You are on page 1of 6

Minerals Engineering 24 (2011) 499–504

Contents lists available at ScienceDirect

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

Fungal pretreatment of sulfides in refractory gold ores


G. Ofori-Sarpong a,⇑, K. Osseo-Asare a,b, M. Tien c
a
Department of Energy and Mineral Engineering, Penn State University, University Park, PA 16802, USA
b
Department of Materials Science and Engineering, Penn State University, University Park, PA 16802, USA
c
Department of Biochemistry and Molecular Biology, Penn State University, University Park, PA 16802, USA

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

Article history: This study assessed the capability of the fungus, Phanerochaete chrysosporium, to decompose pyrite, arse-
Available online 9 March 2011 nopyrite and a sulfide-containing flotation concentrate in an effort to develop a microbial process for pre-
treating refractory gold ores. The extent of biotransformation was monitored by analyzing for iron, sulfur
Keywords: and arsenic in incubation solutions, and for sulfide sulfur in the residual solids. The results were then
Phanerochaete chrysosporium expressed as percentages of the initial weights. For arsenopyrite, 1.5 wt.%, 7.2 wt.% and 10.3 wt.% of iron,
Biotransformation arsenic and sulfur respectively were present as soluble constituents in the incubation solution within
Pyrite
21 days of fungal treatment, whereas for pyrite, there was 1.2 wt.% iron and 6.0 wt.% sulfur. For the same
Arsenopyrite
Flotation concentrate, Refractory gold ores
processing period in the case of the flotation concentrate, 1.8 wt.%, 6.1 wt.% and 10.7 wt.% respectively of
iron, arsenic and sulfur remained in solution. Overall, the decomposition of sulfide sulfur in the samples
was 15 wt.%, 35 wt.% and 57 wt.% respectively for pyrite, arsenopyrite and the flotation concentrate.
Changes in sulfide sulfur concentration and the formation of oxide phases during fungal treatment were
confirmed using Raman spectroscopy and X-ray diffraction analysis. These results suggest that P. chrysos-
porium is a potential microorganism for oxidative decomposition of metal sulfides associated with refrac-
tory gold ores.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction or associated with relatively more porous iron oxides and thus
amenable to cyanidation (Marsden and House, 2006). However,
In sulfidic refractory gold ores, tiny gold particles typically most of the major bacteria meet their carbon requirement by uti-
<1 lm (Marsden and House, 2006) may be highly disseminated lizing carbon dioxide. Consequently, in cases where double refrac-
and locked up within the grain boundaries or fractures of sulfide tory gold ores (DRGO) are treated, the biooxidised concentrate
minerals such as pyrite and arsenopyrite. Thus, decomposition of generally contains substantial levels of carbonaceous matter
the sulfides is required to liberate the gold (Baako, 1972; Boyle, (CM) that is not degraded (Silver, 1970; Glazer and Nikaido,
1979; Komnitsas and Pooley, 1989; Benzaazoua et al., 2007). Com- 1995; Madigan and Martinko, 2006). The CM is therefore routed
mercially, roasting, pressure oxidation and bacterial oxidation are into downstream cyanidation circuits where it adsorbs (preg-robs)
among the processes employed in sulfide oxidation (Arriagada dissolved gold (Brierley and Kulpa, 1992, 1993; Amankwah et al.,
and Osseo-Asare, 1984; Berezowsky et al., 1988; Hutchins et al., 2005; Yen et al., 2008).
1988; Yannopoulos, 1991; Nyavor and Egiebor, 1992; Brierley, Current research efforts have focused on two-stage processes to
1995; Rawlings et al., 2003). oxidize sulfides and deactivate CM (Brierley and Kulpa, 1992,
Biooxidation makes use of iron and sulfur oxidizing bacteria to 1993; Amankwah et al., 2005; Yen et al., 2008). However, using a
catalyze the oxidation of sulfides, liberating gold for subsequent ‘‘one-pot’’ process that can simultaneously oxidize sulfides and
cyanidation (Lundgren and Silver, 1980; Livesey-Goldblatt et al., deactivate CM will be of immense benefit in the treatment of DRGO
1983; Brierley, 1997; Hackl, 1997). The bacteria gain energy by oxi- as it will shorten process time, and thus cut down cost.
dizing ferrous iron and elemental sulfur which generates, in situ, The fungus, Phanerochaete chrysosporium, secretes the oxidative
ferric ions and sulfuric acid. Ferric ions and sulfuric acid are enzymes, lignin peroxidase and manganese peroxidase (Glenn
lixiviants responsible for indirect leaching of the sulfides in the et al., 1983; Tien and Kirk, 1983, 1988), which are capable of
bio-system (Keller and Murr, 1982; Rawlings et al., 1999; Sand degrading aromatic carbonaceous materials. Ofori-Sarpong et al.
et al., 2001; Rohwerder et al., 2003; Madigan and Martinko, (2010) used this fungus to deactivate anthracite-grade
2006). At the end of biooxidation, gold may be totally liberated carbonaceous matter and reduce its gold adsorption ability by
more than 90%. P. chrysosporium also secretes a H2O2-generating
⇑ Corresponding author. enzyme, glyoxal oxidase (Kersten and Kirk, 1987), and hydrogen
E-mail addresses: goforisarp@gmail.com, gad164@psu.edu (G. Ofori-Sarpong). peroxide is known to solubilize pyrite and arsenopyrite as shown

0892-6875/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2011.02.020
500 G. Ofori-Sarpong et al. / Minerals Engineering 24 (2011) 499–504

in Eqs. (1) and (2) (McKibben and Barnes, 1986; Schreiner et al., solution (Ciminelli and Osseo-Asare, 1995; Rait and Aruscavage,
1988; Antonijevic et al., 1997; Jennings et al., 2000). 2006). The mixture was heated at 80 °C for 90 min and then fil-
tered. The filter cake was washed with water and dried at 70 °C.
2FeS2 þ 15H2 O2 ! 2Fe3þ þ 4SO2 þ
4 þ 2H þ 14H2 O ð1Þ Sulfide sulfur in the filter cake was then determined by the volu-
metric combustion technique using the LECO Sulfur determinater
FeAsS þ 7H2 O2 ! Fe3þ þ AsO3 2 þ
4 þ SO4 þ 2H þ 6H2 O ð2Þ SC-4444DR.
The iron (III) and sulfuric acid produced by the reactions in Eqs.
(1) and (2) constitute the main reagents created by the chemolith- 2.4. Characterization of flotation concentrate
otrophic sulfide oxidizing bacteria for the indirect biooxidation of
sulfides (Brierley, 1995; Hackl, 1997; Keller and Murr, 1982; Sand Both X-ray diffraction and Raman spectroscopy were used to
et al., 2001; Holmes and Bonnefoy, 2006). evaluate phase changes in the flotation concentrate before and
This study was therefore aimed at investigating the ability of after treatment with P. chrysosporium. In the case of Raman spec-
the CM-deactivating fungus, P. chrysosporium, to also solubilize/ troscopy, samples were homogenized and mounted on adhesive
oxidize sulfides found in refractory gold ores so as to assess the po- tape attached to a sample holder. Raman spectra were collected
tential of a single-stage process for the pretreatment of double using the Confocal WITec XY Raman spectrometer at 5 min interval
refractory gold ores. with 1 s integration time. For X-ray diffraction (XRD) studies, the
samples were ground to fine powder in an agate mortar and then
sprinkled on the surface of a quartz zero-background sample
2. Experimental investigations holder. The analysis was carried out using PANalytical X’Pert Pro
powder diffractometer with X’celerator detector. A Ni-filtered
2.1. Materials Cu Ka radiation produced at 45 kV and 40 mA was used for the
analysis. The scan was run from 10° to 70° for 2-theta at a scan
Fungal spores of P. chrysosporium ME446, were obtained from speed of 2 deg/min. Data acquisition was done using MDI Jade 9
the Ming Tien lab of the Department of Biochemistry and Molecu- software.
lar Biology, Penn State University. Pyrite and arsenopyrite samples
were supplied by VWR and Ward’s Natural Science Establishment
2.5. Analysis of data
respectively. The sulfide-containing flotation concentrate (FC) al-
ready milled to all passing 75 lm was obtained from the sulfide
Fungal-dissolution of constituents (FDC) during incubation was
treatment plant at Bogoso Mine of Golden Star Prestea–Bogoso Re-
estimated in wt.% as shown in Eq. (3), where V is the volume of
sources, Ghana. The growth media for the fungus, millet and wheat
solution used in fungal incubation, C is the concentration of dis-
bran (MWB) were obtained from Nature’s Pantry, State College, PA,
solved constituent as measured by ICP-AES, Pm is the percentage
whereas reagent grades succinic acid and sodium hydroxide were
of the constituent in the sulfide material and W is the mass of sul-
obtained from Alfa Aesar.
fide material used. Eq. (4) also depicts the percentage conversion of
sulfide sulfur (CoS) in the residual solids, where SI and SF are the
2.2. Medium preparation, incubation and harvesting of treated respective sulfide sulfur contents before and after fungal treat-
material ment. All experiments were carried out in triplicate, and the data
reported in the figures are mean values, with the error bars repre-
The millet and wheat bran (MWB) medium was prepared by senting the standard errors of the means.
using 8 g millet and 2 g wheat bran. Double-distilled (dd) H2O used
for the incubation was buffered with succinic acid and the pH ad- V ðmLÞ  C ðl g=mLÞ
FDC ðwt:%Þ ¼  100% ð3Þ
justed to four using sodium hydroxide. The medium, in Erlenmeyer Pm ð%Þ  W ðgÞ
flasks, was moistened with water and autoclaved at 121 °C for
30 min. On cooling, the medium was inoculated with 1 mL of spore ðSI  SFÞ ðgÞ
CoS ðwt:%Þ ¼  100% ð4Þ
suspension of P. chrysosporium (made by suspending one vial of SI ðgÞ
spores in 25 mL of dd H2O). Ten grams each of pyrite and arseno-
pyrite samples was moistened with 5 mL of water and autoclave-
sterilized for 30 min. The samples at 30% solids were incubated 3. Results and discussion
in triplicates at 37 °C, for up to 21 days on a New Brunswick Series
25 Incubator shaker at 150 rpm. Control experiments were also set 3.1. Transformation of pyrite by P. chrysosporium
up for 14 days under similar conditions. At the end of the incuba-
tion period the samples were washed with water to get rid of the To establish the direct effect of P. chrysosporium on the biotrans-
media and fungal biomass, and then dried at 37 °C for 7 days. formation of sulfides it was necessary to examine the effect of the
growth media, and this was done by running control experiments.
2.3. Determination of dissolved components and sulfide sulfur analysis The results obtained for pyrite samples incubated for 2 weeks both
of fungal-treated solids in the absence and presence of P. chrysosporium are illustrated in
Fig. 1. The figure shows that though there was some amount of dis-
The liquid samples obtained after fungal treatment were tested solution of pyrite in the absence of fungal biomass, the presence of
for total dissolved arsenic, iron and sulfur using a Perkin–Elmer the fungus led to 3–6 fold increase in dissolution of iron and sulfur.
Optima 5300 inductively coupled plasma–atomic emission spec- The minor solubilization in the control experiment might be due to
troscopy (ICP-AES). Before determining sulfide sulfur in both the natural oxidation of sulfides according to the reaction shown in Eq.
as-received and processed solids, sulfate sulfur and elemental sul- (5).
fur, if any, had to be removed. For elemental sulfur (S0), samples
2FeS2 þ 7O2 þ 2H2 O ! 2Fe2þ þ 4SO2
4 þ 4H
þ
ð5Þ
were digested in the ratio of 2 g solid sample to 25 mL tetrachloro-
ethylene for 60 min in a boiling water bath, after which the sample This is, however, a slow reaction and cannot lead to any appre-
was filtered. The sulfate sulfur portion was removed by contacting ciable level of dissolution under atmospheric conditions. In a
the dried filter cake with 3 M HCl in the ratio of 1 g solid to 50 mL strong oxidizing environment containing hydrogen peroxide and
G. Ofori-Sarpong et al. / Minerals Engineering 24 (2011) 499–504 501

sulfates such as Fe2O3, FeOOH and FeSO4, which precipitate on


the surfaces of unreacted material. Other components that may
be present in such a system include sulfites, thiosulfates, and ele-
mental sulfur (Osseo-Asare et al., 1984; Ciminelli and Osseo-Asare,
1995; Wei and Osseo-Asare, 1997). Accumulation of elemental sul-
fur and other precipitates in the system has the potential to passiv-
ate the unreacted surfaces, thus obstructing movement of
reactants and products to and from the reaction sites. This will lead
to reduction in the overall rate of reaction (Lindstrom et al., 1992;
Sasaki et al., 1995; Marsden and House, 2006; Mousavi et al.,
2006).
Stoichiometrically, the molar ratio of sulfur to iron in pyrite is
2:1. However, this ratio does not tally with the measurements
made in solution as the ratio is about 5 mol of sulfur to 1 mol of
iron. The deficiency of iron in the solution is likely due to precipi-
tation as presented above. Consequently, the quantity of dissolved
constituents does not reflect fully the overall biotransformation of
Fig. 1. 2 week biotransformation of pyrite in the absence and presence of sulfide by the fungus. The progressive loss of sulfide sulfur with
P. chrysosporium. The experiment was conducted at 37 °C for 14 days at pH 4 and time during treatment with the fungus was thus determined with
30% solids.
the help of the LECO combustion technique, and used to estimate
the overall oxidation of pyrite with time as shown in Fig. 3.
Within 21 days of fungal treatment of pyrite, an overall oxida-
oxidative enzymes, such as that generated by P. chrysosporium, the
tion of 15% was achieved, with about 90% of the conversion occur-
conversion of ferrous to ferric ion as shown in Eq. (6) is possible
ring within the initial 10 days of treatment. Though the 15% overall
(Keller and Murr, 1982; Marsden and House, 2006).
oxidation appears to be very low, it is substantial considering that
it was achieved for pure sulfides. For refractory sulfidic gold con-
Fe2þ ! Fe3þ þ e ð6Þ centrates with lower sulfide sulfur content, it is possible to realize
higher percentage oxidation.
The fungal-dissolution of pyrite as a function of time is depicted
in Fig. 2. After 21 days of fungal treatment there was 1.2 wt.% iron
3.2. Transformation of arsenopyrite by P. chrysosporium
and 6.0 wt.% sulfur in the incubation solution. It can be seen from
the figure that the rate of dissolution of pyrite is higher at the ini-
The dissolution of the constituents of arsenopyrite in the ab-
tial stages (up to 7 days) but decreases afterwards. In a batch cul-
sence and presence of the fungus, P. chrysosporium is presented
ture such as the one utilized in this study, microbial growth and
in Fig. 4. In the absence of the fungus, dissolution of arsenic, iron
hence activity increases exponentially in the initial stages of
and sulfur was marginal at 0.9 wt.%, 0.16 wt.% and 2.6 wt.% respec-
growth, and retardation takes place afterwards due to exhaustion
tively. When the sample was incubated with P. chrysosporium, dis-
of growth media and generation of waste products of metabolism
solution increased to higher values of 6.3 wt.%, 1.3 wt.% and
(Shuler and Kargi, 1992). The reduction in microbial activity and
9.2 wt.% for arsenic, iron and sulfur respectively. The minor solubi-
thus availability of oxidants may account for the trend observed.
lization in the control experiment might be due to natural oxida-
The pH for optimum operation of P. chrysosporium is 4 (Tien and
tion of sulfides according to the reaction shown in Eq. (7)
Kirk, 1988) and it is at this pH ± 0.3 that the experiment was
(Marsden and House, 2006).
carried out. It is known that in oxidizing leaching of pyrite above
3
pH of 3 there is production of insoluble iron oxides/hydroxides/ 4FeAsS þ 11O2 þ 6H2 O ! 4Fe2þ þ 4SO2 þ
4 þ 12H þ 4AsO3 ð7Þ

Fig. 2. Accumulation of dissolved sulfur and iron during the incubation of pyrite Fig. 3. Biotransformation of sulfide sulfur in pyrite as a function of fungal-
with P. chrysosporium. The experiment was conducted at 37 °C for up to 21 days at treatment time. The experiment was conducted at 37 °C for up to 21 days at pH 4
pH and 30% solids. and 30% solids.
502 G. Ofori-Sarpong et al. / Minerals Engineering 24 (2011) 499–504

Compared to pyrite, a higher overall decomposition was


achieved for arsenopyrite, and this can be attributed to the higher
nobility of pyrite. Microbial attack is naturally initiated from weak-
er zones in a material (Wackett and Ellis, 1999), and hence the
weaker As–S bond is likely to be attacked relatively easier than
the Fe–S bond (Marsden and House, 2006).

3.3. Transformation of flotation (sulfidic gold) concentrates by P.


chrysosporium

Fig. 6 presents the direct dissolution of iron, sulfur and arsenic


from sulfidic gold concentrates by P. chrysosporium after 14 day
contact with MWB medium with and without fungal biomass. As
explained in the case of pyrite and arsenopyrite, the slight dissolu-
tion obtained in the control experiment may be due to natural dis-
solution of sulfides in the presence of water and oxygen, which
may cease in a limited oxidizing environment. However, in the
Fig. 4. 2 week biotransformation of arsenopyrite in the absence and presence of
presence of the fungus and the associated oxidizing environment
P. chrysosporium. The experiment was conducted at 37 °C for 14 days at pH 4 and generated, about 350–500% increase in dissolution was observed.
30% solids. Fig. 7 depicts the dissolution of sulfides from FC by P. chrysospo-
rium as a function of incubation time (Eq. (3)). The dissolution of
The progressive increments of iron, sulfur and arsenic in incu- sulfide begins after the initiation of incubation and rises rapidly
bation solution during the fungal treatment of arsenopyrite over within the first week, beyond which dissolution slows down, end-
different incubation periods are presented in Fig. 5. Over a period ing at 1.5 wt.%, 7.2 wt.% and 10.3 wt.% of dissolved iron, arsenic and
of 21 days, the concentrations of iron, sulfur and arsenic in solution sulfur respectively at the end of 21 days of fungal treatment. The
increased consistently ending at 1.5 wt.%, 7.2 wt.% and 10.3 wt.% figure however indicates that a plateau is yet to be achieved, and
respectively. The trends in Fig. 5 indicate that the concentrations this signifies that given more time for incubation, dissolution will
of the constituents in solution are yet to reach their respective pla- continue.
teaus, and the processing time for maximum degradation extends Fig. 8 presents the transformation of sulfide sulfur (Eq. (4)) in
beyond 21 days. However, with better control of the system, higher the residual solid materials after incubation. About 48% of sulfide
levels of conversion are expected within the given time. sulfur in the flotation concentrate got oxidized within the initial
The stoichiometric ratio of iron, arsenic and sulfur in 1 mol of 10 days of processing. Beyond this period, biotransformation
arsenopyrite is 1:1:1. This ratio changed based on the amount of stalled leading to a further 9% oxidation, and bringing the total
these constituents in incubation solution as the ratio in solution conversion in sulfide sulfur to 57% within 21 days.
was about 0.14 mol of Fe, 0.7 mol of As and 1 mol of sulfur after
21 days of treatment. This implies a deficiency of 0.86 mol of Fe 3.4. Characterization of flotation concentrate
and 0.3 mol of As in solution relative to the as-received material,
as a result of precipitation. The precipitation of ferric arsenate at Many metal sulfides show Raman peaks between 300 and
pH above 2 has been observed by several authors, with concomi- 400 cm1 (Ushioda, 1972; Mernagh and Trudu, 1993; Sasaki
tant decrease in reaction rate (Escobar et al., 2000; Brierley, et al., 1995), and this was confirmed in the current investigation.
2003; Lindstrom et al., 2003; Marsden and House, 2006). Overall, Raman spectroscopy of the as-received and treated flotation
the conversion of sulfide sulfur in arsenopyrite as determined by concentrate indicated a reduction in the intensity of sulfide peaks
the LECO combustion technique was 35 wt.%. after fungal-biotransformation as presented in Fig. 9. These

Fig. 5. Dissolution of iron, sulfur and arsenic from arsenopyrite as a function of Fig. 6. 2 week biotransformation of flotation concentrate in the absence and
incubation time. The experiment was conducted at 37 °C for up to 21 days at pH 4 presence of P. chrysosporium. The experiment was conducted at 37 °C for 14 days at
and 30% solids. pH 4 and 30% solids.
G. Ofori-Sarpong et al. / Minerals Engineering 24 (2011) 499–504 503

300

Intensity (counts)
1 – quartz
2 – pyrite
3 – arsenopyrite
200 4 – hematite
5 – ferric arsenate
6 – muscovite

100

0
10 20 30 40 50 60 70
2 Theta (deg)
Fig. 10. X-ray diffractogram of as-received and fungal-treated flotation concen-
trate. Treatment was conducted for 14 days at 37 °C and pH 4.

observations authenticate the decrease in sulfide sulfur as deter-


Fig. 7. Dissolution of sulfides in flotation concentrate by P. chrysosporium as a
function of time for 21 days. The experiment was conducted at 37 °C for up to
mined by the LECO combustion technique. Further analysis by
21 days at pH 4 and 30% solids. XRD shows other products of sulfide sulfur oxidation (Fig. 10).
By comparing the as-received and fungal-treated samples in
Fig. 10, it was observed that the treated material had two new oxy-
gen-containing phases; hematite (a-Fe2O3) and ferric arsenate
(FeAsO4H2O). In similar systems, ferric arsenate has been observed
earlier by researchers such as Escobar et al. (2000), Brierley (2003)
and Lindstrom et al. (2003). It is possible that other oxides were
precipitated but due to lower resolutions, the peaks might have
been masked by the higher intensity peaks. Also, the diffractome-
ter used for the study had a detection limit of 5%, and thus any
phase with less than 5 wt.% composition may not show a peak
on the diffractogram.
The presence of the oxidation products indicates clearly that P.
chrysosporium has the ability to transform sulfides in refractory
gold ores. The trends in the results signify that by ensuring proper
system controls, higher levels of oxidation will be possible. Oxida-
tion of the sulfides will enhance liberation of the encapsulated
gold, making it more amenable to cyanidation. These results com-
pare well with those obtained by Brierley (2003) in the biooxida-
tion of refractory gold ores containing about 2% sulfide sulfur.
The work was conducted at 20–60 °C for 10 days using mesophilic
and moderate-thermophillic bacteria, and hyper-thermophillic ar-
chea, and the author achieved over 50% increase in the subsequent
Fig. 8. Biotransformation of sulfide sulfur in the flotation concentrate as a function gold recovery by cyanidation. In a related study, gold recovery in-
of incubation time. The experiment was conducted at 37 °C for up to 21 days at pH 4 creased appreciably following fungal treatment (Ofori-Sarpong,
and 30% solids.
2010).

4. Conclusions

The fungus, P. chrysosporium, which has proven its ability to


reduce the preg-robbing capacity of anthracite-grade carbona-
ceous matter, was investigated for its ability to also transform
sulfides, in an effort to develop a microbial process for pretreating
double refractory gold ores. The results show that P. chrysosporium
is capable of solubilizing sulfides. For an incubation period of
21 days, sulfide sulfur decreased by 35% in the treated arsenopyrite
sample and 15% in the pyrite sample. These materials had initial
sulfide sulfur content of 19.6% and 52% respectively. For refractory
sulfidic gold concentrate with about 15% initial sulfide sulfur
content, a higher oxidation level of 57% was realized. Raman
spectroscopy and X-ray diffraction analysis showed reduction in
sulfide sulfur concentration and increase in oxide phases resulting
from fungal treatment. The results demonstrate that the fungus,
Fig. 9. Raman spectrogram of as-received and fungal-treated flotation concentrate. P. chrysosporium, is a potential microorganism for oxidative
Treatment was conducted for 14 days at 37 °C and pH 4. decomposition of metal sulfides associated with refractory gold
504 G. Ofori-Sarpong et al. / Minerals Engineering 24 (2011) 499–504

ores. With improvement in system controls, higher levels of oxida- Komnitsas, C., Pooley, F.D., 1989. Mineralogical characteristics and treatment of
refractory gold ores. Minerals Engineering 2, 449–457.
tion are feasible.
Lindstrom, E.B., Gunneriusson, E., Tuovinen, O.H., 1992. Bacterial oxidation of
refractory sulphide ores for gold recovery. Critical Reviews in Biotechnology 12,
133–155.
Acknowledgment Lindstrom, E.B., Sandstrom, A., Sundkvist, J., 2003. A sequential two-step process
using moderately and extremely thermophilic cultures for biooxidation of
The authors are grateful to the Schlumberger Faculty for the Fu- refractory gold concentrates. Hydrometallurgy 71, 21–30.
Livesey-Goldblatt, E.P., Norman, E.P., Livesey-Goldblatt, D.R., 1983. Gold recovery
ture, Netherlands, for providing supplementary funding for this from arsenopyrite/pyrite ore by bacterial leaching and cyanidation. In: Rossi, G.,
work and to the Golden Star Prestea–Bogoso Resources (GSPBR), Torma, A.E., (Eds.). Recent Progress in Biohydrometallurgy, pp. 627–641.
Ghana, for providing the gold concentrates and assisting with char- Lundgren, D.G., Silver, M., 1980. Ore leaching by bacteria. Annual Review of
Microbiology 34, 263–283.
acterization of the solid samples. The authors also acknowledge the
Madigan, M.T., Martinko, J.M., 2006. Brock Biology of Microorganisms. 11th ed.
assistance offered by Prof. Yaw Yeboah of the Energy and Mineral Pearson Prentice Hall, Upper Saddle River, NJ. pp. 469–472, 691–692.
Engineering Department, Penn State University, USA, and Prof. Marsden, J., House, I., 2006. The chemistry of gold extraction. 2nd ed. Society for
Richard Amankwah of the Mineral Engineering Department, Uni- Mining. Metallurgy and Exploration, Inc. Colorado, pp. 42–44, 111–126, 161–
177, 191–193, 233–263, 297–333.
versity of Mines and Technology, Ghana. The first author is thank- McKibben, M.A., Barnes, H.L., 1986. Oxidation of pyrite in low temperature acidic
ful to the Ghana Education Trust Fund, the University of Mines and solutions: rate laws and surface textures. Geochimica et Cosmochimica Acta 50,
Technology, Ghana, and the PEO International Peace Scholarship, 1509–1520.
Mernagh, T.P., Trudu, A.G., 1993. A laser Raman microprobe study of some
USA, for financial assistance. geologically important sulphide minerals. Chemical Geology 103, 113–127.
Mousavi, S.M., Yaghmaei, S., Salimi, F., Jafari, A., 2006. Influence of process variables
on biooxidation of ferrous sulfate by an indigenous Acidithiobacillus
References ferrooxidans⁄ Part I: Flask experiments. Fuels 85, 2555–2560.
Nyavor, K., Egiebor, N.O., 1992. Application of pressure oxidation pretreatment to a
Amankwah, R.K., Yen, W.T., Ramsay, J., 2005. A two-stage bacterial pretreatment double-refractory gold concentrate. CIM Bulletin 84, 84–90.
process for double refractory gold ores. Minerals Engineering 18, 103–108. Ofori-Sarpong, G. 2010. Simultaneous biotransformation of sulfides and
Antonijevic, M.M., Dimitrijevic, M., Jankovic, Z., 1997. Leaching of pyrite with carbonaceous matter in double refractory gold ores using the fungus,
hydrogen peroxide in sulphuric acid. Hydrometallurgy 46, 71–83. Phanerochaete chrysosporium. PhD Thesis, Pennsylvannia State University,
Arriagada, F.J., Osseo-Asare, K., 1984. Gold extraction from refractory ores: roasting USA. pp. 197.
behavior of pyrite and arsenopyrite. In: Kudryk, V., Corrigan, D., Liang, W.W. Ofori-Sarpong, G., Tien, M., Osseo-Asare, K., 2010. Myco-hydrometallurgy: Coal
(Eds.), Precious Metals: Mining, Extraction and Processing. The Metallurgical model for potential reduction of preg-robbing capacity of carbonaceous gold
Society of AIME, Warrendale, PA, pp. 367–385. ores using the fungus, Phanerochaete chrysosporium. Hydrometallurgy 102, 66–
Baako, A.B., 1972. Mining geology of Prestea gold deposit, Ghana. PhD Thesis, 72.
Universita’ Degli studi di Cagliari, Italy, pp. 80. Osseo-Asare, K., Xue, T., Ciminelli, V.S.T., 1984. Solution chemistry of cyanide
Benzaazoua, M., Marrion, P., Robout, F., Pinto, A., 2007. Gold-bearing arsenopyrite leaching systems. In: Kudryk, V., Corrigan, D., Liang, W.W. (Eds.), Precious
and pyrite in refractory ores: analytical refinements and new understanding of Metals: Mining, Extraction and Processing. The Metallurgical Society of AIME,
gold mineralogy. Mineralogical Magazine 71, 123–142. Warrendale, PA, pp. 173–197.
Berezowsky, R.M.G.S., Haines, A.K., Weir, D.R., 1988. The Sao Bento gold project Rait, N., Aruscavage, P.J., 2006. The determination of forms of sulfur in coal. In:
pressure oxidation process development. In: 18th Annual meeting, Hydromet Golightly, D.W., Simon, F.O., (Eds.). Methods for sampling and inorganic analysis
Section of CIM, CIM, Edmonton, Alberta, pp. 1–24. of coal. US Geological Survey Bulletin 1823. http://www.pubs.usgs.gov/bul/
Boyle, R.W., 1979. The Geochemistry of Gold and its Deposits. Canada Geological b1823/10.htm.
Survey Bulletin, p. 280. Rawlings, D.E., Dew, D., du Plessis, C., 2003. Biomineralization of metal-containing
Brierley, C.L., 1995. Bacterial oxidation. Engineering and Mining Journal 196, 42–44. ores and concentrates. Trends in Biotechnology 21, 38–44.
Brierley, C.L., 1997. Mining biotechnology: research to commercial development Rawlings, D.E., Tributsch, H., Hansford, G.S., 1999. Reasons why Leptospirillum-like
and beyond. In: Rawlings, D.E. (Ed.), Biomining: Theory. Microbes and Industrial species rather than Thiobacillus ferrooxidans are dominant iron-oxidizing
Processes, Springer Verlag, Berlin, Germany, pp. 3–17. bacteria in many commercial processes for the biooxidation for pyrite and
Brierley, J.A., 2003. Response of microbial systems to thermal stress in biooxidation- related ores. Review. Microbiology 145, 5–13.
heap pretreatment of refractory gold ores. Hydrometallurgy 71, 13–19. Rohwerder, T., Gehrke, T., Kinzler, K., Sand, W., 2003. Bioleaching review part
Brierley, J.A., Kulpa, C.F., 1992. Microbial consortium treatment of refractory A: progress in bioleaching: fundamentals and mechanisms of bacterial
precious metal ores. US Patent, 5127,942. metal sulfide oxidation. Applied Microbiology and Biotechnology 63, 239–
Brierley, J.A., Kulpa, C.F., 1993. Biometallurgical treatment of precious metal ores 248.
having refractory carbon content. US Patent, 5, 244,493. Sand, W., Gehrke, T., Jozsa, P.-G., Schippers, A., 2001. (Bio)chemistry of bacterial
Ciminelli, V.S.T., Osseo-Asare, K., 1995. Kinetics of pyrite oxidation in sodium leaching – direct vs. indirect bioleaching. Hydrometallurgy 59, 159–175.
carbonate solutions. Metallurgical and Materials Transactions B: Process Sasaki, K., Tsunekawa, M., Ohtsuka, T., Konno, H., 1995. Confirmation of a sulfur-rich
Metallurgy and Materials Processing Science 26B, 209–218. layer on pyrite after oxidative dissolution by Fe(III) ions around pH 2.
Escobar, B., Huenupi, E., Godoy, I., Wiertz, J.V., 2000. Arsenic precipitation in the Geochimica et Cosmochimica Acta 59 (15), 3155–3158.
bioleaching of enargite by Sulfolobus BC at 70 °C. Biotechnology Letters 22, Schreiner, R.P., Stevens Jr., E., Tien, M., 1988. Oxidation of thianthrene by the
205–209. ligninase of Phanerochaete chrysosporium. Applied and Environmental
Glazer, A.N., Nikaido, A., 1995. Microbial Technology. Freeman and Co., New York. Microbiology 54, 1858–1860.
Glenn, J.K., Morgan, M.A., Mayfield, M.B., Kuwahara, M., Gold, M.H., 1983. An Shuler, M.L., Kargi, F., 1992. Bioprocess Engineering: Basic Concepts. Prentice Hall,
extracellular H2O2-requiring enzyme preparation involved in lignin NJ. pp. 232–310.
biodegradation by the white rot basidiomycete Phanerochaete chrysosporium. Silver, M., 1970. Oxidation of elemental sulfur and sulfur compounds and CO2
Biochemical and Biophysical Research Communications 114, 1077–1083. fixation by Ferrobacillus ferrooxidans (Thiobacillus ferrooxidans). Canadian
Hackl, R.P., 1997. Commercial applications of bacterial-mineral interactions. In: Journal of Microbiology 16, 845–849.
McIntosh, J.M., Groat, L.A. (Eds.), Biological-Mineralogical Interactions. Tien, M., Kirk, T.K., 1983. Lignin-degrading enzyme from the hymenomycete
Mineralogical Association of Canada, pp. 143–167. Phanerochaete chrysosporium burds. Science 221, 661–663.
Holmes, D., Bonnefoy, V., 2006. Insights into iron and sulfur oxidation mechanisms Tien, M., Kirk, T.K., 1988. Lignin peroxidase of Phanerochaete chrysosporium.
of bioleaching organisms. In: Rawlings, D.E., Johnson, B.D. (Eds.), Biomining. Methods in Enzymology 161, 238–249.
Springer-Verlag, Berlin, pp. 281–307. Ushioda, S., 1972. Raman scattering from phonons in iron pyrite (FeS2). Solid State
Hutchins, S.E., Brierley, J.A., Brierley, C.L., 1988. Microbial pretreatment of refractory Communications 10, 307–310.
sulfide and carbonaceous ores improves the economics of gold recovery. Mining Wackett, L.P., Ellis, L.B.M., 1999. Predicting Biodegradation. Environmental
Engineerring 40, 249–254. Microbiology 1, 119–124.
Jennings, S.R., Dollhopf, D.J., Inskeep, W.P., 2000. Acid production from sulfide Wei, D., Osseo-Asare, K., 1997. Semiconductor electrochemistry of particulate
minerals using hydrogen peroxide weathering. Applied Geochemistry 15, 235– pyrite. Mechanisms and products of dissolution. Journal of Electrochemical
243. Society 144, 546–553.
Keller, L., Murr, L.E., 1982. Acid-bacterial and ferric sulfate leaching of pyrite single Yannopoulos, J.C., 1991. The Extractive Metallurgy of Gold. Van Nostrand Reinhold,
crystals. Biotechnology and Bioengineering 24, 83–96. New York. pp. 141–148.
Kersten, P.J., Kirk, T.K., 1987. Involvement of a new enzyme, glyoxal oxidase, in Yen, W.T., Amankwah, R.K., Choi, Y., 2008. Microbial pre-treatment of double
extracellular H2O2 production by Phanerochaete chrysosporium: synthesized in refractory gold ores. In: Young, C.A., Taylor, P.R., Anderson, C.G., Choi, Y. (Eds.),
the absence of lignin in response to nitrogen starvation. Journal of Bacteriology Proceedings of the Sixth International Symposium, Hydrometallurgy 2008,
135, 790–797. Phoenix, USA. SME, Littleton, CO, pp. 506–510.

You might also like