Vous êtes sur la page 1sur 11

canmin.00.00.

1300054

09-06-15 13:5

1
The Canadian Mineralogist
Vol. 00, pp. 1-11 (2015)
DOI: 10.3749/canmin.1300054

THE RAMAN STUDY OF WEATHERING MINERALS FROM THE CORANDA-HONDOL


OPEN PIT (CERTEJ GOLD-SILVER DEPOSIT) AND THEIR PHOTOCHEMICAL
DEGRADATION PRODUCTS UNDER LASER IRRADIATION
ANDREI I. APOPEI

AND

NICOLAE BUZGAR

Alexandru Ioan Cuza University of Iai, Faculty of Geography and Geology, Department of Geology, 20A Carol I Blv.,
700505 Iai, Romania

GHEORGHE DAMIAN
Alexandru Ioan Cuza University of Iai, Faculty of Geography and Geology, Department of Geology, 20A Carol I Blv.,
700505 Iai, Romania and
Tech Univ Cluj Napoca, North University Center of Baia Mare, 62A Dr. Victor Babe Street, 430083 Baia Mare, Romania

ANDREI BUZATU
Alexandru Ioan Cuza University of Iai, Faculty of Geography and Geology, Department of Geology, 20A Carol I Blv.,
700505 Iai, Romania

ABSTRACT
The Coranda-Hondol ore deposit (Certej, Romania) is a sulfide ore deposit that was mined primarily for gold, silver, lead,
and zinc. Secondary minerals were formed through a precipitation process from sulfate solutions with a high concentration of
dissolved metals (especially Fe). These sulfate solutions resulted from acid mine drainage. Fourteen waste samples were analyzed through Raman spectrometry, X-ray diffraction, and scanning electron microscopy. Fe3+-, Fe2+-, Cu-, Zn-, Ca-, Mg-, and
MnAl-hydrated sulfates were identified. All are unstable when exposed to the laser beam of the Raman spectrometer.
Coquimbite, copiapite, ferricopiapite, hydroniumjarosite, and gunningite turn into anhydrous forms or oxides, depending on the
laser power. Gypsum turns into bassanite, while apjohnite loses all water molecules at 53.6 mW laser power on the surface
of the sample. Rhomboclase, melanterite, rozenite, antlerite, and brochantite break down without forming new minerals.
Fe2+-sulfates do not change into hematite under laser irradiation. Epsomite and hexahydrite are stable at 53.6 mW laser power.

Keywords: Raman spectrometry, weathering, sulfates, photochemical degradation, laser irradiation.

INTRODUCTION
Secondary minerals of supergene origin can be
found in various environments, especially in areas of
oxidation of metalliferous ore deposits containing sulfides. In some cases high and varied concentrations of
sulfosalts can form. These supergene alteration minerals
are mainly formed through oxidation of common sulfides (especially pyrite) and then through precipitation
from acid mine drainage (AMD) (Alpers et al. 2003).
Many previous studies have been undertaken by
Raman spectrometry to identify secondary sulfates,
especially those of Fe (Sasaki et al. 1998, Chio et al.

2007, Ling & Wang 2010, Frost 2011), Mg (Makreski


et al. 2005, Wang et al. 2006) and other cations (e.g.,
Ca, Cu, Pb) (Martens et al. 2003b, Makreski et al. 2005,
Buzgar et al. 2009b). These studies aim to understand
the geochemical process and ecological importance of
these alteration products (Buckby et al. 2003, Hammar
strom et al. 2005, Chou et al. 2013). Also, many of
these studies focus on fast and reliable discrimination
between members of a mineralogical series, where
the degree of hydration is the only difference (Wang
et al. 2006, Chio et al. 2007, Ling & Wang 2010,
Chou et al. 2013). All of these studies have reported
Raman spectra at high laser power (from 1 to 300 mW).

Corresponding author e-mail address: andrei_ionut1987@yahoo.com

canmin.00.00.1300054

09-06-15 13:5

THE CANADIAN MINERALOGIST

However, there are some misunderstandings in the


reported Raman spectra, as in some studies, the laser
power is not provided in the experimental section
(nominal or on the surface of the sample). Especially
for in situ and non-contact Raman measurements,
these reference spectra are essential for the correct
use of the information obtained by Raman spectroscopy in mineral identification.
Sample degradation under laser beam was reported
especially for several iron and manganese oxide and
oxyhydroxide species (Thibeau et al. 1978, Bernard
1993, de Faria et al. 1997, Worobiec et al. 2011).
Moreover, it was pointed out that hematite particles
are significantly more sensitive to the laser beam in
vacuum than in ambient air conditions (Worobiec
et al. 2011). The sensitivity to the laser power
depends also on the surface morphology, where flat
surfaces were significantly stable even under high
laser power (de Faria et al. 1997).
To our knowledge, this is the first comprehensive
study showing the transformation of natural secondary sulfates that occur under intense laser irradiation.
The minerals that have been identified so far on the
surface of the Coranda Hondol ore deposit are ferricopiapite, coquimbite, and epsomite (Apopei et al.
2012). Therefore, the goal of this work is to mineralogically identify the secondary minerals of waste samples from the Coranda-Hondol open pit (CHOP) through
Raman spectrometry, XRD, and SEM. In addition,
another goal is to establish the photochemical degradation
intensity thresholds of these minerals when exposed to a
laser beam by means of Raman spectrometry in ambient
air conditions (relative humidity 60% at 24 C). These
changes generally occur when an improper laser power is
used and leads to the formation of new products or, in
some instances, the destruction of the sample.

MATERIALS

AND

METHODS

Study area and sampling


The CHOP is located at the southern extremity of
the Golden Quadrilateral, South Apuseni Mountains
(45 59 37 N, 23 00 27 E, Metaliferi Mts.,
Romania). This area was mined for gold and silver (late
17th century) and recently hosted operations focused
mainly on the extraction of Pb and Zn. Ore from the
CHOP is part of the Certej metalliferous ore deposit.
This deposit is epithermal, of medium-to-high sulfidation type (Pricopie et al. 2004; Forward et al. 2009).
The main metalliferous minerals are galena (PbS), sphalerite (ZnS), and pyrite (FeS2). Chalcopyrite (CuFeS2)
and sulfosalts such as tetrahedrite-tennantite [(Cu,Fe)12
(Sb,As)4S13] and/or bournonite-seligmannite [PbCu(Sb,
As)S3] occur only at the microscopic level. Chalcopyrite
mainly occurs as chalcopyrite disease in sphalerite.
The gangue minerals (of no economic value) are quartz,

calcite, rhodocrosite, barite, and adularia (potassium


aluminosilicate).
Waste samples were randomly collected from the
surface of the CHOP during two seasons: winter 2011
and summer 2012. Weathering products were much
more abundant during the dry summer than the wet
winter. The samples were stored in polypropylene bags
and flasks in order to avoid loss of water that can
lead to mineralogical transformations [e.g., melanterite
Fe2+SO47(H2O) rozenite Fe2+SO44(H2O), within
a short time and in low humidity conditions at room
temperature (Chou et al. 2002)]. The samples were
prepared by hand picking using a binocular stereo
microscope. Mostly monomineral and micron-sized
samples were selected in order to analyze them by
Raman spectrometry, XRD and SEM.
Analytical methods
Raman spectra were obtained at room temperature
with a Raman Spectrograph Horiba Jobin-Yvon RPAHE 532 with a multichannel air cooled (70 C) CCD
detector, using a Nd-Yag laser 532 nm excitation
source and a nominal power of 100 mW. Spectra
were obtained in the spectral range between 210 and
3400 cm1 with a spectral resolution of 3 cm1. The
Raman system includes a Superhead fiber optic Raman
probe for non-contact measurements, with a 50 LWD
Olympus objective, NA = 0.50, WD = 10.6 mm, and
FIB50/10M optical fiber. The laser spot diameter on the
sample surface was approximately 23 m (the minimum theoretical spot diameter is 1.3 m). Sulfur and
ciclohexane were used for the calibration. Data acquisition was performed at a laser power of 153.6 mW on
the surface of the sample. Furthermore, the laser power
was gradually increased by 1% until any photochemical
degradation was observed. Spectra manipulations consist of basic data treatment, such as smoothing adjustments and peak fitting (Lorentz function).
X-ray diffraction analyses were performed using a
Shimadzu X-ray diffractometer LabX XRD-6000 in
order to check the presence of minerals found with
Raman spectrometry. The measurement conditions
were: Cu anode, 40 kV, 30 mA, divergence slit 1.00
deg, scatter slit 1.00 deg, receiving slit 0.30 mm, scan
range 580.0, continuous scan, scan speed 3 deg/min,
sampling pitch 0.02 deg, preset time 0.40 s.
The crystal habit of the identified minerals was
studied by scanning electron microscopy (SEM)
(gold coated samples), using a Tescan Vega II SBH
instrument.

RESULTS

AND

DISCUSSION

A series of 14 secondary minerals were identified.


All are sulfates (Table 1). Rhomboclase was formed
by weathering of pyrite and, to a small extent, chalcopyrite and sulfosalts such as tetrahedrite-tennantite. It

(H3O)Fe3+(SO4)23H2O

Fe3
2 SO4 3 9H2 O
Fe2 Fe3
4 SO4 6 OH2 20H2 O

3
Fe3
2=3 Fe4 SO4 6 OH2 20H2 O

H3 OFe3
3 SO4 2 OH6

Fe2+SO47(H2O)

Fe2+SO44(H2O)

MgSO47(H2O)

MgSO46(H2O)

CaSO42(H2O)

MnAl2 (SO4)422(H2O)

Cu3(SO4) (OH)4

Cu4(SO4) (OH)6

(Zn, Mn)SO4H2O

Fe3+
Rhomboclase

Coquimbite
Copiapite

Ferricopiapite

Hydroniumjarosite

Fe2+
Melanterite

Rozenite

Mg-, Al-, CaEpsomite

Hexahydrite

Gypsum

Apjohnite

Cu and Zn
Antlerite

Brochantite

Gunningite

1601vw, 1171vw, 1136vw, 1077w, 989vs, 928vw, 860vw, 788vw, 759vw,


630w, 604w, 477m, 446w, 420m, 343vw, 301vw, 266vw
3397w, 1598vw, 1126vw, 1094vw, 1077vw, 975vs, 914vw, 903vw, 777vw,
737vw, 625sh, 608w, 506sh, 477s, 423vw, 389w, 320vw, 262sh,
246vw
3230vw, 1492vw, 1192w, 1087w, 1024vs, 884vw, 665vw, 626w, 503w,
423w, 307sh, 277vw, 219vw

3285w, 3219sh, 1668vw, 1136vw, 1097vw, 1062vw, 985vs, 615vw,


449w, 371vw, 246vw
3398w, 3268sh, 1656vw, 1147vw, 1084vw, 985vs, 611vw, 469vw, 449sh,
370vw, 242vw
3398w, 1136w, 1106sh, 1010vs, 671vw, 623vw, 576vw, 495vw, 416w,
312vw, 215vw
3379w, 3299w, 3237sh, 3007sh, 1630vw, 1227sh, 1141vw, 1116vw,
1108vw, 1086vw, 1073sh, 996vs, 978sh, 619w, 529vw, 469w, 432sh,
311vw

3398w, 3350sh, 3235sh, 1653vw, 1139vw, 1100vw, 978vs, 615vw,


452vw, 376vw, 235vw
3388w, 3329sh, 3261sh, 1594vw, 1176sh, 1149vw, 1098sh, 1073vw,
992vs, 658sh, 612vw, 480w, 461sh, 383vw, 284vw, 239sh

2775vw, 2661vw, 1456vw, 1181w, 1081sh, 1028sh, 1014vs, 763vw,


650sh, 622sh, 603w, 472sh, 454w, 381m, 265sh, 242w
1201w, 1167vw, 1097vw, 1024vs, 602vw, 506w, 457vw, 285w
3147w, 1651vw, 1247w, 1143sh, 1113m, 1031s, 999vs, 748vw, 637sh,
609w, 558vw, 477s, 304sh, 274s, 246sh
3098w, 1635w, 1227w, 1119m, 1021vs, 992vs, 749vw, 638sh, 612w,
480s, 304sh, 274s, 250sh
3395m, 1633vw, 1165w, 1105vs, 1014s, 623m, 573w, 452sh, 427s,
361sh, 308vw, 296vw, 227m

v(cm1) (principal Raman bands)a

Notes: a - vsVery strong; sstrong; mmedium; wweak; vwvery weak; shshoulder.


b
- Abundance/Distribution: L - locally distributed; S - spread out; quantitative abundance: from less (+) to very abundant (+++).

Ideal formula

Mineral name

++/S

+/L

+/L

+++/S

++/S

++/S

+++/S

++/S

++/S

++/S

+++/S

++/S
+++/S

++/L

XRD, Raman

Raman, SEM

Raman

XRD, Raman, SEM

XRD, Raman, SEM

XRD, Raman

Raman, SEM

XRD, Raman, SEM

XRD, Raman

XRD, Raman

XRD, Raman, SEM

XRD, Raman, SEM


XRD, Raman, SEM

Raman

Methods

TABLE 1. MINERAL NAMES, IDEAL CHEMICAL FORMULAE, PRINCIPAL RAMAN BANDS, AND DISTRIBUTION AND METHODS OF IDENTIFICATION FOR
THE MINERALS IDENTIFIED IN THIS STUDY

canmin.00.00.1300054
09-06-15 13:5

the raman study of weathering minerals from the coranda-hondol open pit
3

canmin.00.00.1300054

09-06-15 13:5

THE CANADIAN MINERALOGIST

occurs only rarely at CHOP. The Raman spectrum of


rhomboclase (Fig. 1a) shows characteristic spectral
bands at 1014 cm1 and 1028 cm1 (shoulder), which
are due to the (1) symmetric stretching of the SO4
group. Rhomboclase is stable at laser powers less
than 10.5 mW (on the surface of the sample). New
minerals do not form through photochemical degradation at laser powers greater than 10.5 mW (i.e., the
Raman spectra do not contain any Raman lines).
The presence of coquimbite is mainly attributed
to the FeSO4-rich river water (Bigham & Nordstrom
2000). It occurs as tabular hexagonal crystals up to
0.01 mm in size [Fig. S1a (Supporting Information);
available from the Depository of Unpublished Data
on the MAC website, document weathering minerals
CM52_10.3749/canmin.1300054]. The Raman spectrum
of coquimbite (Fig. 1b), acquired at a laser power of
10.5 mW, shows a strong spectral line at 1024 cm1.
It is assigned to the (1) symmetric stretching vibration of the SO4 group. Coquimbite turns into mikasaite [Fe23+(SO4)3] when the laser power is increased
above 14 mW. The resulting Raman spectrum

(Fig. 1c) shows stretching vibrations of the SO4


group, at 1125, 1101, and 1083 cm1, which is in
good agreement with the Raman spectrum of mikasaite from the literature (Ling & Wang 2010).
Copiapite and ferricopiapite are also precipitated
directly from the FeSO4-rich river water (Bigham &
Nordstrom 2000). Copiapite occurs as platy crystals up
to 0.02 mm [Fig. S1b (Supporting Information)] and
ferricopiapite appears with a scaly shape [Fig. S1c
(Supporting Information)]. The Raman spectra of copiapite and ferricopiapite were acquired at a laser power of
7.4 mW and show two peaks assigned to (1) symmetric
stretching of SO4 at 1031 and 999 cm1 for copiapite,
and 1021 and 992 cm1 for ferricopiapite (Fig. 1d, 1e).
Their Raman spectral fingerprints are in good agreement
with other studies (Downs 2006, Ling & Wang 2010,
Frost 2011, Kong et al. 2011).
When the laser power is increased above 18.4 mW
the 1 bands in both spectra trend to higher wavenumbers and form a single broad band with maxima at
1087 cm1 (Fig. 1f, 1g, 1h). The change of Raman
spectra of copiapite and ferricopiapite under laser

FIG. 1. The Raman spectra of rhomboclase at 10.5 mW (a), coquimbite at 10.5 mW (b), mikasaite at 14 mW (c), copiapite
at 7.4 mW (d), ferricopiapite at 7.4 mW (e), intermediate phases between copiapite (ferricopiapite) and mikasaite (f, g),
and mikasaite at 18.4 mW (h).

canmin.00.00.1300054

09-06-15 13:5

the raman study of weathering minerals from the coranda-hondol open pit
beam exposure may have two explanations: (1) at laser
power values above 18.4 mW, the ferricopiapite
and copiapite dehydrate into quasi-amorphous iron
hydroxide sulfate Fe(OH)SO4 (Lu & Wang 2012);
(2) these minerals turn into mikasaite via dehydration.
The second hypothesis is more plausible because the
Raman spectrum acquired (Fig. 1h) fits very well with
that of mikasaite (Fig.1c), much better than with that
of a quasi-amorphous phase.
Hydroniumjarosite is commonly associated with the
oxidation zones of sulfide deposits (Brophy & Sheridan
1965, Frost et al. 2006). The Raman spectral fingerprint
of this mineral can be clearly observed in Figure 2a.
The spectrum was acquired at 3.90 mW laser power.
Discrimination of Raman spectra of the members of the
jarosite-natrojarosite-hydroniumjarosite solid solution
series can be achieved using the following peaks:
3400 cm1, 1014 cm1, and 430 cm1 (Sasaki
et al. 1998, Frost et al. 2006, Lu & Wang, 2012).
When the laser power is increased above 14.3 mW
the spectrum of hydroniumjarosite changes into a
new spectrum (Fig. 2b) with the main Raman bands at

1308, 1064, 608, 495, 408, 293, and 223 cm1.


These bands clearly indicate the presence of hematite
and this is in good agreement with the literature (de
Faria et al. 1997, Shim & Duffy 2002, Downs 2006,
Buzgar et al. 2009a).
The Fe2+-bearing sulfates melanterite and rozenite
occur in large quantities at CHOP. The natural conversion of melanterite to rozenite is well-known to be
very fast and may occur at 30 C within 23 days
(Zodrow & McCandlish 1978, Chou et al. 2002, Chou
et al. 2013). The Raman spectra of melanterite and
rozenite (Fig. 2c, 2d) were acquired at a laser power
of 10 mW. Vibrational modes of symmetric stretching
of the SO4 tetrahedra appear at 978 cm1 for melanterite and 992 cm1 for rozenite. The linear correlation
between Raman wavenumbers of 1, 4 and the number of water molecules has been noted previously in
many studies of hydrated sulfates (Wang et al. 2006,
Chio et al. 2007, Chou et al. 2013).
In order to cause photochemical degradation under
laser beam exposure, the power of the laser was gradually
increased to the maximum 53.6 mW. Around 14 mW,

FIG. 2. The Raman spectra of hydroniumjarosite at 3.9 mW (a), hematite at 14.3 mW (b), melanterite at 5 mW (c), rozenite
at 5 mW (d), synthetic rozenite at 5 mW (e), szomolnokite obtained by heating of synthetic rozenite at 105 C for 30 min
(f), mikasaite obtained by heating of synthetic rozenite at 700 C for 30 min (g), and hematite obtained by heating of
synthetic rozenite at 800 C for 30 min (h); 3 magnication of the spectrum in a specic region is shown.

canmin.00.00.1300054

09-06-15 13:5

THE CANADIAN MINERALOGIST

the crystal structure of both minerals starts to break


down. The instability of rozenite under laser irradiation
was reported at 7.37 mW (Buzatu et al. 2012). This
laser power difference is most probably due to the size
of the crystals. In our case, melanterite did not register
as turning into rozenite as we would have expected.
This transformation is probably very fast and followed
by immediate breakdown of the rozenite. Further
increasing the laser power did not result in new mineral formation as in the case of hydroniumjarosite.
Furthermore, a sample of synthetic rozenite was used
to provide an accurate Raman spectrum (Fig. 2e),
which broke down around 15 mW laser power. Also,
no new mineral appeared up to maximum laser power.
However, by heating synthetic rozenite in an electric
oven, it turned into szomolnokite (105 C, 30 min,
Fig. 2f), mikasaite hematite (700 C, 30 min,
Fig. 2g), and hematite (800 C, 30 min, Fig. 2h). In
the Raman spectrum of mikasaite, hematite was identified based on two bands: 1315 cm1 and 293 cm1
(Fig. 2g). Therefore, melanterite and rozenite break
down under laser beam exposure, but they do not turn

into new minerals (mikasaite and after that hematite)


because the amount of heat transferred from the laser
beam to the sample is much lower than that of the electric oven and was insufficient for Fe2+ oxydation.
Natural rozenite also does not show transformation
into szomolnokite under increased laser power. Most
probably this transformation is also very fast and followed by immediate szomolnokite breakdown (as in
the case of the melanterite-rozenite transformation).
Apjohnite was found in several areas of the CHOP.
Crystals occur as elongated prisms up to 0.1 mm in
length [Fig. S1d, S1e (Supporting Information)]. The
Raman spectrum clearly shows its presence (Fig. 3a).
The main Raman feature consists of the very strong
peak located at 996 cm1 accompanied by a shoulder
at 978 cm1. Both are due to the (1) symmetric
stretching mode of the SO4 group. This mineral is
more stable under laser beam exposure and breaks
down only at maximum laser power (53.6 mW).
The main Raman line shifts from 996 cm1 up to
1055 cm1 (Fig. 3b). This is in agreement with the
general rule of hydrated sulfates, which postulates that

FIG. 3. The Raman spectra of apjohnite at 22.9 mW (a), anhydrous form of apjohnite at 53.6 mW (b), gypsum at 14.3 mW
(c), bassanite at 35 mW (d), hexahydrite at 53.6 mW (e), and epsomite at 53.6 mW (f); 3, 4, and 5 magnication of
the spectrum in a specic region is shown.

canmin.00.00.1300054

09-06-15 13:5

the raman study of weathering minerals from the coranda-hondol open pit
the v1 Raman mode trends to higher wavenumbers
with dehydration. For this reason it can be considered
that apjohnite turns into the anhydrous form.
Gypsum is widespread in waste samples from the
CHOP. It occurs as tabular monoclinic crystals up to
0.01 mm in length [Fig. S1a (Supporting Information)].
Its presence is due to calcite, one of main gangue minerals in the CHOP ore. Calcite breaks down and gypsum
appears in the acidic environment of the oxidation zones
of the sulfide deposits. Conversion of gypsum into bassanite as a function of temperature is irreversible and
well established (Sarma et al. 1998). Conversion of gypsum into bassanite under laser irradiation occurs after
increasing laser power above 35 mW (Fig. 3c, 3d).
In some cases gypsum turns into bassanite only at maximum laser power (53.6 mW). Crystal size certainly
influences the power required.
Hexahydrite and epsomite occur as very fine crystalline aggregates [Fig. S1f (Supporting Information)].
Raman spectra are shown in Figures 3e and 3f. The (v1)
symmetric stretching vibrations of both minerals appear
at 985 cm1. Taking into account our spectra and data
from the literature (Wang et al. 2006, Frezzotti et al.
2012), the distinction between hexahydrite and epsomite
can be done using the 3 bands of SO4 and the vibrational modes of H2O. The Raman bands at 1147 and
1084 cm1 are assigned to hexahydrite, while the (3)
antisymmetric stretching mode of epsomite appears at
1136, 1097, and 1062 cm1. The vibrational modes of

H2O are represented by bands at 3400 cm1 (hexahydrite)


and 3300 cm1 and a shoulder at 3219 cm1 (epsomite).
Taking into account the principle of repeatability,
multiple Raman spectra were acquired, and none (at
53.6 mW laser power on the surface of the sample)
show photochemical degradation of these Mg-sulfates.
This could be due to the fact that magnesium sulfates
have a refractive character.
The basic Cu sulfates brochantite and antlerite were
found locally in small amounts and are very easy to
distinguish by their green color. The Raman spectrum
of brochantite (Fig. 4a) shows the (1) symmetric
stretching mode at 975 cm1. Antlerite (Fig. 4b) has
the (1) symmetric stretching mode at 989 cm1.
Another distinguishing spectral feature between these
two basic Cu sulfates is (3) antisymmetric stretching
modes, for brochantite at: 1126, 1094, and 1077 cm1,
and for antlerite at 1171, 1136, and 1077 cm1. In the
low-wavenumber region, brochantite and antlerite have
spectra that are roughly similar, especially for the (2)
vibrational modes. The Raman spectra of brochantite
and antlerite are in good agreement with those reported
in the literature (Martens et al. 2003a, Downs 2006,
Buzgar et al. 2009a).
Brochantite and antlerite are stable at 5 and 3.9 mW,
respectively. Breakdown of the crystal structure results
after increasing laser power above these values. Another
study pointed out that these basic Cu sulfates had undergone changes under near infrared laser beam exposure

FIG. 4. The Raman spectra of brochantite at 5 mW (a), antlerite at 3.9 mW (b), and of both minerals after exposure under
intense laser beam at 53.6 mW (c); Raman band at 1591cm1 belongs to carbon.

canmin.00.00.1300054

09-06-15 13:5

THE CANADIAN MINERALOGIST

(785 nm) of 1 mW (Martens et al. 2003a). As previously noted, crystal size influences the laser power at
which these minerals break down. No new minerals
appear after the breakdown of these minerals (Fig. 4c).
The Raman line at 1591 cm1 belongs to carbon, which
comes from organic matter in the sample that is subjected to intense laser irradiation.
Gunningite was found on the surface of the sphalerite crystals, in particular. Goslarite (ZnSO47H2O) and
bianchite [(Zn,Fe2+)SO46H2O] are more widespread in
the oxidation zone of the metalliferous ore deposits containing sulfides (Chou et al. 2013) but at CHOP only
gunningite was identified. Its presence was confirmed
by XRD [Fig. S2 (Supporting Information)]. The
Raman spectrum of gunningite from CHOP (Fig. 5a)
fits very well with the Raman spectrum of synthetic
zinc sulfate monohydrate (Fig. 5b). A strong band
occurs at 1024 cm1 in both Raman spectra and is due
to (1). Two bands of the (3) antisymmetric stretching
mode are observed at 1192 and 1087 cm1. The bands
at 503 and 423 cm1 are assigned to the (2) bending

mode. For the (4) bending mode two bands located at


665 and 626 cm1 are observed. The peaks below
300 cm1 are assigned to the ZnO vibration. The (1)
symmetric stretching mode of water molecules is missing from the recorded range (2103400 cm1). The
Raman line at 1492 cm1 is assigned to the (2) bending
mode of H2O, while the broad band around 3230 cm1 is
considered to be from the stretching of H2O involved
in H bonding (Rudolph et al. 1999, Downs 2006).
Gunningite loses water and turns into zincosite
under laser beam exposure of ca. 35 mW. The Raman
spectrum (Fig. 5g) also contains Raman lines for pyrite
(378 and 343 cm1) and carbon (1600 cm1). The
Raman spectrum of synthetic zincosite (Fig. 5f)
obtained by heating synthetic gunningite in an electric
oven at 400 C fits well enough with the Raman spectrum of gunningite photochemically decomposed under
intense laser irradiation (Fig. 5g). The Raman spectrum
of synthetic zincosite shows (1) stretching vibrations
of SO4 at 1098, 1049, and 996 cm1. These values are
in accordance with the literature (Rudolph et al. 1999).

FIG. 5. The Raman spectrum of gunningite from CHOP at 7.37 mW (a); synthetic gunningite at 7.37 mW (b); gunningite
heated at 105 C for 30 min (c); synthetic gunningite heated at 800 C for 30 min (zinc oxysulfate) (d); synthetic gunningite heated at 900 C for 30 min (zinc oxide) (e); synthetic gunningite heated at 400 C for 30 min (zincosite) (f); photochemical degradation of gunningite (zincosite) at 35.8 mW (g); 3 magnication of the spectrum in a specic region
is shown.

canmin.00.00.1300054

09-06-15 13:5

the raman study of weathering minerals from the coranda-hondol open pit
Triple degeneration of the stretching vibrations modes
suggests a large variation in SO4 tetrahedral distortion
(different OSO) due to lattice distortion which can
be attributed to the high temperature.
At high temperature, synthetic gunningite turns into
zinc oxysulfate (ZnO2ZnSO4) (Fig. 5d) and zinc
oxide (Fig. 5e) (Mu & Perlmutter 1981, Sharma &
Exarhos 1997, Siriwardane et al. 1999). These compounds do not form from photochemical degradation
of natural gunningite under laser beam exposure.

.
.
.
.
.
.

CONCLUSIONS
Pyrite is the main sulfide mineral from the Coranda
Hondol open pit which turns into secondary minerals
through oxidation and hydration. It is followed by
sphalerite with a low content of Fe (13% Fe, after
Udubaa et al. 1982) and chalcopyrite. Galena,
although widespread, did not lead to the formation of
secondary lead minerals (e.g., anglesite). This can be
explained by the fact that lead sulfates (especially
anglesite) have very high solubility in acidic waters,
and their precipitation is not possible under these
circumstances (Dove & Czank 1995).
Based on the main sulfides which generate oxidation
reactions, neutralization, hydration, and/or dehydration
and lead to the formation of supergene alteration, secondary minerals can be divided into the following categories (in order of abundance from most to least):
. Fe(III)-sulfates rhomboclase, copiapite, ferricopiapite, coquimbite, and hidroniumjarosite;
these are the most abundant secondary minerals,

mainly from the copiapite-group, and occur on


the walls of benches;
Fe(II)-sulfates melanterite and rozenite,
found in high concentrations at the bottom of
the open pit;
Zn-sulfates gunningite generally occurs on the
walls of the benches, as a white powder, but only
where sphalerite mineralization is also present;
Cu-sulfates antlerite and brochantite;
Ca-sulfate exclusively gypsum;
Mg-sulfates epsomite and hexahydrite;
MnAl-sulfates apjohnite.

In this manuscript, we show that some common


secondary sulfate minerals undergo changes in the
laser beam if too high a laser power is used (generally
between 715 mW on the sample surface). Some of
them turn into other minerals. This photochemical
breakdown is generally linked to dehydration and
oxidation of iron(II) to iron(III). Other minerals are
unstable and suffer lattice destruction without turning
into other minerals. Hydroniumjarosite is the only mineral that turns into hematite at maximum laser power.
In the case of Fe2+ sulfates, laser power is insufficient
to cause Fe2+ oxydation and hematite formation.
The laser intensity thresholds and optimum experimental conditions for the non-destructive identification
of the sulfates found at CHOP by Raman spectrometry
are summarized in Table 2. For all samples analyzed
by Raman spectrometry, the damage only occurred at
higher laser powers.
This work will significantly improve the understanding of how the Raman spectra of sulfates are affected by

TABLE 2. DAMAGE THRESHOLDS IN LASER RAMAN SPECTROMETRY OF SULFATE


MINERALS IDENTIFIED IN THIS STUDY
Mineral
Rhomboclase
Coquimbite
Copiapite
Ferricopiapite
Hydroniumjarosite
Melanterite
Rozenite
Epsomite
Hexahydrite
Gypsum
Apjohnite
Antlerite
Brochantite
Gunningite

Laser power threshold


10.5 mW
14.3 mW
18.4 mW
18.4 mW
14.3 mW
14.3 mW
14.3 mW
*
*
35 mW
53.6 mW
3.9 mW
5 mW
35 mW

Photochemical degradation product


x
Mikasaite
Mikasaite
Mikasaite
Hematite
x
x

Bassanite
anhydrous Apjohnite
x
x
Zincosite

Notes: * do not turn into a new mineral or break down at the maximum power of 53.6 mW;
x break down above the specified laser power threshold.

canmin.00.00.1300054

09-06-15 13:5

10

THE CANADIAN MINERALOGIST

differences in instrument parameters (in principal laser


power on the surface of the sample). Also, the results of
this paper contribute to improved decision making
regarding the laser power used for analysis, which is
important for suitable conservation interventions and
also for proper identification of mineralogical phases.

ACKNOWLEDGMENTS
The authors gratefully acknowledge the staff of
S.C. DevaGold S.A. for access to the Coranda Hondol
open pit and their support at the sampling site. Thanks
are also extended to Prof. Dr. Essaid Bilal and to
anonymous reviewers for valuable suggestions and
comments to improve the quality of the paper.

REFERENCES
ALPERS, C.N., NORDSTROM, D.K., & SPITZLEY, J. (2003) Extreme
acid mine drainage from a massive sulfide deposit: The Iron
Mountain end-member. In Environmental Aspects of Mine
Wastes (J.L. Jambor, D.L. Blowes, & A.I.M. Ritchie, eds.).
Mineralogical Association of Canada Short Course Series
31, 407430.
APOPEI, A.I., DAMIAN, G., & BUZGAR, N. (2012) A preliminary
Raman and FT-IR spectroscopic study of secondary
hydrated sulfate minerals from the Hondol open pit
(Metaliferi Mts., Romania). Romanian Journal of Mineral
Deposits 85(2), 16.
BERNARD, M.-C. (1993) Electrochromic Reactions in
Manganese Oxides. Journal of The Electrochemical
Society 140(11), 3065.
BIGHAM, J.M. & NORDSTROM, D.K. (2000) Iron and Aluminum
Hydroxysulfates from Acid Sulfate Waters. Reviews in
Mineralogy and Geochemistry 40(1), 351403.
BROPHY, G.P. & SHERIDAN, M.F. (1965) Sulfate Studies IV:
The Jarosite - Natrojarosite - Hydronium Jarosite solid
solution series. American Mineralogist 50, 15951607.
BUCKBY, T., BLACK, S., COLEMAN, M.L., & HODSON, M.E. (2003)
Fe-sulphate-rich evaporative mineral precipitates from the
Rio Tinto, southwest Spain. Mineralogical Magazine 67(2),
263278.
BUZATU, A., DAMIAN, G., & BUZGAR, N. (2012) Raman and
Infrared studies of weathering products from Baia Sprie ore
deposit (Romania). Romanian Journal of Mineral Deposits
85(2), 710.
BUZGAR, N., APOPEI, A.I., & BUZATU, A. (2009a) Romanian
Database of Raman Spectroscopy. http://rdrs.uaic.ro.
BUZGAR, N., BUZATU, A., & SANISLAV, I.V. (2009b) The Raman
study on certain sulfates. Analele tiinifice ale Universitatii
Alexandru Ioan Cuza Iai, Seria Geologie (AUI-G)
LV(1).

CHIO, C.H., SHARMA, S.K., & MUENOW, D.W. (2007) The


hydrates and deuterates of ferrous sulfate (FeSO4): a Raman
spectroscopic study. Journal of Raman Spectroscopy 38(1),
8799.
CHOU, I.-M., SEAL, R.R.I., & HEMINGWAY, B.S. (2002)
Determination of melanterite-rozenite and chalcanthitebonattite equilibria by humidity measurements at 0.1
MPa. American Mineralogist 87(1), 108114.
CHOU, I.M., SEAL, R.R., & WANG, A. (2013) The stability of
sulfate and hydrated sulfate minerals near ambient conditions and their significance in environmental and planetary sciences. Journal of Asian Earth Sciences 62,
734758.
DE

FARIA, D.L.A., VENNCIO SILVA, S., & DE OLIVEIRA, M.T.


(1997) Raman microspectroscopy of some iron oxides
and oxyhydroxides. Journal of Raman Spectroscopy
28(11), 873878.

DOVE, P.M. & CZANK, C.A. (1995) Crystal chemical controls on


the dissolution kinetics of the isostructural sulfates: Celestite,
anglesite, and barite. Geochimica et Cosmochimica Acta
59(10), 19071915.
DOWNS, R.T. (2006) The RRUFF Project: an integrated study
of the chemistry, crystallography, Raman and infrared
spectroscopy of minerals. Program and Abstracts of the
19th General Meeting of the International Mineralogical
Association in Kobe, Japan, O03-13.
FORWARD, P., LIDDELL, N., & JACKSON, T. (2009) Certej
Updated Definitive Feasibility Study Summary Technical
Report. European Goldfields Limited, Romania.
FREZZOTTI, M.L., TECCE, F., & CASAGLI, A. (2012) Raman
spectroscopy for fluid inclusion analysis. Journal of
Geochemical Exploration 112, 120.
FROST, R.L. (2011) A Raman spectroscopic study of copiapites
Fe2+Fe3+(SO4)6(OH)220H2O: environmental implications.
Journal of Raman Spectroscopy 42(5), 11301134.
FROST, R.L., WILLS, R.A., KLOPROGGE, J.T., & MARTENS, W.N.
(2006) Thermal decomposition of hydronium jarosite
(H3O)Fe3(SO4)2(OH)6. Journal of Thermal Analysis and
Calorimetry 83(1), 213218.
HAMMARSTROM, J.M., SEAL, R.R., MEIER, A.L., & KORNFELD, J.M.
(2005) Secondary sulfate minerals associated with acid
drainage in the eastern US: recycling of metals and acidity in surficial environments. Chemical Geology 215(14),
407431.
KONG, W.G., WANG, A., FREEMAN, J.J., & SOBRON, P. (2011) A
comprehensive spectroscopic study of synthetic Fe2+, Fe3+,
Mg2+ and Al3+copiapite by Raman, XRD, LIBS, MIR and
vis-NIR. Journal of Raman Spectroscopy 42(5), 11201129.
LING, Z.C. & WANG, A.L. (2010) A systematic spectroscopic
study of eight hydrous ferric sulfates relevant to Mars.
Icarus 209(2), 422433.

canmin.00.00.1300054

09-06-15 13:5

the raman study of weathering minerals from the coranda-hondol open pit
LU, Y. & WANG, A. (2012) Synthesis and spectral characterization of OH-bearing ferric sulfates. Lunar and Planetary
Science Conference, XXXXIII, Houston, Texas, United
States.
MAKRESKI, P., JOVANOVSKI, G., & DIMITROVSKA, S. (2005)
Minerals from Macedonia. Vibrational Spectroscopy
39(2), 229239.
MARTENS, W., FROST, R.L., KLOPROGGE, J.T., & WILLIAMS, P.A.
(2003a) Raman spectroscopic study of the basic cop
per sulphatesimplications for copper corrosion and
bronze disease. Journal of Raman Spectroscopy 34,
145151.
MARTENS, W., FROST, R.L., KLOPROGGE, J.T., & WILLIAMS,
P.A. (2003b) Raman spectroscopic study of the basic
copper sulphatesimplications for copper corrosion and
bronze disease. Journal of Raman Spectroscopy 34(2),
145151.
MU, J. & PERLMUTTER, D.D. (1981) Thermal decomposition
of inorganic sulfates and their hydrates. Industrial &
Engineering Chemistry Process Design and Development
20(4), 640646.
PRICOPIE, M., TUA, L., CRISTEA, P., CPRARU, N., & MARTON, I.
(2004) Geology of the Certej project area and a new model
for high-grade gold mineralisation hosted within the Dealul
Grozii - Hondol Perimeter (Certej deposit). In Gold-SilverTelluride Deposits of the Golden Quadrilateral, South
Apuseni Mts., Romania (C. Ciobanu, N.J. Cook, G. Damian,
F. Damian, & G. Buia, eds.). Guidebook of the International
Field Workshop of IGCP project 486, Alba Iulia, Romania,
31st August 7th September 2004.
RUDOLPH, W.W., BROOKER, M.H., & TREMAINE, P.R. (1999)
Raman Spectroscopy of Aqueous ZnSO4 Solutions under
Hydrothermal Conditions: Solubility, Hydrolysis, and
Sulfate Ion Pairing. Journal of Solution Chemistry 28(5),
621630.
SARMA, L.P., PRASAD, P.S.R., & RAVIKUMAR, N. (1998)
Raman spectroscopic study of phase transitions in natural
gypsum. Journal of Raman Spectroscopy 29(9), 851856.

11

SASAKI, K., TANAIKE, O., & KONNO, H. (1998) Distinction


of jarosite-group compounds by Raman spectroscopy.
Canadian Mineralogist 36, 12251235.
SHARMA, S.K. & EXARHOS, G.J. (1997) Raman Spectroscopic
Investigation of ZnO and Doped ZnO Films, Nanoparticles
and Bulk Material at Ambient and High Pressures. Solid
State Phenomena 55, 3237.
SHIM, S.H. & DUFFY, T.S. (2002) Raman spectroscopy of Fe2O3
to 62 GPa. American Mineralogist 87(23), 318326.
SIRIWARDANE, R.V., POSTON, J.A., JR., FISHER, E.P., SHEN, M.-S.,
& MILTZ, A.L. (1999) Decomposition of the sulfates of
copper, iron (II), iron (III), nickel, and zinc: XPS, SEM,
DRIFTS, XRD, and TGA study. Applied Surface Science
152(34), 219236.
THIBEAU, R.J., BROWN, C.W., & HEIDERSBACH, R.H. (1978)
Raman Spectra of Possible Corrosion Products of Iron.
Applied Spectroscopy 32(6), 532535.
UDUBAA, G., ISTRATE, G., & VLUREANU, M. (1982)
Metallogenesis of Coranda - Hondol region, Metaliferi
Mountains (Metalogeneza regiunii Coranda - Hondol,
Munii Metaliferi). D.S. Inst. Geol. LXVII(2), 197232
(in Romanian).
WANG, A., FREEMAN, J.J., JOLLIFF, B.L., & CHOU, I.M. (2006)
Sulfates on Mars: A systematic Raman spectroscopic study
of hydration states of magnesium sulfates. Geochimica et
Cosmochimica Acta 70(24), 61186135.
WOROBIEC, A., DARCHUK, L., BROOKER, A., POTGIETER, H., &
VAN GRIEKEN, R. (2011) Damage and molecular changes
under a laser beam in SEM-EDX/MRS interface: a case study
on iron-rich particles. Journal of Raman Spectroscopy 42(4),
808814.
ZODROW, E.L. & MCCANDLISH, K. (1978) Hydrated Sulfates
in the Sydney Coalfield, Cape Breton, Nova Scotia.
Canadian Mineralogist 16, 1722.
Received December 13, 2013, revised manuscript accepted
November 24, 2014.

Vous aimerez peut-être aussi