Sei sulla pagina 1di 13

Minerals Engineering 69 (2014) 120–132

Contents lists available at ScienceDirect

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

Adsorption of O-isopropyl-N-ethyl thionocarbamate


on Cu sulfide ore minerals
Alan N. Buckley a,⇑, Gregory A. Hope b, Kenneth C. Lee c, Eddie A. Petrovic d, Ronald Woods b
a
School of Chemistry, The University of New South Wales, Sydney, NSW 2052, Australia
b
School of Biomolecular and Physical Sciences, Griffith University, Nathan, QLD 4111, Australia
c
Orica Mining Chemicals, 16-20 Beauchamp Road, Matraville, NSW 2036, Australia
d
Orica Mining Chemicals, 1 Nicholson Street, Melbourne, VIC 3000, Australia

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

Article history: The adsorption of the flotation collector O-isopropyl-N-ethyl thionocarbamate on air-exposed Cu metal,
Received 18 May 2014 cuprite, chalcocite, covellite, cubanite, chalcopyrite, bornite and pyrite at pH 6 has been investigated
Accepted 1 August 2014 primarily by X-ray photoelectron spectroscopy. It was found that CuI oxide and sulfide surfaces condi-
tioned in the collector solution were hydrophobic, regardless of whether the collector adsorbed as the
anion (IPETC), in protonated form (HIPETC), or as multilayer CuIPETC patches. The CuI native oxide layer
Keywords: on the metal was largely unaffected by the adsorbed collector. For surfaces at which chemisorption
Thionocarbamate flotation collectors
occurred, interaction of the S in the collector was directly with Cu in the metal or mineral surface, rather
Copper sulfide ore minerals
XPS
than via O. Apart from chalcocite, the main species adsorbed from the collector solution on to CuI oxide
Adsorption surfaces were different from those adsorbed on the sulfide minerals, in that the N was predominantly
deprotonated in the former and protonated in the latter. On chalcocite, comparable amounts of IPETC
and HIPETC were adsorbed. Adsorbed HIPETC was in good electrical contact with the surface of cubanite,
chalcopyrite and bornite, presumably because of direct interaction with the mineral rather than adsorp-
tion on top of chemisorbed IPETC or FeOOH. Collector coverage of the Cu minerals correlated with the
availability of CuAOHsurf with which HIPETC could react. Coverage by collector adsorbed as HIPETC on
pyrite surfaces was low, but not negligible, nevertheless conditioned abraded pyrite surfaces were not
obviously hydrophobic.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction effective collector for all forms of copper mineralisation and zinc
that offers improved selectivity over pyrite. Notwithstanding the
The structure of O-isopropyl-N-ethyl thionocarbamate, also consensus on the ores for which IPETC should be used, since the
known (more correctly) as O-isopropyl-N-ethyl thiocarbamate, is mid-1960s there have been numerous studies of the interaction
shown in Fig. 1 in its protonated, molecular form (HIPETC). The col- of dialkyl thionocarbamates with Cu metal, Cu sulfide minerals
lector is commonly referred to as IPETC, though more specifically, and pyrite, but agreement has yet to be reached on the extent of
IPETC represents the deprotonated, anionic form. adsorption, the collector species adsorbed and the mechanism of
When IPETC was initially used for sulfide mineral flotation their adsorption.
under the name Z-200, its advantages over xanthates and dithio- Finkelstein and Goold (1972) reported that the adsorption of Z-
phosphates were claimed to be higher selectivity against pyrite 200 on sulfide minerals was low, and they were unable to obtain IR
and higher stability at low pH (Harris and Fischback, 1954). Those spectra from surface extracts to identify the adsorbed species. The
advantages have since been validated, and there is general agree- fraction of a theoretical monolayer abstracted from solution in
ment on the ore types that benefit from IPETC use. Avotins et al. 15 min at pH 8 was 0.16 for chalcocite, 0.11 for chalcopyrite,
(1994) described the major attributes of IPETC as its hydrolytic sta- 0.05 for covellite, 0.04 for bornite and 0.03 for pyrite. These esti-
bility, shelf life and selectivity against gangue iron sulfides. The mates were surprising, in that adsorption by pyrite did not appear
Orica analogue DSP 009 is described by that company as an to be markedly less than adsorption by covellite or bornite. Less
surprising was the microflotation data for IPETC reported by
Ackerman et al. (1999), which showed that at pH 5, recovery of
⇑ Corresponding author. chalcopyrite, chalcocite and covellite was above 90%, while that
E-mail address: a.buckley@unsw.edu.au (A.N. Buckley).

http://dx.doi.org/10.1016/j.mineng.2014.08.002
0892-6875/Ó 2014 Elsevier Ltd. All rights reserved.
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 121

S which results in a metal thiol compound being formed on the min-


eral surface:

MS ! Mnþ þ S0 þ ne ð1Þ


O N
H Mnþ þ nðTCÞ ! MðTCÞn ð2Þ

Fig. 1. Structure of O-isopropyl-N-ethyl thionocarbamate. where TC is the thionocarbamate anion. The coupled reactions
would occur at a potential greater than or equal to the thermody-
namic potential for the (bulk) metal thionocarbamate formation.
Those authors pointed out that, since their proposed mechanism
of bornite and pyrite was 50–90%. At pH 8.5, the recovery of chal-
involved an electrochemical step followed by a chemical reaction,
cocite and covellite was above 90%, but that for chalcopyrite was
it was best described as an EC mechanism (Zanello, 2003). Yoon
only just below 90%, while recovery for bornite was lower and pyr-
and Basilio argued that dissolution–precipitation was more likely
ite 59%.
than chemisorption for thionocarbamate collectors because of the
The mechanism of IPETC adsorption remains unresolved,
very high pK values for the metal thionocarbamates. Nevertheless,
despite the research effort. From an IR spectroscopic study of O-
even when the production of S0 in Eq. (1) is interpreted as the for-
butyl-N-ethyl thionocarbamate, its Cu salt, and its interaction with
mation of a metal-deficient (or S-rich) sulfide surface layer, there
chalcopyrite and a Cu metal film, Bogdanov et al. (1976) concluded
appears to be a dearth of convincing experimental support for dis-
that the thionocarbamate attached to the surface of a Cu film in
solution–precipitation as an alternative to chemisorption.
ionic (deprotonated) form, but by ‘chemisorption’ in molecular
Rather than precipitation as CuIPETC, Woods and Hope (1999)
(protonated) form to the surface of chalcopyrite. In each case, it
and Hope et al. (2000) provided evidence for chemisorption of
was deduced that both the S and N of the collector formed a bond
IPETC on Cu metal. They investigated the interaction of IPETC with
with the same Cu atom. For pyrite, both chemisorption and phys-
an electrochemically roughened Cu metal electrode by means of
ical adsorption were deduced. On the other hand, Glembotskii
in-situ and ex-situ surface-enhanced Raman scattering (SERS)
(1977) argued that in the dialkyl thionocarbamates, both the elec-
spectroscopy. The SERS spectra from emersed electrodes that had
tron density on the hetero atom, and its steric accessibility,
been held in de-aerated IPETC solution of pH 6 at a potential below
decreased in the order S > O > N. Accordingly, if the chemisorption
that at which multilayer CuIPETC would be formed were broadly
of the collector to Cu in the surface of Cu sulfides was going to
similar to the spectrum for CuIPETC, in which the IPETC ligand
involve a hetero atom in addition to the S, it was more likely to
was deprotonated. The similarity of the spectra suggested that
be the O rather than the N. Adsorption on pyrite, by contrast,
any interaction between the O or N of IPETC and surface Cu was
would be predominantly physical in nature.
similar to that in the bulk complex. The spectra were consistent
Ackerman et al. (1984) noted that while the S was the most acces-
with the collector being chemisorbed through the S to surface Cu
sible and most reactive atom in a dialkyl thionocarbamate molecule,
atoms in the anodic process shown in Eq. (3).
the N must also be accessible for chelate ring formation. Thus, it was
implicitly assumed that if a heteroatom other than S was involved in Cu þ HIPETC ! CuIPETC þ Hþ þ e ð3Þ
the interaction, it would be the N rather than O. Furthermore,
Ackerman et al. (1999) claimed that for sulfide flotation, ‘‘the most In the flotation of a Cu sulfide, an anodic process analogous to
effective collectors, in terms of high flotation recoveries and rates Eq. (3) could be coupled with the cathodic reduction of oxygen,
with good pyrite rejection, are those compounds containing N and but IPETC interaction would not necessarily be represented by such
S as functional groups for chelation’’. With an even greater emphasis a mechanism for mineral surfaces that had been exposed to air
on the N in IPETC, Solozhenkin et al. (2013), in their molecular mod- before contact with collector.
elling of the in vacuo interaction of IPETC with pyrite (analogous to In the present investigation, the unresolved and contentious
the chemisorption of IPETC on pyrite), compared monodentate issues concerning the interaction of IPETC with Cu sulfides and
bonding of the collector N to a pyrite Fe atom with bidentate bonding pyrite have been addressed. In particular, specific answers have
of the collector N and S to the Fe. However, they did not model biden- been sought to questions concerning (a) the identity of the collec-
tate bonding of the collector S and O to the pyrite Fe, or calculate the tor species adsorbed on the minerals of interest, (b) the adsorption
energy for an HIPETC/pyrite complex. mechanisms of the collector species, and (c) whether there is a cor-
From an in-situ FTIR investigation of the adsorption of IPETC on relation between the collector species adsorbed and observed flo-
chalcocite, chalcopyrite and pyrite, Leppinen et al. (1988) found atability. Characterisation of collector species adsorbed at
that, except at very low pH, adsorption of IPETC on chalcocite conditioned mineral surfaces was performed by means of X-ray
was 3–20 times higher than on chalcopyrite (or pyrite). They con- photoelectron spectroscopy (Buckley, 2010).
sidered their IR spectra obtained at pH 4 to be consistent with
CuAS bonding, but it was not clear whether there was any coordi- 2. Experimental details
nation via N. At pH 6, new bands appeared in the IR spectrum that
were thought to indicate coordination via O as well as S. They rea- 2.1. IPETC collector and mineral specimens
soned, like Glembotskii (1977), that steric accessibility and elec-
tron density considerations favoured additional coordination Commercial DSP 009 collector supplied by Orica Mining Chem-
through O rather than N, and concluded that below pH 6, IPETC icals was used without purification. According to the SDS informa-
adsorbed on chalcocite through the S only, but above pH 6, adsorp- tion, DSP 009 consists of at least 95% O-isopropyl ethyl
tion was through the S and O. They found that IPETC did not thionocarbamate (MW 147.24), with less than 2% 1,3-diethyl thio-
remove sulfate-like oxidation products from the surface of pyrite, urea and less than 2% isopropyl alcohol. Manufacturers of IPETC
and hypothesised that this might account for the flotation selectiv- collector claim no more than 1% diethyl thiourea, no more than
ity against pyrite. They noted that ethyl xanthate does remove the 2% isopropanol and 1% maximum moisture. The collector is
oxidation products from pyrite. described as being slightly soluble in water. The pH of the undi-
Yoon and Basilio (1993) claimed that thionocarbamates adsorb luted DSP 009 was near 7, consequently with decreasing concen-
on sulfide minerals via a dissolution–precipitation mechanism tration, the pH of the diluted collector approached that (5.5) of
122 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

the water in equilibrium with air used for dilution. All conditioning spectrometer using monochromatised Al Ka X-rays focused to a
was carried out at the unadjusted pH of the collector solution, and spot size of 0.5 mm, and an electron analyser pass energy of
with physical agitation to simulate the turbulent environment in a 20 eV for narrow range scans. The magnetic lens mode of the ana-
flotation cell. A freshly diluted collector solution was prepared lyser was used in all cases. Included in the binding energies
before each conditioning. Collector concentrations between 104 employed for calibration were 83.96 eV for Au 4f7/2 of metallic gold
and 102 M were used, with conditioning periods between 1 and and 932.6 eV for Cu 2p3/2 of Cu metal. The pressure in the analysis
10 min. chamber was better than 5  107 Pa during spectral acquisition.
Fairthorne et al. (1996) confirmed that IPETC, with a pKa of 12, The data were collected and processed under Thermo Scientific
was present in molecular form in near-neutral aqueous solution, Avantage 4.58 and 4.54 software. The elemental composition
and had a solubility at pH 7.7 of 2.6  102 mol dm3 at 25 °C. That (excluding H) of the 2–4 nm surface layer characterised by XPS
solubility is slightly higher than the 5  103 mol dm3 value was estimated, where appropriate, from the Cu 2p3/2, Fe 2p3/2, S
determined by Poling (1976) for Z-200. Poling had noted that 2p, O 1s, N 1s and C 1s peak intensities above a Shirley background
because Z-200 was an oily liquid with limited solubility in water, in each narrow range spectrum, and the theoretical photo ionisa-
it was often added at the grinding circuit. Livshits et al. (1968) tion cross-sections incorporated in the Avantage software. The esti-
had measured the pKa of Z-200 to be 11.52 ± 0.06. mated composition could contain a small systematic error arising
The chalcocite and covellite investigated were from Butte, Mon- from 2p3/2 peak background selection.
tana, USA, the chalcopyrite was from Kadina, Yorke Peninsula, In most cases, photoelectron spectra were obtained initially
South Australia, the cubanite was from the Henderson #2 mine, without the use of a flood-gun, and subsequently under the influ-
Chibougamau, Quebec, Canada, and the bornite was from Kipuchi, ence of flood-gun electrons with the specimen earthed (suite 2) or
Zaire. The cuprite was a crystalline specimen from the Red Dome the specimen floating (suite 3). However, the frozen DSP 009 spec-
Mine, Chillagoe, Queensland, Australia, and its interaction with imen required the flood-gun and spectra were determined with the
dialkyl dithiophosphate collectors has been studied previously specimen floating. With the specimen holder floating, in minimis-
(Buckley et al., 2009). The pyrite was from a 2.3 cm cube from ing peak broadening from a non-uniform potential within the spec-
Navajún, La Rioja, Spain. imen surface layer, the low energy electrons ‘over-compensate’ for
Mineral particle surfaces produced during plant-scale commi- charging and the measured photoelectron binding energies are
nution (prior to flotation collector addition) typically display frac- typically lower than their correct energies by a value approaching
ture and abrasion characteristics, depending on the degree of the energy of the electrons. The measured binding energies were
hardness and brittleness of the particular mineral. To emulate corrected by assuming that they all experienced the same shift,
those characteristics in this investigation, mineral surfaces were and adjusting them so that the hydrocarbon C 1s binding energy
prepared by fracture or abrasion in air, and as the composition of was 285.0 eV. Where appropriate, spectra in the figures have been
abraded and fracture surfaces can be subtly different, collector shown with binding energy adjusted in that way. Whenever the
interaction with both is of relevance. Also, provided mineral spec- flood-gun was used, the possibility of beam damage by the low
imens are of adequate purity and do not fracture along a pre-oxi- energy electrons was monitored, but in all cases, spectra were
dised interface, fracture surfaces are inherently cleaner than obtained as quickly as possible at the expense of signal-to-noise
abraded surfaces, and subsequent surface characterisation is less in order to minimise any damage arising from the secondary elec-
complicated by extraneous contamination. Either P1200 silicon trons associated with the X-ray photoemission.
carbide paper or 0.3 lm alumina powder moistened with distilled
water was used in the preparation of surfaces by abrasion. 2.3.2. Collector coverage information from Cu L3M4,5M4,5 kinetic
energies
2.2. Assessment of hydrophobicity of conditioned metal and mineral It is argued that just as the onset of molecular Cu or Ag xanthate
specimens or dithiophosphate adsorption on Cu or Ag sulfide can be detected
from the initial appearance of a significantly lower kinetic energy
A conditioned Cu metal or mineral surface was assessed as peak in the X-ray excited Cu L3M4,5M4,5 or Ag M4N4,5N4,5 spectrum
hydrophobic if the water used to rinse any adhering collector solu- (Buckley and Woods, 1993, 1995), so the onset of Cu thionocarba-
tion from the specimen, when it was emersed at the end of the mate adsorption on a Cu sulfide should be able to be detected from
conditioning period, did not leave a film that wet the surface. For the initial appearance of a Cu L3M4,5M4,5 peak at 915 eV. Cu 2p or
abraded surfaces, a contact angle greater than 50o for any adhering Ag 3d binding energy shifts are too small to differentiate the chem-
droplets of water could usually be determined by visual examina- isorbed monolayer from any adsorbed multilayer CuI or AgI thio-
tion. Mineral fracture surfaces were typically too rough to make late. Unlike the situation for Cu thiolate adsorption, there should
such a visual assessment. However, any fine particles, which had be no change in the Cu L3M4,5M4,5 spectrum when a thiol collector
been broken off the edges of mineral specimens by agitation during is chemisorbed on a Cu sulfide mineral surface. Since the main Cu
the conditioning period, accumulated at the liquid/air interface L3M4,5M4,5 peak for CuI in an oxide and sulfide lattice is at a com-
when those particles had been rendered hydrophobic. For fracture parable kinetic energy (916.6 and 917.2 eV, respectively), the
surfaces of some minerals, such behaviour provided an alternative chemisorption of IPETC on the native oxide layer of Cu metal
indicator of hydrophobicity. should also give rise to no noticeable change in the Cu L3M4,5M4,5
spectrum, as an additional, low intensity component near 917 eV
2.3. X-ray photoelectron spectroscopy (XPS) would not be resolved from the main oxide peak at 916.6 eV. The
relevant, previously observed Cu L3M4,5M4,5 peak kinetic energies
2.3.1. Acquisition and processing of XPS data are summarised in Table 1.
X-ray photoelectron spectra were collected from a surface of
single piece mineral or Cu metal specimens of size approximately 2.3.3. Interpretation of N 1s spectral components
5 mm  5 mm  1 mm mounted in electrical contact with the The N 1s spectra from HIPETC (Section 3.1) and multilayer CuI-
sample stub. Each fracture surface, or surface abraded until rela- PETC (Section 3.3) confirmed the assignment of components at
tively smooth to the unaided eye, was rinsed with pure water if 400.0 ± 0.2 and 398.6 ± 0.3 eV to protonated and deprotonated
conditioned in collector solution and characterised by XPS at ambi- thionocarbamate N, respectively. These binding energies are close
ent temperature. XPS data were obtained on an ESCALAB 250Xi to those observed for protonated and deprotonated N in analogous
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 123

Table 1
Previously reported Cu L3M4,5M4,5 peak kinetic energies.

Cu environment Example Cu L3M4,5M4,5 kinetic energy Reference


(eV)
CuI coordinated by O in lattice Cuprite 916.6 Buckley et al. (2009)
CuI coordinated by O in lattice and S by chemisorbed Thiol chemisorbed on Cu in 917 Buckley et al. (2009)
collector cuprite
CuII coordinated by O in adsorbed molecule Multilayer Cu hydroxamate 915.0 ± 0.5 Hope et al. (2011)
CuI coordinated by S in lattice Chalcocite 917.2 ± 0.1 Buckley and Woods (1993)
CuI coordinated by S in lattice and S in chemisorbed Thiol chemisorbed on Cu in 917.2 ± 0.3 Buckley and Woods (1993)
collector chalcocite
CuI coordinated by S in adsorbed molecule Multilayer Cu thiolate 915.0 ± 0.2 Buckley et al. (2009) and Buckley and Woods
(1993)

materials, and their assignments are not contentious. However, the (a) (b)
origin of an unresolved component near 401 eV required to fit
some spectra is less clear cut. A component between 400.8 and
401.4 eV in the N 1s spectrum determined with the flood-gun off
and the specimen earthed has been interpreted as arising from
charge-shifted protonated N in adsorbed HIPETC in poor electrical
contact with the surface. The justification for this assignment is the
absence of this 401 eV component, with a concomitant increase
in the intensity of the 400 eV component, when the flood-gun
is on while the specimen is floating (e.g., Section 3.4.3), or the shift-
ing of the 401 eV component by the energy of the flood-gun elec-
trons when the specimen is earthed (e.g., Section 3.4.5).
Accordingly, the observation of such a component is considered a 291 288 285 282 291 288 285 282
diagnostic indicator, along with a more intense N 1s component
Binding Energy (eV) Binding Energy (eV)
at 400 eV and a C 1s component near 288.4 eV from HIPETC in
good electrical contact with the specimen surface, for the presence (c) (d)
of multilayer adsorbed HIPETC.

3. Results

Conditioned surface elemental concentrations (excluding H),


peak intensity ratios, as well as N 1s fitted components and line-
widths, are listed in the Supplementary Data. Also included with
the Supplementary Data is the detailed reasoning leading to the
deduction, from the photoelectron spectra shown in Figs. 2–8, of
collector species adsorbed at each conditioned surface. The reason-
ing is similar to that described in Section 3.3 for CuIPETC adsorbed
on a Cu metal surface conditioned in 0.01 M DSP 009. For other
291 288 285 282 291 288 285 282
conditioned surfaces, only a summary of the species adsorbed is
Binding Energy (eV) Binding Energy (eV)
provided in this section.
Fig. 2. C 1s spectrum from (a) frozen HIPETC; (b) multilayer CuIPETC on Cu metal;
3.1. XPS characterisation of frozen commercial IPETC collector (c) Cu metal native oxide conditioned 10 min in 104 M collector; (d) chalcocite
conditioned 10 min in 104 M collector with flood-gun on and specimen floating.
DSP 009 (95% HIPETC) is a liquid with a sufficiently high
vapour pressure to require cooling for a small droplet of it to be
C3 H8 OCðSÞNHC2 H5 þ H2 O ! C3 H8 OCðOÞSH þ C2 H5 NH2 ð4Þ
retained as a specimen in the UHV of a photoelectron spectrome-
ter. When such a specimen was cooled to 150 K under dry nitro- The C 1s spectrum (Fig. 2a) could be fitted with four components
gen gas, the only foreign element detected was a low concentration of linewidth less than 1.2 eV; a major component assigned to a
of Si, most probably from the droplet substrate. The principal com- binding energy of 285.0 eV accounting for slightly more than half
ponent in the O 1s spectrum was at 533.4 eV and accounted for 86% the intensity, and three components each accounting for about
of the intensity. An obvious shoulder on the low binding energy 14% of the intensity at 286.1, 286.85 and 288.4 eV. The main com-
side of the main peak could be fitted with a component at ponent at 285.0 eV would have arisen from the three C atoms at
532 eV to account for the remaining intensity. A binding energy the ends of the molecule, as well as the majority of the C in the
of 533.4 eV is near to that expected for HIPETC, as a value of impurities. The other components, in increasing binding energy,
533.7 eV has been reported for an O adjacent to a C@O group in can be attributed to the ethyl C bonded to N, the isopropyl C bonded
a polymer (López et al., 1991), a close analogy to an O adjacent to O, and the C bonded to N, S and O. The assignment of the ethyl C
to the C@S in HIPETC. For the carbonyl oxygen in an ester, López atom bonded to N is on the basis of the observed binding energy of
et al. (1991) reported an O 1s binding energy of 532.2 ± 0.2 eV, so 285.6 eV for poly(ethyleneimine), [ACH2CH2NHA]n, (Beamson and
the minor component in the spectrum from DSP 009 would be con- Briggs, 1992) while the assignment of the isopropyl C atom bonded
sistent with some isopropyl monothiocarbonate (carbonothioic to O is supported by the measurements carried out by López et al.
acidoisopropyl ester) produced by IPETC decomposition: (1991). The assignment is also supported by the 13C NMR spectra
124 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

for HIPETC dissolved in chloroform (Woods and Hope, 1999). (a) (b)
Observed values for the C 1s binding energy in thiourea,
H2NAC(@S)ANH2, have been 288.0 and 288.1 eV (Lee and
Rabalais, 1977; Srinivasan and Walton, 1977). The highest binding
energy component at 288.4 eV from C bonded to the S, O and pro-
tonated N is observed at a significantly lower binding energy
(287.4 eV) in the deprotonated adsorbate (Section 3.3). It would
appear that the C bonded to the S and two O atoms in any mono-
thiocarbonate decomposition product must have had a 1s binding
energy close to that (288.4 eV) for the C bonded to N, S and O in
IPETC.
The N 1s spectrum determined at the outset (Fig. 3a) comprised
a single peak at 399.9 eV of width 1.1 eV, but it was slightly 172 168 164 160 156 172 168 164 160 156
broader (1.25 eV) at the end of the spectral suite. This binding Binding Energy (eV) Binding Energy (eV)
energy was close to the value expected on the basis of those
(399.2 and 400.1 eV) observed for the N in lactams (cyclic
(c) (d)
ACAC(@O)ANHACA) and amides, 399.6–399.9 eV (e.g., Salyn
et al., 1978; Jansen and van Bekkum, 1995), and was consistent
with a N that was protonated (ANHA). As discussed further below,
any ethyl amine retained as a result of IPETC decomposition, reac-
tion (4), must have been of negligible concentration, as a compo-
nent constrained to be at the binding energy expected for amine
(399.1 eV) could not be fitted with intensity greater than zero.
The S 2p spectrum (Fig. 4a) could be fitted predominantly with a
doublet at a 2p3/2 binding energy of 162.2 eV and with a relatively
narrow component linewidth (1.0 eV). Near the start of the spec-
tral suite, a minor doublet of similar linewidth accounting for 5%
of the overall S 2p intensity at 163.8 eV was required for an ade- 172 168 164 160 156 172 168 164 160 156
quate fit. This minor doublet is expected to have originated from an Binding Energy (eV) Binding Energy (eV)
impurity or decomposition product (Section 3.2). Towards the end
of the spectral suite, the spectrum was similar, but the linewidth Fig. 4. S 2p spectrum from (a) frozen HIPETC; (b) multilayer CuIPETC on Cu metal;
(c) Cu metal native oxide conditioned 10 min in 104 M collector; (d) chalcocite
conditioned 10 min in 104 M collector with flood-gun on and specimen floating.
(a) (b)
was slightly broader (1.1 eV). The binding energy of the principal
doublet was close to that expected for a >C@S functional group
adjacent to an ether and amine group based on the values of
162.0 and 162.4 eV observed for thiourea (Lee and Rabalais,
1977; Srinivasan and Walton, 1977).
The C:O:N:S atomic ratio estimated from the C 1s, O 1s, N 1s and
S 2p intensities using theoretical photoionisation cross-sections in
the Avantage data processing software was 7.2:1.2:1.0:1.2, close to
the expected 6:1:1:1 when the influence of impurities and system-
atic errors arising from the theoretical cross-sections were taken
408 405 402 399 396 408 405 402 399 396
into account. The photoelectron spectra showed that the purity
of the commercial collector was acceptable for the purposes of this
Binding Energy (eV) Binding Energy (eV)
investigation. In particular, there was no evidence that DSP 009
was a mixture of isopropyl xanthate and ethyl amine as had been
(c) (d) raised as a possibility for Z-200 by Ackerman et al. (1984). The N 1s
binding energy for any ethyl amine present would be 399.1 eV
(Barber et al., 1973) rather than the 399.9 eV value observed, and
a significant concentration of a species with N 1s binding energy
0.8 eV lower than that for IPETC would give rise to a discernible
asymmetry to the photoelectron peak. The N 1s spectrum observed
was essentially symmetrical.

3.2. Residue of collector allowed to evaporate under UHV

In order to obtain data for any low vapour pressure IPETC


decomposition products or impurities, a droplet of the commercial
408 405 402 399 396 408 405 402 399 396
IPETC collector (DSP 009) was allowed to evaporate on an Al sub-
Binding Energy (eV) Binding Energy (eV) strate at ambient temperature under the vacuum of the photoelec-
Fig. 3. N 1s spectrum from (a) frozen HIPETC; (b) multilayer CuIPETC on Cu metal;
tron spectrometer introduction chamber. Spectra from the residue
(c) Cu metal native oxide conditioned 10 min in 104 M collector; (d) chalcocite were determined with the flood-gun on. The C:O:N:S atomic ratio
conditioned 10 min in 104 M collector with flood-gun on and specimen floating. was different for different evaporated droplets, but was within the
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 125

range 47 ± 7:9 ± 2:1:5. Hence, the relative intensity of the N 1s (a) (b)
peak was markedly lower than for the frozen collector, indicating
at least one low vapour pressure residue that contained little or
no N. When adjusted for the shift caused by the flood-gun elec-
trons, the C 1s spectrum was broadly similar to that for the frozen
collector apart from the high binding energy region. Instead of a
component near 288.4 eV of intensity similar to that at
286.85 eV, there was a well-resolved component at 289.3 eV
accounting for 8% of the intensity in addition to a minor (2%)
unresolved component near 288 eV. The S 2p spectrum could be
fitted adequately with two components, one at 162.7 eV (13%)
and the other at 164.3 eV (87%). The N 1s spectrum consisted of
a single peak at 400.5 eV. The observation that there was no
908 912 916 920 924 928 908 912 916 920 924 928
noticeable S 2p3/2 component at 162.2 eV or N 1s component at
399.9 eV indicated that there was minimal HIPETC remaining in Kinetic Energy (eV) Kinetic Energy (eV)
the residue. The O 1s spectrum consisted primarily of two compo-
nents; one near 533.7 eV (40%) consistent with ether type O and (c) (d)
the other near 532.25 eV, a binding energy the same as that
observed for carbonyl O in an ester (López et al., 1991).
The minor residue species was broadly consistent with the
declared impurity 1,3-diethyl-2-thiourea, which has a melting
point of 77 °C, although the N 1s binding energy was 0.2 eV
higher than might have been expected. The C 1s, N 1s and S 2p3/2
binding energies for thiourea are 288.1, 400.1 and 162.4 eV
(Srinivasan and Walton, 1977), and the N 1s binding energy for
diethyl thiourea might be expected to be similar. The major species
was likely to have contained S and O but not N, and with compara-
ble concentrations of O in two different environments such as
those in an ester. From the S 2p binding energy near 164.3 eV it 908 912 916 920 924 928 908 912 916 920 924 928
is also likely that the species contained a disulfide group, and Kinetic Energy (eV) Kinetic Energy (eV)
hence the most probable major residue species would have been
isopropyl ‘carbonate disulfide’. On that basis, the most likely Fig. 5. Cu L3M4,5M4,5 spectrum from (a) Cu metal native oxide conditioned 10 min
decomposition of IPETC would have been initially to the monothio- in 104 M collector; (b) multilayer CuIPETC on Cu metal; (c) unconditioned abraded
chalcocite; (d) chalcocite conditioned 10 min in 104 M collector.
carbonate, reaction (4), followed by oxidation of the monothiocar-
bonate to the ‘carbonate disulfide’ (Murphy and Winter, 1973):

2C3 H8 OCðOÞSH þ 0:5O2 ! C3 H8 OCðOÞSASðOÞCOC3 H8 þ H2 O 3 nm depth analysed by those Auger electrons, and the O 1s spec-
ð5Þ trum showed that the native oxide layer was also a minor compo-
nent within the depth analysed. The main Cu L3M4,5M4,5 peak was
The fact that a S 2p component near 164 eV was not observed
at a kinetic energy of 915.0 ± 0.2 eV. The elemental composition of
for any of the specimens conditioned in DSP 009 establishes that
the surface layer, estimated on the basis of theoretical photoionisa-
any carbonate disulfide oxidation product present in the collector
tion cross-sections, was consistent with a stoichiometry of CuIPETC
would not have affected the findings of the investigation reported
apart from a high O concentration that would have been aug-
here. However, it is likely that a minor concentration of isopropyl
mented by some hydroxylated native oxide within the depth ana-
monothiocarbonate was present as a decomposition product in
lysed. Clearly, adsorption of the collector had not removed the
the collector, and therefore a low concentration of monothiocar-
surface oxide, despite the high collector concentration.
bonate might have co-adsorbed with the IPETC, thereby adding
The C 1s spectrum (Fig. 2b) was fitted with four components of
to the complexity of the O 1s (but not N 1s) spectrum, and making
equal width, and when corrected for charge-shifting of the thick
minor contributions to the C 1s and S 2p spectra. In particular,
layer of CuIPETC, the components were at 285.0 (assigned),
some intensity near 532.2 eV in the O 1s spectrum and near
285.75 ± 0.15, 286.7 ± 0.1 and 287.35 ± 0.05 eV, with the uncer-
289.3 eV in the C 1s spectrum could have arisen from minor co-
tainties depending on the constraints placed upon the relative
adsorbed monothiocarbonate, but C 1s intensity near 289 eV could
intensities of the three highest binding energy components. The
also have been from charge-shifted adsorbed HIPETC in poor elec-
energy values were not very sensitive to the intensity of the high-
trical contact with a conditioned surface.
est binding energy component being constrained to the same as
only the next highest, the next two highest (shown in Fig. 2b), or
3.3. Multilayer CuIPETC no other C 1s component. The most important difference between
the C 1s spectrum for HIPETC and multilayer CuIPETC is the posi-
Multilayer CuIPETC was prepared by treating a Cu metal sur- tion of the highest binding energy component, with that of the for-
face, freshly polished with alumina in air, with 0.01 M aqueous mer being 1 eV higher than for the latter.
DSP 009 for 10 min. It is pertinent to note that when the treated The N 1s spectrum determined at the outset was essentially a
surface was rinsed with water prior to characterisation it was single, narrow (1.2 eV) peak at 398.7 eV (Fig. 3b), corresponding
clearly hydrophobic. Multilayer coverage was evident from the to a single N electronic environment and confirming that the N
photoelectron and X-ray excited Auger electron spectra deter- in CuIPETC was deprotonated. Less than 3% of the N 1s intensity
mined in suites 1 and 2. The Cu L3M4,5M4,5 spectrum (Fig. 5b) con- was in a component at 401.2 eV. The S 2p spectrum, too, indi-
firmed, from a barely discernible component at 918.3 eV, that cated a single S electronic environment; the spectrum could be fit-
only a very low concentration of Cu metal remained within the ted with a principal doublet with relatively narrow component
126 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

linewidth (<1.2 eV) at a 2p3/2 binding energy of 162.6 eV (Fig. 4b). shifted by 0, 0.9, 1.9, 3.0 and 4.1 eV, but because the shifted com-
An attempt to fit a second doublet at 162.2 eV indicated that neg- ponents were not resolved, the fits would not have been unique.
ligible HIPETC could have been adsorbed. Also, notwithstanding
the substantial multilayer coverage, the absence of a S 2p doublet
near 164 eV revealed that there was no (IPETC)2 adsorbed at the 3.4. Surfaces conditioned in IPETC solutions under conditions of
solid/vacuum interface (and there had been no evidence for the relevance to flotation
loss of a high vapour pressure adsorbate species during specimen
evacuation). The O 1s spectrum was dominated by a component While it is the adsorption of IPETC, or any other sulfhydryl col-
at 533.25 eV that could be assigned to adsorbed IPETC. A compo- lector, on sulfide minerals that is of principal industrial interest,
nent near 530 eV that would have arisen from oxide in the native the adsorption on a surface of the relevant metal freshly prepared
oxide layer accounted for only 6% of the O 1s intensity, but a com- in air can provide information that assists the interpretation of
ponent at 531.6 eV can be attributed to hydroxyl groups at the sur- spectra from sulfide minerals conditioned in that collector. In the
face of the native oxide layer. particular case of IPETC, it is its adsorption on a freshly-formed
In the second spectral suite with the flood-gun on and specimen native CuI oxide layer on Cu metal, or on cuprite (Cu2O), that would
earthed, all spectra were shifted to lower apparent binding energy be expected to be of use in helping to elucidate the mechanism of
by different amounts, up to the 4 eV energy of the flood-gun elec- interaction with CuI in the Cu sulfide minerals such as chalcocite
trons, depending on the degree of electrical contact between the Cu and chalcopyrite. It is clear from §3.3 that IPETC adsorbs readily
metal substrate and the CuIPETC layers. For example, the Cu 2p on air-exposed Cu metal. The adsorption of IPETC on pyrite is also
spectrum (Fig. 6a) showed that about 20% of the Cu, within the relevant given that pyrite is a common sulfide gangue mineral
2 nm depth analysed by the 555 eV kinetic energy photoelec- from which the Cu sulfides should be separated. In general, low
trons, gave rise to 2p intensity that was essentially unshifted, concentrations of sulfhydryl collectors resulting in sub-monolayer
and therefore would have been in Cu metal, Cu native oxide, or coverage are of greatest industrial relevance, but for research pur-
Cu to which IPETC was chemisorbed. The remaining 80% was poses it is important to also include higher than normal concentra-
shifted by 0.9–4.1 eV, and it is likely that the intensity shifted by tions to investigate the effect of any multilayer collector
0.9 eV was from the CuIPETC adsorbed immediately on top of the adsorption. Another reason multilayer coverage is of interest is in
chemisorbed IPETC, while the intensity shifted by 4 eV was from the determination of whether any (IPETC)2 is formed and adsorbed.
CuIPETC several layers remote from the chemisorbed collector. All It is possible that the vapour pressure of (IPETC)2 might be suffi-
spectra could be fitted with five components of typical width ciently high at ambient temperature for any adsorbed at low col-
lector coverage to be lost from the solid/vacuum interface in the
spectrometer. However, for substantial multilayer coverage, some
of any adsorbed (IPETC)2 would be expected to desorb slowly
(a) (b) enough to be detected at the start of the analysis period.

3.4.1. Cu metal
A copper metal surface, freshly polished in air with alumina,
was conditioned for 10 min in an aliquot of the same 104 M IPETC
solution as used for abraded chalcocite (Section 3.4.3). The rinsed
surface was clearly hydrophobic. The photoelectron spectra
(Figs. 2c and 4c) and Cu L3M4,5M4,5 Auger electron spectrum
(Fig. 5a) indicated that collector coverage was marginally higher
(4.1 at.% N) than for the similarly treated chalcocite (3.6 at.% N),
and that the N was predominantly deprotonated in IPETC chemi-
sorbed to Cu atoms with some molecular CuIPETC on top of the
965 955 945 935 925 965 955 945 935 925 chemisorbed collector. The spectra also showed that a substantial
Binding Energy (eV) Binding Energy (eV) native oxide layer remained.
Other Cu metal surfaces freshly polished in air were condi-
tioned for 10 min in 103 M IPETC solution. After being rinsed
(c) (d)
with water, the surfaces were clearly hydrophobic, and with
4.9 at.% N, collector coverage was greater than for the surface con-
ditioned for the same period in 104 M solution, but the native
oxide layer had a thickness similar to that for the specimen condi-
tioned in 104 M solution. The spectral behaviour under the influ-
ence of the flood-gun electrons showed that the adsorbed species
would have been mostly multilayer CuIPETC on top of chemisorbed
IPETC with only a minor concentration of HIPETC.

3.4.2. Cuprite (Cu2O)


Cuprite fracture surfaces conditioned for 10 min in 103 M
IPETC solution were too rough for their hydrophobicity to be
965 955 945 935 925 965 955 945 935 925
assessed by visual examination. The N and S concentrations (3.2–
Binding Energy (eV) Binding Energy (eV) 3.7 at.%) at these surfaces were comparable, as expected, but there
was evidence for preferential desorption of N relative to S with
Fig. 6. Cu 2p spectrum determined in (a) suite 2 or (b–d) suite 1 from (a) multilayer
CuIPETC; (b) chalcocite conditioned 10 min in 104 M collector; (c) covellite
increasing analysis time. The spectra indicated that most N was
conditioned 10 min in 103 M collector; (d) chalcopyrite conditioned 10 min in deprotonated in chemisorbed IPETC, with only minor HIPETC
103 M collector. adsorbed.
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 127

3.4.3. Chalcocite (Cu2S) (a) (b)


Freshly abraded chalcocite surfaces conditioned for 10 min in
104 M IPETC solution appeared hydrophobic when rinsed with
water. The surface layer elemental composition included
3.6 at.% N. The spectra (Figs. 3d, 5c, d and 6b) revealed the pres-
ence of comparable concentrations of chemisorbed IPETC and
HIPETC. The absence of adsorbed molecular CuIPETC was con-
firmed by the similarity, within experimental variability, of the
Cu L3M4,5M4,5 spectra for the unconditioned and conditioned chal-
cocite surfaces (Fig. 5c and d), in conjunction with less surface sen-
sitive S 2p and N 1s components from chemisorbed collector. The
spectra also showed that for no greater than monolayer coverage,
the collector S interacted with a Cu atom in the mineral surface,
not a S atom as argued by Crozier (1991). 740 730 720 710 740 730 720 710
Freshly abraded chalcocite surfaces conditioned for 10 min in Binding Energy (eV) Binding Energy (eV)
103 M and 102 M IPETC solutions also appeared to be hydropho-
bic after being rinsed with water. For the 103 M solution, the sur-
(c) (d)
face N concentration was 6.4 at.%. The spectra indicated that there
was some HIPETC adsorbed as well as chemisorbed IPETC and
minor CuIPETC.

3.4.4. Covellite (CuS)


Surfaces prepared initially by cleavage parallel or perpendicular
to the basal {0 0 1} plane, and subsequently abraded, might be
expected to have retained some preferential orientation rather
than have become completely random. The S(2)AS(2) distance in
covellite is 0.207–0.209 nm, significantly shorter than the
0.214 nm in pyrite, so the covellite SAS bonds are expected to be
stronger than the CuAS bonds and hence be less likely to cleave 740 730 720 710 740 730 720 710
(Rosso and Hochella, 1999). On that basis, a basal plane fracture Binding Energy (eV) Binding Energy (eV)
surface would be expected to be dominated by broken CuAS bonds
parallel to the c-axis rather than broken disulfide bonds, and hence Fig. 7. Fe 2p spectrum from (a) chalcopyrite or (b–d) pyrite (b) abraded or (a, c and
significant reconstruction of the two different opposing fracture d) fracture surface conditioned for 10 min in (a–c) 103 M collector; (d) 102 M
collector.
surfaces would be anticipated. However, Ohmasa et al. (1977)
argued that the highly perfect cleavage characteristic of covellite however, that even unconditioned abraded chalcopyrite surfaces
could be ascribed to the short (0.219 nm) trigonal Cu(1)AS(1) are usually hydrophobic to some extent because of the sulfur-rich
bonds being stronger than the more typical tetrahedral 0.234 nm CuFe1xS2 regions formed during air exposure (Zachwieja et al.,
Cu(2)AS(1) and 0.231 nm Cu(2)AS(2) bonds. They also noted that 1989). The N concentration at the conditioned surfaces was
the S(2)AS(2) distance was longer than for VS4 (0.203–0.204 nm), 2.2 at.%, higher than for fresh fracture surfaces treated for the
but they did not speculate on the more likely bond broken during same period in collector solutions of similar concentration. The
cleavage of covellite. spectra showed that there was no molecular CuIPETC adsorbed,
Freshly abraded covellite surfaces, that initially had been either but there was clear evidence for monolayer and multilayer HIPETC,
parallel or perpendicular to the {0 0 1} cleavage plane, conditioned the predominant adsorbed collector species. Similar species were
for 10 min in 103 M IPETC solution and subsequently rinsed with adsorbed on freshly abraded chalcopyrite surfaces conditioned
water, appeared to be slightly hydrophobic. For the conditioned for 1 min in 103 M IPETC solution or 10 min in 102 M solution.
abraded edges, the N concentration was 1.9 at.%, and the spectra In each case, the conditioned surface was hydrophobic.
(including Fig. 6c) showed that, surprisingly, the predominant A chalcopyrite fracture surface, exposed to air for 60 min prior
adsorbed species was HIPETC rather than IPETC chemisorbed to to conditioning for 10 min in another aliquot of the same 103 M
Cu atoms. IPETC solution that had been used for conditioning a fresh fracture
The spectra from conditioned abraded surfaces predominantly surface for 10 min, was too rough to assess its hydrophobicity. Col-
parallel to the basal plane indicated very low (1.3 at.%) initial lector coverage was low (1.7 at.% N), comparable with that
concentrations of N, predominantly in adsorbed HIPETC as well observed for a chalcopyrite fresh fracture surface that had been
as minor chemisorbed IPETC. conditioned for the same period in another aliquot of the same
dilution. The spectra showed that more than 80% of the collector
3.4.5. Chalcopyrite (CuFeS2) was strongly adsorbed as HIPETC, with minor IPETC chemisorbed
A chalcopyrite fresh fracture surface conditioned for 10 min in to Cu atoms.
103 M IPETC solution, then rinsed with water, was too rough for Another chalcopyrite fracture surface exposed to air for 7 d was
a visual assessment of its hydrophobicity. Collector adsorption conditioned for 10 min in 102 M IPETC solution. The surface was
was quite low (1.6 at.% N), and the spectra (including Figs. 6d, too rough to assess its hydrophobicity. The spectra indicated that
7a and 8a) indicated that about 70% of the collector was adsorbed collector coverage was very low (1.2 at.% N), and were consistent
as HIPETC. The HIPETC was in good electrical contact with the with collector that was strongly adsorbed primarily as HIPETC.
surface, perhaps because it was adsorbed directly on the FeAO or
CuFe 1xS2 rather than on top of chemisorbed IPETC.
Freshly abraded chalcopyrite surfaces were conditioned for 3.4.6. Cubanite (CuFe2S3)
10 min in 103 M IPETC solution, and when subsequently rinsed The reactivity of freshly exposed cubanite surfaces has been
with water, the surfaces appeared to be slightly hydrophobic. Note, found to be comparable with that for chalcopyrite, and the Cu, Fe
128 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

and S core electron binding energies are similar to those for chal- (a) (b)
copyrite (Goh et al., 2010). While the (Cu + Fe):S ratio is 1:1 in both
minerals, the Cu concentration in cubanite is lower, and hence the
CuAOHsurf concentration for air-exposed surfaces might also be
expected to be lower for cubanite because of the comparable rates
of FeAO formation.
Cubanite surfaces that had been abraded in air, conditioned for
10 min in 103 M IPETC solution and subsequently rinsed with
water were moderately hydrophobic. This hydrophobicity was
despite a significant concentration of FeAO species within the
depth analysed, but no SAO species were retained at the rinsed
surface. The surface N concentration was 1.8 at.%, and 80% of
the adsorbed collector was present as HIPETC in good electrical 171 168 165 162 159 156 171 168 165 162 159 156
contact with the mineral substrate. There was no evidence for Binding Energy (eV) Binding Energy (eV)
any adsorbed molecular CuIPETC.
(c) (d)
3.4.7. Bornite (Cu5FeS4)
Freshly abraded bornite surfaces were conditioned for 10 min in
103 M IPETC solution, and after being rinsed with water, were
clearly hydrophobic. The obvious hydrophobicity was surprising
given the low N concentration (1.2 at.%) and hence low collector
coverage. More than 80% of that coverage had been adsorbed as
HIPETC, and the spectra also revealed the presence of FeOOH
patches rather than a uniform layer of FeAO species across the sur-
face. No molecular CuIPETC was detected.
Abraded bornite surfaces exposed to air for 1 h prior to condi-
tioning for 10 min in 103 M collector solution were clearly hydro-
phobic. The surface N concentration was 1.8 at.%, and the spectra 171 168 165 162 159 156 171 168 165 162 159 156
were consistent with the adsorption of predominantly HIPETC,
Binding Energy (eV) Binding Energy (eV)
with less than 5% as chemisorbed IPETC.
A bornite fracture surface exposed to air for 25 h was condi- Fig. 8. S 2p spectrum from (a) chalcopyrite or (b–d) pyrite (b) abraded or (a, c and
tioned for 10 min in 103 M collector solution. The fracture surface d) fracture surface conditioned for 10 min in (a–c) 103 M collector; (d) 102 M
was rough, but fine particles broken off by the agitation during collector.

conditioning accumulated at the liquid/air interface and hence


were deduced to be hydrophobic. Almost all of the Fe within the
formed after the surface had been conditioned, as negligible SAO
depth analysed by the Fe 2p photoelectrons was present as FeAO
species were retained at other surfaces conditioned in IPETC
species. The N concentration was 2.0 at.%. The spectra indicated
solution.
about 70% of the adsorbed collector was present as HIPETC in good
For the fracture surface conditioned in 102 M IPETC solution,
electrical contact with the mineral surface; the other 30% was pres-
collector coverage was found to be low, but not negligible at
ent as chemisorbed IPETC and molecular CuIPETC with some of the
1.8 at.% N, with 85% present as HIPETC. The collector coverage
multilayer CuIPETC in poor electrical contact with the surface.
on a Cu metal surface polished in air and treated with another ali-
quot of the same IPETC solution under exactly the same conditions
3.4.8. Pyrite (FeS2)
had been substantial (6.6 at.% N). In contrast to the fracture surface
Freshly abraded surfaces of pyrite conditioned for 10 min in
conditioned in 103 M solution, this surface layer contained a
102 M or 103 M IPETC solution were not obviously hydrophobic.
greater concentration of FeAO species (with Fe 2p3/2 peak maxi-
After conditioning in 103 M solution, the surface N concentration
mum at 711.0 eV) but negligible SAO species (Figs. 7d and 8d). Oxi-
was 1.5 at.%, with the collector adsorbed largely as monolayer
dation of the pyrite surface while in the collector solution would
HIPETC. The concentration of FeAO species at the rinsed surface
probably not have occurred if the fracture surface had been ‘pro-
was relatively low (Fig. 7b), and there were no SAO species evident
tected’ by an appreciable coverage of collector.
(Fig. 8b).
Fresh fracture surfaces of pyrite conditioned for 10 min in 103
or 102 M IPETC solution were too rough to reliably assess their 4. Discussion
hydrophobicity. Fracture surfaces of this relatively pure Navajún
pyrite exposed only to air for 10 min became oxidised to a negligi- 4.1. Identity of collector species adsorbed
ble extent, but surfaces conditioned in the collector solution
became substantially oxidised, and at least some of the oxidation The predominant collector species adsorbed on each (air-
products were retained at the surface after rinsing with water. exposed) surface investigated, as deduced from the electron spec-
For the fracture surface conditioned in 103 M IPETC solution, the tra, are summarised in Table 2. The N concentration resulting from
N surface concentration (0.7 at.%) revealed that collector coverage conditioning for 10 min in 103 M collector solution is also tabu-
was considerably lower than for similarly conditioned abraded sur- lated as an indicator of the relative degree of coverage of each sur-
faces. The predominant adsorbed collector species was HIPETC. The face by all adsorbed collector species, but note that the different
concentrations of FeAO and SAO species were both appreciable mineral compositions and extents of surface oxidation should be
(Figs. 7c and 8c), however, the retention of such a high concentra- taken into account in any direct comparison of coverage based
tion of SAO species was unusual, and it is possible that they had on the N concentration. The predominant collector species varied
been present as an inclusion at the fresh fracture surface prior to from chemisorbed IPETC and multilayer CuIPETC for Cu metal
conditioning. Note that the SAO species would not have been native CuI oxide and cuprite, through comparable concentrations
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 129

Table 2
Summary of species adsorbed on substrate surface (freshly abraded, 10 min conditioning in 103 M collector, unless otherwise specified) from IPETC solution. In all cases where
HIPETC was the main adsorbed species, it was mostly not adsorbed on top of chemisorbed IPETC.

Substrate Main adsorbed collector species at.% N Coverage (0.001 M collector solution)
Cu native oxide Chemisorbed IPETC and CuIPETC 4.9 Multilayer
Cuprite (fracture) Chemisorbed IPETC 3.7 Monolayer
Chalcocite Comparable CuIPETC on chemisorbed IPETC and HIPETC 6.4 Multilayer
Covellite HIPETC 1.9 Monolayer
Chalcopyrite HIPETC 2.2 Multilayer
Chalcopyrite (fracture) HIPETC 1.7 Monolayer
Chalcopyrite (fracture; 7 d air-exposure; 102 M) HIPETC 1.2 Monolayer
Cubanite HIPETC 1.8 Monolayer
Bornite HIPETC 1.2 Monolayer
Bornite (1 h air-exposure) HIPETC 1.8 Monolayer
Bornite (fracture; 25 h air) HIPETC (with minor CuIPETC on chemisorbed IPETC) 2.0 Multilayer
Pyrite HIPETC 1.5 Monolayer
Pyrite (fracture) HIPETC 0.7 Monolayer

of HIPETC and CuIPETC on chemisorbed IPETC for chalcocite, to adsorbed on chemisorbed IPETC remained a minority adsorbed col-
HIPETC for covellite, the CuAFe sulfides and pyrite. This progres- lector species. The adsorption of predominantly HIPETC on chalco-
sion in predominant adsorbed collector species from chemisorbed pyrite was in partial agreement with the conclusions of Fairthorne
IPETC and CuIPETC to HIPETC correlates with a decrease in the sur- et al. (1997); they estimated that approximately two layers of
face availability of CuAOHsurf to facilitate reaction (6): IPETC were adsorbed on chalcopyrite at pH 5.

CuAOHsurf þ HIPETC ! CuAIPETCads þ H2 O ð6Þ


4.2. IPETC collector adsorption mechanisms
where CuAIPETCads represents the collector chemisorbed to surface
Cu atoms. In air, hydroxyl groups chemisorbed to Cu atoms at the
In order to unequivocally corroborate the chemisorption of
surface of a Cu sulfide mineral such as chalcocite could result from
IPETC to Cu atoms in a mineral surface, rather than collector bond-
the reaction of the mineral surface with oxygen:
ing to surface S atoms, or interaction by a dissolution–precipitation
Cu2 S þ 0:25xO2 ! Cu2x S þ 0:5xCu2 Osurf ð7Þ mechanism, it would be necessary to limit coverage to monolayer
at most, but to have sufficient coverage for both the N 1s and S 2p
followed by the reaction of the Cu2O at the solid/air interface with spectra from the adsorbed collector to be well-resolved, and to
water vapour: confirm equal surface concentrations of N and S from the adsorbed
collector. Equal concentrations of N and S would be necessary to
Cu2 Osurf þ H2 O ! 2CuAOHsurf ð8Þ
eliminate the possibility of collector decomposition. The N and C
While such a dependence of adsorbed collector species on the 1s spectra would need to establish that a substantial proportion
availability of CuAOHsurf might be obvious in the case of pure pyr- of the collector had been deprotonated (and hence not adsorbed
ite that had not been Cu-activated, it is not at all self-evident for a as HIPETC). For the S 2p component from IPETC to be resolved from
mineral such as covellite. Also, it might not have been expected the S 2p spectrum from a sulfide mineral substrate, the latter
that the species adsorbed on chalcocite would be different from would need to have no components with 2p3/2 binding energy
those adsorbed on freshly exposed surfaces of the Cu-rich mineral above about 161.8 eV. Under those sub-monolayer conditions,
bornite. Furthermore, there appeared to be a second-order differ- the absence of a Cu L3M4,5M4,5 Auger component near 915 eV
ence in the HIPETC physically adsorbed as a multilayer species would confirm the absence of adsorbed CuIPETC, and hence pre-
on top of chemisorbed IPETC (such as on chalcocite), and HIPETC clude a dissolution–precipitation mechanism. Those criteria were
adsorbed in good electrical contact with the surface (and hence best satisfied by a chalcocite substrate conditioned in collector
most probably adsorbed strongly) as a monolayer species (such solution more dilute than 103 M.
as on pyrite). For none of the surfaces investigated was there any When chalcocite was conditioned for 10 min in 104 M IPETC
evidence for the adsorption of (IPETC)2, however, it is possible that solution, the C 1s, N 1s and S 2p spectra confirmed the adsorption
when the coverage was low, this species might have been adsorbed of both protonated (HIPETC) and deprotonated (IPETC) collector,
at the solid/solution interface but, because of an insufficiently low without the formation of S-S bonds, and with concentrations of N
vapour pressure, lost from the solid/vacuum interface at ambient and S that were equal within experimental error. However, the
temperature. Cu L3M4,5M4,5 Auger spectrum showed that no CuIPETC had been
The predominant collector species (HIPETC) adsorbed on exten- formed, therefore the deprotonated collector would have been
sively air-exposed chalcopyrite or moderately air-exposed bornite IPETC chemisorbed to Cu atoms in the chalcocite surface. The Cu
was similar to that on fresh chalcopyrite or bornite surfaces. This is L3M4,5M4,5 Auger electrons (915 eV kinetic energy) are more sur-
somewhat surprising, as it might have been expected that there face sensitive than the S 2p photoelectrons (1325 eV kinetic
would have been a difference similar to that exhibited by CuAFe energy), therefore the absence of a component at 915 eV kinetic
sulfide surfaces towards hydroxamate collector adsorption, which energy cannot be due to insufficient sensitivity. These observations
was influenced by increased air exposure leading to FeAO patches for chalcocite constitute definitive evidence for the chemisorption
and more exposed Fe-deficient Cu sulfide regions between the of IPETC rather than the adsorption of sub-monolayer CuIPETC, and
oxide patches (Buckley et al., 2014). The only sign of that behaviour hence definitive evidence for the inapplicability of a dissolution–
was displayed by the bornite fracture surface exposed to air for precipitation mechanism for this system. Similar behaviour was
25 h. At that surface, essentially all the Fe would have been present observed for at least some of the other Cu sulfide minerals, how-
in patches of FeAO species, thereby exposing Fe-depleted Cu sul- ever the evidence was not as definitive, as although chemisorption
fide regions that probably would have presented CuAOHsurf for of IPETC was revealed by the N 1s spectrum, and the absence of
interaction with HIPETC during the conditioning. Even so, CuIPETC CuIPETC confirmed by the Cu L3M4,5M4,5 Auger spectrum, the
130 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

adsorption of IPETC could not be unequivocally corroborated by (0.306 to 0.352) calculated by Liu et al. (2008) and did not
the S 2p spectrum because of the complexity of the component appear to be comparable.
from the mineral substrate. Broadly similar behaviour was The N 1s binding energy in multilayer CuIPETC, 398.7 eV, is
observed for cuprite and native CuI oxide on Cu metal, too, but con- quite low and characteristic of N that has not only been deproto-
ditioning of the latter for 10 min in 104 M solution resulted in suf- nated but also is not interacting with a metal atom. If there had
ficient coverage for minor multilayer CuIPETC to be adsorbed on been significant interaction between deprotonated N and Cu, a N
top of the major chemisorbed IPETC. 1s binding energy near 399.4 eV would have been expected. There-
Controversially, Crozier (1991) had argued that sulfhydryl col- fore there is no evidence for significant interaction between either
lector–sulfide mineral bonding was a S-to-S bond rather than a the O or N and the Cu to which the S is bonded or to a Cu in a neigh-
S-to-metal bond, and that such SAS bonds could not be distin- bouring molecule in the multilayer.
guished spectroscopically from that in dithiolates. While the latter The O and N 1s binding energies observed for chemisorbed
statement might be true, when dithiolates are not present, the IPETC were within the ranges 533.1–533.5 eV and 398.3–
absence of SAS bonds (as observed in the present investigation) 398.8 eV. Each energy spread probably reflected the uncertainty
provides confirmation that collector chemisorption is via bond for- in fitting a minor unresolved component of the O 1s or N 1s spec-
mation between the S in the collector and a metal atom in the min- trum rather than interaction with the particular mineral, neverthe-
eral surface. less, the N 1s component was usually near 398.3 eV for
The C 1s, N 1s and S 2p spectra from CuI native oxide on Cu chalcopyrite, 398.5 eV for chalcocite and covellite, 398.6 eV for
metal or cuprite conditioned in dilute IPETC solution revealed the cuprite, and 398.8 eV for bornite and pyrite. Within experimental
adsorption of predominantly deprotonated IPETC, without the uncertainty, however, the O and N 1s binding energies were similar
wholesale removal of the native oxide layer. This observation for IPETC chemisorption as for multilayer CuIPETC, indicating that
was in agreement with previous reports. For cuprite, the Cu there was no significant interaction between the O or N and a Cu
L3M4,5M4,5 Auger spectrum revealed the presence of either IPETC atom in that sub-monolayer species either. This finding for pH
chemisorbed to Cu at the surface of the oxide lattice or molecular 6 is broadly in agreement with the conclusions of Woods and
CuIPETC, or both, as those two species could not be differentiated Hope (1999) and Leppinen et al. (1988).
on the basis of the Auger spectrum alone. The fate of O atoms at It would be expected that when HIPETC adsorbed on top of
the surface of Cu metal, CuI oxide or air-exposed Cu sulfide miner- chemisorbed IPETC, an alkyl group of the HIPETC would interact
als during conditioning raises the question as to whether the col- physically with an alkyl group of the chemisorbed IPETC. That rel-
lector chemisorbs directly to a Cu in the oxidised metal or atively weak interaction would be consistent with the observed
cuprite (and possibly Cu sulfide mineral) surface, as in reaction poor or moderate electrical contact between the HIPETC and the
(6), or whether it forms a SAOACu bond. The S 2p spectrum, with substrate. On the other hand, when minimal IPETC was chemi-
a 2p3/2 binding energy of 162.6 ± 0.1 eV, shows unequivocally that sorbed, and HIPETC adsorbed directly on to a mineral surface, elec-
the SACu bond is direct. The 2p3/2 binding energy for a SAS bond trical contact between the HIPETC and the mineral surface
(in [DTP]2, for example) is no lower than 164.0 eV, and the corre- appeared to be good. It is possible that even though the HIPETC
sponding binding energy for a SAO bond would be at least as high. remained protonated, interaction with the mineral surface was
A corollary of direct SACu bond formation with predominant via the S atom. Such an interaction would be weaker than the CuAS
native CuI oxide retention is collector adsorption in patches, with bond in chemisorbed IPETC, but significantly stronger than the
adsorption of multilayer CuIPETC without the completion of a uni- alkyl–alkyl interaction for HIPETC physically adsorbed on top of
form monolayer of chemisorbed IPETC. chemisorbed IPETC.
The increase in the S 2p3/2 binding energy from 162.2 eV in In their molecular modelling of the in vacuo interaction of IPETC
HIPETC to 162.6 eV in multilayer CuIPETC and the chemisorbed collector with pyrite (represented by an isolated FeS2 or cyclic
collector indicated a reduction in the electron density at S atoms FeS2AFeS2), Solozhenkin et al. (2013) found that in monodentate
in IPETC, consistent with the formation of a CuAS bond. Accord- bonding of the collector N with the pyrite Fe, charge was trans-
ingly, the XPS data supported the conclusion reached in almost ferred from the mineral to the collector S. By contrast, in bidentate
all previous studies of thionocarbamate adsorption that the princi- bonding of the collector N and S with the Fe, charge was trans-
pal interaction between the surface and the collector is through its ferred from the collector to the mineral. The situation for bidentate
S. However, the question remains as to whether there is also a sig- bonding of the collector S and O with the Fe, or for interaction of
nificant interaction of either N or O with Cu atoms. HIPETC with pyrite, was not reported, nevertheless their calcula-
For bulk HIPETC, the 1s binding energy for the C bonded to N, S tions suggested that IPETC should interact with the Fe in pyrite.
and O is 288.4 eV, and the O 1s binding energy is 533.4 eV. For mul- In principle, the direction of any charge transfer should be evident
tilayer CuIPETC, in which the N is deprotonated, the 1s binding from the collector S 2p3/2 binding energy, but in practice such
energy for C bonded to N, S and O is 1 eV lower (287.4 eV), indicat- information would not be available for pyrite because the mineral
ing that the electron density on that C atom is higher. A higher S 2p3/2 binding energy is almost the same as that of any adsorbed
electron density on the C would mean that the electron density IPETC. The direction of any charge transfer should also be evident
on the O would be unlikely to be lower, and could be slightly in the 1s binding energy of the C bonded to the collector S. For pyr-
higher, than in HIPETC, so that the O 1s binding energy would be ite, that C 1s binding energy was 288.3 eV, consistent with
expected to be no greater or slightly lower than 533.4 eV unless adsorbed HIPETC rather than IPETC chemisorbed to the pyrite Fe.
there was significant interaction between the O and a metal atom
to offset any increase in the electron density that might have 4.3. Correlation of adsorbed collector species with observed floatability
derived from the increase on the C atom. The observation of a mar- and implications for flotation
ginally lower O 1s binding energy in CuIPETC than in HIPETC sug-
gests no more than minor interaction between the O and Cu. Ab Apart from pyrite, conditioned surfaces that were smooth
initio calculations of the charge on the S atom in HIPETC were enough to assess by visual inspection appeared to be hydrophobic.
reported by Liu et al. (2008), but corresponding charges for the For not extensively oxidised abraded surfaces, the order of decreas-
other atoms in HIPETC and for IPETC were not. Solozhenkin ing coverage was chalcocite > native CuI oxide > cuprite > chalco-
et al. (2013) calculated atomic charges for HIPETC, but their value pyrite > covellite ffi cubanite > bornite ffi pyrite. For the sulfide
for S (-0.465) was significantly more negative than the values minerals, that order was broadly in agreement with the
A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132 131

microflotation data of Ackerman et al. (1999), and the (pH 8) the oxide layer from the air-exposed mineral surface. HIPETC could
adsorption data of Finkelstein and Goold (1972). It should be noted be adsorbed as a multilayer species on top of chemisorbed IPETC,
that for the mineral surfaces investigated in both forms, the cover- or when chemisorbed IPETC coverage was low, as a monolayer spe-
age on fracture surfaces was significantly different from that on cies in good contact with the surface. Surfaces bearing such
abraded surfaces. Coverage on an extensively oxidised surface adsorbed monolayer HIPETC were hydrophobic.
was usually significantly different from that on a freshly exposed The decrease in coverage of chemisorbed IPETC on the different
surface. Cu sulfide ore minerals was broadly in agreement with their previ-
The question of whether the coverage, and predominant ously observed decreasing floatability. However, the lack of obvi-
adsorbed collector species, on pyrite is consistent with its lack of ous hydrophobicity and the poor floatability of pyrite appeared
obvious hydrophobicity and the low recovery observed for that to be anomalous, in that the coverage of HIPETC on pyrite was
mineral relative to the Cu sulfides is a vexed one. Under the pH not markedly different from that on the CuAFe sulfides such as
conditions used, adsorption on pyrite was low, but perhaps not bornite. It was surmised that this anomaly might be attributable
as low as might have been expected relative to adsorption on born- to the different oxidation and alteration products retained at the
ite, especially as Navajún pyrite is not a particularly reactive pyrite. surface of conditioned pyrite.
The lack of obvious hydrophobicity for pyrite was surprising given
the hydrophobicity observed for bornite surfaces bearing compara- Acknowledgements
ble coverages of HIPETC. Even the form in which the collector was
adsorbed, strongly adsorbed HIPETC, was similar for pyrite and the The authors are grateful to Dr. Bill Gong (UNSW Analytical Cen-
CuAFe sulfides. Flotation tests have shown that at pH 6, IPETC col- tre) for assistance in obtaining the XPS data, and to Prof. Bill Skin-
lector affords good separation of chalcopyrite from pyrite, but the ner (Ian Wark Research Institute, University of South Australia) for
origin of that differentiation might be more than merely one of col- providing the cuprite and cubanite specimens.
lector coverage. It is surmised that reasons for the difference in flo-
atability might include the rate of collector adsorption as well as Appendix A. Supplementary material
the eventual coverage or the particular species adsorbed.
Leppinen et al. (1988) had suggested that the differentiation might Supplementary data associated with this article can be found, in
have been the inability of IPETC to remove oxidation products from the online version, at http://dx.doi.org/10.1016/j.mineng.20
the surface of pyrite, and while oxidation products are certainly 14.08.002.
not removed from pyrite surfaces, the photoelectron spectra have
shown that IPETC doesn’t remove oxidation products from the Cu References
sulfides or Cu metal either. It is true that the nature of the surface
oxidation products formed on pyrite and the CuAFe sulfides is dif- Ackerman, P.K., Harris, G.H., Klimpel, R.R., Aplan, F.F., 1984. Effect of alkyl
ferent, but it would be the nature of the alteration products substituents on performance of thionocarbamates as copper sulphide and
pyrite collectors. In: Jones, M.J., Oblatt, R. (Eds.), Reagents in the Minerals
retained at the surface under flotation conditions that would be Industry. Institute of Mining and Metallurgy, London, pp. 69–78.
of relevance. For pyrite, freshly-formed SAO species would nor- Ackerman, P.K., Harris, G.H., Klimpel, R.R., Aplan, F.F., 1999. Use of chelating agents
mally be expected to enter the aqueous phase, whereas some as collectors in the flotation of copper sulfides and pyrite. Miner. Metall.
Process. 16, 27–35.
freshly-formed FeAO species could be retained at the surface. For Avotins, P.V., Wang, S.S., Nagaraj, D.R., 1994. Recent advances in sulfide collector
the CuAFe sulfides, some freshly-formed FeAO species would enter development. In: Mulukutha, P.S. (Ed.), Reagents for Better Metallurgy. Society
the aqueous phase and some would be retained, but SAO species of Mining Engineers, Littleton, Colorado, pp. 47–56.
Barber, M., Connor, J.A., Guest, M.F., Hillier, I.H., Schwarz, M., Stacey, M., 1973.
are not formed initially, and patches of metastable Fe-deficient sul- Bonding in some donor–acceptor complexes involving boron trifluoride. Study
fide would be more prevalent at the mineral/solution interface by means of ESCA and molecular orbital calculations. J. Chem. Soc. Faraday
than for pyrite. As noted in Section 3.4.5, there is strong evidence Trans II 69, 551–558.
Beamson, G., Briggs, D., 1992. High Resolution XPS of Organic Polymers: The Scienta
for the S-rich (Fe-deficient) sulfide surface layer to be hydrophobic, ESCA300 Database. Wiley, Chichester, 295 p.
so the macroscopic mineral surface hydrophobicity is probably Bogdanov, O.S., Vainshenker, I.A., Podnek, A.K., Ryaboi, V.I., Yanis, N.A., 1976. Trends
determined by both the adsorbed collector and the hydrophobicity in the search for effective collectors. Tsvetn. Met. (English Translation) 17 (4),
79–85.
or hydrophilicity of the mineral surface regions not bearing
Buckley, A.N., 2010. Surface chemical characterisation for identifying and solving
adsorbed collector. Therefore while removal of oxidation products problems within base metal sulfide flotation plants. In: Greet, C.J. (Ed.),
by the collector does not appear to be a factor in determining the Flotation Plant Optimisation. AusIMM, Melbourne, pp. 137–153 (chapter 8).
effective hydrophobicity of conditioned sulfide minerals, the alter- Buckley, A.N., Woods, R., 1993. Underpotential deposition of dithiophosphate on
chalcocite. J. Electroanal. Chem. 357, 387–405.
ation products retained at the surface do appear to be important. Buckley, A.N., Woods, R., 1995. Identifying chemisorption in the interaction of thiol
collectors with sulfide minerals by XPS: adsorption of xanthate on silver and
5. Conclusions silver sulfide. Colloids Surf., A 104, 295–305.
Buckley, A.N., Goh, S.W., Skinner, W.M., Woods, R., Fan, L.-J., 2009. Interaction of
cuprite with dialkyl dithiophosphates. Int. J. Miner. Process. 93, 155–164.
Under the conditions investigated, the predominant thionocar- Buckley, A.N., Hope, G.A., Parker, G.K., 2014. A spectroscopic investigation of the
bamate collector species adsorbed on Cu sulfide ore minerals were interaction of n-octanohydroxamate collector with CuAFe sulfide minerals.
Miner. Eng. 64, 23–37.
found to vary from chemisorbed IPETC (and molecular CuIPETC for Crozier, R.D., 1991. Sulphide collector mineral bonding and the mechanism of
sufficiently high concentrations) on cuprite, through comparable flotation. Miner. Eng. 4, 839–858.
concentrations of HIPETC and chemisorbed IPETC on chalcocite, Fairthorne, G., Fornasiero, D., Ralston, J., 1996. Solution properties of
thionocarbamate collectors. Int. J. Miner. Process. 46, 137–153.
to HIPETC adsorbed on covellite, the CuAFe sulfides and pyrite. Fairthorne, G., Fornasiero, D., Ralston, J., 1997. Interaction of thionocarbamate and
This progression from chemisorbed IPETC to HIPETC correlated thiourea collectors with sulphide minerals: a flotation and adsorption study.
with a decrease in CuAOHsurf availability for exchange of surface Int. J. Miner. Process. 50, 227–242.
Finkelstein, N.P., Goold, L.A., 1972. The reaction of sulphide minerals with thiol
OH with IPETC. No evidence was observed for the adsorption
compounds. Report No. 1439, National Institute for Metallurgy, South Africa, 9
of (IPETC)2. p.
When monolayer IPETC was adsorbed, it was chemisorbed to Glembotskii, A.V., 1977. Theoretical bases for the prediction and modification of the
the Cu atoms in the mineral surface rather than to surface S atoms. properties of collectors. Tsvetn. Met. (English Translation) 18 (9), 68–72.
Goh, S.W., Buckley, A.N., Skinner, W.M., Fan, L.-J., 2010. An X-ray photoelectron and
Monolayer IPETC adsorption was not in the form of molecular CuI- absorption spectroscopic investigation of the electronic structure of cubanite,
PETC, and did not lead to the concomitant wholesale removal of CuFe2S3. Phys. Chem. Miner. 37, 389–405.
132 A.N. Buckley et al. / Minerals Engineering 69 (2014) 120–132

Harris, G.H., Fischback, B.C., 1954. Process for the manufacture of dialkyl Poling, G.W., 1976. Reactions between thiol reagents and sulphide minerals. In:
thionocarbamates. U.S. Patent No. 2,691,635. Fuerstenau, M.C. (Ed.), Flotation – A.M. Gaudin Memorial Volume, vol 1. AIME,
Hope, G.A., Woods, R., Watling, K., 2000. Surface enhanced Raman scattering New York, pp. 334–363 (chapter 11).
spectroscopic studies of the adsorption of flotation collectors. Proc. Rosso, K.M., Hochella Jr., M.F., 1999. A UHV STM/STS and ab initio investigation of
Electrochem. Soc. 14, 48–59. covellite 001 surfaces. Surf. Sci. 423, 364–374.
Hope, G.A., Woods, R., Parker, G.K., Buckley, A.N., McLean, J., 2011. Spectroscopic Salyn, Ya.V., Nefedov, V.I., Maiorova, A.G., Kuznetsova, G.N., 1978. X-ray electron
characterisation of copper acetohydroxamate and copper n- study of the compounds of platinum(III) with acetamide. Zh. Neorg. Khim. 23,
octanohydroxamate. Inorg. Chim. Acta 365, 65–70. 829–831 (Russian J. Inorg. Chem., 23, 458–460).
Jansen, R.J.J., van Bekkum, H., 1995. XPS of nitrogen-containing functional groups on Solozhenkin, P.M., Kondrat’ev, S.A., Angelova, E.I., 2013. Quantum-mechanics of
activated carbon. Carbon 33, 1021–1027. pyrite flotation. J. Mining Sci. 49, 819–830.
Lee, T.O., Rabalais, J.W., 1977. X-ray photoelectron spectra and electronic structure Srinivasan, V., Walton, R.A., 1977. X-ray photoelectron spectra of inorganic
of some diamine compounds. J. Electron Spectrosc. 11, 123–127. molecules. XX. Observations concerning the sulfur 2p binding energies in
Leppinen, J.O., Basilio, C.I., Yoon, R.H., 1988. FTIR study of thionocarbamate metal complexes of thiourea. Inorg. Chim. Acta 25, L85–L86.
adsorption on sulfide minerals. Colloids Surf. 32, 113–125. Woods, R., Hope, G.A., 1999. A SERS spectroelectrochemical investigation of the
Liu, G., Zhong, H., Dai, T., Xia, L., 2008. Investigation of the effect of N-substituents interaction of O-isopropyl-N-ethylthionocarbamate with copper surfaces.
on performance of thionocarbamates as selective collectors for copper sulfides Colloids Surf., A 146, 63–74.
by ab initio calculations. Miner. Eng. 21, 1050–1054. Yoon, R-H., Basilio, C.I., 1993. Adsorption of thiol collectors on sulphide minerals
Livshits, A.K., Glembotskii, A.V., Gurvich, S.M., 1968. Possible forms of O-alkyl N- and precious metals – a new perspective. In: Proc. XVIII Int. Miner. Process.
alkylthionocarbamates in acidic and alkaline aqueous media. Tsvetn. Met. Congress, Sydney, pp. 611–617.
(English Translation) 9 (2), 27–29. Zachwieja, J.B., McCarron, J.J., Walker, G.W., Buckley, A.N., 1989. Correlation
López, G.P., Castner, D.G., Ratner, B.D., 1991. XPS O 1s binding energies for polymers between the surface composition and collectorless flotation of chalcopyrite. J.
containing hydroxyl, ether, ketone and ester groups. Surf. Interface Anal. 17, Colloid Interface Sci. 132, 462–468.
267–272. Zanello, P., 2003. Inorganic Electrochemistry: Theory, Practice and Application. The
Murphy, C.N., Winter, G., 1973. Monothiocarbonates and their oxidation to Royal Society of Chemistry, Cambridge, UK, 606 p.
disulphides. Aust. J. Chem. 26, 755–760.
Ohmasa, M., Suzuki, M., Takeuchi, Y., 1977. A refinement of the crystal structure of
covellite, CuS. Mineral. J. 8, 311–319.

Potrebbero piacerti anche