Next Article in Journal
Synthesis, Property and Mechanism Analysis of a Novel Polyhydroxy Organic Amine Shale Hydration Inhibitor
Next Article in Special Issue
Hydrothermal Chromitites from the Oman Ophiolite: The Role of Water in Chromitite Genesis
Previous Article in Journal
Adsorption of Salmonella in Clay Minerals and Clay-Based Materials
Previous Article in Special Issue
Multivariate Statistical Analysis of Trace Elements in Pyrite: Prediction, Bias and Artefacts in Defining Mineral Signatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Integrated Study of the Serpentinite-Hosted Hydrothermal System in the Pollino Massif (Southern Apennines, Italy)

1
Department of Sciences, University of Basilicata, 85100 Potenza, Italy
2
Departamento de Petrología y Geoquímica, Universidad Complutense, 28040 Madrid, Spain
3
Istituto Nazionale di Geofisica e Vulcanologia, Sezione di Palermo, 90146 Palermo, Italy
4
Istituto di Metodologie per l’Analisi Ambientale, Consiglio Nazionale delle Ricerche, 85050 Tito Scalo (PZ), Italy
*
Author to whom correspondence should be addressed.
Minerals 2020, 10(2), 127; https://doi.org/10.3390/min10020127
Submission received: 18 November 2019 / Revised: 18 January 2020 / Accepted: 29 January 2020 / Published: 31 January 2020
(This article belongs to the Special Issue Feature Papers in Mineral Geochemistry and Geochronology 2019)

Abstract

:
A comprehensive study of the serpentinite and associated veins belonging to the Frido Unit in the Pollino Massif (southern Italy) is presented here with the aim to provide new constraints about the hydrothermal system hosted by the accretionary wedge of the southern Apennines. The studied serpentinites are from two different sites: Fosso Arcangelo and Pietrapica. In both sites, the rocks show mylonitic-cataclastic structures and pseudomorphic and patch textures and are traversing by pervasive carbonate and quartz-carbonate veins. The mineralogical assemblage of serpentinites consists of serpentine group minerals (with a predominance of lizardite), amphiboles, pyroxene, chlorite, titanite, magnetite, and talc. In some samples, hydro-garnet was also detected and documented here for the first time. As for cutting veins, different mineralogical compositions were observed in the two sites: calcite characterizes the veins from Fosso Arcangelo, whereas quartz and dolomite are the principal minerals of the Pietrapica veins infill, suggesting a different composition of mineralizing fluids. Stable isotopes of C and O also indicate such a different chemistry. In detail, samples from the Pietrapica site are characterized by δ13C fluctuations coupled with a δ18O shift documenting calcite formation in an open-system where mixing between deep and shallow fluids occurred. Conversely, δ13C and δ18O of the Fosso Arcangelo veins show a decarbonation trend, suggesting their developing in a closed-system at deeper crustal conditions. Precipitation temperature calculated for both sites indicates a similar range (80 °C to 120 °C), thus suggesting carbonate precipitation within the same thermal system.

1. Introduction

Mantle peridotites are exposed on the seafloor at slow and ultraslow spreading mid-ocean ridges [1]. Seafloor spreading developed in areas characterized by tectonic extension and detachment faults, allowing uplift and exposure of mantle peridotite in oceanic core complexes [2,3,4,5,6,7]. Due to the interaction with seawater and deep hydrothermal fluids, seafloor peridotites are readily serpentinized over a wide range of conditions [8]. Serpentinites occurring in ophiolitic complexes are important for volatile cycling and fluid mobile elements acting as a source for water, carbon, sulfur, chlorine, boron, arsenic, and nitrogen [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26]. These rocks, in fact, in the last years have been widely studied to gain a better understanding of their reactivity with respect to hydrothermal fluids promoting carbonation processes and mineral immobilization of large amounts of CO2 [27,28].
Carbonates, usually as serpentinite matrix and/or vein infill within an extensive network, are a common feature of altered oceanic lithosphere, although ocean-floor metasomatism is not the only process responsible for their presence in these rocks. Carbonates are common minerals that form in a number of geological settings and upon different environmental conditions. As the other chemically-derived precipitates, carbonate composition mirrors the physico-chemical properties and composition of the mineralizing fluid, a feature that helps to constrain the source fluid and environmental conditions that promote carbonate precipitation also under a complex tectonic scenario [29,30,31,32,33,34,35,36].
In southern Apennines, serpentinites occur as slices within the geological and structural frame of the Pollino Massif ophiolites (southern Ligurian Tethys) in the Ligurian Accretionary Complex (LAC) [37]. These rocks are traversed by a complex network of veins with different textural and macroscopic features. Their chemical and mineralogical compositions have never been studied before, although could better constrain the geological evolution of the serpentinites in the LAC of the Frido Unit.
With this in mind, we present here, for the first time, a comprehensive study of the serpentinites and associated carbonate and quartz-carbonate veins from the Pollino Massif, based on field observations, petrographic and mineralogical data, and carbon (δ13CV-PDB) and oxygen (δ18OV-SMOW) stable isotopes, to provide new constraints about the hydrothermal system hosted by the accretionary wedge of southern Apennines. Our principal goals are to define (1) source and composition of mineralizing fluids and (2) processes leading to mobilization, fractionation, and redistribution of chemical elements during the emplacement of the Frido Unit serpentinites within the accretionary wedge.

2. Geological Background

The Pollino Massif is located in the southern Apennines at the Calabria-Lucania border zone (Figure 1). It consists of tectonically juxtaposed thrust belts derived from the deformation of the African passive margin [38]. The deformation occurred between Oligocene and Pleistocene and involved the Ligurian ocean ophiolitic crust and its sedimentary cover [38,39]. In this area, the Liguride Complex [38], also defined as Liguride Units [40], is well exposed and is located in the highest position in the tectonics edifice of the southern Apennines. The Liguride Complex is derived from the NW subduction of the Tethyan ocean–continent transition zone and was divided into different tectonic units, where fragments of Jurassic oceanic crust [38], associated with slices of continental crust rocks, are preserved [38,39,41,42]. These terrains consist of a Mesozoic to Cenozoic flysch and a series of ophiolitic nappes, widely exposed along the whole Apennine Chain and in Calabria [38].
The Liguride Complex has been subdivided into two units, the metamorphosed Frido Unit and the non-metamorphic Calabro-Lucano Flysch [39] or North Calabria Unit [40]. The Frido Unit forms the uppermost thrust sheet, and tectonically overlies the North Calabrian Units, which in turn is split in a number of thrust sheets [39].
The Frido Unit consists of a metasedimentary sequence (phyllite, meta-arenite, quartzite, and isolated bodies of meta-limestone or calcschist and metapelite) [37,43,44,45,46] with blocks of both oceanic and continental rocks [38,41,42]. Oceanic lithosphere in the Frido Unit [47] is represented by tectonized serpentinite [48,49,50,51], metabasalt [52], metagabbro, metapillow lavas [37], and dismembered metadoleritic dykes [53,54,55,56]. Continental crust rocks mainly consist of weathered granofels, garnet gneiss, garnet–biotite gneiss, leucocratic biotite gneiss and lenticular bodies of amphibolite [41,42] often cross-cut by basic dykes [41].
According to several studies [38,39,41,53,54,57,58], the Frido Unit underwent a polyphase blueschist to greenschist facies metamorphism developed in the deeper parts of the Liguride accretionary wedge. The blueschist metamorphism in mafic rocks developed at peak pressure conditions of 0.6–0.8 GPa and temperatures of 350 °C, whereas the subsequent greenschist facies overprint took place at P = 0.4 GPa and T = 300–350 °C [37,59]. Cavalcante et al. [60] interpreted that also the nearby metasediments were affected by HP–LT conditions using illite crystallinity data and the b0 parameter of K-white mica in phyllite. Similar metamorphic conditions have also been documented by Invernizzi et al. [61] and Laurita and Rizzo [62] for the metabasites of the Frido Unit (200–300 °C and 0.6–0.8 GPa; 300–400 °C and 0.8–1.2 GPa, respectively).
During blueschist facies conditions, crystallization of glaucophane occurred in metabasites [63], and of magnesio-riebeckite in schist associated with metabasites [64] and carpholite in some veins in metapelites and phyllites [37,65]. Afterward, in a brittle deformation regime, the most obvious structures are characterized by cataclastic bands well developed in the serpentinites [63].

3. Sampling and Analytical Method

A total of 26 representative samples of serpentinites dominated by carbonate-veins (SpFA) and serpentinites dominated by quartz-carbonate veins (SpPP) were collected at the Fosso Arcangelo site, near the San Severino Lucano village, and in the Pietrapica site, at the Calabria-Lucanian boundary (Figure 1) (Table 1).
The serpentinites outcropping at the Fosso Arcangelo site (SpFA) are brownish-grey-green and appear intensely reworked by strong brittle and ductile deformation, evidenced by several slip surfaces and by the presence of isoclinal and tight folds and some intrafolial folds often associated with crenulation cleavage. The fractures are commonly filled by carbonate minerals and display changes in both thickness and length (Figure 2a).
Serpentinites in this site crosscut by metadolerite dykes that were affected by metamorphism under relatively HP/LT (blueschist facies) conditions during the formation of the Apennine accretionary prism [45,53,54,55]. The Pietrapica site is characterized by a complex exposure of dark-green cataclastic serpentinites (SpPP) with several slip surfaces and badland-like morphology. The serpentinite shows quartz-carbonate veins with talc-rich domains (Figure 2b). At the surrounding of the Pietrapica area, no rocks with any experience of HP/LT metamorphism occur.
Petrographic characterization of all samples was carried out by optical microscopy on thin sections of rock samples oriented following their foliations and lineations.
Mineralogy analyses were performed on randomly oriented powdered samples of both host-rocks and vein infill by using X-ray powder diffraction (XRPD) and μ-Raman techniques at the Department of Sciences, University of Basilicata (Potenza, Italy). XRPD analyses were accomplished by means of Siemens D5000 equipment with CuKα radiation, 40 kV and 32 mA, 2 s per step, and a step scan of 0.02° 2θ. Data were recorded between 5 and 70° 2θ for the bulk rock samples and from 15 to 70° 2θ for the vein infill. The mineral phase identification was carried out by means of the X’Pert HighScore Plus software (PANalytical 2001, Version 01 using the PDF-2 (2005) database. The μ-Raman analyses were carried out using a Horiba Jobin-Yvon LabRam HR800 spectrometer equipped with a HeNe laser source with a wavelength of 633 nm, a CCD detector operating at –70 °C, and an edge filter that excludes from detection shift below 150 cm−1. A spectral resolution of 4 cm−1 was obtained by a holographic grating with 600 lines/mm. Correct calibration of the instrument was verified checking the position of the Si band at ±520.7 cm−1. Output laser power was 20 mV, and measurements were performed using an optical microscope Olympus with objectives of 10×, 50×, and 100×. A laser beam spatial resolution of 1 µm was obtained with the 100× objective. Spectra results were from the average of 5 acquisitions of 10 s to optimize the signal/noise ratio. Two regions of the Raman spectra were investigated: 1200–150 cm−1 for structural bonding characterization and 3800–3500 cm−1 for the characterization of the hydroxyl groups. The minerals were identified based on the data reported in the online RUFF database.
Mineral chemistry was determined at the Centro Nacional de Microscopía Electrónica (CNME) of the Universidad Complutense (Madrid, Spain) by electron microprobe (EMP) analyses on the serpentinites and associated veins, using a JEOL Superprobe JXA-8900M equipped with four wavelength dispersive spectrometers. Silicate and oxide analyses were conducted at an accelerating voltage of 15 kV, an electron beam current of 20 nA, and a beam diameter of 5 µm. An accelerating voltage of 20 kV, an electron beam current of 10 nA, and a beam diameter of 5 µm were used for carbonate minerals. Each element was counted for 15 s. The following minerals were used as standards: sillimanite for the Al, albite for the Si and Na, almandine for the Mn and Fe, kaersutite for Mg, Ti and Ca, microcline for the K, fluorapatite for the P, Ca, F, Cl, and Ni, Cr pure metals. Corrections were made using the ZAF (Z: atomic number; A: absorption; F: fluorescence) method. The estimation uncertainties for major and minor elements were determined for each analysis, which have uncertainties from ±0.8% to ±5%. The structural formula of amphiboles was recalculated on the basis of 23 oxygens and classified by using the amphiboles nomenclature suggested by Leake et al. [66,67]. The andradite structural formula was recalculated on the basis of 24 oxygens. The structural formula of pyroxene was recalculated on the basis of 6 oxygens.
A total of 21 vein infills were selected for stable oxygen and carbon isotope analysis including 15 samples from the SpFA and 6 samples from the SpPP. In detail, about 0.1 mg of powder samples were put in a 12 mL screw cap Exetainer(R) vial and then flushed with pure helium to remove the air in the headspace. Subsequently, about 50 μL of 100% H3PO4 was added to each sample for the conversion to carbon dioxide. The analyses were performed using a Thermo GB-II peripheral coupled with a Thermo Delta V Plus CF-IRMS at “Istituto Nazionale di Geofisica e Vulcanologia-Sezione di Palermo (Italy) laboratories” (Palermo, Italy). The results are reported in delta per mil (‰) values relative to the Vienna Pee Dee Belemnite (V-PDB) international scale. The δ18O values on the V-PDB scale are converted to the Vienna Standard Mean Ocean Water (V-SMOW) using the equation of Coplen et al. [68]:
δ18OV-SMOW = 1.03091* δ18OV-PDB + 30.91
and are expressed in delta per mil (‰) relative to the Vienna Standard Mean Ocean Water (V-SMOW) value. A Carrara Marble internal standard (MAB) and two international standards (NBS-18 and NBS-19) were used for calibration. Analytical reproducibility was better than 0.1‰ and 0.15‰ for δ13C and δ18O, respectively.
Some quartz-carbonate veins of the SpPP were selected for the study of fluid inclusions and for the microthermometric determinations performed at the Departamento de Estratigrafía, Universidad Complutense, (Madrid, Spain). Doubly polished 100 μm-thick sections were prepared for the samples without any heating and glued to frosted glass with cyanoacrylate. The microthermometric study was performed on the selected portions of samples in a linkam THMSG-600 heating and freezing stage. The stage was calibrated with synthetic fluid inclusions, including the triple point of CO2, melting point of H2O, and critical point of H2O. The melting point of H2O standards show that the accuracy for low-temperature measurements is better than ±0.1 °C. Critical point standards show that accuracy for high-temperature measurements is better ±0.1 °C. Homogenization temperatures (Th) have been interpreted as minimum entrapment temperatures. In this case, no pressure corrections were applied because a pressure determination would involve too many error-prone assumptions without an independently obtained value of pressure. The interpretation of Th as minimum entrapment temperatures is a typical procedure in working with Th data [69,70]. To interpret salinity, a NaCl-H2O model (using the equations from [71]) was used on the basis of the observed ice-melting temperatures of the last ice crystal Tm(Ice) from fluid inclusions.

4. Results

4.1. Petrography

4.1.1. The SpFA

The SpFA consist of serpentinite breccias with locally protomylonitic fabric and crosscutting carbonate veins. The serpentinite breccias (Figure 3a) are characterized by angular and irregularly shaped serpentinite grains embedded in the carbonate matrix. In the protomylonitic portion, breccias show differentiated crenulation cleavage with antisymmetric microfolds. The main foliation is defined by cleavage domains (limbs of microfolds) and microlithons (fold hinge areas). The main foliation is well-developed and is marked by hydro-andradite. This schistosity refolded the previous foliation (Figure 3b) defined by recrystallized serpentine, metamorphic clinopyroxene, and accessory minerals such as titanite and magnetite.
The SpFA samples consist of a primary mineral assemblage made up of olivine, orthopyroxene and clinopyroxene (augite), and a serpentinite assemblage consisting of serpentine group minerals (mainly lizardite and a minor amount of chrysotile and antigorite), tremolite, diopside, and clinochlore. Accessory minerals are garnet, titanite, magnetite, and carbonate phases. As recently documented by Dichicco et al. [73], locally edenite amphibole may occur in chlorite-free samples.
Lizardite + magnetite mesh texture or hourglass structures [74,75,76] usually occur in the SpFA showing cores of relict olivine grains replaced by calcite that locally are cross-cut by carbonate veins (Figure 3c). Along cleavage planes, lamellae and fibers of lizardite and chrysotile are spread on bastite pseudomorphs after pyroxene.
Orthopyroxene is replaced by bastite pseudomorphs with exsolution lamellae of lizardite or fine-grained diopside aggregates. Primary magmatic clinopyroxene (augite) is preserved or replaced by metamorphic diopside. Spinel is replaced by Cr and Al-magnetite at the core and clinochlore at the rim.
Several types of carbonate veins have been recognized in the SpFA samples. As shown in Figure 3d,e, veins crosscut the bulk rock and are distinguished in sheeted (micrite-filled veins, carbonate veins with serpentine, carbonate veins with amphibole, fibrous calcite veins) and carbonate types displaying different thickness. Veins occasionally form an anastomosing network with acicular, fibrous, and radial serpentine and amphibole crystals (Figure 3f).

4.1.2. The SpPP

The SpPP are characterized by serpentinite breccias with quartz-carbonate rich veins. The serpentinite assemblage is made up of lizardite and minor chrysotile and antigorite, carbonate minerals, amphibole minerals (actinolite, tremolite), clinochlore, Cr-spinel, quartz, and talc. Magnetite is the only accessory mineral. Similarly to the SpFA, in the SpPP, lizardite occurs in the relict mesh texture and in the matrix.
An irregular patchy texture with carbonates growing after serpentine (Figure 4a) characterizes the SpPP samples. The carbonates occur as microcrystals, together with talc and rare fibrous tremolite. They occur as elongated rombohedric crystals (Figure 4b) only in the veins. Locally, the SpPP samples are characterized by talc-rich domains where quartz and carbonate minerals are also present. Quartz is in micrometer sub-grains, showing undulatory extinction and dynamic recrystallization [77] and intergrowths with carbonate crystals (Figure 4c). Talc, mostly associated with serpentine (lizardite and chrysotile) and chlorite, occurs as massive coarse- to medium-grained aggregates, fine fibers, and/or tabular crystals with perfect cleavage on the [001] plane (Figure 4d).

4.2. Mineral Chemistry

To better characterize silicate and carbonate minerals, an EMPA analysis was performed on selected samples of serpentinites and associated veins, from Fosso Arcangelo (SpFA5 and SpFA39) and Pietrapica (SpPP34A) sites. Amphiboles were analyzed in the host rock and associated veins, pyroxene and garnet were analyzed in the host rock, carbonate minerals were analyzed in the veins only (Table 2, Table 3, Table 4 and Table 5).
The chemical composition of amphiboles is Ca-amphiboles [66], including tremolite and magnesio-ferro-hornblende (Table 2). EMP analyses of edenite crystals are reported by [73].
Garnet is a Ti-rich hydro-andradite (TiO2 = 0.75–3.59 wt %) with high Fetot, Ca, and Mg contents (FeOtot = 17.49–23.41 wt %; CaO = 27.57–32.88 wt %; MgO = 1.41–6.50 wt %) and low Al content (Al2O3 = 0.95–1.97 wt %) (Table 3, Figure 5).
Clinopyroxene is characterized by a homogeneous composition rich in the diopside end-member [78,79], with high Fe content (FeO = 0.88–2.85 wt %) and usually low Al (Al2O3 = 0.03–1.53 wt %) and Cr amounts (Cr2O3 = 0.01–0.16 wt %) (Table 4, Figure 5).
Carbonates consist of calcite (CaO = 52.05–62.38 wt %), Mg-calcite (CaO = 54.81–55.55 wt %; MgO = 1.26 wt %), and dolomite (CaO = 29.60–29.79 wt %; MgO = 19.06–19.23 wt %; and FeOtot = 2.17–3.68 wt %) (Table 5).

4.3. Mineralogy

According to petrographic observations, the XRD analysis revealed that serpentinite samples from both studied sites are made up of serpentine polimorphs, mainly lizardite and, in minor amounts, chrysotile and antigorite, amphibole-like minerals, mainly actinolite and tremolite, clinochlore, magnetite, and calcite. Diopside and hydro-andradite are also present in the SpFA only.
As for veins, a different mineralogical composition was detected for the two analyzed sample groups. The vein infill of the SpFA samples consists of prevalent calcite and traces of aragonite and rhodochrosite. In these samples, traces of silicate minerals, such as serpentine, actinolite, and tremolite, were also detected. The infill of veins traversing the SpPP is dominated by dolomite and Mg-calcite, with quartz as the sole silicate phase.
μ-Raman spectroscopy has been used as a complementary technique to the X-ray diffraction analysis to better identify the carbonate minerals (calcite, aragonite, and dolomite) of both serpentinites and crosscutting veins. The optical vibrations are internal vibrations of the CO3 group (three Raman bands lying between 1500 and 700 cm−1) and external or lattice vibrations involving translation and librations of the CO3 groups relative to the Ca or Mg atoms (500–100 cm−1) [80]. In our samples, calcite is characterized by a dominant Raman band at 1091 cm−1, minor bands at 713, 280, and 155 cm−1 and a very weak band at 1439 cm−1 (Figure 6a). The Raman spectrum for aragonite signals are detected for a dominant Raman band at 1086 cm−1, two strong bands at 212 and 150 cm−1, and three weak bands at 703, 250, and 180 cm−1 (Figure 6b). In dolomite, the main peak in the Raman spectrum occurs at 1103 cm−1, whereas the weak peaks are at 730, 305, and 180 cm−1 (Figure 6c).

4.4. Carbon and Oxygen Stable Isotope Analyses

Results of isotope analyses of carbonate phases in the veins of SpFA and SpPP are presented in Table 6 and Table 7.
The C and O isotope ratios of carbonates from the veins of the SpFA have two distinct ranges. The δ13C values range from −0.81‰ to +2.16‰ and from −2.78‰ to −3.66‰ vs. V-PDB, corresponding δ18O values are between +15.02‰ and +16.38‰ vs. V-SMOW, except for three samples (SpFA7av, SpFA30v, SpFA10v) having δ18O values slightly more positive (between +18.36‰ and +21.53‰ vs. V-SMOW). The δ13C values of carbonate veins of the SpPP are in the range from −1.78‰ to −3.60‰ vs. V-PDB, and the corresponding δ18O average value is +21.3‰ (1σ = 0.18; n = 6) vs. V-SMOW.
Based on the XRPD results, calcite and dolomite have been considered as dominant carbonate phases in the veins of the SpFA and SpPP, respectively. Accordingly, equilibrium temperatures for SpFA veins were computed from δ18O data and considering the calcite-water fractionation curves of O’Neil et al. [81], Friedman and O’Neil [82], and Kim and O’Neil [83]. Instead, for vein samples of SpPP, dolomite-water fractionation curves of Schmidt et al. [84] and Horita [85] have been considered. In both areas, the oxygen isotope composition of water (δ18O) has been assumed to be 0‰ similar to modern seawater, as used by Agrinier et al. [86]. Depending on the fractionation factors available from the literature, for the vein samples of SpFA we have obtained an equilibrium temperature ranging from 83 to 117°C (Δ18Ocalcite from [81]), from 81 to 121 °C (Δ18Ocalcite from [82]), and from 76 to 107 °C (Δ18Ocalcite from [83]).
Similarly, for the vein samples of SpPP, the equilibrium temperature was inferred in the range between 105 and 110 °C (Δ18Odolomite from [84]) and between 83 and 87 °C (Δ18Odolomite from [85]).
The pristine δ13CCO2 (gas) values of the fluid from which veins were formed were computed from the δ13CCaCO3 values and the calculated carbonate deposition temperature for each carbonate sample. We have assumed the achievement of the isotope equilibrium between fluid CO2 and a carbonate mineralogical phase (calcite in the SpFA veins and dolomite in SpPP veins) and applied the following equation:
δ13CCO2 = δ13Ccarb.min–Δcarb.min–CO2
where δ13Ccarb.min is the isotope composition of the considered carbonate mineralogical phase, and Δcarb.min-CO2 is the equilibrium fractionation factor for carbon between the carbonate mineral and CO2 and calcite [87] and between CO2 and dolomite [85] (Table 6 and Table 7). Based on the equilibrium temperature estimated using the fractionation factors computed by [81], [82], and [83], the average δ13CCO2 values of SpFA veins is in the range from −4.36‰ to −5.02‰. Slightly more negative average δ13CCO2 values (from −6.86‰ to −8.32‰) were obtained for SpPP veins if we consider the equilibrium temperature estimated following [84] and [85], respectively.

4.5. Fluid Inclusions Hosted by Quartz in Sppp Veins

Most of the fluid inclusions in the quartz are arranged along lines of crystal growth, and thus, they are considered as primary and/or pseudosecondary fluids according to the criteria defined by [88] (Figure 7a). Some fluid inclusions occur along secondary trails and necking down is occasionally seen. In Table 8, the primary or secondary origin of each fluid inclusion is indicated. When possible, fluid inclusion assemblages [89] have been analyzed. Two major types of fluid inclusions were recognized at room temperature: predominantly all-liquid fluid inclusions (single-phase inclusions: LH2O) and in less amount liquid-vapor inclusions (biphasic inclusions: LH2O + VH2O) (Figure 7b). All-liquid fluid inclusions usually nucleated a small gas bubble with little heating (around 50 °C) (Figure 7c). This means that they are in a metastable state out of their stability field. Only the smallest ones (usually <5 µm) remain all-liquid after heating. These ones cannot be used for microthermometric studies because a bubble is required for temperature determinations.
Collectively, the two types of fluid inclusions have rounded and sub-rounded shapes and exhibit a relatively wide range of liquid/vapor volume ratios (some with ratios around 95:5 and others with ratios between 50:50 and 90:10), indicating a heterogeneous entrapment.
We did not find any evidence for the presence of CO2- and CH4-phases (Tm = −56.6 and −147.0 °C for CO2- and CH4, respectively). In fact, during the cooling phase, we went down until liquid nitrogen temperatures and no melting process different than the melting of ice was observed. This means that the gas is likely water vapor (Figure 7d). However, the presence of another gas different from water vapor cannot be discarded. If present, this gas would have very low density (H2 or He), so that would be undetectable by microthermometry. The temperatures of final ice-melting Tm(Ice) values range from −0.3 to −1.9 °C. Many inclusions present positive ice-melting temperatures, which means that they are under high pressure out of their stability field. Temperatures of the first melting (eutectic temperature, Te) were observed around −30 °C, which is the metastable temperature of the H2O + NaCl system.
Based on the Tm(Ice) and taking into account the H2O + NaCl system, the biphasic inclusions are found to be of low salinity, between 0.53 and 3.23 NaCl mass % equivalent (using equation by [71]). All the biphasic inclusions homogenize to liquid, with the final homogenization temperature (Th) present in two different ranges of temperature: the first one from 93 to 140 °C and the second one from 185 to 335 °C.

5. Discussions

5.1. Mineral Assemblage

As reported by previous studies, the mineral assemblages and texture of serpentinites of the Frido Unit show evidence of ocean floor metamorphism [54,55]. In addition to minerals typical of worldwide serpentinites, including serpentine minerals, amphiboles, pyroxene, chlorite, titanite, and magnetite, the Frido Unit serpentinites are characterized by talc and hydro-garnet. In particular, the presence of garnet in the serpentinitic rocks from the studied area has been documented here for the first time.
Hydro-andradite, containing variable amounts of TiO2 (0.75 to 3.60 wt %, Table 3), occurs in several mineral assemblages of serpentinites, among which the “serpentine + diopside + magnetite” association is one of the most common. Hydro-garnet is stable in these rocks over a wide range of oxygen fugacities and Ca activities, and its stability is controlled by the following reaction [90]:
3Ca3Fe2Si3O12 + 9SiO2 + Fe2O3 = 9CaFeSi2O6 + 2O2
andradite + silica + magnetite = pyroxene + fluid
The titanian hydro-andradite may form in both magmatic and hydrothermal systems. According to [91], in fact, the presence of such a mineral has been documented in rocks associated with silica undersaturated magmatic systems (carbonatites, kimberlites, alkaline intrusions) as well as with hydrothermally alterated oceanic lithosphere (for example the Sanbagawa metamorphic complex, the mid-Atlantic ridge, the Nagaland ophiolite belt), testifying intermediate to low (150 to 300 °C) temperature fluids with low SiO2 activity.
The hydrothermal activity is also thought of as responsible for the talc occurrence in the studied serpentinites. Frost et al. [90] stated that hydrothermal fluids in equilibrium with basic rocks may have high enough silica activity to alter serpentine to talc following the reactions below [16]:
Mg3Si2O5(OH)4 + 2SiO2(aq) → Mg3Si4O10(OH)2 + H2O
serpentine + silica-saturated fluid → talc + aqueous fluid
and/or
2Mg3Si2O5(OH)4 + 3CO2 + 3CaCO3 → Mg3Si4O10(OH)2 + 3CaMg(CO3)2 + 3H2O
serpentine + CO2-rich fluid + calcite → talc + dolomite + aqueous fluid
Based on the mineralogical composition of studied samples, as also confirmed by the petrographic observations (Figure 4d), the above reactions mirror the rock-fluid interaction processes that have involved samples from the Pietrapica site only, in which talc-rich domains and dolomite have been detected. However, in such samples along with talc and carbonate minerals, quartz is present as well. According to Moore and Rymer [92], a large amount of dissolved silica may be supplied to serpentinite rocks during the hydrothermal alteration of serpentine. Therefore, the quartz in the studied samples could represent the result of direct precipitation from a silica-rich fluid derived from the breakdown of serpentinite-forming silicates (serpentine).
The mode of occurrence of quartz in the SpPP veins suggests a further consideration of the chemistry of the mineralizing fluid in the Pietrapica area. The petrographical study of veins, in fact, has shown that intergrowth structures characterize quartz and dolomite in those samples, and this is consistent with the hypothesis of a contemporaneous formation of silicate and carbonate phases from the same source fluid. As a consequence, the chemical features of fluid inclusions in the quartz can be assumed as representative of the chemistry of the whole mineralizing fluid.
As for veins from the Fosso Arcangelo site, no hypothesis can be made on the basis of their mineralogical composition only. These veins are dominated by calcite as the principal carbonate phase. Calcite is a common mineral because it may form in a great variety of geological settings. It represents an important rock-forming mineral in sedimentary rocks, can be an essential component of metamorphic and igneous rocks, and is common in hydrothermal environments [93]. In particular, in geothermal systems, the calcite formation is chiefly favored by boiling, dilution, and condensation processes that control its occurrence, distribution, and stable isotope composition [94]. Further, during serpentinization, the mineralogical and geochemical processes transforming the oceanic lithosphere usually produce Ca-rich fluids that can migrate in the hydrothermal system and promote carbonates precipitation (mainly calcite) as serpentinite matrix and/or infill of veins and veinlets [35,95,96]. In the SpFA veins, calcite locally is in association with serpentine and amphibole crystals. Habitus of crystals and the lack of a preferential orientation of these silicates allow us to suppose that they were englobed into the hydrothermal fluid during its migration through the serpentinite host rocks. The lobate contacts between serpentine and/or amphibole crystals and calcite support this hypothesis.

5.2. Temperature of Precipitation, Fluid Composition, and Sources

Stable isotope (carbon and oxygen) geochemistry provides relevant constraints about metasomatic processes involving carbonates [97,98,99,100,101]. As previously stated, based on δ18O data and assuming the water/mineral isotope equilibrium, equilibrium temperature fluctuates in a narrow range for both the vein samples (T = 80–120 °C for SpFA and T = 80–110 °C for SpPP). Nonetheless, the veins belonging to different outcrops show different prevalent carbonate minerals (calcite in SpFA and dolomite for SpPP) and occurred under similar thermal regimes. The lack of a positive correlation between δ13CCaCO3 and δ18OCaCO3 (Figure 8) seems to indicate that the depositional temperature controls, exclusively, the isotope signature of the carbonate veins. Therefore, the computed δ13CCO2-gas values are representative of the original CO2 supplied during vein formation.
The average fluid δ13CCO2 inferred from the fluid-carbonates isotope equilibrium range from −4.36‰ to −5.32‰ in the veins of the SpFA and from −6.86‰ to −8.32‰ in the veins of the SpPP.
The range of the C-isotope composition of both vein groups is slightly more negative than that of seawater carbonates (δ13CCO2 around 0‰), whereas it lies fully within the range of δ13CCO2 values typically associated to mantle carbon (−8‰< δ13CCO2< −4‰; [102,103,104]. Degassing of CO2-rich fluids during their rising from the underlying lithospheric mantle might lead to the precipitation of calcite [105]. Thus, the inferred δ13CCO2 values are consistent with a magmatic CO2 component in the hydrothermal fluids.
However, the available data do not allow us to rule out that carbon in these veins might derive from other sources than the mantle. In fact, the mixing between fluids having different isotope carbon isotope composition (e.g., seawater and carbon derived from organic rich sediments or from the oxidation of methane having δ13CCO2 < −15‰) would reproduce δ13CCO2 values in the same range to those inferred for the SpFA and SpPP veins.
Alternatively, decarbonation of marine sediments having a typical isotope signature (e.g., δ13C close to 0‰, [106] would produce a lowering of δ13C, with or without change in δ18O, and an O shift can be observed only when decarbonation is driven by infiltration by externally derived H2O-rich fluids [34].
As shown in Figure 8, samples from the two studied sites fall in different fields of the δ18O vs. δ13C diagram. In detail, the SpFA veins overlap the pathway typical of the decarbonation process as suggested by [35] and references therein. Such a process likely developed at depth in the crust in a “closed system” wherein no external fluid supply can occur.
On the contrary, for the SpPP site, vein samples are characterized by an oxygen isotope shift toward more positive δ18O values. Moreover, significant differences in the mineralogical assemblage were also found in SpFA and SpPP veins, being calcite prevalent in the former veins and dolomite and quartz in the latter ones. Intergrowth structures between dolomite and quartz crystals suggest the hypothesis of a contemporaneous formation of silicate and carbonate phases in the SpPP veins.
Data of fluid inclusions in the quartz (in SpPP veins) show abundant aqueous (LH2O-VH2O) and low salinity features (between 0.53 and 3.23 NaCl mass % equivalent). Further, the decrease of salinity associated with the decrease of homogenization temperature (335–185 °C and 140–93 °C) may be related to a large infiltration of shallow, diluted, and fresh waters that also led to a progressive cooling of the hydrothermal system. Therefore, it seems that SpPP veins are consistent with a crystallization in an open-system at shallower crustal conditions.
Accordingly, all these features, even not well-constrained would suggest the hypothesis that SpFA and SpPP veins were deposited under different boundary conditions (e.g., temperature), and/or from parental fluids having different chemical composition.
Further detailed investigations on the fluid inclusions composition as well as on the oxygen isotopes of the silicate are requested to clarify the formation process of these veins and to fully understand the role and the compositions of the parental fluid(s).

6. Conclusions

Our study demonstrates that, in serpentinites of the Frido Unit (southern Apennines), different types of veins occur recording fluid production and migration in the accretionary wedge of the southern Apennines. We envisage at least two formation episodes for veins that crosscut serpentines from the studied sites (Fosso Arcangelo and Pietrapica) within the same thermal system (Figure 9a). We have identified a first vein group formed by decarbonation of serpentinites in a closed hydrothermal system that generated Ca- and CO2-rich fluids from which carbonates of the Fosso Arcangelo veins derive. Even if not well-constrained, such fluids could migrate toward shallower depths, modifying their composition by interaction with serpentinite host rocks. Then, fluid migration ended in an open hydrothermal system where mixing between deep Si and Mg enriched fluids and externally derived H2O-rich fluids occurred, promoting the quartz-carbonate vein formation in the Pietrapica serpentinites. Subsequently, serpentinite slices and associated veins were involved by tectonic activity related to the formation of the Liguride Complex and finally exhumed (Figure 9b).
As a consequence, our study contributes to a better understanding of processes leading to mobilization, fractionation, and redistribution of chemical elements within subduction zones, with particular attention to the southern Apennines.

Author Contributions

Conceptualization, G.R. and G.M.; Methodology, M.C.D.; Validation, P.C., F.G. and M.P.; Formal Analysis, M.C.D., F.G., M.P.; Investigation, M.C.D., S.L., R.S.; Data Curation, M.C.D. and S.L.; Writing-Original Draft Preparation, M.C.D; Writing–Review & Editing, G.R. and R.S.; Supervision, G.M.; Funding Acquisition, G.M. and P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly funded by Mongelli G. grant (RIL2016) and by the project CGL2016-78560-P of the Spanish Ministerio de Economía y Competitividad.

Acknowledgments

The authors wish to thank González Acebrón Laura (Departamento de Estratigrafía, Universidad Complutense, Madrid, España) for fluid inclusions studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alt, J.C.; Shanks III, W.C.; Crispini, L.; Gaggero, L.; Schwarzenbach, E.M.; Früh-Green, G.L.; Bernasconi, S.M. Uptake of carbon and sulphur during seafloor serpentinization and the effects of subduction metamorphism in Ligurian peridotites. Chem. Geol. 2012, 322, 268–277. [Google Scholar] [CrossRef]
  2. Tucholke, B.E.; Lin, J. A geological model for the structure of ridge segments in slow spreading ocean crust. J. Geophys. Res. Space Phys. 1994, 99, 11937–11958. [Google Scholar] [CrossRef]
  3. Cannat, M.; Mevel, C.; Maia, M.; Deplus, C.; Durand, C.; Gente, P.; Agrinier, P.; Belarouchi, A.; Dubuisson, G.; Humler, E.; et al. Thin crust, ultramafic exposures, and rugged faulting patterns at the Mid-Atlantic Ridge (22°–24°N). Geology 1995, 23, 49–52. [Google Scholar] [CrossRef]
  4. Cannat, M.; Fontaine, F.; Escartin, J. Serpentinization and associated hydrogen and methane fluxes at slow spreading ridges. In Sea Ice; American Geophysical Union (AGU): Washington, DC, USA, 2010; Vol. 188, pp. 241–264. [Google Scholar]
  5. Cann, J.R.; Blackman, D.K.; Smith, D.K.; McAllister, E.; Janssen, B.; Mello, S.; Avgerinos, E.; Pascoe, A.R.; Escartin, J.; Cann, D.K.B.J.R. Corrugated slip surfaces formed at ridge–transform intersections on the Mid-Atlantic Ridge. Nature 1997, 385, 329–332. [Google Scholar] [CrossRef]
  6. Blackman, D.K.; Janssen, B.; Smith, D.K.; Cann, J.R. Origin of extensional core complexes: Evidence from the Mid-Atlantic Ridge at Atlantis Fracture Zone. J. Geophys. Res. Space Phys. 1998, 103, 21315–21333. [Google Scholar] [CrossRef]
  7. Smith, D.K.; Cann, J.R.; Escartín, J. Widespread active detachment faulting and core complex formation near 13° N on the Mid-Atlantic Ridge. Nature 2006, 442, 440–443. [Google Scholar] [CrossRef]
  8. Schwarzenbach, E.M.; Früh-Green, G.L.; Bernasconi, S.M.; Alt, J.C.; Iii, W.C.S.; Gaggero, L.; Crispini, L. Sulfur geochemistry of peridotite-hosted hydrothermal systems: Comparing the Ligurian ophiolites with oceanic serpentinites. Geochim. Et Cosmochim. Acta 2012, 91, 283–305. [Google Scholar] [CrossRef]
  9. Alt, J.C.; Shanks, W.C. Sulfur in serpentinized oceanic peridotites: Serpentinization processes and microbial sulfate reduction. J. Geophys. Res. Space Phys. 1998, 103, 9917–9929. [Google Scholar] [CrossRef]
  10. Alt, J.C.; Shanks, W.C. Serpentinization of abyssal peridotites from the MARK area, Mid-Atlantic Ridge: sulfur geochemistry and reaction modeling. Geochim. Et Cosmochim. Acta 2003, 67, 641–653. [Google Scholar] [CrossRef]
  11. Alt, J.C.; Shanks, W.C.; Bach, W.; Paulick, H.; Garrido, C.J.; Beaudoin, G. Hydrothermal alteration and microbial sulfate reduction in peridotite and gabbro exposed by detachment faulting at the Mid-Atlantic Ridge, 15°20′N (ODP Leg 209): A sulfur and oxygen isotope study. Geochem. Geophys. Geosystems 2007, 8. [Google Scholar] [CrossRef]
  12. Alt, J.C.; Schwarzenbach, E.M.; Früh-Green, G.L.; Shanks, W.C.; Bernasconi, S.M.; Garrido, C.J.; Crispini, L.; Gaggero, L.; Padron-Navarta, J.A.; Marchesi, C.; et al. The role of serpentinites in cycling of carbon and sulfur: Seafloor serpentinization and subduction metamorphism. Lithos 2013, 178, 40–54. [Google Scholar] [CrossRef]
  13. Bach, W.; Garrido, C.J.; Paulick, H.; Harvey, J.; Rosner, M. Seawater-peridotite interactions: First insights from ODP Leg 209, MAR 15°N. Geochem. Geophys. Geosystems 2004, 5. [Google Scholar] [CrossRef]
  14. Barnes, J.; Sharp, Z. Achlorine isotope study of DSDP/ODP serpentinized ultramafic rocks: Insights into the serpentinization process. Chem. Geol. 2006, 228, 246–265. [Google Scholar] [CrossRef]
  15. Bonatti, E.; Lawrence, J.R.; Morandi, N. Serpentinization of oceanic peridotites: temperature dependence of mineralogy and boron content. Earth Planet. Sci. Lett. 1984, 70, 88–94. [Google Scholar] [CrossRef]
  16. Boschi, C.; Früh-Green, G.L.; Escartín, J. Occurrence and significance of serpentinite-hosted, talc-and amphibole-rich fault rocks in modern oceanic settings and ophiolite complexes: An overview. Ofioliti 2006, 31, 129–140. [Google Scholar]
  17. Boschi, C.; Dini, A.; Früh-Green, G.L.; Kelley, D.S. Isotopic and element exchange during serpentinization and metasomatism at the Atlantis Massif (MAR 30°N): Insights from B and Sr isotope data. Geochim. Et Cosmochim. Acta 2008, 72, 1801–1823. [Google Scholar] [CrossRef]
  18. Delacour, A.; Früh-Green, G.L.; Bernasconi, S.M. Sulfur mineralogy and geochemistry of serpentinites and gabbros of the Atlantis Massif (IODP Site U1309). Geochim. Et Cosmochim. Acta 2008, 72, 5111–5127. [Google Scholar] [CrossRef]
  19. Delacour, A.; Früh-Green, G.L.; Bernasconi, S.M.; Schaeffer, P.; Kelley, D.S. Carbon geochemistry of serpentinites in the Lost City Hydrothermal System (30°N, MAR). Geochim. Et Cosmochim. Acta 2008, 72, 3681–3702. [Google Scholar] [CrossRef]
  20. Delacour, A.; Früh-Green, G.L.; Bernasconi, S.M.; Kelley, D.S. Sulfur in peridotites and gabbros at Lost City (30°N, MAR): Implications for hydrothermal alteration and microbial activity during serpentinization. Geochim. Et Cosmochim. Acta 2008, 72, 5090–5110. [Google Scholar] [CrossRef]
  21. Früh-Green, G.L.; Connolly, J.A.; Plas, A.; Kelley, D.S.; Grobéty, B.; Wilcock, W.S.; Delong, E.F.; Baross, J.A.; Cary, S.C. Serpentinization of oceanic peridotites: Implications for geochemical cycles and biological activity. Subseafloor Biosph. Mid-Ocean Ridges 2004, 144, 119–136. [Google Scholar]
  22. Kendrick, M.A.; Scambelluri, M.; Honda, M.; Phillips, D. High abundances of noble gas and chlorine delivered to the mantle by serpentinite subduction. Nat. Geosci. 2011, 4, 807–812. [Google Scholar] [CrossRef]
  23. Paulick, H.; Bach, W.; Godard, M.; De Hoog, J.; Suhr, G.; Harvey, J. Geochemistry of abyssal peridotites (Mid-Atlantic Ridge, 15°20′N, ODP Leg 209): Implications for fluid/rock interaction in slow spreading environments. Chem. Geol. 2006, 234, 179–210. [Google Scholar] [CrossRef] [Green Version]
  24. Scambelluri, M.; Fiebig, J.; Malaspina, N.; Muntener, O.; Pettke, T. Serpentinite Subduction: Implications for Fluid Processes and Trace-Element Recycling. Int. Geol. Rev. 2004, 46, 595–613. [Google Scholar] [CrossRef]
  25. Schwarzenbach, E.M. Serpentinization, fluids and life: Comparing carbon and sulfur cycles in modern and ancient environments. Doctoral dissertation, ETH Zurich, Zurich, Switzerland, 2011; 240p. [Google Scholar]
  26. Vils, F.; Muntener, O.; Kalt, A.; Ludwig, T. Implications of the serpentine phase transition on the behaviour of beryllium and lithium–boron of subducted ultramafic rocks. Geochim. Et Cosmochim. Acta 2011, 75, 1249–1271. [Google Scholar] [CrossRef]
  27. Kelemen, P.B.; Matter, J.; Streit, E.E.; Rudge, J.F.; Curry, W.B.; Blusztajn, J. Rates and Mechanisms of Mineral Carbonation in Peridotite: Natural Processes and Recipes for Enhanced, in situ CO2 Capture and Storage. Annu. Rev. Earth Planet. Sci. 2011, 39, 545–576. [Google Scholar] [CrossRef]
  28. Lacinska, A.M.; Styles, M.T.; Bateman, K.; Hall, M.R.; Brown, P.D. An Experimental Study of the Carbonation of Serpentinite and Partially Serpentinised Peridotites. Front. Earth Sci. 2017, 5, 1–20. [Google Scholar] [CrossRef] [Green Version]
  29. Miller, J.A.; Cartwright, I.; Buick, I.S.; Barnicoat, A.C. An O-isotope profile through the HP-LT Corsican ophiolite, France and its implications for fluid flowduring subduction. Chem. Geol. 2001, 178, 43–69. [Google Scholar] [CrossRef]
  30. Eickmann, B.; Bach, W.; Rosner, M.; Peckmann, J. Geochemical constraints on the modes of carbonate precipitation in peridotites from the Logatchev Hydrothermal Vent Field and Gakkel Ridge. Chem. Geol. 2009, 268, 97–106. [Google Scholar] [CrossRef]
  31. Bach, W.; Rosner, M.; Jöns, N.; Rausch, S.; Robinson, L.F.; Paulick, H.; Erzinger, J. Carbonate veins trace seawater circulation during exhumation and uplift of mantle rock: Results from ODP Leg 209. Earth Planet. Sci. Lett. 2011, 311, 242–252. [Google Scholar] [CrossRef] [Green Version]
  32. Schwarzenbach, E.M.; Früh-Green, G.L.; Bernasconi, S.M.; Alt, J.C.; Plas, A. Serpentinization and carbon sequestration: A study of two ancient peridotite-hosted hydrothermal systems. Chem. Geol. 2013, 351, 115–133. [Google Scholar] [CrossRef]
  33. Cook-Kollars, J.; Bebout, G.E.; Collins, N.C.; Angiboust, S.; Agard, P. Subduction zone metamorphic pathway for deep carbon cycling: I. Evidence from HP/UHP metasedimentary rocks, Italian Alps. Chem. Geol. 2014, 386, 31–48. [Google Scholar] [CrossRef]
  34. Collins, N.C.; Bebout, G.E.; Angiboust, S.; Agard, P.; Scambelluri, M.; Crispini, L.; John, T. Subduction zone metamorphic pathway for deep carbon cycling: II. Evidence from HP/UHP metabasaltic rocks and ophicarbonates. Chem. Geol. 2015, 412, 132–150. [Google Scholar] [CrossRef] [Green Version]
  35. Piccoli, F.; Brovarone, A.V.; Beyssac, O.; Martinez, I.; Ague, J.J.; Chaduteau, C. Carbonation by fluid–rock interactions at high-pressure conditions: Implications for carbon cycling in subduction zones. Earth Planet. Sci. Lett. 2016, 445, 146–159. [Google Scholar] [CrossRef] [Green Version]
  36. De Felipe, I.; Pedreira, D.; Pulgar, J.A.; Iriarte, E.; Mendia, M. Mantle exhumation and metamorphism in the Basque-Cantabrian Basin (N Spain): Stable and clumped isotope analysis in carbonates and comparison with ophicalcites in the North-Pyrenean Zone (Urdach and Lherz). Geochem. Geophys. Geosystems 2017, 18, 631–652. [Google Scholar] [CrossRef] [Green Version]
  37. Structural and petrological analyses of the Frido Unit (southern Italy): New insights into the early tectonic evolution of the southern Apennines–Calabrian Arc system. Lithos 2013, 168, 219–235.
  38. Knott, S.D. Structure, kinematics and metamorphism in the Liguride Complex, southern Apennines, Italy. J. Struct. Geol. 1994, 16, 1107–1120. [Google Scholar] [CrossRef]
  39. Monaco, C.; Tortorici, L. Tectonic role of ophiolite-bearing terranes in the development of the Southern Apennines orogenic belt. Terra Nova 1995, 7, 153–160. [Google Scholar] [CrossRef]
  40. Bonardi, G.; Amore, F.O.; Ciampo, G.; De Capoa, P.; Micconet, P.; Perrone, V. II complesso Liguride. Auct: Stato delle conoscenze e problemi aperti sulla sua evoluzione pre-appenninica ed i suoi rapporti con l’Arco Calabro. Mem. Della Soc. Geol. Ital. 1988, 41, 17–35. [Google Scholar]
  41. Spadea, P. Continental rocks associated with ophiolites in Lucanian Apennine, southern Italy. Ofioliti 1982, 7, 501–522. [Google Scholar]
  42. Laurita, S.; Prosser, G.; Rizzo, G.; Langone, A.; Tiepolo, M.; Laurita, A. Geochronological study of zircons from continental crust rocks in the Frido Unit (southern Apennines). Acta Diabetol. 2014, 104, 179–203. [Google Scholar] [CrossRef]
  43. Sansone, M.T.C.; Rizzo, G. Rodingites from Frido Unit: Evidences for metasomatic alteration. Rend. Online Della Soc. Geol. Ital. Abstr. 2010, 11, 106–107. [Google Scholar]
  44. Rizzo, G.; Sansone, M.T.C.; Perri, F.; Laurita, S. Mineralogy and petrology of the metasedimentary rocks from the Frido Unit (southern Apennines, Italy). Period. Miner.. 2016, 85, 153–168. [Google Scholar]
  45. Rizzo, G.; Canora, F.; Dichicco, M.C.; Laurita, S.; Sansone, M.T.C. P–T estimates from amphibole and plagioclase pairs in metadolerite dykes of the Frido unit (southern Apennines-Italy) during the ocean-floor metamorphism. J. Mediterr. Earth Sci. 2019, 11. [Google Scholar] [CrossRef]
  46. Sinisi, R.; Mongelli, G.; Perri, F.; Rizzo, G. The braunite (3Mn2O3–MnSiO3)-rich mineralization in the metasedimentary succession from southern Apennines (Italy): Genesis constraints. Ore Geol. Rev. 2018, 94, 1–11. [Google Scholar] [CrossRef]
  47. Rizzo, G.; Laurita, S.; Altenberger, U. The Timpa delle Murge ophiolitic gabbros, southern Apennines: Insights from petrology and geochemistry and consequences to the geodynamic setting. Period. Miner.. 2018, 87, 5–20. [Google Scholar]
  48. Dichicco, M.C.; Laurita, S.; Paternoster, M.; Rizzo, G.; Sinisi, R.; Mongelli, G. Serpentinite Carbonation for CO2 Sequestration in the Southern Apennines: Preliminary Study. Energy Procedia 2015, 76, 477–486. [Google Scholar] [CrossRef] [Green Version]
  49. Dichicco, M.C.; De Bonis, A.; Mongelli, G.; Rizzo, G.; Sinisi, R. μ-Raman spectroscopy and X-ray diffraction of asbestos’ minerals for geo-environmental monitoring: The case of the southern Apennines natural sources. Appl. Clay Sci. 2017, 141, 292–299. [Google Scholar] [CrossRef]
  50. Dichicco, M.C.; Laurita, S.; Sinisi, R.; Battiloro, R.; Rizzo, G. Environmental and Health: The Importance of Tremolite Occurence in the Pollino Geopark (Southern Italy). Geosciences 2018, 8, 98. [Google Scholar] [CrossRef] [Green Version]
  51. Dichicco, M.C.; Castiñeiras, P.; Francisco, C.G.; Acebrón, L.G.; Grassa, F.; Laurita, S.; Paternoster, M.; Rizzo, G.; Sinisi, R.; Mongelli, G. Genesis of carbonate-rich veins in the serpentinites at the Calabria-Lucania boundary (southern Apennines). Rend. Online della Soc. Geol. Ital. 2018, 44, 143–149. [Google Scholar] [CrossRef] [Green Version]
  52. Mazzeo, F.; Zanetti, A.; Aulinas, M.; Petrosino, P.; Arienzo, I.; D’Antonio, M. Evidence for an intra-oceanic affinity of the serpentinized peridotites from the Mt. Pollino ophiolites (Southern Ligurian Tethys): Insights into the peculiar tectonic evolution of the Southern Apennines. Lithos 2017, 284, 367–380. [Google Scholar] [CrossRef]
  53. Sansone, M.T.C.; Rizzo, G.; Mongelli, G. Petrochemical characterization of mafic rocks from Ligurian ophiolites, southern Apennines. Int. Geol. Rev. 2011, 53, 130–156. [Google Scholar] [CrossRef]
  54. Sansone, M.T.C.; Prosser, G.; Rizzo, G.; Tartarotti, P. Spinel peridotites of the Frido unit ophiolites (southern Apennines Italy): Evidence for oceanic evolution. Period. Miner.. 2012, 81, 35–59. [Google Scholar]
  55. Sansone, M.T.C.; Tartarotti, P.; Prosser, G.; Rizzo, G. From ocean to subduction: The polyphase metamorphic evolution of the Frido unit metadolerite dykes (southern Apennine, Italy). J. Virtual Explor. 2012, 41, 3. [Google Scholar] [CrossRef]
  56. Sansone, M.T.C.; Rizzo, G. Pumpellyite veins in the metadolerite of the Frido unit (southern Apennines-Italy). Period. Miner.. 2012, 81, 75–92. [Google Scholar]
  57. Spadea, P. Contributo alla conoscenza dei metabasalti ofiolitici della Calabria settentrionale e centrale e dell’Appennino lucano. Rend. Soc. Ital. Mineral. Petrol. 1979, 35, 251–276. [Google Scholar]
  58. Knott, S.D. The Liguride Complex of Southern Italy—A Cretaceous to Paleogene accretionary wedge. Tectonophys. 1987, 142, 217–226. [Google Scholar] [CrossRef]
  59. Beccaluva, L.; Maciotta, G.; Spadea, P. Petrology and geodynamic significance of the calabria–lucania ophiolites. Rend. Soc. Ital. Min.. Pet.. 1982, 38, 973–987. [Google Scholar]
  60. Cavalcante, F.; Belviso, C.; Laurita, S.; Prosser, G. P–T constraints from phyllosilicates of the Liguride complex of the Pollino area (southern Apennines, Italy): Geological inferences. Ofioliti 2012, 37, 65–75. [Google Scholar]
  61. Invernizzi, C.; Bigazzi, G.; Corrado, S.; Di Leo, P.; Schiattarella, M.; Zattin, M. New thermobaric constraints of the exhumation history of the Liguride accretionary wedge, Southern Italy. Ofioliti 2008, 33, 21–32. [Google Scholar]
  62. Laurita, S.; Rizzo, G. Blueschist metamorphism of metabasite dykes in the serpentinites of the Frido Unit, Pollino Massif. Rend. Online Della Soc. Geol. Ital. 2018, 45, 129–135. [Google Scholar] [CrossRef]
  63. Cirrincione, R.; Monaco, C. Evoluzione tettonometamorfica dell’Unità del Frido (Appennino Meridionale). Mem. della Soc. Geol. Ital. 1996, 51, 83–92. [Google Scholar]
  64. Laurita, S.; Rizzo, G. The First Occurrence of Asbestiform Magnesio-Riebeckite in Schists in the Frido Unit (Pollino Unesco Global Geopark, Southern Italy). Fibers 2019, 7, 79. [Google Scholar] [CrossRef] [Green Version]
  65. Laurita, S. IL prisma di accrezione liguride affiorante al confine Calabro-Lucano: studio termocronologico e strutturale. Ph.D Thesis, Università degli Studi della Basilicata, Potenza, Italy, 2008; p. 225. [Google Scholar]
  66. Leake, B.E.; Woolley, A.R.; Arps, C.E.S.; Birch, W.D.; Gilbert, M.C.; Grice, J.D.; Hawthorne, F.C.; Kato, A.; Kisch, H.J.; Krivovichev, V.G.; et al. Nomenclature of amphiboles: Report of the Subcommittee on Amphiboles of the International Mineralogical Association, Commission on New Minerals and Mineral Names. Am. Miner. 1997, 82, 1019–1037. [Google Scholar]
  67. Leake, B.E.; Woolley, A.R.; Birch, W.D.; Burke, E.A.J.; Ferraris, G.; Grice, J.D.; Hawthorne, F.C.; Kisch, H.J.; Krivovichev, V.G.; Schumacher, J.C.; et al. Nomenclature of amphiboles: Additions and revisions to the International Mineralogical Association’s amphibole nomenclature. Am. Mineral. 2004, 89, 883–887. [Google Scholar]
  68. Coplen, T.B.; Kendall, C.; Hopple, J. Comparison of stable isotope reference samples. Nature 1983, 302, 236–238. [Google Scholar] [CrossRef]
  69. Goldstein, R.H.; Rossi, C. Recrystallization in quartz over-growths. J. Sediment Res. 2002, 72, 432–440. [Google Scholar] [CrossRef]
  70. Goldstein, R.H. Fluid inclusion geothermometry in sedimentary system: From paleoclimate to hydrothermal. In SEPM Special Publication, Thermal History Analysis of Sedimentary Basins; Harris, N.B., Peters, K.E., Eds.; Society for Sedimentary Geology: Tulsa, OK, USA, 2012; pp. 45–63. [Google Scholar]
  71. Bodnar, R.J. Revised equation and Table for determining the freezing point depression of H2O–NaCl solution. Geochim. Cosmochim. Acta 1993, 57, 683–684. [Google Scholar] [CrossRef]
  72. Whitney, D.L.; Evans, B.W. Abbreviations for names of rock-forming minerals. Am. Miner.. 2010, 95, 185–187. [Google Scholar] [CrossRef]
  73. Dichicco, M.C.; Paternoster, M.; Rizzo, G.; Sinisi, R. Mineralogical Asbestos Assessment in the Southern Apennines (Italy): A Review. Fibers 2019, 7, 24. [Google Scholar] [CrossRef] [Green Version]
  74. Wicks, F.J.; Whittaker, E.J.W. Serpentine textures and serpentinization. Can. Miner.. 1977, 15, 459–488. [Google Scholar]
  75. Wicks, F.J.; Plant, A.G. Electron-microprobe and X-ray microbeam studies of serpentine textures. Can. Miner.. 1979, 17, 785–830. [Google Scholar]
  76. Blaise, S.; Auvray, B. Serpentinization in the Archean komatitic rocks of the kuhmo greenstone belt, eastern Finland. Can. Miner.. 1990, 28, 56–66. [Google Scholar]
  77. Passchier, C.W.; Trouw, R.A.J. Microtectonics; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
  78. Morimoto, N. Nomenclature of pyroxenes. Mineral. Petrol. 1988, 39, 55–76. [Google Scholar] [CrossRef]
  79. Morimoto, N. Nomenclature of pyroxenes. Mineral. J. 1989, 14, 198–221. [Google Scholar] [CrossRef] [Green Version]
  80. Perraki, M.; Proyer, A.; Mposkos, E.; Kaindl, R.; Hoinkes, G. Raman micro-spectroscopy on diamond, graphite and other carbon polymorphs from the ultrahigh-pressure metamorphic Kimi Complex of the Rhodope Metamorphic Province, NE Greece. Earth Planet. Sci. Lett. 2006, 241, 672–685. [Google Scholar] [CrossRef]
  81. O’Neil, J.R. Oxygen Isotope Fractionation in Divalent Metal Carbonates. J. Chem. Phys. 1969, 51, 5547–5558. [Google Scholar] [CrossRef]
  82. Friedman, I.; O’Neil, J.R. Data of Geochemistry: Compilation of Stable Isotope Fractionation Factors of Geochemical Interest; US Government Printing Office: Washington, DC, USA, 1977; Volume 440.
  83. Kim, S.T.; O’Neil, J.R. Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochim. Cosmochim. Acta 1997, 61, 3461–3475. [Google Scholar] [CrossRef]
  84. Schmidt, M.; Xeflide, S.; Botz, R.; Mann, S. Oxygen isotope fractionation during synthesis of CaMg-carbonate and implications for sedimentary dolomite formation. Geochim. Et Cosmochim. Acta 2005, 69, 4665–4674. [Google Scholar] [CrossRef] [Green Version]
  85. Horita, J. Oxygen and carbon isotope fractionation in the system dolomite–water–CO2 to elevated temperatures. Geochim. Et Cosmochim. Acta 2014, 129, 11–124. [Google Scholar] [CrossRef]
  86. Agrinier, P.; Cornen, G.; Beslier, M.O. Mineralogical and oxygen isotopic features of serpentinites recovered from the ocean/continent transition in the Iberia Abyssal Plain. In Proceedings of the ocean drilling program scientific results; National science foundation: Alexandria, VA, USA, 1996; pp. 541–552. [Google Scholar]
  87. Ohmoto, H.; Rye, R.O. Isotopes of sulfur and carbon. Geochemistry of Hydrothermal Ore Deposits; Barnes, H.L., Ed.; John Wiley & Sons: New York, NY, USA, 1979; pp. 509–567. [Google Scholar]
  88. Roedder, E. Fluid inclusion analysis—prologue and epilogue. Geochim. Et Cosmochim. Acta 1990, 54, 495–507. [Google Scholar] [CrossRef]
  89. Goldstein, R.H.; Reynolds, T.J. Fluid inclusion microthermometry. In Systematics of Ffluid Inclusions in Diagenetic Minerals; Society for Sedimentation Geology: Tulsa, OK, USA, 1994; Short course 31; pp. 87–121. [Google Scholar]
  90. Frost, B.R.; Beard, J.S. On silica activity and serpentinization. J. Petrol. 2007, 48, 1351–1368. [Google Scholar] [CrossRef] [Green Version]
  91. Ghosh, B.; Morishita, T.; Ray, J.; Tamura, A.; Mizukami, T.; Soda, Y.; Ovung, T.N. A new occurrence of titanian (hydro) andradite from the Nagaland ophiolite, India: Implications for element mobility in hydrothermal environments. Chem. Geol. 2017, 457, 47–60. [Google Scholar] [CrossRef]
  92. Moore, D.E.; Rymer, M.J. Talc-bearing serpentinite and the creeping section of the San Andreas fault. Nature 2007, 448, 795. [Google Scholar] [CrossRef] [PubMed]
  93. Klein, C. Mineralogia; Zanichelli: Bologna, Italy, 2004. [Google Scholar]
  94. Simmons, S.F.; Christenson, B.W. Origins of calcite in a boiling geothermal system. Am. J. Sci. 1994, 294, 361–400. [Google Scholar] [CrossRef]
  95. Piccoli, F.; Brovarone, A.V.; Ague, J.J. Field and petrological study of metasomatism and high-pressure carbonation from lawsonite eclogite-facies terrains, Alpine Corsica. Lithos 2018, 304, 16–37. [Google Scholar] [CrossRef]
  96. Scambelluri, M.; Bebout, G.E.; Belmonte, D.; Gilio, M.; Campomenosi, N.; Collins, N.; Crispini, L. Carbonation of subduction-zone serpentinite (high-pressure ophicarbonate; Ligurian Western Alps) and implications for the deep carbon cycling. Earth Planet. Sci. Lett. 2016, 441, 155–166. [Google Scholar] [CrossRef] [Green Version]
  97. Valley, J.W. Stable isotope geochemistry of metamorphic rocks. Rev. Mineral. Geochem. 1986, 16, 445–489. [Google Scholar]
  98. Wang, Q.; Rumble, D. Oxygen and carbon isotope composition from the UHP Shuanghe marbles, Dabie Mountains, China. Sci. China Ser. D Earth Sci. 1999, 42, 88–96. [Google Scholar] [CrossRef]
  99. Ague, J.J.; Nicolescu, S. Carbon dioxide released from subduction zones by fluid mediated reactions. Nat. Geosci. 2014, 7, 355. [Google Scholar] [CrossRef]
  100. Galvez, M.E.; Beyssac, O.; Martinez, I.; Benzerara, K.; Chaduteau, C.; Malvoisin, B.; Malavieille, J. Graphite formation by carbonate reduction during subduction. Nat. Geosci. 2013, 6, 473–477. [Google Scholar] [CrossRef]
  101. Galvez, M.E.; Martinez, I.; Beyssac, O.; Benzerara, K.; Agrinier, P.; Assayag, N. Metasomatism and graphite formation at a lithological interface in Malaspina (Alpine Corsica, France). Contrib. Mineral. Petrol. 2013b, 166, 1687–1708. [Google Scholar] [CrossRef]
  102. Des Marais, D.J.; Moore, J.G. Carbon and its isotopes in mid-oceanic basaltic glasses. Earth Planet. Sci. Lett. 1984, 69, 43–57. [Google Scholar] [CrossRef]
  103. Marty, B.; Jambon, A.; Sano, Y. Helium isotopes and CO2 in volcanic gases of Japan. Chem. Geol. 1989, 76, 25–40. [Google Scholar] [CrossRef]
  104. Sano, Y.; Williams, S.N. Fluxes of mantle and subducted carbon along convergent plate boundaries. Geophys. Res. Lett. 1996, 23, 2749–2752. [Google Scholar] [CrossRef]
  105. Zheng, Y.F. Carbon-oxygen isotopic covariation in hydrothermal calcite during degassing of CO2. Miner. Depos. 1990, 25, 246–250. [Google Scholar] [CrossRef]
  106. Hoefs, J. Stable Isotope Geochemistry; Springer: Berlin/Heidelberg, Germany, 2018; Volume XXVI, p. 437. [Google Scholar]
Figure 1. (a) Synthetic geological map of the Liguride Complex in the Pollino Massif and location of the study area, (b) tectonic and stratigraphic sequence with the ultramaphic and continental rocks contact (not to scale).
Figure 1. (a) Synthetic geological map of the Liguride Complex in the Pollino Massif and location of the study area, (b) tectonic and stratigraphic sequence with the ultramaphic and continental rocks contact (not to scale).
Minerals 10 00127 g001
Figure 2. (a) Carbonate veins in the serpentinite from the Fosso Arcangelo site (SpFA); (b) Cataclastic serpentinites with quartz-carbonate veins and talc-rich domains in the Pietrapica quarry (SpPP).
Figure 2. (a) Carbonate veins in the serpentinite from the Fosso Arcangelo site (SpFA); (b) Cataclastic serpentinites with quartz-carbonate veins and talc-rich domains in the Pietrapica quarry (SpPP).
Minerals 10 00127 g002
Figure 3. Photomicrographs of SpFA samples: (a) The serpentinite breccia characterized by serpentinite fragments in a carbonate matrix, plane polarized light, 4×; (b) schistosity with crenulation cleavage, crossed polarized light, 4×; (c) relict olivine grains in the core of the mesh-texture. Olivine is replaced by calcite, the mesh-texture are cross-cut by carbonate veins, crossed polarized light, 4×; (d) sub-parallel carbonate veins, crossed polarized light, 4×; (e) vein infill with carbonate and serpentine minerals, crossed polarized light, 4×; (f) carbonate veins with anastomosing network, crossed polarized light, 4×. Cal = calcite; Srp = serpentine. Abbreviation after [72].
Figure 3. Photomicrographs of SpFA samples: (a) The serpentinite breccia characterized by serpentinite fragments in a carbonate matrix, plane polarized light, 4×; (b) schistosity with crenulation cleavage, crossed polarized light, 4×; (c) relict olivine grains in the core of the mesh-texture. Olivine is replaced by calcite, the mesh-texture are cross-cut by carbonate veins, crossed polarized light, 4×; (d) sub-parallel carbonate veins, crossed polarized light, 4×; (e) vein infill with carbonate and serpentine minerals, crossed polarized light, 4×; (f) carbonate veins with anastomosing network, crossed polarized light, 4×. Cal = calcite; Srp = serpentine. Abbreviation after [72].
Minerals 10 00127 g003
Figure 4. Photomicrographs of SpPP samples: (a) irregular patchy texture, crossed polarized light, 4×; (b) carbonate veins with elongated rhombohedral crystals, plane polarized light, 4×; (c) dynamic recrystallization of quartz, crossed polarized light, 4×; (d) fibers of talc associated with serpentine and quartz, crossed polarized light, 4×. Qtz = quartz; Srp = serpentine; Tlc = talc. Abbreviation after [72].
Figure 4. Photomicrographs of SpPP samples: (a) irregular patchy texture, crossed polarized light, 4×; (b) carbonate veins with elongated rhombohedral crystals, plane polarized light, 4×; (c) dynamic recrystallization of quartz, crossed polarized light, 4×; (d) fibers of talc associated with serpentine and quartz, crossed polarized light, 4×. Qtz = quartz; Srp = serpentine; Tlc = talc. Abbreviation after [72].
Minerals 10 00127 g004
Figure 5. Secondary-electron image of serpentinite SpFA showing the presence of andradite (Adr), diopside (Di), and small crystals of magnetite (Mag) analyzed by EMPA. Adr = andradite; Cal = calcite; Di = diopside; Mag = magnesite; Serp = serpentine.
Figure 5. Secondary-electron image of serpentinite SpFA showing the presence of andradite (Adr), diopside (Di), and small crystals of magnetite (Mag) analyzed by EMPA. Adr = andradite; Cal = calcite; Di = diopside; Mag = magnesite; Serp = serpentine.
Minerals 10 00127 g005
Figure 6. Raman spectra at low wavenumbers of carbonate minerals: (a) calcite; (b) aragonite; (c) dolomite in selected veins from SpFA and SpPP.
Figure 6. Raman spectra at low wavenumbers of carbonate minerals: (a) calcite; (b) aragonite; (c) dolomite in selected veins from SpFA and SpPP.
Minerals 10 00127 g006
Figure 7. Photomicrographs of fluid inclusions in quartz of veins from SpPP. Primary (red dotted line) and secondary (green dotted line) fluid inclusions (a); fluid inclusions at room temperature (b) and at ~50°C (c); monophasic (LH2O) and biphasic fluid inclusions (LH2O + VH2O) (d).
Figure 7. Photomicrographs of fluid inclusions in quartz of veins from SpPP. Primary (red dotted line) and secondary (green dotted line) fluid inclusions (a); fluid inclusions at room temperature (b) and at ~50°C (c); monophasic (LH2O) and biphasic fluid inclusions (LH2O + VH2O) (d).
Minerals 10 00127 g007
Figure 8. δ13CV-PDB vs. δ18OV-SMOW plot of calcite/dolomite in the serpentinites from the Fosso Arcangelo site (SpFA = green diamonds) and Pietrapica quarry (SpPP = red squares). Geochemical trends of carbonates affected by dissolutions [99], carbonate reduction [101], and carbonation and decarbonation [35] are also shown.
Figure 8. δ13CV-PDB vs. δ18OV-SMOW plot of calcite/dolomite in the serpentinites from the Fosso Arcangelo site (SpFA = green diamonds) and Pietrapica quarry (SpPP = red squares). Geochemical trends of carbonates affected by dissolutions [99], carbonate reduction [101], and carbonation and decarbonation [35] are also shown.
Minerals 10 00127 g008
Figure 9. Bi-dimensional sketch displaying the tectonic evolution of the Liguride Complex. Inset (a) shows the position of serpentinites and cutting veins of Fosso Arcangelo (SpFA) and Pietrapica (SpPP) sites in the subduction zone. In the red box is shown a scheme illustrating the different hydrothermal systems of both studied sites along with the pathway of migrating fluids (not in scale). Inset (b) displays the actual position of studied rocks in the accretionary wedge of the Pollino Massif. See the text further details.
Figure 9. Bi-dimensional sketch displaying the tectonic evolution of the Liguride Complex. Inset (a) shows the position of serpentinites and cutting veins of Fosso Arcangelo (SpFA) and Pietrapica (SpPP) sites in the subduction zone. In the red box is shown a scheme illustrating the different hydrothermal systems of both studied sites along with the pathway of migrating fluids (not in scale). Inset (b) displays the actual position of studied rocks in the accretionary wedge of the Pollino Massif. See the text further details.
Minerals 10 00127 g009
Table 1. GPS coordinates, texture, and mineral assemblages of the serpentinite samples.
Table 1. GPS coordinates, texture, and mineral assemblages of the serpentinite samples.
Sample No.Sample CodeGPS CoordinatesTextureMineral Assemblage
Samples from the Fosso Arcangelo site (SpFA)
1SpFA241°01’40.2’’ N 16°08’09.6’’ EBreccia, pseudomorphic and veins textureSrp–Cal–Am–Pren–Chl–Mag and Ol–Opx relict
2SpFA540°01’39.2’’ N 16°08’09.6’’ EBreccia, veins textureSrp–Cal–Am–Chl–Mag–Ttn
3SpFA6.340°01’38.3’’ N 16°08’10.4’’ EBreccia, pseudomorphic, veins and patch textureSrp–Cal–Am–Chl–Mag–Pmp and Ol–Opx–Cpx relict
4SpFA740°01’37.6’’ N 16°08’10.7’’ EPseudomorphyc and veins textureSrp–Cal–Am–Chl–Pmp–Ol–Opx–Cpx relict–Mag
5SpFA7a40°01’37.6’’ N 16°08’10.7’’ EPseudomorphyc, veins and patch textureSrp–Cal–Am–Chl–Mag and Ol–Opx–Cpx relict
6SpFA940°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins textureSrp–Cal–Di–Adr–Mag
7SpFA3740°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
8SpFA3840°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
9SpFA3940°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
10SpFA4040°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
11SpFA4140°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage and veins textureSrp–Cal–Di–Adr–Mag
12SpFA4240°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
13SpFA4340°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulations cleavage and veins textureSrp–Cal–Di–Adr–Mag
14SpFA4440°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
15SpFA4540°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulation cleavage, veins texture and protomylonitic fabricSrp–Cal–Di–Adr–Mag
16SpFA4640°01’37.0’’ N 16°08’11.9’’ ESchistosity with crenulations cleavage and veins textureSrp–Cal–Di–Adr–Mag
17SpFA2840°01’45.3’’ N 16°0.8’27.3’’ EPseudomorphyc texture–veins texture Patch textureSrp–Cal–Am–Chl–Mag and Ol–Opx–Cpx relict
18SpFA2940°01’45.3’’ N 16°0.8’27.3’’ EPseudomorphyc texture Veins texture Patch textureSrp–Cal–Am–Chl–Mag and Ol–Opx–Cpx relict
19SpFA3040°01’45.3’’ N 16°0.8’27.3’’ EPseudomorphyc texture Veins texture Patch textureSrp–Cal–Am–Chl–Mag and Ol–Opx–Cpx relict
20SpFA1040°02’56,8’’ N 16°09’01.6’’ EPseudomorphyc texture Veins texture Patch textureSrp–Cal–Am–Chl–Mag and Ol–Opx–Cpx relict
Samples from the Pietrapica quarry (SpPP)
21SpPP3140°04’08.6’’ N 16°0.9’19.6’’ EPatch texture, veins textureSrp–Cal ± Am–Chl–Mag
22SpPP3240°04’08.6’’ N 16°0.9’19.6’’ EBrecciated texturesTlc–Qtz–Cal
23SpPP3340°04’08.6’’ N 16°0.9’19.6’’ EPatch texture, veins textureSrp–Cal ± Am–Chl–Mag
24SpPP3440°04’08.6’’ N 16°0.9’19.6’’ EBrecciated texturesTlc–Qtz–Cal
25SpPP3540°04’08.6’’ N 16°0.9’19.6’’ EPatch texture, veins textureSrp–Cal ± Am–Chl–Mag
26SpPP3640°04’08.6’’ N 16°0.9’19.6’’ EBrecciated texturesTlc–Qtz–Cal
Table 2. Chemical analyses of amphiboles in serpentinites and associated veins (v subscript) of Fosso Arcangelo (SpFA39, SpFA5).
Table 2. Chemical analyses of amphiboles in serpentinites and associated veins (v subscript) of Fosso Arcangelo (SpFA39, SpFA5).
Sample CodeSpFA39vSpFA39vSpFA5vSpFA5vSpFA5vSpFA5vSpFA39SpFA39SpFA5vSpFA5v
No. Analysis73767778799198102109130
Oxides (wt %)
SiO254.58857.39255.67452.0453.73557.54751.65755.26355.01557.337
P2O50.031n.d.n.d.0.010.0160.028n.d.0.0050.0240.057
TiO20.1580.0110.0590.4330.2680.0750.4820.0650.059n.d.
Al2O32.7980.4791.5595.2823.5431.3695.4841.8712.509n..d.
Cr2O30.2240.0070.0920.4250.4860.0090.5020.0060.225n.d.
MnO0.1380.0890.0450.0260.1190.0820.030.170.0820.022
FeO3.9253.1472.4753.2313.0642.6632.8717.0652.4012.012
NiO0.1080.090.050.0820.0760.0450.1390.05n.d.n.d.
MgO23.123.41524.40822.60223.23623.52421.8120.99923.62323.445
CaO11.42713.57412.84511.95912.06312.27312.4219.6812.52313.653
Na2O1.1920.090.4591.1871.1440.3581.2961.9950.8130.07
K2O0.002n.d.0.0140.0080.0140.0140.0280.0150.0160.022
Fn.d.0.093n.d.0.0390.023n.d.n.d.n.d.0.0370.032
Cln.d.0.0180.0190.0040.020.0110.0270.0030.0060.01
Sum97.69198.36297.69597.31197.79297.99696.74197.18697.31696.645
Fe3+/ΣFe used1.0000.9701.0001.0001.0000.7981.0001.0001.0000.827
Mn3+/ΣMn used1.0000.0001.0001.0001.0000.0001.0001.0001.0000.000
Final wt %
MnO0.000.090.000.000.000.080.000.000.000.02
Mn2O30.150.000.050.030.130.000.030.190.090.00
FeO0.000.090.000.000.000.540.000.000.000.35
Fe2O3 4.363.392.753.593.412.363.197.852.671.85
H2O+2.142.142.172.052.102.182.052.152.152.18
Sum 100.28100.88100.1599.74100.26100.4299.12100.1499.76
GroupOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,ClOH,F,Cl
Subgroup of (OH,F,Cl)CaCaCaCaCaCaCaCaCaCa
SpeciesTrTrTrTrTrTrMg-Fe2-HblMg-Fe2-HblTrTr
T (ideally 8 apfu)
Si7.4937.8007.6197.2077.3897.8137.2097.6287.5617.907
P0.0020.0000.0000.0010.0010.0020.0000.0000.0010.003
Al0.4530.0770.2510.7920.5740.1850.7910.3040.4060.000
Ti0.0160.0010.0060.0000.0280.0000.0000.0070.0060.000
Fe3+0.0360.1230.1230.0000.0080.0000.0000.0610.0250.090
T subtotal8.0008.0017.9998.0008.0008.0008.0008.0007.9998.000
C (ideally 5 apfu)
Ti0.0000.0000.0000.0450.0000.0080.0510.0000.0000.000
Al0.0000.0000.0000.0700.0000.0340.1110.0000.0000.000
Cr0.0240.0010.0100.0470.0530.0010.0550.0010.0240.000
Mn3+0.0160.0000.0050.0030.0140.0000.0040.0200.0100.000
Fe3+0.4140.2250.1600.3740.3440.2410.3350.7550.2510.102
Ni0.0120.0100.0060.0090.0080.0050.0160.0060.0000.000
Mn2+0.0000.0100.0000.0000.0000.0000.0000.0000.0000.003
Fe2+0.0000.0110.0000.0000.0000.0000.0000.0000.0000.000
Mg4.5344.7444.8194.4524.5814.7114.4294.2194.7154.820
C subtotal5.0005.0015.0005.0005.0005.0005.0015.0015.0004.965
B (ideally 2 apfu)
Mn2+0.0000.0000.0000.0000.0000.0090.0000.0000.0000.000
Fe2+0.0000.0000.0000.0000.0000.0610.0000.0000.0000.000
Mg0.1930.0000.1600.2140.1820.0500.1080.1020.1250.000
Ca1.6811.9771.8401.7751.7771.7851.8571.4321.8442.000
Na0.1260.0230.0000.0110.0410.0940.0350.4670.0310.000
B subtotal2.0002.0002.0002.0002.0001.9992.0002.0012.0002.000
A (from 0 to 1 apfu)
Ca0.0000.0000.0440.0000.0000.0000.0000.0000.0000.017
Na0.1910.0000.1220.3080.2640.0000.3160.0670.1860.019
K0.0000.0000.0020.0010.0020.0020.0050.0030.0030.004
A subtotal0.1910.0000.1680.3090.2660.0020.3210.0700.1890.040
O (non-W)22.00022.00022.00022.00022.00022.00022.00022.00022.00022.000
OH1.9671.9541.9831.8921.9301.9821.8921.9861.9701.984
F0.0000.0400.0000.0170.0100.0000.0000.0000.0160.014
Cl0.0000.0040.0040.0010.0050.0030.0060.0010.0010.002
O0.0330.0020.0120.0900.0560.0150.1010.0140.012n.d.
W subtotal2.0002.0001.9992.0002.0012.0001.9992.0011.9992.000
Sum T,C,B,A15.19115.00215.16715.30915.26615.00115.32215.07215.18815.005
Note: Hbl = hornblende; Tr = tremolite.
Table 3. Chemical analyses of Ti-rich hydro-andradites in a sample (SpFA39) of Fosso Arcangelo serpentinite.
Table 3. Chemical analyses of Ti-rich hydro-andradites in a sample (SpFA39) of Fosso Arcangelo serpentinite.
Sample codeSpFA39SpFA39SpFA39SpFA39SpFA39
N. Analysis818262948
Oxides (wt %)
SiO234.9033.6834.2834.1734.67
TiO23.552.753.13.563.41
Al2O33.902.952.943.653.97
MnO0.060.120.070.09b.l.d.
Fe2O320.2723.4121.4921.3220.88
MgO0.940.410.500.850.64
CaO33.4733.4934.5733.8733.85
Na2O0.010.020.010.010.03
K2On.d.n.d.b.l.d.b.l.d.b.l.d.
H2O*1.451.361.341.371.40
Sum98.5598.1998.3098.8998.85
Structural formula
Si5.6725.5105.6105,5825.650
Ti0.4700.4120.4450.4720.467
Al0.7400.6400.6340.7050.745
Fe2.7702.9882.8992.8312.779
Mn0.0100.0200.0110.0130.000
Mg0.1720.1060.1100.1560.131
Ca5.8785.8986.0025.9505.932
H/40.3490.3290.3190.3340.343
Sum16.06115.90316.03016.04316.047
SpeciesHy-adrHy-adrHy-adrHy-adrHy-adr
Note: n.d. = not-detected; b.l.d. = below limit detection. H2O* determined from stoichiometry. Hy-adr = hydro-andradite.
Table 4. Analyses of clinopyroxene in serpentinite of Fosso Arcangelo (SpFA39, SpFA5).
Table 4. Analyses of clinopyroxene in serpentinite of Fosso Arcangelo (SpFA39, SpFA5).
Sample CodeSpFA39SpFA39SpFA39SpFA39SpFA39SpFA39SpFA39SpFA5SpFA5SpFA5
N.Analysis5101921505153575861
Oxides (wt %)
SiO254.4854.6452.1454.4353.3554.4552.4253.0753.4553.39
P2O50.010.020.03b.l.d.0.010.02n.d.0.020.060.01
TiO20.050.10.110.040.01b.l.d.n.d.0.020.02n.d.
Al2O30.240.461.040.030.960.261.601.530.860.75
Cr2O3n.d.0.050.160.040.01n.d.b.l.d.0.02n.d.0.03
MnO0.050.170.300.250.150.060.120.180.080.04
FeO0.882.023.811.371.261.271.152.301.771.34
NiO0.02n.d.0.01n.d.0.010.040.05n.d.0.050.06
MgO17.3817.2317.0717.7617.6817.0814.9916.5116.9217.51
CaO25.8624.4223.5725.4423.6425.3424.4824.5324.8324.72
Na2O0.010.030.11n.d.0.100.060.050.140.090.07
K2Ob.l.d.0.010.040.020.050.03n.d.0.040.040.01
F0.09n.d.0.01n.d.n.d.n.d.0.010.01n.d.n.d.
Cl0.010.010.04b.l.d.0.020.020.010.070.030.01
Sum99.0299.1698.4199.4197.2498.6294.8998.4398.1997.92
Structural formula
Si1.991.99921.9861.9842.0021.9981.9651.981.987
Ti0.0010.0030.0030.00100000.0010
Al0.010.020.0440.0020.0420.0110.0720.0670.0370.033
Cr0.0000.0020.0040.0010.0000.0000.0000.0000.0000.001
Fe+30.030.0000.0000.00230.0000.0000.0000.0000.0000.000
Fe+20.0040.0620.1150.0190.0390.0390.0370.0710.0550.042
Mn0.0010.0050.0090.0080.0050.0020.0040.0060.0030.001
Mg0.9460.940.9180.9660.980.9360.8520.9110.9840.967
Ca1.0120.9570.9110.9950.9420.99810.9730.9850.981
Na0.0010.0000.0000.0000.0000.0000.0000.0000.0000.000
K0.0000.0010.0020.0000.0020.00000.0020.0020
H0.0000.0000.0000.0000.2480.0000.0000.0000.0000.000
P+50.0000.0000.0000.0000.0000.0000.0000.0000.0010.000
Cl0.0000.0000.0020.0000.0010.0010.0010.0050.0020.001
Ni0.0010.0000.0000.0000.0000.0010.0020.0000.0010.002
F0.010.0000.0040.0000.0000.0000.0010.0020.0000.000
Sum43.988443.9963.9923.9664.00244.005
SpeciesDiDiAug-DiDiAug-DiDiAugAug-DiAug-DiAug-Di
Sample CodeSpFA39SpFA39SpFA39SpFA39SpFA39SpFA39SpFA39SpFA39
N.Analysis62636970789798104
Oxides (wt %)
SiO249.3654.1352.9554.2953.8854.5953.3253.82
P2O50.010.010.01b.l.d.0.050.05n.d.0.08
TiO2b.l.d.0.010.010.020.02n.d.0.050.03
Al2O31.750.470.780.380.390.030.560.14
Cr2O3n.d.0.010.01b.l.d.0.01n.d.n.d.0.01
MnO0.090.120.120.140.120.100.310.27
FeO1.661.851.961.432.851.072.111.95
NiO0.02n.d.0.050.030.010.02n.d.n.d.
MgO16.8816.7216.7017.1016.9917.9917.0616.67
CaO24.0525.2324.9525.5125.4025.6924.3225.69
Na2O0.110.080.060.100.06n.d.0.05n.d.
K2O0.04n.d.0.03n.d.0.010.010.010.01
F0.020.02n.d.n.d.n.d.0.020.030.02
Cl0.020.01b.l.d.0.010.01b.l.d.0.01n.d.
Sum94.0398.6897.6599.0299.79499.5697.8298.69
Structural formula
Si1.9181.9941.9761.9931.9761.9891.9831.989
Ti0.0000.0000.0000.0010.0010.0000.0010.001
Al0.080.0210.0350.0170.0170.0010.0250.006
Fe+20.0540.0570.0610.0440.0870.0320.0660.06
Mn0.0030.0040.0040.0040.0040.0030.010.009
Mg0.9780.9180.9290.9360.9290.9770.9460.918
Ca1.0010.9960.9981.0030.9981.0030.9691.017
K0.0020.0000.0010.0000.0000.0010.0000.001
P+50.0000.0000.0000.0000.0010.0010.0000.001
Cl0.0020.0010.0000.0010.0000.0000.0010.000
Ni0.0010.0000.0010.0010.0000.0010.0000.000
F0.0030.0030.0000.0000.0000.0020.0030.002
Sum4.0423.9954.0063.9984.0144.014.0044.005
SpeciesAugDiAug-DiDiAug-DiDiAug-DiAug-Di
Note: n.d., not-detected; b.l.d., below limit detection. Di = diopside; Aug = augite
Table 5. Analyses of carbonate minerals in veins of Fosso Arcangelo (SpFA39v, SpFA5v) and Pietrapica (SpPP34Av) sites.
Table 5. Analyses of carbonate minerals in veins of Fosso Arcangelo (SpFA39v, SpFA5v) and Pietrapica (SpPP34Av) sites.
Sample CodeSpFA394vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39v
N.Analysis12341112142324313239404154
Oxides (wt %)
P2O5n.d.n.d.n.d.n.d.0.0350.0550.0140.0470.0270.0710.0780.0580.0490.0350.035
TiO2n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.0.058n.d.n.d.0.0530.020.008
Al2O3n.d.n.d.n.d.n.d.0.030.016n.d.n.d.0.0140.01n.d.n.d.0.0020.04n.d.
Cr2O3n.d.n.d.n.d.n.d.n.d.0.0310.076n.d.0.0450.003n.d.0.002n.d.n.d.0.001
MnO0.1140.055n.d.n.d.n.d.n.d.0.018n.d.n.d.0.015n.d.n.d.0.0330.0240.004
FeO1.78n.d.n.d.0.0050.113n.d.n.d.n.d.n.d.n.d.n.d.0.1030.042n.d.n.d.
NiOn.d.n.d.n.d.n.d.n.d.0.095n.d.0.0170.017n.d.n.d.0.0120.0410.0050.032
MgO18.706n.d.n.d.n.d.n.d.n.d.0.004n.d.n.d.n.d.n.d.n.d.n.d.n.d.0.059
CaO25.19656.11857.27755.39455.81356.9857.2754.87656.91356.18862.38255.59855.654.75254.193
SrOn.d.0.2030.1930.218n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.n.d.
Na2On.d.n.d.n.d.n.d.0.015n.d.0.030.0250.0060.018n.d.n.d.n.d.0.0410.013
K2On.d.n.d.n.d.n.d.0.005n.d.0.0320.002n.d.0.0170.009n.d.0.0090.01n.d.
Fn.d.n.d.n.d.n.d.n.d.n.d.n.d.0.03n.d.n.d.n.d.0.0180.0260.0490.003
Cln.d.n.d.n.d.n.d.0.0030.0030.0150.0020.005n.d.0.0050.0020.009n.d.0.013
Sum45.79656.37657.4755.61756.01357.17957.45654.98657.02656.3862.47355.78555.85154.95554.357
CO2 *41,3644,1645,0343,5744.0344.86n.d.43.1744.7144.1548.9843.7543.8243.1742.66
Structural formula
Sr0.0000.010.010.010.010.0000.0000.0000.0000.040.0000.0000.0000.0000.000
Ti0.0000.0000.0000.0000.0000.0000.0000.0000.0000.0010.0000.0000.0010.0010.000
Al0.0000.0000.0000.0000.0010.0010.0000.0000.0000.0000.0000.0000.0000.0020.000
Cr0.0000.0000.0000.0000.0000.0000.0020.0000.0010.0000.0000.0000.0000.0000.000
Fe+20.0000.0000.0000.0000.0030.0000.0000.0000.0000.0000.0000.0030.0010.0000.000
Mn0.0030.020.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.0010.0010.000
Mg0.9880.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.003
Ca0.9561.9931.9951.9941.9841.9941.9951.9991.9971.9831.9981.9941.9911.991.994
K0.0000.0000.0000.0000.0000.0000.0010.0000.0000.0010.0000.0000.0000.0000.000
P+50.0000.0000.0000.0000.0000.0010.0000.0000.0000.0010.0010.0010.0010.0010.001
Cl0.0000.0000.0000.0000.0000.0000.0010.0000.0000.0000.0000.0000.0010.0000.001
Ni0.0000.0000.0000.0000.0000.0020.0000.0000.0000.0000.0000.0000.0010.0000.001
F0.0000.0000.0000.0000.0000.0000.0000.0030.0000.0000.0000.0020.0030.0050.000
C21.9991.9991.9991.99921.99921.9991.9861.9991.9991.9991.9992
Sum44.0014.0014.0024.0013.998443.9994.0123.9993.9993.9993.9993.999
SpeciesDolCalCalCalCalCalCalCalCalCalCalCalCalCalCal
Sample CodeSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA39vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5v
N. Analysis5556596066738587924651657086
Oxides (wt %)
P2O50.0320.060.0570.0650.0520.0760.0540.0540.0110.0780.0650.0340.0670.075
TiO20.0480.0170.014n.d.0.1410.04n.d.n.d.0.0170.0360.0280.024n.d.0.045
Al2O30.012n.d.n.d.n.d.n.d.n.d.0.226n.d.n.d.0.05n.d.0.592n.d.0.021
Cr2O3n.d.0.0030.0950.0330.02n.d.0.019n.d.n.d.n.d.n.d.0.012n.d.0.016
MnOn.d.0.046n.d.0.0130.040.023n.d.0.035n.d.n.d.0.023n.d.0.0260.004
FeO0.014n.d.0.0130.0170.006n.d.0.0170.130.2270.0990.0480.0250.003n.d.
NiOn.d.n.d.0.0050.0390.005n.d.n.d.n.d.n.d.n.d.0.0520.06n.d.0.008
MgOn.d.n.d.n.d.n.d.n.d.n.d.0.1440.1090.0181.265n.d.n.d.n.d.n.d.
CaO54.68654.33254.80454.07453.6455.50255.1252.04744.39454.81357.31455.6356.34456.791
Na2O0.010.0160.030.028n.d.0.012n.d.0.0370.0370.0190.0120.0590.0010.026
K2O0.015n.d.n.d.0.0130.0150.0170.0010.0040.009n.d.0.036n.d.n.d.0.009
F0.0160.0250.028n.d.0.04n.d.0.015n.d.0.016n.d.0.022n.d.0.014n.d.
Cl0.0160.0070.005n.d.0.048n.d.0.005n.d.0.3730.012n.d.0.006n.d.n.d.
Sum54.83854.49355.03854.28253.97955.6755.64252.41645,10257.24157.59156.44156.44956.995
CO2 *43.0342.7243.1642.5242.3843.62n.d.41.1045.01145.1845.1644.0444.2844.68
Structural formula
Ti0.0010.0000.0000.0000.0040.0010.0000.0000.0010.0010.0010.0010.0000.001
Al0.0000.0000.0000.0000.0000.0000.0000.0000.0000.0020.0000.0230.0000.001
Cr0.0000.0000.0030.0010.0010.0000.0000.0000.0000.0000.0000.0000.0000.000
Fe+20.0000.0000.0000.0000.0000.0000.0000.0040.0080.0030.0010.0010.0000.000
Mn0.0000.0010.0000.0000.0010.0010.0000.0010.0000.0000.0010.0000.0010.000
Mg0.0000.0000.0000.0000.0000.0000.0000.0060.0010.0610.0000.0000.0000.000
Ca1.9941.9961.9921.9951.9851.9961.9811.9871.9611.8941.9921.9751.9971.994
K0.0000.0010.0000.0010.0010.0010.0000.0000.0000.0000.0010.0000.0000.000
P+50.0000.0010.0010.0010.0010.0010.0010.0010.0000.0010.0010.0000.0010.001
Cl0.0010.0000.0000.0000.0030.0000.0000.0000.0260.0010.0000.0000.0000.000
Ni0.0000.0000.0000.0010.0000.0000.0000.0000.0000.0000.0010.0020.0000.000
F0.0020.0000.0000.0000.0040.0000.0000.0000.0020.0000.0020.0000.0010.001
C1.9991.9991.9991.9991.9981.9991.996221.9891.9991.9921.9991.999
Sum3.9993.9993.9993.9993.9973.9993.9973.9993.9993.9793.9993.9953.9993.998
SpeciesCalCalCalCalCalCalCalCalCalMg-CalCalCalCalCal
Sample CodeSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpFA5vSpPP34AvSpPP34AvSpPP34AvSpPP34Av
N.Analysis879611512112212512913113213315163741
Oxides (wt %)
P2O50.0680.0160.0220.0530.0410.0170.0550.0270.0440.0470.034n.d.0.0350.007
TiO20.035n.d.0.022n.d.0.04n.d.n.d.0.013n.d.n.d.0.0010.0360.025n.d.
Al2O30.0160.02n.d.0.020.001n.d.0.007n.d.0.014n.d.0.447n.d.0.253n.d.
Cr2O3n.d.n.d.n.d.0.016n.d.0.049n.d.n.d.0.0430.122n.d.n.d.0.0280.009
MnO0.116n.d.0.045n.d.0.041n.d.0.0720.0230.004n.d.0.2230.1670.3550.231
FeO0.1060.0790.018n.d.0.010.010.008n.d.0.0330.032.3033.6852.1662.405
NiOn.d.n.d.0.0520.0440.0410.010.018n.d.n.d.0.0210.031n.d.n.d.0.049
MgO1.264n.d.n.d.0.13n.d.n.d.n.d.n.d.n.d.n.d.19.17819.23219.12419.062
CaO55.55156.64255.90456.89655.96857.2158.5656.41356.83157.20529.59829.65329.79429.685
Na2O0.014n.d.0.0240.0340.0120.019n.d.n.d.0.0050.0260.0370.054n.d.n.d.
K2O0.0130.004n.d.0.006n.d.0.0350.029n.d.0.0290.0130.0140.008n.d.n.d.
F0.012n.d.0.0320.039n.d.0.143n.d.n.d.0.0350.053n.d.n.d.0.0130.076
Cln.d.0.0050.0020.001n.d.0.018n.d.n.d.n.d.0.0060.0050.013n.d.n.d.
Sum57.1956.76556.10857.22356.15457.44758.74956.47657.02357.551.8752.84551.78851.492
CO2 *45.1944.5244.0545.6044.245.3146.0444.30 44.75 45.12 45.9846.7145.9845.94
Structural formula
Ti0.0010.0000.0010.0000.0010.0000.0000.0000.0000.0000.0000.0010.0010.000
Al0.0010.0010.0000.0010.0000.0000.0000.0000.0010.0000.0170.0000.0090.000
Cr0.0000.0000.0000.0000.0000.0010.0000.0000.0010.0030.0000.0000.0010.000
Fe+20.0030.0020.0010.0000.0000.0000.0000.0000.0010.0010.0610.0970.0580.064
Mn0.0030.0000.0010.0000.0010.0000.0020.0010.0000.0000.0060.0040.010.006
Mg0.0610.0000.0000.0340.0000.0000.0000.0000.0000.0000.9080.8990.9070.906
Ca1.9281.9971.9921.9581.9951.9811.9961.9981.9931.9891.0080.9971.0151.014
K0.0010.0000.0000.0000.0000.0010.0010.0000.0010.0010.0000.0000.0000.000
P+50.0010.0000.0000.0010.0010.0000.0000.0000.0010.0010.0010.0000.0000.000
Cl0.0000.0000.0000.0000.0000.0010.0000.0000.0000.0000.0000.0010.0000.000
Ni0.0000.0000.0010.0010.0010.0000.0000.0000.0000.0010.0010.0000.0000.001
F0.0010.0000.0030.0040.0000.0150.0000.0000.0040.0050.0000.0000.0010.008
C1.999221.9991.99921.99921.9991.9991.99421.9962
Sum3.99943.9993.9993.999443.99943.9993.8573.9993.9984
SpeciesMg-CalCalCalCalCalCalCalCalCalCalDolDolDolDol
Note: n.d. = non-detected; CO2 * = calculated from stoichiometry. Cal = calcite; Dol = dolomite.
Table 6. C and O isotope data and calculated fluid isotopic composition for the calcite in serpentinites from the Fosso Arcangelo site.
Table 6. C and O isotope data and calculated fluid isotopic composition for the calcite in serpentinites from the Fosso Arcangelo site.
Sample No.Sample CodeCarbonate Phaseδ13CPDB (‰)δ18OPDB (‰)δ18OSMOW (‰)a)T °C b)T °C c)T °C d)δ13CCO2 fluid (‰) e)δ13CCO2 fluid (‰) f)δ13CCO2 fluid (‰) g)
1SpFA2vCal–Arg2.16–15.0715.37113.47116.99104.09–1.01–1.21–1.76
2SpFA6.3vCal–Arg–0.12–14.9815.47112.68116.00103.36–3.53–3.35–4.08
3SpFA7avCal–Arg0.96–12.1718.3689.5288.5181.88–3.89–3.96–4.42
4SpFA9vCal–Arg–3.66–14.9415.51112.33115.57103.04–7.09–6.91–7.64
5SpFA38vCal–Arg–0.36–14.4116.05107.74109.9298.80–4.06–3.93–4.60
6SpFA39vCal–Arg–3.17–15.4115.02116.49120.79106.88–6.37–6.14–6.92
7SpFA40vCal–Arg–3.11–15.0615.38113.38116.88104.01–6.48–6.29–7.03
8SpFA42vCal–Arg0.22–15.0215.42113.03116.44103.69–3.17–2.98–3.72
9SpFA44vCal–Arg–0.48–14.3216.15106.98108.9998.09–4.22–4.11–4.77
10SpFA45vCal–Arg–2.79–14.9515.50112.42115.68103.12–6.22–6.04–6.76
11SpFA28vCal0.45–14.3216.15106.98108.9998.09–3.29–3.18–3.84
12SpFA29vCal0.32–14.0916.38105.03106.6396.28–3.54–3.44–4.08
13SpFA29.1vCal–0.4–14.3716.09107.40109.5198.48–4.12–4.00–4.66
14SpFA30vCal–0.81–11.3919.1783.5781.8676.35–6.07–6.19–6.59
15SpFA10vCal–Arg1.12–11.9718.5787.9786.7780.45–3.83–3.92–4.36
Average 105.93107.9790.17–4,39–4.30–4.90
Note: n.d., non-detected. aδ18OV-PDB converted in δ18OV-SMOW using the equation: δ18OV-SMOW = 1.03091*δ18OV-PDB + 30.91 from [68]. b Temperature calculated using the equation proposed by [81].c Temperature calculated using the equation proposed by [81]; [82].d Temperature calculated using the equation proposed by [83]. e δ13CCO2 fluid calculated using the equation proposed by [87] considering b. f δ13CCO2 fluid calculated using the equation proposed by [87] considering c. g δ13CCO2 fluid calculated using the equation proposed by [87] considering d.
Table 7. C and O isotope data and calculated fluid isotopic composition for the dolomite in serpentinites from the Pietrapica site.
Table 7. C and O isotope data and calculated fluid isotopic composition for the dolomite in serpentinites from the Pietrapica site.
Sample No.Sample CodeCarbonate Phaseδ13CPDB (‰)δ18OPDB (‰)δ18OSMOW (‰)a)T °C b)T °C c)δ13CCO2 fluid (‰) d)δ13CCO2 fluid (‰) e)
1SpPP33vDol–Mg-Cal–2.69–9.1221.51105.0483.77–6.97–8.40
2SpPP33.1vDol–Mg-Cal–3.37–9.1721.45105.5784.15–7.61–9.05
3SpPP33DvDol–Ank–2.7–9.3421.28107.3985.43–6.83–8.28
4SpPP33D.1vDol–Ank–3.6–9.421.22108.0485.88–7.69–9.15
5SpPP35vDol–Cal–1.79–9.4921.13109.0286.57–5.83–7.29
6SpPP35.1vDol–Cal–2.26–9.5721.04109.9087.18–6.24–7.72
Average 107.4985.50–6.86–8.32
Note: n.d = non-detected. aδ18OV-PDB converted in δ18OV-SMOW using the equation: δ18OV-SMOW = 1.03091*δ18OV-PDB + 30.91 from [68]. b Temperature calculated using the equation proposed by [84]. c Temperature calculated using the equation proposed by [85]. d δ13CCO2 fluid calculated using the equation proposed by [85] considering b. e δ13CCO2 fluid calculated using the equation proposed by [85] considering c.
Table 8. Microthermometric data of fluid inclusions in quartz crystals in a quartz-carbonate vein from the Pietrapica quarry (SpPP).
Table 8. Microthermometric data of fluid inclusions in quartz crystals in a quartz-carbonate vein from the Pietrapica quarry (SpPP).
IFFIAOriginTh (°C)Tm(Ice) (°C)Tn (°C)Size (µm)L:VSalinity (NaCl Mass %)
SpPP36.1_1vn.d.P198–200–0.8–301880/201.4
SpPP36.1_2v1S121–124+–33895/5n.d.
SpPP36.1_3v1S125–130–0.832695/51.4
SpPP36.1_4v2Pn.d.–0.6–365090/101.05
SpPP36.1_5v2S120–125+–40795/5n.d.
SpPP36.1_6vn.d.S185–188–0.3–34490/100.53
SpPP36.1A_7vn.d.n.d.n.d.+–30n.d.n.d.n.d.
SpPP36.1A_8vn.d.n.d.112–125+–30895/5n.d.
SpPP36.1A_10vn.d.n.d.n.d.+–284085/15n.d.
SpPP36.1A_11vn.d.n.d.190–196–0.8–291090/101.4
SpPP36.1A_1vn.d.n.d.n.d.+–29695/5n.d.
SpPP34.1_13v3P>200(–1.9) (–1.6)–301770/303.23–2.74
SpPP34.1_14v3P138–140+–31.3795/5n.d.
SpPP34.1_15v4S118–120+–32.5495/5n.d.
SpPP34.1_16v5P>200–1.9–32690/53.23
SpPP34.1_17v6S>200–1.8n.d.490/103.06
SpPP34.1_18v6S118–120–0.8–311095/51.4
SpPP34.1_19v7Sn.d.+–29.7795/5n.d.
SpPP34.1_20vn.d.P105–110–0.7–30.81190/101.23
SpPP34.1_21v8Sn.d.+–34.8695/5n.d.
SpPP34.4_22v9Pn.d.(–0.9) (–0.7)–311170/301.57–1.23
SpPP34.4_23v9Pn.d.–0.9–301090/101.57
SpPP34.4_24v8Sn.d. –2810n.d.n.d.
SpPP34.4_25v8Sn.d.–0.3–29595/50.53
SpPP34.4_26v10Pn.d.+–30.8795/5n.d.
SpPP34.4_27v10P280–288(–0.7) (–0.5)–33.6890/101.23–0.88
SpPP34.4_28v11S>200–0.7–34.2590/101.23
SpPP34.4_29v11S120.8+–31.8495/5n.d.
SpPP34.4_30v12S210–216–0.7–34.2790/101.23
SpPP34.4_31v13S288–293–0.4–34.8550/500.71
SpPP34.4_32v14P93–0.6–34.2890/101.05
SpPP34.4_33vn.d.P120–130n.d.n.d.595/5n.d.
SpPP34.4_34vn.d.P330n.d.n.d.710/90n.d.
SpPP34.4_35vn.d.P225–235n.d.n.d.450/50n.d.
SpPP34.4_36vn.d.P>395n.d.n.d.n.d.10/90n.d.
SpPP34.4_37vn.d.P335n.d.n.d.n.d.10/90n.d.
Note: n.d. = not-detected; + = positive value; FIA = fluid inclusions association; Th (°C) = homogenization temperatures; Tm(Ice) (°C) = temperatures of final ice-melting; Tn (°C) = nucleation temperature; L:V = liquid vapor ratio; Salinity = wt. % NaCl calculated from Tm(Ice).

Share and Cite

MDPI and ACS Style

Rizzo, G.; Dichicco, M.C.; Castiñeiras, P.; Grassa, F.; Laurita, S.; Paternoster, M.; Sinisi, R.; Mongelli, G. An Integrated Study of the Serpentinite-Hosted Hydrothermal System in the Pollino Massif (Southern Apennines, Italy). Minerals 2020, 10, 127. https://doi.org/10.3390/min10020127

AMA Style

Rizzo G, Dichicco MC, Castiñeiras P, Grassa F, Laurita S, Paternoster M, Sinisi R, Mongelli G. An Integrated Study of the Serpentinite-Hosted Hydrothermal System in the Pollino Massif (Southern Apennines, Italy). Minerals. 2020; 10(2):127. https://doi.org/10.3390/min10020127

Chicago/Turabian Style

Rizzo, Giovanna, Maria Carmela Dichicco, Pedro Castiñeiras, Fausto Grassa, Salvatore Laurita, Michele Paternoster, Rosa Sinisi, and Giovanni Mongelli. 2020. "An Integrated Study of the Serpentinite-Hosted Hydrothermal System in the Pollino Massif (Southern Apennines, Italy)" Minerals 10, no. 2: 127. https://doi.org/10.3390/min10020127

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop