Вы находитесь на странице: 1из 12

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/236861821

Zincian dolomite related to supergene


alteration in the Iglesias mining district (SW
Sardinia)
ARTICLE JANUARY 2013
DOI: 10.1007/s00531-012-0785-0

CITATIONS

READS

79

5 AUTHORS, INCLUDING:
Nicola Mondillo

Giuseppina Balassone

University of Naples Federico II

University of Naples Federico II

24 PUBLICATIONS 74 CITATIONS

56 PUBLICATIONS 323 CITATIONS

SEE PROFILE

SEE PROFILE

Michael M. Joachimski

Abner Colella

Friedrich-Alexander-University of Erlangen-

University of Naples Federico II

194 PUBLICATIONS 4,219 CITATIONS

50 PUBLICATIONS 278 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Michael M. Joachimski


Retrieved on: 02 January 2016

Int J Earth Sci (Geol Rundsch) (2013) 102:6171


DOI 10.1007/s00531-012-0785-0

ORIGINAL PAPER

Zincian dolomite related to supergene alteration in the Iglesias


mining district (SW Sardinia)
M. Boni N. Mondillo G. Balassone
M. Joachimski A. Colella

Received: 13 August 2011 / Accepted: 21 April 2012 / Published online: 29 May 2012
Springer-Verlag 2012

Abstract One of the main effects of supergene alteration


of ore-bearing hydrothermal dolomite in areas surrounding
secondary zinc orebodies (Calamine-type nonsulfides) in
southwestern Sardinia (Italy) is the formation of a broad
halo of Zn dolomite. The characteristics of supergene Zn
dolomite have been investigated using scanning electron
microscopy and qualitative energy-dispersive X-ray spectroscopy, thermodifferential analysis, and stable isotope
geochemistry. The supergene Zn dolomite is characterized
by variable amounts of Zn, and low contents of Pb and Cd
in the crystal lattice. It is generally depleted in Fe and Mn
relative to precursor hydrothermal dolomite (Dolomia
Geodica), which occurs in two phases (stoichiometric
dolomite followed by Fe-Mn-Zn-rich dolomite), well distinct in geochemistry. Mg-rich smithsonite is commonly
associated to Zn dolomite. Characterization of Zn-bearing
dolomite using differential thermal analysis shows a drop
in temperature of the first endothermic reaction of dolomite
decomposition with increasing Zn contents in dolomite.
The supergene Zn dolomites have higher d18O but lower
d13C values than hydrothermal dolomite. In comparison
with smithsonite-hydrozincite, the supergene Zn dolomites
have higher d18O, but comparable d13C values. Formation
of Zn dolomite from meteoric waters is indicated by low
d13C values, suggesting the influence of soil-gas CO2 in
near-surface environments. The replacement of the dolomite host by supergene Zn dolomite is interpreted as part of
M. Boni (&)  N. Mondillo  G. Balassone  A. Colella
Dipartimento di Scienze della Terra, Universita` di Napoli,
Via Mezzocannone 8, 80134 Naples, Italy
e-mail: boni@unina.it
M. Joachimski
GeoZentrum Nordbayern University of Erlangen-Nuremberg,
Schlossgarten 5, 91054 Erlangen, Germany

a multistep process, starting with a progressive zincitization of the dolomite crystals, followed by a patchy
dedolomitization s.s. and potentially concluded by the
complete replacement of dolomite by smithsonite.
Keywords SW Sardinia  Zn dolomite  Supergene 
Nonsulfides

Introduction
In the first decades of the twentieth century, with more than
50 active mines of lead, zinc, and barium, the IglesienteSulcis (SW Sardinia) was one of the most important mining
districts in Europe (Fig. 1). The metallic ores were hosted
mainly in a Lower Cambrian calcareous formation (Ceroide Limestone), which is largely replaced by epigenetic
hydrothermal dolomite, considered to be of late- to postVariscan age because of its crosscutting relationships to
both sedimentary and tectonic structures (Boni et al. 2000).
This dolomite forms large-scale bodies, which can be
clearly identified on outcrop due to their yellowbrown
color, caused by supergene oxidation of Fe2? contained in
the dolomite lattice (Fig. 2a, b). The carbonate-hosted ZnPb sulfide ores have also been altered in the oxidation zone,
resulting in the so-called Calamine or nonsulfide ores (Boni
et al. 2003).
In addition to precipitating typical ore carbonates
(smithsonite and hydrozincite) or silicates (hemimorphite),
the supergene alteration has also promoted a widespread
replacement of previously deposited dolomites by new
zincian dolomite phases (Boni et al. 2011). The formation
of a broad halo of Zn dolomite, spottily replacing
the previous hydrothermal dolomite along fractures and
discontinuities, is one of the main effects of supergene

123

62

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

Fluminimaggiore
Nonsulfide Ore Deposits
& Zn-Dolomite District

Buggerru

Sar

din
ia

4
1 3
Gonnesa

Iglesias

Mines & Outcrops

1 Monteponi

IGLESIENTE

2 San Giovanni
3 Campo Pisano
4 Nebida
5 BuggerruMalfidano

Carbonia

SULCIS

2
3
4
5
6
7
8
9

10 Km

10

Fig. 1 Geological sketch map of southwestern Sardinia with the


location of the main Zn nonsulfide orebodies (15), surrounded by
Zn dolomite areas of variable extension. 1 Overthrust; 2 normal
fault; 3 Cenozoic; 4 Mesozoic; 5 Variscan granites; 6 Paleozoic

(allochthonous); 7 Ordovician to Devonian succession; 8 Iglesias Group


(Middle Cambrian-Lower Ordovician); 9 Gonnesa Group (Lower
Cambrian); 10 Nebida Group (Lower Cambrian) (modified from Boni
et al. 2003)

alteration in the areas surrounding the Calamine orebodies.


This replacement process is relatively common in several
mining districts subjected to supergene weathering, even
though the extent of substitution of Zn for Mg (and Pb
for Ca) in the dolomite structure is not easy to quantify.
The supergene zincitization of the dolomite has been
described at the Jabali (Yemen) and Yanque (Peru) minesites by Boni et al. (2011) and in the Polish ore district by
Zabinski (1959, 1980). This phenomenon is not restricted
to the mentioned localities only, but the precipitation of Zn
dolomite may be characteristic around most dolomitehosted sulfide concentrations undergoing supergene alteration. However, the genetic relationships between host
rock, primary sulfide mineralization, and the newly formed
nonsulfide Zn phases (including Zn dolomite) need better

clarification. Therefore, the aim of this study has been the


characterization of the supergene Zn dolomite in southwest
Sardinia, and its relationships with both the primary
hydrothermal dolomite and the Calamine-type Zn nonsulfide ores.

123

Geological setting
The geology of SW Sardinia is largely dominated by
Paleozoic (mainly Cambro-Ordovician) rocks of sedimentary as well as igneous origin, belonging to the so-called
external zones of the Variscan orogen (Carmignani et al.
1994) (Fig. 1). The Lower Cambrian succession is subdivided into the basal Nebida Group and the overlying

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

63

Fig. 2 Yellow (oxidized) hydrothermal dolomite in the Iglesiente


district. a Scoglio Il Morto, Nebida: Yellow dolomite replacing
Cambrian limestone in a carbonate block along the coast; b Hills east
of the Nebida village: unreplaced limestone areas in Dolomia Gialla;
c Cungiaus open pit of the Monteponi mine (Iglesias): the limestone
has been completely replaced by Yellow, locally Zn-rich dolomite;

d Malfidano open pit of the Buggerru mine (Fluminimaggiore): the


limestone is dolomitized and patchily enriched in Zn dolomite;
e Canale San Giuseppe (Nebida): Yellow Zn-rich dolomite hosting
nonsulfide Zinc ores; f San Giovanni Mount (Gonnesa): apophyses of
oxidized hydrothermal dolomite, patchily replacing the Cambrian
limestone

Gonnesa Group, which consists of siliciclastic sedimentary


rocks with carbonate intercalations toward the top and of
tidal dolomites and limestones, respectively (Bechstadt and
Boni 1994). Middle and Upper Cambrian to Lower Ordovician strata are represented by nodular limestones (Campo
Pisano Formation, Iglesias Group) and slates (Cabitza
Formation, Iglesias Group), respectively. Upper Ordovician and Silurian lithologies are separated by an angular

unconformity from the underlying series due to partial


erosion of Cambrian and Lower Ordovician sediments.
The pre-Variscan, stratiform, and/or stratabound Zn
PbBa orebodies are hosted in the Lower Cambrian carbonates (Boni 1985). Two groups of genetically distinct ore
types are known: SEDEX and MVT-type ores (Boni et al.
1996). Part of the orebodies are enclosed within an
epigenetic hydrothermal dolomite (Dolomia Geodica).

123

64

Contrary to most MVT deposits, this dolomitization phase


clearly postdates both the emplacement of the stratabound
ores and Variscan deformation (Boni et al. 1992). The
epigenetic replacive dolomitization affected the Cambrian
limestones as well as the early diagenetic dolomites in
large areas (more than 500 km2 in outcrop) of the Iglesiente-Sulcis district. The dolomitization process is especially pervasive in southern Iglesiente, where only limited
parts of the Ceroide limestone are unaffected and occur as
gray, isolated spots within a sea of weathered, brownyellowish dolomite (Fig. 2a, b, f). The large-scale relationship between dolomite and limestone clearly suggests
a post-deformational origin of the dolomite, since the
dolomite bodies clearly crosscut the vertical foliation and
are apparently controlled by the former as well as later
extensional faults. The epigenetic dolomite is informally
known as Dolomia Geodica (=Geodic Dolomite, due to its
vuggy appearance) and/or Dolomia Gialla (=Yellow
Dolomite; Brusca and Dessau 1968). Boni et al. (2000)
demonstrated that Dolomia Gialla and Dolomia Geodica
are different names for basically the same bodies of Febearing dolomite. The Dolomia Gialla appears yellow
brown on outcrop and for several hundred meters underground, due to oxidation of Fe2? contained in the dolomite lattice.
The relative age of the Dolomia Geodica event can be
tentatively inferred from the crosscutting relationship with
the host rocks and tectonic lineaments. This age can be
bracketed between the Late Carboniferous and Middle
Permian, as it has been reported in other European late
Variscan domains as well (Gasparrini et al. 2006).
The nonsulfide ores of the Iglesiente district, derived from
repeated weathering episodes, are generally hosted within the
Dolomia Gialla (hydrothermal Dolomia Geodica weathered
to brownish, rusty colors; Fig. 2ce). Smithsonite, hydrozincite, and hemimorphite are the principal Zn-bearing minerals in the nonsulfide zinc (?lead) deposits (Boni et al. 2003).
Cerussite and anglesite also occur, generally associated with
nodules of remnant or supergene galena, iron and manganese
oxy-hydroxides, and clay minerals.
The extent of the oxidized ore zones in the mining
district, which reach deep below the surface, is generally
independent of the present-day water table and highly
variable in different areas of the mining district. These
differences may be related to several distinct phases of
block faulting that displaced mature oxidation profiles
(Boni et al. 2003). The vertical tectonic movements
occurred during both the Tertiary and Quaternary periods.
The base of the oxidation profiles containing nonsulfide Zn
minerals can be both elevated above or submerged below
the recent water table, and the supergene alteration of the
primary ores is considered to be related to fossil, locally
reactivated, oxidation processes (Boni et al. 2003).

123

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

However, the Zn nonsulfide ore shoots in most mines are


roughly located within the lower vadose zone of a paleokarstic system that is hundreds of meters deep, but above
the water-filled conduits of the phreatic saturated zone. The
mineralization is considered to be the result of in situ
oxidation of the primary sulfide ores by increasingly acidic
meteoric fluids that circulated through the carbonates
(Moore 1972; Boni et al. 2003).
The formation of a broad halo of zincian dolomite,
spottily replacing the previous hydrothermal dolomite
along fractures and discontinuities, is another effect of the
supergene alteration in the areas surrounding the supergene
orebodies (Boni et al. 2011).
Based on geological and paleomagnetic arguments, it
was hypothesized that the Middle Eocene to Plio-Pleistocene represents the most probable age interval for the
formation of the supergene nonsulfide Zn-Pb ores, as well
as for the weathering and reddening of the hydrothermal
Dolomia Geodica to Dolomia Gialla (Boni et al. 2003,
2005).

Analytical methods
To investigate the mineralogy of the zincian dolomites, we
studied 25 samples from the Iglesiente district (Table 1)
using thin sections, scanning electron microscopy (SEM),
and qualitative energy-dispersive X-ray spectroscopy
(EDS). SEM examination was carried out using a Jeol JSM
5310 instrument at the University of Napoli (CISAG).
Element mapping and EDS spectra were obtained by the
INCA microanalysis system (Oxford Instruments). X-ray
diffraction analyses were performed on all samples using a
Philips PW 3020 automated diffractometer (XRD) at the
University of Heidelberg (CuKa radiation, 40 kV and
30 mA, 10 s/step, and a step scan of 0.022h; data were
collected from 3 to 1102h.
Stable carbon and oxygen isotopes were measured on
five samples of non-oxidized and three samples of oxidized
hydrothermal Dolomia Geodica, two samples of smithsonite from the Monteponi and Buggerru mines, and seven
samples containing larger amounts of zincian dolomite
located around the mines of Monteponi (Fig. 2c), Buggerru
(Fig. 2d), Nebida (Fig. 2e), San Giovanni (Fig. 2f), and
Planu Sartu. We were not able to separate the supergene
zincian dolomite from the hydrothermal dolomite, the two
phases being strictly intergrown, but we took care of
choosing those samples in which Zn dolomite was most
abundant, as well as some samples with only traces of Zn
dolomite detected by SEM analysis. All samples were
treated with EDTA solution to eliminate calcite.
Carbonate powders for stable isotope analyses were
collected with a dental drill and reacted with 103 %

Int J Earth Sci (Geol Rundsch) (2013) 102:6171


Table 1 Dolomite, Zn
dolomite, and smithsonite
samples from several localities
of southwestern Sardinia

Sample

65

Location

Mineral species

Bugr

Buggerru

Smithsonite

CP 2

Campo Pisano-Iglesias

Saddle dolomite

Cung

Monteponi-Cungiaus

Smithsonite

Cung 2

Monteponi-Cungiaus

Smithsonite, Zn dolomite, dolomite

DG 5B

San Giovanni-Gonnesa

Saddle dolomite

GT25-B

San Giovanni-Gonnesa

White Saddle dolomite

GT25-GR

San Giovanni-Gonnesa

Gray dolomite

GT26-B

San Giovanni-Gonnesa

White Saddle dolomite

GT26-GR

San Giovanni-Gonnesa

Gray dolomite

MP-TC

Monteponi-Iglesias

Saddle dolomite

M Poni 2

Monteponi-Iglesias

Dolomite [ Zn dolomite

Malf

Buggerru-Malfidano

Dolomite [ Zn dolomite

Malf 5

Buggerru-Malfidano

Dolomite [ Zn dolomite

NEB1

Nebida

Zn dolomite [ dolomite

NEB6
NEB7

Nebida
Nebida

Dolomite [ Zn dolomite
Zn dolomite [ dolomite

PS ? 55

Buggerru-Planu Sartu

Zn dolomite [ dolomite

PSV1

Buggerru-Planu Sartu

Zn dolomite [ dolomite

PSV2

Buggerru-Planu Sartu

Zn dolomite [ dolomite

PSV3

Buggerru-Planu Sartu

Zn dolomite [ dolomite

PSV4

Buggerru-Planu Sartu

Zn dolomite [ dolomite

SG-GON-DG1

San Giovanni-Gonnesa

Saddle dolomite, Fe-hydroxides

SG-GON-DG2

San Giovanni-Gonnesa

Saddle dolomite, Fe-hydroxides

S MAR

Santa Margherita-Nebida

Dolomite, Zn dolomite

S MAR 2

Santa Margherita-Nebida

Dolomite, Zn dolomite

phosphoric acid at 70 C using a Gasbench II connected to


a Thermo Finnigan Five Plus mass spectrometer (University of Erlangen-Nuremberg). All values are reported in
per mil relative to V-PDB by assigning a d13C value of
?1.95 % and a d18O value of -2.20 % to NBS19.
Reproducibility was checked by replicate analysis of laboratory standards and was better than 0.07 % (1r) for
both carbon and oxygen isotope analyses. Oxygen isotope
values of dolomite and smithsonite were corrected using
the phosphoric acid fractionation factors given by Kim
et al. (2007), Rosenbaum and Sheppard (1986), and Gilg
et al. (2008).
Differential thermal analysis (DTA) is a quick method to
provide additional information on carbonate minerals
assemblages (Zabinski 1959; Mondillo et al. 2011), since it
allows distinguishing between pure dolomite and Zn
dolomite on the basis of the temperature of the first
endothermic reaction. This reaction concerns MgCO3 decarbonatization occurring in the dolomite crystal structure,
and its temperature is subjected to variations induced by
substitution of metals, in particular Zn, in place of Mg in
the lattice (Zabinski 1959, 1980; Mondillo et al. 2011).
Thermal analysis was performed on few samples that were
analyzed for stable isotopes as well. The analyses were

conducted at the CISAG Laboratory of the University of


Napoli using a multiple thermoanalyzer Netzsch STA 409
(sample mass of 100 mg, air atmosphere, continuous
heating from room temperature to 1,100 C at 10 C
min-1). Elemental composition of the whole rock was
obtained, analyzing powder pellets at the CISAG Laboratory with a Philips PW1400 X-ray fluorescence spectrometer, following the methods described by Melluso et al.
(2005). LOI (weight loss on ignition) was measured
gravimetrically igniting the samples at 1,100 C.

Results
X-ray analyses of most samples show the occurrence of
dolomite with its usual peak at about 312h. No shifting or
doubling of this peak could be observed in the samples
containing important amounts of Zn dolomite (detected by
SEM). Goethite and haematite occur as well. Small
amounts of quartz and barite have been detected in few
dolomite samples from the San Giovanni mining area.
Variable amounts of smithsonite were identified in the
Cungiaus (Monteponi mine) and Buggerru (Malfidano
mine) samples.

123

66

Most specimens were studied under SEM and analyzed


by EDS (Fig. 3). The nonweathered, well-formed crystals
of white, saddle hydrothermal dolomites (specimens
GT25-B, GT26-B, MP-TC), sampled from underground
mines at several hundred meters depth, are mineralogically
not uniform as initially reported (Boni et al. 2000). In fact,
the dolomite crystals consist of two optically and geochemically distinct phases (Table 2a, b). The most abundant phase is an almost stoichiometric CaMg dolomite
with small amounts of Fe and Mn (phase 1; Fig. 3a). Phase
1 dolomite is locally cut and partly replaced along the
crystal growth zones by a Zn-Fe-rich dolomite phase
(phase 2, up to 17 wt% ZnO, up to 12 wt% FeO), which is
particularly abundant at the crystal boundaries, as shown in
Fig. 3b. Both dolomite phases are ascribed to successive
flows of hydrothermal fluids. Local oxidation and dedolomitization of both dolomite types resulted in the precipitation of calcite and FeMn(hydr)oxides (Fig. 3c, d).
Fe(hydr)oxides are never pure goethite, but may contain
ZnO (up to 12 wt%), PbO (up to 7 wt%), SiO2 (up to 6
wt%) together with Fe. Mn(hydr)oxides also contain Pb
and Fe in variable proportions (*30 wt% MnO, *35 wt%
PbO, *6 wt% FeO).
The supergene zincian dolomite (phase 3) is different
from both phase 1 Dolomia Geodica and phase 2 Zn dolomite
precipitated at the crystal boundaries. Phase 3 dolomite is
generally much richer in zinc, is patchily distributed
throughout the crystals (Fig. 3d, e), and commonly associated with Fe(hydr)oxides. ZnO contents vary between 3 and
21 wt%, PbO can reach 4 wt%, and around 1 wt% CdO has
been locally measured. Low FeO and MnO values from the
original hydrothermal dolomites have been detected as well
(Table 2c; Boni et al. 2011). Pure minrecordite (dolomite
where Mg is almost totally replaced by 29 wt% Zn, Garavelli
et al. 1982) has never been recorded in Sardinia. The precipitation of Zn dolomite is followed, eventually, by sparry
calcite. Some of the smithsonite generations, occurring in the
nonsulfide orebodies, may contain variable amounts of Mg
(up to 6 wt% MgO) (Fig. 3f). We argue that part of the
magnesium derived from dolomite replacement was incorporated in the smithsonite lattice.
Oxygen and carbon isotope data of both unweathered phase
1 and 2 hydrothermal dolomites and of supergene Zn dolomite
(phase 3) are presented in Table 3, together with two newly
measured smithsonite samples. The comparison with carbon
and oxygen isotope data of early diagenetic tidal dolomites
outcropping in the whole Iglesiente district (Santa Barbara
Formation of the Gonnesa Group), hydrothermal Dolomia
Geodica (Boni et al. 2000), and smithsonite-hydrozincite
(Boni et al. 2003; Gilg et al. 2008) is shown in Fig. 4. Samples
consisting predominantly of supergene phase 3 Zn dolomite
have d18O and d13C values ranging between -5 and -8 %
VPDB and ?1 and -6.5 % VPDB, respectively.

123

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

Differential thermal analyses of three Sardinia dolomite


samples are shown in Fig. 5, together with the DTA patterns of a typical Triassic dolomite from Southern Italy and
a Zn dolomite from Tsumeb (Hurlbut 1957). Sample MPTC consists in most part of phase 1 and locally of phase 2
hydrothermal dolomite (determined by SEMEDS). Sample MALF contains remnants of phase 1 dolomite and is
locally dedolomitized resulting in a mixture of calcite and
phase 3 Zn dolomite (SEMEDS). Sample NEB1 is composed mainly of phase 3 Zn dolomite, with only few
remnants of phase 1 hydrothermal dolomite (SEMEDS).
The DTA patterns show two important endothermic reaction peaks in all measured samples. However, the first
endothermic reaction occurs at about 780 C in samples
MP-TC and MALF, whereas in sample NEB1 it occurs at
about 710720 C. In all the three samples, the second
endothermic reaction occurs at about 890 C. Heating of
sample NEB1 results as well in two separate endothermic
reactions, both observed below 800 C.
Whole rock chemical analyses of samples MP-TC,
MALF, and NEB1 (Table 4) show Zn contents of 0.98 wt%
ZnO, 3.50 wt% ZnO, and 11.25 wt% ZnO, respectively.
FeO contents are 5.40 wt%, 5.52 wt%, and 9.68 wt%,
respectively. The high Fe content detected in NEB1 is
related to the higher abundance of Fe-hydroxides in this
sample than in the other two. The geochemical composition of the dolomite samples (from SEMEDS analyses) is
illustrated in the triangular diagram of Fig. 6. It becomes
obvious that sample NEB1 contains a Zn dolomite, which
is, on average, richer in Zn than in samples MALF and MPTC. The comparison of samples MALF and NEB1 shows
that both contain phase 3 Zn dolomite, but that the dolomite is less enriched in Zn in sample MALF than in sample
NEB1. Sample MP-TC is characterized by phase 2 Zn
dolomite variably enriched in Zn but, as mentioned above,
most part of the sample MP-TC is represented by stoichiometric phase 1 dolomite.

Discussion
The supergene Zn dolomites (phase 3) in the nonsulfide
mining districts of southwestern Sardinia are widely distributed (from hundreds of cubic meters to several kilometers around the main deposits). They can reach at least
five hundreds meters in depth (palaeoweathering), always
in association with Zn-Pb nonsulfide ores. The time constraints for the formation of the nonsulfide ores and of the
supergene Zn dolomites (phase 3) are still unclear, owing
to multiple oxidation events through time.
The Zn dolomites replace the predecessor dolomite phases. From the geochemical point of view, they are characterized by variable amounts of Zn, and very low amounts of

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

67

Table 2 Selected chemical analyses of three dolomite phases from southwestern Sardinia mining district
(a)

(b)

(c)

CaO

31.59

31.82

30.33

28.55

27.70

27.71

29.80

29.19

29.07

27.77

MgO

20.83

20.38

20.78

13.63

12.86

10.97

16.14

16.38

15.01

12.17

MnO

N.D.

N.D.

0.42

1.39

0.09

N.D.

N.D.

0.24

0.19

0.34

FeO

N.D.

N.D.

3.35

12.18

N.D.

0.38

0.40

0.22

2.44

1.17

ZnO

N.D.

N.D.

N.D.

N.D.

14.95

17.33

4.42

6.95

8.18

12.79

CdO

N.D.

N.D.

N.D.

N.D.

N.D.

N.D.

0.30

0.04

N.D.

0.20

PbO

N.D.

N.D.

N.D.

N.D.

N.D.

N.D.

1.12

0.15

0.42

0.31

CO2*

47.48

47.92

45.05

44.27

44.37

43.59

47.78

46.80

44.70

45.34

Total

99.90

100.12

99.92

100.01

99.97

99.98

99.96

99.97

100.00

100.10

All the analyses were performed by energy-dispersive spectroscopy and all the chemical compositions are in oxide weight percentage.
N.D. = not determined
* Calculated from stoichiometry
(a) Stoichiometric Phase 1 dolomite (hydrothermal)
(b) Fe- to Zn-rich Phase 2 dolomites (hydrothermal)
(c) Phase 3 Zn dolomites related to supergene processes

Fig. 3 a Monteponi mine.


Stoichiometric hydrothermal
CaMg dolomite (hypogene,
phase 1), with a border of
(hypogene?) zincian dolomite
(phase 2). b San Giovanni mine.
Stoichiometric CaMg dolomite
(hypogene, phase 1), with a
border of hydrothermal ferroan
dolomite (phase 2). c Malfidano
mine. Stoichiometric CaMg
and low ferroan dolomite
(hypogene), patchily
dedolomitized. d Malfidano
mine. Stoichiometric CaMg
and low ferroan dolomite
(hypogene), patchily replaced
by calcite and Fe
Mn(hydr)oxides, and locally, by
supergene Zn dolomite (phase
3). e Nebida mine. Hypogene
dolomite (phase 1), diffusely
replaced by supergene Zn
dolomite (phase 3); the white
spots are Fe(hydr)oxides.
f Nebida mine. Mg-rich
smithsonite concretions
(supergene), associated with
phase 3 Zn dolomite

stoichiometric
dolomite

Low Fe-Mn
dolomite

Zn-rich
dolomite

30 m

20 m

Zn-Fe-rich
dolomite

remnant
dolomite

Low Fe-Mn
dolomite
calcite/
dolomite

calcite

Fe-hydroxides
30 m

40 m

supergene
Zn dolomite

30 m

supergene
Zn dolomite

remnant
dolomite

Mg smithsonite

smithsonite
15 m

123

68

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

Table 3 Carbon and oxygen isotope ratios of hydrothermal Dolomia


Geodica, Zn dolomite, and smithsonite from southwestern Sardinia
Sample

d13C (% V-PDB)

d18O (% V-PDB)

Bugr old1 (Boni et al. 2003)

-0.73

-8.80

Bugr old2 (Boni et al. 2003)

-3.73

-6.98

Bugr

-3.56

-2.90

Cung

-6.46

-2.37

GT25-B

-0.02

-9.91

GT25-GR

0.33

-9.86

GT26-B
GT26-GR

0.65
0.63

-10.60
-10.67

MP-TC

1.06

-9.16

M Poni 2

0.66

-8.40

Malf

1.71

-8.11

0.96

-8.52

NEB1 a

Malf 5

-5.93

-6.39

NEB1 b

-6.41

-7.30

NEB6 a

-1.96

-9.26

NEB6 b

-1.44

-10.03

NEB7

-2.04

-5.04

PS ? 55

-1.53

-6.36

PSV2

-5.34

-5.73

18O PDB
-14.0

-12.0

-10.0

-8.0

-6.0

-4.0

-2.0

Santa Barbara Fm.


Cambrian tidal dolomite

0.0
4.0

2.0

0.0

-2.0

-4.0

-6.0

13C PDB

Dolomia Geodica

-8.0

smithsonite

-10.0

-12.0
GT25-B, GT25-GR, GT26-B , GT26-GR, MP-TC: hydrothermal dolomite
M Poni 2, Malf, Malf 5, Bugr old1, NEB6 a, NEB6 b: hydrothermal dolomite > Zn dolomite
Bugr old2, NEB1 a, NEB1 b, NEB7, PS+55, PSV2: Zn dolomite > hydrothermal dolomite
Bugr, Cung: smithsonite

Fig. 4 Plot of d13C versus d18O for dolomites, smithsonites, and Zn


dolomites from southwestern Sardinia. Zincian dolomite is cogenetic
with smithsonite. The Tidal Dolomite Field comprises the values
published in Boni et al. (1988), the Dolomia Geodica Field comprises
the values published in Boni et al. (2000), and the smithsonite Field
comprises the values published in Boni et al. (2003). The Buggerru
dolomite values have been published in Boni et al. (2003). Symbols as
in Table 1

123

Pb and Cd in the crystal lattice. They are also generally


depleted in Fe and Mn relative to precursor phases 1 and 2 of
hydrothermal dolomite (Dolomia Geodica).
The carbon and oxygen isotope ratios of unweathered
(or slightly weathered) hydrothermal dolomite vary
between -1.5 and ?1.0 % VPDB and from -7.0 to
-10.0 % VPDB, respectively, and confirm earlier published
values (Boni et al. 2000). The oxygen isotopes values are
depleted in 18O with respect to d18O values of Cambrian
early diagenetic intertidal dolomites and limestone (Boni
et al. 1988; 2000), but the d13C values are never lower than
-2 % VPDB. The oxygen isotope ratios, together with the
rather uniform cathodoluminescence pattern, indicate a
water-dominated fluid-flow system (Boni et al. 2000). The
smithsonite d18O values are within the range of the published d18O (average value: -4 % VPDB) and d13C values
(-10.4 to -0.6 % VPDB; Boni et al. 2003), with the low
d13C values being characteristic for supergene smithsonites
(Gilg et al. 2008). The d13C and d18O values of supergene
phase 3 Zn dolomite plot between the d13C/d18O field of
hydrothermal dolomite and that of smithsonite-hydrozincite with the d13C values of zincian dolomite being comparable to those of smithsonite and hydrozincite. The large
variation in the carbon isotope values of zincian dolomite
(similar to those of smithsonite) combined with the
restricted range of oxygen isotope values suggests that only
meteoric waters were involved in the oxidation.
Differential thermal analysis confirms the results of
Zabinski (1980) and Mondillo et al. (2011). The dissociation of a stoichiometric dolomite is characterized by two
endothermic reactions representing the decomposition of
the MgCO3 component at around 800 C and the CaCO3
component at around 900 C (Fig. 5; Webb and Kruger
1970; Smykatz-Kloss 1974; Gunasekaran and Anbalagan
2007). However, the substitution of Mg2? by Zn2? in the
dolomite structure causes a decrease of the first endothermic reaction by about one hundred degrees in comparison
with pure dolomite (Hurlbut 1957). The DTA traces of
samples MALF and MP-TC (where Zn dolomite is scarce)
show the dissociation reaction of the MgCO3 at a temperature of about 780 C, slightly less than the dehydration
value of stoichiometric dolomite (Gunasekaran and Anbalagan 2007). In the NEB1 sample, where the Zn value is
higher (Table 4; Fig. 6), this reaction occurs at about
710720 C (Fig. 5). The difference of temperature of the
first endothermic reaction between the MALF and NEB1
samples should not only be due to the different Zn amount
registered by the chemical analysis (Table 4), but also to
the higher Zn grade of the NEB1 Zn dolomite, relative
to the MALF Zn dolomite (Fig. 6).
On the base of textural evidence, the zincian dolomite
phases from the supergene zone of sulfide zinc deposits in
the Iglesiente mining district are interpreted as a possible

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

69

missing link between dolomite and smithsonite. This


interpretation is supported by the carbon isotope ratios of
the samples rich in zincian dolomite, which are very similar to those of the Iglesiente smithsonites (Boni et al.
2003). The observation that samples poor in zincian
dolomite plot close to the Dolomia Geodica d18O field,
whereas samples rich in zincian dolomite are generally
enriched in 18O and plot closer to the smithsonite d18O field
suggests that the variation in zincian dolomite d18O can be
explained by different proportions of precursor dolomite
relative to newly formed zincian dolomite. It must be taken
into account that it was impossible to separate completely
the different dolomite phases before isotope analyses.
However, since the oxygen as well as carbon isotope ratios
of zincian dolomite-rich samples and smithsonite are
comparable, the formation of zincian dolomite is suggested
to have occurred under conditions that are comparable with
those of the formation of smithsonite. Formation from
meteoric waters is supported by low d13C values indicative
of the influence of soil-gas CO2 in near-surface environments (Gilg et al. 2008). Temperatures during smithsonite
precipitation are estimated to have been between 11 and
23 C, assuming an oxygen isotopic composition of precipitation fluid in southwestern Sardinia of -6.5 %
VSMOW (Gilg et al. 2008).

Fig. 5 Differential thermal analysis curves of selected dolomite


samples from the Iglesiente mining district. In the box at the top of the
figure are highlighted two DTA reference traces: * Triassic dolomite
from Southern Italy, ** Zn dolomite from Tsumeb (Hurlbut 1957).
MALF Malfidano mine, MP-TC Monteponi mine, NEB1 Nebida mine

Table 4 Whole rock chemical analysis of the MP-TC, MALF, and


NEB1 samples
MP-TC

MALF

NEB1

SiO2

0.55

0.56

0.68

TiO2
Al2O3

0.01
0.00

0.06
0.19

0.01
0.00

FeO

5.40

5.52

9.68

MnO

0.22

0.44

1.16

MgO

13.85

9.72

7.53

CaO

32.34

34.63

27.37

Na2O

0.10

0.10

0.10

K2O

0.01

0.12

0.06

P2O5

0.01

0.21

0.01

ZnO

0.98

3.50

11.25

LOI

45.52

44.93

42.16

Total

100.00

100.00

100.00

Compositions are in oxide weight percentage. LOI loss of ignition

Conclusions
The occurrence of zincian dolomites in the oxidation zone
of base metal sulfide deposits in SW Sardinia confirms the
supergene origin of these carbonates. There is strong evidence that the oxidation profiles and related nonsulfide
mineral deposits evolved throughout late Tertiary and were
later displaced and rejuvenated by younger block tectonics.
The precipitation temperature of the Zn dolomite is interpreted to correspond to the temperature of the meteoric
fluids during the main weathering periods, when the sulfide
deposits were oxidized. We interpret the replacement of the
dolomite host as a multistep process, starting with a progressive zincitization of the dolomite crystals, followed by a
patchy dedolomitization (resulting in the formation of calcite
and Fe-Mn-hydroxides), potentially concluded by the complete replacement of dolomite by smithsonite (Fig. 7).
This progressive zincitization phenomenon has been
described also in other dolomite-hosted zinc deposits, as
Jabali (Yemen) and Yanque (Peru) (Boni et al. 2011;
Mondillo et al. 2011). As it is the case in the above-mentioned mining districts, the extent of the replacement
bodies of Zn dolomite may be highly significant for the
exploration of nonsulfide Zn ores (Boni et al. 2011). In fact,
the amount of the total Zn contained in Zn dolomite

123

70

Int J Earth Sci (Geol Rundsch) (2013) 102:6171

CaZn(CO3)2
Malfidano - MALF
M. Poni - MP-TC

Ca(Mg,Zn)(CO3)2

Nebida - NEB1 Boni et al. 2011

CaZn(CO3)2

CaMg(CO3)2

CaMg(CO3)2

Ca(Fe,Mn)(CO3)2

Ca(Fe,Mn)(CO3)2

Fig. 6 Composition of dolomites of the MP-TC, MALF, and NEB1 samples, in the system CaMg(CO3)2Ca(Fe,Mn)(CO3)2CaZn(CO3)2
(SEMEDS analyses)

Pre-Variscan

Permian(?)

Tertiary

Recent

References

Dolomite ph1
Dolomite ph2
Zn dolomite ph3
Calcite
Hydroxides
Smithsonite

Variscan compressive tectonics

Primary Sulfides

Mg smithsonite

Fig. 7 Paragenesis of the main mineralogical phases observed in the


SW Sardinia nonsulfide Zn district

(concealed Zn), together with Zn measured from the


processable ore minerals (i.e., smithsonite and hydrozincite), might lead to a strong overestimation of the metallic
resources calculated from the assay data only (Mondillo
et al. 2011). This fact has to be kept in mind when
exploring for Zn nonsulfide ores in dolomite host rocks.
Acknowledgments This study has been carried out partly with a
PhD bursary of the University of Napoli to Nicola Mondillo. The
authors would like to thank R. de Gennaro (CISAG Napoli) for his
support during SEMEDS analysis, L. Franciosi and L. Melluso for
their help with XRF analyses. A special thank is reserved to an
anonymous reviewer and to the Editor of the Journal, whose criticism
greatly improved the quality of the paper.

123

Bechstadt T, Boni M (eds) (1994) Sedimentological, stratigraphical and


ore deposits field guide of the autochthonous Cambro-Ordovician of
Southwestern Sardinia. Servizio Geologico dItalia Memorie
Descrittive Carta Geologica dItalia, v. XLVIII, 434 pp
Boni M (1985) Les gisements de type Mississippi Valley du Sud
Ouest de la Sardaigne (Italie): une synthe`se. Chronique recherches minie`res BRGM 489:734
Boni M, Iannace A, Pierre C (1988) Stable isotopes in the lower
Cambrian ore deposits and their host rocks in SW Sardinia.
Chem Geol (Isotope Geoscience) 72:267282
Boni M, Iannace A, Koppel V, Hansmann W, Fruh-Green G (1992)
Late- to post Hercynian hydrothermal activity and mineralization
in SW Sardinia. Econ Geol 87:21132137
Boni M, Iannace A, Balassone G (1996) Base metal ores in the lower
Palaeozoic of South-Western Sardinia. Econ Geol 75th Anniv
Vol, Spec Publ 4:1828
Boni M, Parente G, Bechstadt T, De Vivo B, Iannace A (2000)
Hydrothermal dolomites in SW Sardinia (Italy): evidence for
a widespread late-Variscan fluid flow event. Sed Geol 131:181200
Boni M, Gilg HA, Aversa G, Balassone G (2003) The calamine of
SW Sardinia: geology, mineralogy and stable isotope geochemistry of a supergene Zn mineralization. Econ Geol 98:731748
Boni M, Dinares-Turell J, Sagnotti L (2005) A first attempt to
paleomagnetic dating of non-sulfide Zn ores in SW Sardinia
(Italy). Ann Geophys 48(2):112
Boni M, Mondillo N, Balassone G (2011) Zincian dolomite: a
peculiar dedolomitization case? Geology 39:183186
Brusca P, Dessau G (1968) I giacimenti piombo-zinciferi di S.Giovanni
(Iglesias) nel quadro della geologia del Cambrico sardo. LIndustria
Mineraria 19(9):470494, (10):533556, (11):597609
Carmignani L, Carosi R, Di Pisa A, Gattiglio M, Musumeci G,
Oggiano G, Pertusati PC (1994) The Hercynian chain in
Sardinia. Geodin Acta 7:3147

Int J Earth Sci (Geol Rundsch) (2013) 102:6171


Garavelli CG, Vurro F, Fioravanti GC (1982) Minrecordite, a new
mineral from Tsumeb. Mineralog Rec 13:131136
Gasparrini M, Bechstadt T, Boni M (2006) Massive hydrothermal
dolomites in the southwestern Cantabrian Zone (Spain) and their
relation to the Late Variscan evolution. Mar Petrol Geol 23:543568
Gilg HA, Boni M, Hochleitner R, Struck U (2008) Stable isotope
geochemistry of carbonate minerals in supergene oxidation
zones of Zn-Pb deposits. Ore Geol Rev 33:117133
Gunasekaran S, Anbalagan G (2007) Thermal decomposition of
natural dolomite. Bull Mater Sci 30:339344
Hurlbut CS Jr (1957) Zincian and plumbian dolomite from Tsumeb,
South-West Africa. Am Mineral 42:798803
Kim ST, Mucci A, Taylor BE (2007) Phosphoric acid fractionation
factors for calcite and aragonite between 25 and 75 C: revisited.
Chem Geol 246:135146
Melluso L, Morra V, Brotzu P, Tommasini S, Renna MR, Duncan RA,
Franciosi L, DAmelio F (2005) Geochronology and petrogenesis
of the Cretaceous Antampombato-Ambatovy complex and associated dyke swarm, Madagascar. J Petrol 46:19631996

71
Mondillo N, Boni M, Balassone G, Grist B (2011) In search of the lost
zinc: a lesson from the Jabali (Yemen) nonsulfide zinc deposit.
J Geoch Expl 108:209219
Moore JMcM (1972) Supergene mineral deposits and physiographic
development in southwest Sardinia, Italy. Trans Inst Min and
Metallurgy (Sect B: Appl Earth Sci) B81:5966
Rosenbaum J, Sheppard SM (1986) An isotopic study of siderites,
dolomites and ankerites at high temperatures. Geochim Cosmochim Acta 50:11471150
Smykatz-Kloss W (1974) Differential thermal analysis: application
and results in mineralogy. Springer, NY
Webb TL, Kruger JE (1970) Carbonates. In: Mackenzie RC (ed)
Differential thermal analysis. Academic Press, London, pp 303341
Zabinski W (1959) Zincian dolomite from the Warynski mine, Upper
Silesia. Bulletin de lAcademie Polonaise des Sciences, S Sci
Chim, Geol et Geogr 7:355358
Zabinski W (1980) Zincian dolomite: the present state of knowledge.
Mineralog Pol 11:1931

123

Вам также может понравиться