the feed remains largely unchanged through both
flotation and sample preparation under the experimental conditions.
The advancing contact angle values of the flotation products were
measured, and, as observed in the case of the chalcopyrite-DTP system,
the first concentrate was more hydrophobic than the second concentrate.
The tailing, which is dominated by the non-floatable component, had
lower advancing contact angles than either concentrates. Differences in
advancing contact angles of the flotation products demonstrate
conclusively that there is distribution of advancing contact angles within
each size fraction of the feed, as observed in the case of the chalcopyrite-
DTP system. Advancing contact angle heterogeneity may therefore be a
characteristic of all mineral ensembles such that when an advancing
contact angle is measured on a bed of particles, it is actually the surface
area weighted mean of a statistical distribution.
There appears to be a difference in the recovery–time profile for
pyrite as compared to chalcopyrite. The recovery–time curve plateaus at
shorter flotation times for chalcopyrite (Fig. 1) than pyrite (Fig. 5). This
phenomenon is more apparent when the pyrite feed contact angle is low
(Fig. 5c). This suggests that the rate of pyrite flotation is lower than
chalcopyrite for the same feed advancing contact angle, an issue that is
discussed further below.
Comparison was also made between the maximum recovery of
chalcopyrite and pyrite for each size fraction exhibiting the same contact
angle. The maximum recovery of chalcopyrite for each size fraction and
feed contact angle measured on pyrite (Table 8, Tests 5–7) was obtained
from previous work, which generated maximum recovery data for each
size fraction and contact angle range (Muganda et al., 2011). The
maximum recovery for pyrite of a given size fraction and advancing
contact angle is less than that for chalcopyrite, for advancing contact
angle values less than 80° (Table 8, Fig. 6). Considering that the flotation
tests were carried out under the same hydrodynamic conditions,
possible reasons for the apparent discrepancy between chalcopyrite and
pyrite are (i) the pyrite size fractions have a wider distribution of
advancing contact angle values about the mean, such that a greater
proportion of pyrite particles are below the critical contact angle value
required for stable bubble–particle attachment, and/or (ii) the pyrite
size fractions have higher critical advancing contact angle values than
chalcopyrite. This may explain why pyrite has lower recovery than
chalcopyrite for similar advancing contact angles within a specific size
fraction. These possibilities are discussed further below.
The inference of an operational advancing contact angle based on the
undistributed rate constant may provide further insight into the
flotation behaviour of pyrite compared to chalcopyrite. Empirical
equations (Eq. (7)) derived from the chalcopyrite-amyl xanthate system
(Muganda et al., 2011) were used to calculate an operational advancing

Fig. 5. Cumulative recovery with time for pyrite size fractions (a) KAX, Test 5—low
contact angle regime, (b) KAX, Test 6—high contact angle regime, (c) DTP, Test 7. The
continuous lines are for Eq. (3), after subtraction of recovery by entrainment.

important to note firstly that the recalculated feed advancing contact Fig. 6. Comparison of maximum recoveries (%) for chalcopyrite and pyrite for each size
angle values in Table 8, based on the weighted measurements on the fraction. The values are for advancing feed contact angles for pyrite and chalcopyrite in
flotation products, is the same, within experimental error (±2.5°), of Table 7.
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 191

contact angle for pyrite from the undistributed rate constant values. Table 9
Generally, the inferred operational advancing contact angles are Advancing and receding contact angles of feed chalcopyrite and pyrite conditioned with
KAX, 75–105 μm size fractions.
lower than the measured feed advancing contact angles (Fig. 7) in all
tests. θadv, ECP θrec, ECP Washburn θadv, ± 2.5°
Lower inferred operational advancing contact angle implies that the Chalcopyrite 65 ± 8 61 ± 3 57
calibration curves generated with the chalcopyrite-amyl xanthate 77 ± 6 75 ± 7 82
system (Muganda et al., 2011) may not hold for pyrite due to mineral Pyrite 58 ± 6 55 ± 5 65
85 ± 3 83 ± 5 87
specific factors such as specific gravity, shape, and roughness, etc., or due
to a wider distribution of contact angle values about the mean value. The θadv, θrec are the advancing and receding contact angles, respectively.
roughness of the mineral surface affects its wetting characteristics
(Sedev et al., 2004; Krasowska and Malysa, 2007). In this study, the
advancing contact angle was generally used to characterize hydropho- subtracting the receding from the advancing contact angle. There does
bicity. However, for successful bubble–particle attachment, the receding not appear to be significant difference in contact angle hysteresis for
contact angle may play a significant role in determining the rate of both minerals, suggesting that the wetting and dewetting behaviour are
intervening liquid film thinning, rupturing, and receding. This particular similar, based on advancing contact angle data for the 75–105 μm size
issue is tackled in the following section. fraction (Table 9). This result appears to rule out the possibility that
differences in flotation behaviour between the two minerals are due to
3.4.1. Wetting and dewetting behaviour of chalcopyrite and pyrite differences in receding contact angle. Further, this exercise provided
It was decided to compare the wetting and dewetting behaviour of opportunity for comparison of the advancing contact angle values
chalcopyrite and pyrite size fractions exhibiting the same advancing measured by the two common methods for powdered particles. There is
contact angle, within experimental error, of the Washburn technique general agreement in the advancing contact angles measured, given the
(±2.5°). The high and low advancing contact angle regimes were relatively large error in contact angle measurement for the ECP method.
probed for contact angle hysteresis of each mineral, obtained by Copper activated pyrite conditioned with DTP did not achieve high
advancing contact angles. The maximum advancing contact angles
achieved were in the low to medium range (50–60°). The measured
feed and inferred operational advancing contact angles show scatter,
especially in the intermediate size range (Fig. 7c). Further discussion
on the benchmarking of pyrite flotation response against chalcopyrite
follows below.

3.4.2. Discussion
From the literature (Brito et al., 2010), individual particle advancing
contact angles in a size fraction may be statistically distributed around a
mean. In previous work (Muganda et al., 2011), a gamma (Γ) statistical
distribution of individual particle advancing contact angles around the
measured mean was assumed following the work of Brito E Abreu et. al.
(2010). Brito E Abreu et al. (2010) used the same particles as those used
in the current study of the chalcopyrite-DTP system. The model
developed for chalcopyrite can predict the maximum recovery of any
size fraction when the feed advancing contact angle is known, assuming
a particular statistical distribution of advancing contact angle values.
The statistical distribution of advancing contact angles around the mean
value may be particle size-dependent and/or mineral-specific. Two
assumptions, in the consideration of pyrite, are considered in this
section: (1) pyrite size fractions have the same critical advancing

Fig. 7. Operational advancing contact angle compared with feed advancing contact angle Fig. 8. Variation of the mean and mode advancing contact angle values for a sample,
for (a) pyrite-xanthate system with low advancing contact angle—Test 5, (b) pyrite- assuming gamma statistical distribution of individual particle contact angles (Muganda
xanthate system with high advancing contact angle—Test 6, (c) pyrite-DTP system with et al., 2010). The mean is the value measured by the Washburn technique. The mean and
low advancing contact angle—Test 7. mode contact angle values determine the contact angle distribution in a feed sample.
192 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

have different critical advancing contact angle values than chalcopyrite


size fractions, but the same statistical distribution (i.e., the same
relationship between the measured mean value and the mode of the
contact angle distribution in Fig. 8) can be used to fit the flotation data
for both minerals.
Pyrite flotation behaviour was compared to that of chalcopyrite
assuming a gamma statistical distribution of contact angle within a
size fraction and fitting the pyrite flotation data using the same
parameters determined previously (i.e., assumption 1). For the feed
mean advancing contact angles shown in Table 8, a mode was
determined using Fig. 8, based on relationships derived from
chalcopyrite-amyl xanthate system (Muganda et al., 2011). The
procedure for the model fit has been dealt with in detail in a previous
publication (Muganda et al., 2011).
When assumption 1 was applied, the model fit to the pyrite
flotation data overestimated the maximum recovery for the mean
contact angles measured using the Washburn method (Fig. 9a), with
an R2 value of only 0.84. This finding may suggest that either pyrite
has a wider distribution of advancing contact angles around the mean
value than chalcopyrite (i.e., Fig. 8 does not hold for pyrite) and/or
pyrite has a higher critical contact angle possibly due to specific
gravity differences between the two minerals. The latter possibility
seems more likely given that chalcopyrite and pyrite particles exhibit
similar contact angle hysteresis for the same mean advancing contact
angle (Table 9), suggesting the same statistical distribution of contact
angle values around the mean.
Fig. 9. Experimental versus model maximum recovery for pyrite flotation, 2% solids.
The possibility of similar advancing contact angle distributions in the
(a) Assumption 1, critical advancing contact angle for pyrite is the same as for feed for chalcopyrite and pyrite size fractions was evaluated by
chalcopyrite, R2 = 0.84, and (b) Assumption 2, pyrite has a different critical advancing comparing the cumulative masses of tail, second and first concentrates
contact angle than chalcopyrite, R2 = 0.97. (Tables 6 and 8), in that order, and the corresponding cumulative mean
contact angle for specific size fractions of the two minerals exhibiting
the same mean contact angle in the feed, within the experimental error
contact angles as those calculated for chalcopyrite, and the distribution of the advancing contact angle measurement (±2.5°). There is close
of individual particle advancing contact angle around the mean is the agreement between the recalculated cumulative mass and cumulative
same irrespective of particle size fraction; (2) pyrite size fractions may mean advancing contact angle relationships for the two different

Fig. 10. Cumulative mass recovered into the tail, second, and third concentrate (total = recalculated feed) against the cumulative mean contact angle for pyrite (O) and chalcopyrite
(□) for (a) −20 (Tests 2 and 6), (b) 38–53 (Tests 2 and 6), (c) 53–75 (Tests 1 and 6), (d) + 300 (Tests 2 and 6) μm size fractions from Tables 5 and 7. Tests compared with similar
mean contact angles values in the feed.
S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194 193

minerals (Fig. 10), providing demonstration that the same statistical


function may be applied to pyrite flotation data as for the chalcopyrite
(i.e., Fig. 8 applies for both minerals).
Fitting to the pyrite flotation data under assumption 2 was
attempted by allowing the critical contact angle to vary to obtain the
best fit. A least squares method that minimised the difference between
the model and experimental Rmax values was used. An example of the
fitting procedure for the 53–75 and +300 μm size fractions is shown in
Table 10. The model was successful in predicting pyrite recovery
(Fig. 9b) with an R2 value of 0.97 across all size fractions.
These results appear to support the assumption that the advancing
contact angles are statistically distributed around the measured mean
contact angle in the feed in the same way for both pyrite and
chalcopyrite and that the pyrite size fractions generally have a higher
critical advancing contact angles than chalcopyrite. Thus, the lower
recovery of pyrite compared with chalcopyrite, for the same feed
advancing contact angle (Fig. 6), in the range less than 80°, may be Fig. 11. Variation of the critical advancing contact angle with particle size for pyrite and
explained by the higher critical contact angle for stable bubble–particle chalcopyrite.
attachment in the case of pyrite. Pyrite has a higher specific gravity and
therefore lower stability efficiency than chalcopyrite. Evidently, this
difference in critical advancing contact angle may not be driven by tackled in another publication (Muganda et al., submitted for
differences in surface functional groups but by differences in mineral publication).
specific gravity.
A higher advancing contact angle is required to recover pyrite 4. Conclusions
particles in flotation at the same rate as chalcopyrite particles within
the same size fraction. The critical advancing contact angles for pyrite Chalcopyrite particles exhibiting the same mean advancing
and chalcopyrite particles have been determined in this work under contact angle in the feed (within experimental error of the Washburn
the same specific hydrodynamic conditions (Fig. 11). It should be technique), but with different functional groups on their surfaces (i.e.,
noted that there is a significant difference in the critical contact angle amyl xanthate and dicresyl dithiophosphate) are recovered with a
between chalcopyrite and pyrite for all size fractions greater than the similar undistributed rate constant (within experimental error). This
−20 μm size fraction. This supports the hypothesis that the difference finding further validates the calibration developed using the chalco-
in critical contact angle between chalcopyrite and pyrite is probably pyrite-amyl xanthate system and extends its application to a different
due to the difference in specific gravity between the minerals. collector on chalcopyrite. The conservation of the advancing contact
For the −20 μm fraction, effects due to differences in specific gravity angle through flotation and sample preparation for a given size
would be minimal, and there is only a small difference in the critical fraction, for both chalcopyrite and pyrite, was demonstrated by the
contact angle. Fine particles less than 20 μm, for both chalcopyrite and very close agreement between the measured and recalculated feed
pyrite, have a higher critical contact angle than intermediate size contact angle values under the conditions of this study. The
fractions to affect stable bubble–particle attachment due probably to heterogeneity of contact angles in each feed size fraction, for both
their lower kinetic energy (Scheludko et al., 1976). The reason fine chalcopyrite and pyrite, was confirmed by measurements of contact
particles generally have lower maximum recovery than intermediate angle on the flotation products. The concentrate contact angles values
sized particles is their higher critical contact angle and the difficulty in were always higher than the tailings contact angle values, which, in
furnishing mean contact angle values well above the critical value. As turn, were always at, or lower, than the critical contact angle that is
the particle size increases above approximately 53 μm, the critical specific for each mineral. Under the conditions of the present study,
contact angle again increases for both chalcopyrite and pyrite due to the flotation response of chalcopyrite particles at pulp densities of 2%
decreases in the stability efficiency of the bubble–particle aggregate and 30% solids, with quartz as non-interacting gangue, was the same
(Schulze, 1977). within experimental error. This finding may allow benchmarking of
Changes in the hydrodynamic conditions may give different critical chalcopyrite flotation behaviour from an ore against the calibration
contact angle values. If these results reflect the general behaviour of developed for the chalcopyrite-amyl xanthate system.
mineral particles, then it should be possible to determine the maximum The undistributed rate constants of pyrite were lower than for
recovery of any fully liberated mineral in a specific size fraction when chalcopyrite exhibiting the same feed advancing contact angle, within
the feed mean contact angle and critical advancing contact angle values experimental error, while the maximum recovery of pyrite was also
are known and assuming the same statistical distribution of contact lower for the same feed advancing contact angle, in the range less than
angle values as for the chalcopyrite-amyl xanthate system. The 80°. This difference was attributed to differences in the critical
methodology needs to be tested against a natural ore, a difficult task advancing contact angle for pyrite and chalcopyrite due to differences

Table 10
Pyrite flotation data model fitting to a gamma statistical distribution with mean (measured) and mode contact angle. The critical contact angle, fixed for each size fraction, was
allowed to vary to obtain the best fit.

Size fraction, μm Mean contact angle (°) Mode contact angle (°) Rmax (model) (%) Rmax (Exptl, %) (ΔRmax)2 (model-Exptl) Σ(ΔRmax)2 Critical contact angle (°)

53–75 38 18 20 28 60 62 57
56 57 47 47 1
74 93 82 81 1
+ 300 61 73 15 15 1 7 82
69 87 35 33 3
90 100 83 85 3
194 S. Muganda et al. / International Journal of Mineral Processing 98 (2011) 182–194

in mineral specific gravity. The same statistical function may be applied Krasowska, M., Terpilowski, K., Chibowski, E., Malysa, K., 2006. Apparent contact angles
and time of the three phase contact formation by the bubble colliding with teflon
to both pyrite and chalcopyrite flotation data to determine the surfaces of different roughness. Physicochemical Problems of Mineral Processing
distribution of contact angle values given a mean contact angle 40, 293–306.
measured in the feed. It may be possible to predict the maximum Krasowska, M., Zawala, J., Malysa, K., 2009. Air at hydrophobic surfaces and kinetics of
three phase contact formation. Advances in Colloid and Interface Science 147–148,
recovery of a liberated mineral as a function of size fraction when the 155–169.
feed contact angle is known or, conversely, to infer an operational Lynch, A.J., Johnson, N.W., Manlapig, E.V., Thorne, C.G., 1981. Mineral and Coal Flotation
contact angle based on particle size, flotation rate constant, and critical Circuits: Their Simulation and Control. Elsevier Scientific, Amsterdam.
Miller, J.D., Veeramasuneni, S., Drelich, J., Yalamanchili, M.R., 1996. Effect of roughness
contact angle of a mineral under the specific test conditions used in this as determined by atomic force microscopy on the wetting properties of ptfe thin
study. films. Polymer Engineering And Science 36, 1849–1855.
Muganda, S., Zanin, M., Grano, S., 2008. Flotation behaviour of sulphide mineral size
fractions with controlled contact angle. Chemeca Conference Proceedings: Towards
Acknowledgments
a sustainable Australasia. Newcastle City Hall, New South Wales, Australia,
Engineers Australia.
Financial support from the sponsors of the AMIRA P260E project Muganda, S., Zanin, M., Grano, S.R., 2011. Influence of particle size and contact angle on
and the Australian Research Council is gratefully acknowledged. the flotation of chalcopyrite in a laboratory batch flotation cell. International
Journal of Mineral Processing 98, 150–162.
Muganda, S., Zanin, M., Grano, S.R., submitted for publication. Benchmarking flotation
References performance: natural ore.
Najafi, A.S., Xu, Z., Masliyah, J., 2008. Measurement of sliding velocity and induction
Agar, G.E., Stralton-Crawly, R., Bruce, T.J., 1986. Optimising the design of flotation time of a single micro-bubble under an inclined collector surface. Canadian Journal
circuits. CIM Bulletin 73, 173–181. of Chemical Engineering 86, 1001–1010.
Bloom, F., Heindel, T.J., 2003. Modelling flotation separation in a semi-batch process. Piantadosi, C., 2001. Competitive collector adsorption in the selective flotation of galena
Chemical Engineering Science 58, 353–365. and chalcopyrite from iron sulphide minerals. Ian Wark Research Institute.
Brito, E., Abreu, S., Brien, C., Skinner, W., 2010. ToF-SIMS as a new method to determine University of South Australia, Adelaide, South Australia.
the contact angle of mineral surfaces. Langmuir 26 (11), 8122–8130. Piantadosi, C., Jasieniak, M., Skinner, W.M., Smart, R.S.C., 2000. Statistical comparison of
Cassie, A.B.D., Baxter, S., 1944. Wettability of porous surfaces. Transcripts of the Faraday surface species in flotation concentrates and tails from ToF-SIMS evidence.
Society 40, 547–551. Minerals Engineering 13, 1377–1394.
Chen, F., Gomez, C.O., Finch, J.A., 2001. Technical note bubble size measurement in Pyke, B., Fornasiero, D., Ralston, J., 2003. Bubble particle heterocoagulation under
flotation machines. Minerals Engineering 14, 427–432. turbulent conditions. Journal of Colloid and Interface Science 265, 141–151.
Dai, Z., Dukhin, S., Fornasiero, D., Ralston, J., 1998. The inertial hydrodynamic Ralston, J., Dukhin, S.S., Mishchuk, N.A., 1999a. Inertial hydrodynamic particle–bubble
interaction of particles and rising bubbles with mobile surfaces. Journal of Colloid interaction in flotation. International Journal of Mineral Processing 56, 207–256.
and Interface Science 197, 275–292. Ralston, J., Fornasiero, D., Hayes, R., 1999b. Bubble–particle attachment and detachment
Derjaguin, B.V., Dukhin, S.S., 1961. Theory of flotation of small and medium-size in flotation. International Journal of Mineral Processing 56, 133–164.
particles. Transactions of the Institution of Mining and Metallurgy 70, 221–246. Ross, V.E., 1991. Comparison of methods for evaluation of true flotation and
Diggins, D., Ralston, J., 1993. Particle wettability by equilibrium capillary pressure entrainment. Trans. Instn. Min. Metall. (Sect. C: Mineral Process. Extr. Metall.):
measurements. Coal Preparation 13, 1–19. The Institution of Mining and Metallurgy, 100, pp. C121–C126.
Drelich, J., 2001. Contact angles measured at mineral surfaces covered with adsorbed Runge, K.C., Franzidis, J.P., Manlapig, E.V., 2003. A study of the flotation characteristics of
collector layers. Minerals and Metallurgical Processing 18, 31–37. different mineralogical classes in different streams of an industrial circuit.
Duan, J., Fornasiero, D., Ralston, J., 2003. Calculation of the flotation rate constant of Proceedings: XXII International Mineral Processing Congress: Cape Town, South
chalcopyrite particles in an ore. International Journal of Mineral Processing 72, 227–237. Africa, 29 September–3 October 2003.
Dunstan, D., White, L.R., 1986. A capillary pressure method for measurement of contact Savassi, O.N., Alexander, D.J., Franzidis, J.P., Manlapig, E.V., 1998. An empirical model for
angles in powders and porous media. Journal of Colloid and Interface Science 111, entrainment in industrial flotation plants. Minerals Engineering 11, 243–256.
60–64. Scheludko, A., Toshev, B.V., Bojadjiev, D.T., 1976. Attachment of particles to a liquid
Fairthorne, G., Fornasiero, D., Ralston, J., 1997. Effect of oxidation on the collectorless surface (capillary theory of flotation). Journal of the Chemical Society, Faraday
flotation of chalcopyrite. International Journal of Mineral Processing 49, 31–48. Transactions 1: Physical Chemistry in Condensed Phases 72, 2815–2828.
Finch, J.A., Dobby, G.S., 1990. Column Flotation. Pergamon Press, Elmsford, New York. Schulze, H.T., 1977. New theoretical and experimental investigations on stability of
Fuerstenau, M.C., Huiatt, J.L., Kuhn, M.C., 1971. Dithiophosphate vs. xanthate flotation of bubble/particle aggregates in flotation—a theory on the upper particle size of
chalcocite and pyrite, minerals beneficiation. AIME 250. floatability. International Journal of Mineral Processing 3, 241–259.
Gorain, B.K., Franzidis, J.P., Manlapig, E.V., 1997. Studies on impeller type, impeller Sedev, R., Fabretto, M., Ralston, J., 2004. Wettability and surface energetics of rough
speed and air flow rate in an industrial scale flotation cell. Part 4: effect of bubble fluoropolymer surfaces. Journal of Adhesion 80, 497–520.
surface area flux on flotation performance. Minerals Engineering 10, 367–379. Smart, R.S.C., Amarantidis, J., Skinner, W., Prestidge, C.A., Lavanier, L., Grano, S., 1998.
Grau, R. A. (2006) An investigation of the effect of physical and chemical variables on Surface analytical studies of oxidation and collector adsorption in sulphide mineral
bubble generation and coalescence in laboratory scale flotation cells. PhD Thesis. flotation. Scanning Microscopy 12, 553–583.
Department of Materials Science and Engineering. Helsinki, Finland, Helsinki Song, S., Lopez-Valdivieso, A., Ojeda-Escamilla, M.C., 2001. Electrophoretic mobility
University of Technology (Espoo). study of the adsorption of alkyl xanthate ions on galena and sphalerite. Journal of
Gu, G., Xu, Z., Nandakumar, K., Masliyah, J., 2003. Effects of physical environment on Colloid and Interface Science 237, 70–75.
induction time of air–bitumen attachment. International Journal of Mineral Sripriya, R., Rao, P.V.T., Roy Choudhury, B., 2003. Optimisation of operating variables of
Processing 69, 235–250. fine coal flotation using a combination of modified flotation parameters and
Gu, G., Sanders, R.S., Xu, Z., Nandakumar, K., Masliyah, J., 2004. A novel experimental statistical techniques. International Journal of Mineral Processing 68, 109–127.
technique to study single bubble-bitumen attachment in flotation. International Stevens, N.I. (2005) Contact angle measurements on particulate systems. Ian Wark
Journal of Mineral Processing 69, 235–250. Research Institute. Adelaide, PhD Thesis. University of South Australia.
Harris, M.C., Runge, K.C., Whiten, W.J., Morrison, R.D., 2002. Jksimfloat as a practical tool Subrahmanyam, T.V., Monte, M.B.M., Middea, A., Valdiviezo, E., Lins, F.F., 1999. Contact
for flotation process design and optimization. Mineral Processing Plant Design, angles of quartz by capillary penetration of liquids and captive bubble techniques.
Practice, and Control: Proceedings. SME. Minerals Engineering 12, 1347–1357.
Hart, B., Biesinger, M., Smart, R.S.C., 2006. Improved statistical methods applied to Vera, M.A., Franzidis, J.P., Manlapig, E.V., 1999. Simultaneous determination of
surface chemistry in minerals flotation. Minerals Engineering 19, 790–798. collection zone rate constant and froth zone recovery in a mechanical flotation
Hernandez-Aguilar, J.R., Rao, S.R., Finch, J.A., 2005. Testing the k–sb relationship at the environment. Minerals Engineering 12, 1163–1176.
microscale. Minerals Engineering 18, 591–598. Warren, L.J., 1985. Determination of the contributions of true flotation and entrainment
Jameson, G.J., Nam, S., Young, M., 1997. Physical factors affecting recovery rates in in batch flotation tests. International Journal of Mineral Processing 14, 33–44.
flotation. Mineral Science Engineering 9 (3), 103–118. Washburn, E.W., 1921. The dynamics of capillary flow. Physical Review XVII (3),
Krasowska, M., Malysa, K., 2007. Kinetics of bubble collision and attachment to 273–283.
hydrophobic solids: I. Effect of surface roughness. International Journal of Mineral
Processing 81, 205–216.

Das könnte Ihnen auch gefallen