SlideShare a Scribd company logo
1 of 18
Download to read offline
A buried volcano in the Calabrian Arc (Italy) revealed
by high‐resolution aeromagnetic data
R. De Ritis,1
R. Dominici,2
G. Ventura,1
I. Nicolosi,1
M. Chiappini,1
F. Speranza,1
R. De Rosa,2
P. Donato,2
and M. Sonnino2
Received 26 November 2009; revised 8 June 2010; accepted 18 June 2010; published 4 November 2010.
[1] Aeromagnetic data collected between the Aeolian volcanoes (southern Tyrrhenian
Sea) and the Calabrian Arc (Italy) highlight a WNW‐ESE elongated positive magnetic
anomaly centered on the Capo Vaticano morphological ridge (Tyrrhenian coast of
Calabria), characterized by an apical, subcircular, flat surface. Results of forward
and inverse modeling of the magnetic data show a 20 km long and 3–5 km wide
magnetized body that extends from sea floor to about 3 km below sea level. The magnetic
properties of this body are consistent with those of the medium to highly evolved
volcanic rocks of the Aeolian Arc (i.e., dacites and rhyolites). In the Calabria mainland,
widespread dacitic to rhyolitic pumices with calc‐alkaline affinity of Pleistocene age
(1–0.7 Ma) are exposed. The tephra falls are related to explosive activity and show a
decreasing thickness from the Capo Vaticano area southeastward. The presence of lithics
indicates a provenance from a source located not far from Capo Vaticano. The combined
interpretation of the magnetic and available geological data reveal that (1) the Capo
Vaticano WNW‐ESE elongated positive magnetic anomaly is due to the occurrence
of a WNW‐ESE elongated sill; (2) such a sill represents the remnant of the plumbing
system of a Pleistocene volcano that erupted explosively producing the pumice tephra
exposed in Calabria; and (3) the volcanism is consistent with the Aeolian products,
in terms of age, magnetic signature, and geochemical affinity of the erupted products,.
The results indicate that such volcanism developed along seismically active faults
transversal to the general trend of the Aeolian Arc and Calabria block, in an area where
uplift is maximized (∟4 mm/yr). Such uplift could also be responsible for fragmentation
of the upper crust and formation of transversal faults along which seismic activity and
volcanism occur.
Citation: De Ritis, R., R. Dominici, G. Ventura, I. Nicolosi, M. Chiappini, F. Speranza, R. De Rosa, P. Donato, and M. Sonnino
(2010), A buried volcano in the Calabrian Arc (Italy) revealed by high‐resolution aeromagnetic data, J. Geophys. Res., 115,
B11101, doi:10.1029/2009JB007171.
1. Introduction
[2] Aeromagnetic surveying is a relatively new technique for
studying volcanic or active tectonic areas through improve-
ments in data acquisition, GPS measurements, and data‐
processing methods [e.g., Finn et al., 2001; Lenat et al., 2001;
Chiappini et al., 2002; De Ritis et al., 2005]. In volcanic
areas, aeromagnetic surveys allow researchers to decipher
the inner structure of eruptive centers and, in marine environ-
ments, to detect submerged, hidden volcanoes and vents
[Rollin et al., 2000; Blanco‐Montenegro et al., 2007; De
Ritis et al., 2007]. Due to the occurrence of widespread vol-
canic deposits of uncertain origin in southern Calabria (Italy),
particularly in the Capo Vaticano (CV) promontory [De Rosa
et al., 2001, 2008], a recent high‐resolution aeromagnetic
survey has been conducted by the Airborne Geophysics Sci-
ence Team of Istituto Nazionale di Geofisica e Vulcano-
logia (INGV), covering the offshore and onshore areas
between CV and the calc‐alkaline to shoshonitic Stromboli
and Panarea volcanoes of the Aeolian Islands (Figures 1
and 2). Results of the aeromagnetic survey reveal a 25 km
WNW‐ESE elongated magnetic anomaly covering the CV
on land and in offshore areas. This anomaly is noteworthy
to be investigated because of (1) the vicinity of the Aeolian
volcanic Arc, (2) the lack of high‐susceptibility geologic
units outcropping in the CV area, and (3) the presence of
volcanic products of unknown origin outcropping in western
Calabria and mainly on CV. A geological model has been
built on the basis of inverse and forward magnetic models
together with the geotectonic and seismic knowledge of the
area. Inversion of magnetic data set was carried out to define
a 3‐D anomalous susceptibility contrast model. To further
constrain the obtained model, a 2.75‐D forward modeling
was also made that included the lithological and structural
1
Istituto Nazionale di Geofisica e Vulcanologia, Rome, Italy.
2
Department of Earth Science, UniversitĂ  della Calabria, Arcavacata di
Rende, Italy.
Copyright 2010 by the American Geophysical Union.
0148‐0227/10/2009JB007171
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B11101, doi:10.1029/2009JB007171, 2010
B11101 1 of 18
information of the study area. Rock magnetic properties
have been measured on samples representative of the main
local lithological units. The results highlight the presence of
a previously unknown volcanic body emplaced along the
main fault located at the southern edge of the CV prom-
ontory. This body could represent the source area of the
widespread volcanic deposits exposed in CV and western
Calabria. The results, discussed in light of the available
geological and geophysical information, shed new light on
(1) the tectonics of the Calabrian Arc and (2) the general
control of preexisting structures and, in particular, of trans-
versal structures in subduction setting on the emplacement
of volcanoes.
2. Geodynamic and Geological Setting
2.1. Structure
[3] The Calabrian Arc, Sicilian Maghrebides, and Apen-
nines are mountain belts bounding the southern and eastern
sector of the Tyrrhenian Sea back‐arc basin (Figure 1). The
evolution of the Calabrian Arc and Tyrrhenian Sea back arc
during Neogene and Quaternary periods was driven by the
southeastward retreat of the Ionian slab [Malinverno and
Ryan, 1986; Jolivet and Faccenna, 2000; Faccenna et al.,
2001]. The Calabrian Arc is presently located on the top
of a 200 km wide, subducting slab dipping 70° toward NE
(Figure 1) [Spakman et al., 1993; Chiarabba et al., 2008;
Neri et al., 2009]. Earthquake distribution and seismic tomog-
raphy indicate that the northeastern boundary of the sub-
ducting slab roughly corresponds on the surface to the
boundary between southern Apennines and Calabrian Arc
[Chiarabba et al., 2008, and references therein]. The western
boundary of the slab occurs in correspondence of the NNW‐
SSE‐striking Tindari‐Letojanni fault system, which crosses
the eastern margin of Sicily (Figure 1) [Barberi et al., 1994;
De Astis et al., 2003]. This fault system constitutes the
northern extent of the Malta Escarpment, which represents a
transtensional (right‐lateral) transfer zone between the con-
tinental crust of eastern Sicily (west) and the oceanic Ionian
basin (east) [Doglioni et al., 2001, and references therein].
The Malta Escarpment transfer zone, along which the Mt.
Etna volcano is emplaced, is characterized by vertical slip‐
rates of up to 2 mm/yr [Azzaro et al., 2000]; its movement is
related to the larger retreat of the subduction hinge in the
Ionian Sea with respect to the Sicily mainland.
[4] The Aeolian Arc, which forms a ring‐like chain of
volcanic islands and seamounts in the southern margin of
the southern Tyrrhenian Sea oceanic basin (Marsili basin)
[Nicolosi et al., 2006], formed during the last 1 Myr, and
the volcanoes of Lipari, Vulcano, Panarea, and Stromboli
Figure 1. Tectonic scheme of Italy (modified from Chiarabba et al. [2008]; Rosenbaum et al. [2008]).
Dashed lines with numbers identify the depth of the slab below the Calabrian Arc as deduced from
seismic data. TL, the Tindari‐Letojianni fault system; ME, the Malta Escarpment. Lines with triangles
indicate reverse faults; lines with perpendicular marks indicate normal faults.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
2 of 18
are still active today (Figure 2a). The evolution of the
Aeolian volcanism, which includes calc‐alkaline to shosh-
onitic and K‐alkaline products, has been related to both
the subduction of the Ionian lithosphere [Chiarabba et al.,
2008] and a postcollisional, rift‐type volcanism associated
with the extensional tectonics that have affected the Calab-
rian Arc in last 1–0.7 Myr [De Astis et al., 2003, and refer-
ences therein]. Recently, Rosenbaum et al. [2008] also
proposed that slab tear faults, which reflect the segmentation
of the Apennines–Calabrian Arc–Tyrrhenian Sea subduction,
control the emplacement of some Quaternary volcanoes and
identify, in the Aeolian Islands–Calabrian Arc region, the
Tindari‐Letojanni fault system as one of these slab tear faults
(Figure 1). The internal structure and recent tectonics of the
Calabria block is debated, and two main end‐member models
have been proposed in the last years: (1) Calabria is frag-
mented into several blocks undergoing differential displace-
ments toward the trench and separated by NW‐SE striking
faults [Knott and Turco, 1991; Van Dijk, 1994; Guarnieri,
2006; Tansi et al., 2007; Del Ben et al., 2008]; (2) alterna-
tively, the Calabria block is solely characterized by exten-
sional faulting, as syn‐sedimentary extensional features are
documented in the whole middle to upper Miocene to Pleis-
tocene sedimentary succession [Mattei et al., 1999, 2002;
Monaco and Tortorici, 2000; Cifelli et al., 2007a, 2007b].
[5] The Calabrian Fore‐arc basin (CFB), located in the
central‐southern sector of the Calabrian Arc, is composed
by several NE‐SW elongated basins fed by sediments
derived from Serre, Poro, and Aspromonte crystalline Mas-
sifs (Figure 2b). The geometry of these basins is controlled
by late Pliocene–Holocene NE‐SW and minor NW‐SE
striking faults. CFB consists of different allochthonous struc-
tural units (Figure 2b). A continuous crustal section from
granulite grade through granitoids and amphibolite‐facies
gneisses into unmetamorphosed Paleozoic and younger
sediments crop out in the Serre Massif [Schenk, 1984, 1990;
Amodio Morelli et al., 1976]. The granulite‐facies is over-
lain by gneiss and paragneiss with granofels and siliceous
marble (Polia‐Copanello unit). In the middle sector of the
Serre and Poro Massifs, granitoids crop out over an area of
about 1250 km2
. The granitoids overlie the Polia‐Copanello
unit and intrude into the overlying ortho‐ and paragneisses
(Aspromonte unit) and lower‐grade Paleozoic rocks (Stilo
unit) [Schenk, 1984; Ayuso et al., 1994; Graessner and Schenk,
2001; Caggianelli and Prosser, 2002]. Along the south-
eastern side of the Serre Massif, the granites come into
contact with Mammola and Pazzano complexes [Colonna
et al., 1973], which are made up by micaschists and para-
gneisses with interbedded Ordovician tholeiitic metabasalts
[Acquafredda et al., 1994]. The Pazzano complex consists
of phyllites, with interbedded graywackes and marbles
[Bouillin et al., 1984, 1987; Acquafredda et al., 1987, 1989].
A Mesozoic sequence made up of dolostone, calcirudite and
calcarenite, breccia, and reef limestone [Roda, 1965; Bonardi
et al., 1984] lies on the Stilo–Pazzano and Mammola com-
plexes. Along the Ionian coast of Calabria, a succession
of latest Chattian–Quaternary age rocks [Patterson et al.,
1995] overlies the Stilo unit and Mesozoic carbonates. The
Stilo–Capo d’Orlando formation is composed of red con-
glomerate, sandstone, and pelites [Cavazza, 1989; Cavazza
and DeCelles, 1993]. The unit is conformably overlain by
a chaotic melange of pelitic matrix that encloses olistoliths
of calciturbidites and Chattian–early Miocene quartzarenitic
turbidites [Cavazza et al., 1997; Bonardi et al., 2001].
[6] The Serravallian–Tortonian sedimentary sequence con-
sists of continental conglomerate, marine sandstone, marl,
and shale, as well as minor shales, Messinian limestone,
and gypsum. This unit is overlain by alluvial conglomerate and
discontinuous shallow‐marine to continental sandstone and
pelite. The upper portion of the basin filling is represented by
marls of the Trubi formation overlain by shallow‐marine and
transitional‐continental facies of late Pliocene–Quaternary age.
Figure 2. (a) Location of the Capo Vaticano ridge. (b) Geological map of southern Calabria showing
outcrops of the volcaniclastic deposits (stars) and WNW‐ESE geological section (profile A‐B). The
bathymetry of the Capo Vaticano ridge is reported along with offshore active faults [data are from
Cifelli et al., 2007a; De Rosa et al., 2008; Tortorici et al., 2003; and Milia et al., 2009). (c) Southern
Calabria and northeastern Sicily magnetic anomaly field sketch map from Caratori Tontini et al. [2004].
The black dotted line is the geological profile shown in Figure 2b.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
3 of 18
Figure
2.
(continued)
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
4 of 18
[7] Several orders of marine terraces distributed from
about 50 up to 1,000 m above sea level (asl) indicate an
uplift of the southern and western sectors of Calabria from
lower to middle Pleistocene times [Ghisetti, 1981; Dumas et
al., 1982; Mulargia et al., 1984; Westaway, 1993; Miyauchi
et al., 1994; Tortorici et al., 1995; Antonioli et al., 2006].
The estimated uplift velocity is generally between 1.0 and
2.1 mm/yr in the last 124 kyr. In the CV area, however, the
uplift velocity has reached 4.0 mm/yr [Westaway, 1993;
Miyauchi et al., 1994]. Westaway [1993] estimated an uplift
velocity of the Calabria area of 1 mm/yr for the last 0.9 Myr.
Normal, active faults trending WNW or NE affect the onshore
Figure 2. (continued)
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
5 of 18
and offshore sectors of the CV promontory. A major WNW‐
ESE fault shows late Quaternary slip‐rates of up to 2.5 mm/yr
and is responsible for the northeastward tilting of the CV
promontory (Figure 2b) [Tortorici et al., 2003]. This fault
also controls the morphology of the CV southern coastline
and its offshore sector. The CV offshore is characterized by
a WNW‐ESE elongated ridge with a flat, subcircular trun-
cated top at 200 m below sea level (bsl) (Figure 2b). A
seismic profile crossing the northern flank of the ridge
evidences a chaotic basement covered by a thin succession
of sediments younger than 0.7 Myr [Milia et al., 2009].
Tomographic studies [Barberi et al., 2004] reveal the
occurrence of a subcircular body located between 0 and
5 km depth with Vp/Vs values (1.8) higher than those of the
surrounding areas (Vp/Vs < 1.6–1.7). This body is centered
on the CV morphological ridge.
2.2. Volcaniclastic Sedimentation in the Calabria
Fore‐Arc Basins
[8] The CFB and CV sedimentary evolution is charac-
terized by an important middle Pleistocene volcanic and
volcaniclastic input. Volcaniclastic deposits consist of iso-
lated pumice swarms, ash layers, and thick pumiceous lapilli
sequences occurring as distinct lithofacies within sedimen-
tary successions of shallow marine environments [De Rosa
et al., 2001]. According to De Rosa et al. [2008], these
sequences have the characteristics of proximal tephra fall
deposits from Plinian columns reworked in shore environ-
ments. In the Mesima basin and Gioaia Tauro plain, the
deposits have reached 6 and 21 m in thickness, respectively
(Figure 2b). In the CV area, the occurrence of impact sags in
the volcanic deposits attests the proximity of the vent area.
The compositional homogeneity of fragments (calc‐alkaline
dacites and rhyolites) and the lack of nonvolcanic detritus,
soils, or organic matter indicate that the deposits were
quickly reworked after their primary emplacements [De
Rosa et al., 2001, 2008]. K‐Ar determinations of the glass
from the pumices give ages in the range 0.67–1.07 Myr
[Cornette et al., 1987]. The mineralogical assemblage and
the trace element geochemistry of the pumices are consistent
with a provenance from magmatic‐arc volcanism [De Rosa
et al., 2008]. The decrease in the maximum size of pumice
from Mesima basin‐CV area to the Gioia Tauro plain and
Reggio Calabria basin (Figure 2b) and the presence of lithics
of lavas suggest a provenance from a source located off-
shore, not far from the present CV promontory [De Rosa
et al., 2001, 2008]. De Rosa et al. [2008] propose two
hypotheses for the source areas of the pumice deposits of
CFB and CV:
[9] 1. A provenance from Panarea Island (Aeolian Arc)
during an early highly explosive phase, leading to the col-
lapse of the summit of the first edifice. This hypothesis
remains to be confirmed and does not agree with the lack of
rhyodacitic, pumiceous deposits on other Aeolian terrains
with age >100 kyr. In addition, the occurrence of impact
sags in the CV deposits [De Rosa et al., 2001] is not con-
sistent with a source located at a distance of about 70 km
(i.e., the distance between Panarea and the CV promontory).
[10] 2. The tephra falls are related to the explosive activity
of a “buried” calc‐alkaline volcano located in the Tyr-
rhenian Sea. According to De Rosa et al. [2001, 2008], the
CV pumice deposits could represent the only evidence of a
large explosive eruption in the southern Tyrrhenian Sea, the
very proximal equivalent of which is not preserved.
3. Regional Magnetic Anomaly Field
[11] To get a reliable interpretation of the CV magnetic
anomaly system, it is necessary to arrange the study area in
the frame of the regional knowledge described in section 2.
Geologic and magnetic field maps [Agip, 1981; Chiappini
et al., 2000; Caratori Tontini et al., 2004; Finetti, 2005]
provide a general outline of the geostructural characteristics
and long‐wavelength magnetic anomalies (Figure 2c) of
southern Calabria and the southern Tyrrhenian Sea. Fifty
kilometers west of the CV, the Aeolian Arc (Panarea Island
and Lamentini seamount; Figure 2a) is set up by strongly
magnetic Holocene–Pleistocene volcanics [Zanella, 1995]
overlying the barely magnetic units of the Calabrian Arc
[Barberi et al., 1994]. Consistently, this western region is
characterized by high‐intensity positive magnetic anoma-
lies corresponding to the volcanic islands and seamounts
[Chiappini et al., 2000; De Ritis et al., 2005]. Eastward,
the Calabride units [Schenk, 1990], characterized by low to
null magnetic anomaly values, are exposed in the CV and
Le Serre horsts and in the Mesima graben (Figure 2b).
Westward, in the CV offshore sector, the Calabrian units are
intensely faulted and lowered by normal faults, which have
produced the nearby Paola and Gioia Tauro sedimentary
basins. These depressions are filled with as much as 5,000 m
of middle Miocene to Pleistocene sedimentary sequences
that can be considered of low susceptibility, according to
measurements carried out on similar sediments exposed
elsewhere in Italy [Sagnotti et al., 1998]. In fact, in this sector,
marine terrigenous series above the Calabrian metamorphic
units have been interpreted through seismic profiles and dril-
lings starting from the Serravallian age [Milia et al., 2009]. A
wide region characterized by low to negative values of the
magnetic anomaly field exists as a result of this regional
setting (Figure 2c). It extends from the Stromboli canyon to
the Ionian Sea (E‐W) and from the Messina Strait to the
Catanzaro plain (N‐S).
[12] Toward North and Northeast, two long‐wavelength,
low‐intensity linear positive anomalies develop parallel to the
main regional tectonic features of the Catena Costiera and
Sila (Figure 2c) [Van Dijk and Okkes, 1991]. The anomalies
express the presence of contrasts characterized by very low
magnetization values of the Calabrian Arc units.
[13] Southward, the Etna long‐wavelength magnetic anom-
aly negative lobe contributes to lowering of the regional
magnetic anomalies field (Figure 2c). Negative lobes of other
high‐intensity magnetic anomalies lying in the Stromboli
canyon affect the CV magnetic anomaly system. By means
of synthetic models, we estimated the total entity of this
magnetic anomaly lowering to be in the range of 15–20 nT.
[14] The CV high‐intensity positive magnetic anomaly
(Figures 2c, 3a, and 3b) lies in correspondence of the CV
morphological ridge (Figure 2b), highlighting the presence
of a noticeable magnetization contrast in an area strongly
affected by tectonic structures.
[15] Because of the lack of susceptibility measurements
available in the literature for the Calabrian Arc, we carried
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
6 of 18
out sampling of the main lithological units of the area. Then,
we measured rock sample susceptibility in the paleomag-
netic laboratory of INGV (Rome) to get reliable modeling
and effective interpretation of magnetic anomalies. In addi-
tion, because crustal magnetic anomalies are originated by
3‐D variations of rock properties at depth, we reconstructed
the stratigraphic sequence reported in the geological pro-
file of Figure 2b and sampled the most representative and
widespread lithologies of the region. Southern Calabrian
structures, from the Ionian to the Tyrrhenian sides (NW‐SE
direction), are composed by conglomerates, sandstones, and
clays of the Calabride units (Tortonian to lower Pliocene),
granites, granodiorites, and tonalities of the Aspromonte
Peloritani units (Permian–Carboniferous), and terrigenous
marine deposits and calcarenites lower Messinian to lower
Pleistocene in age. The CV promontory, in particular, is
made up by intrusive igneous rocks and terrigenous marine
deposits (Figure 2b). Therefore, it is reasonable to infer that
these units are the representative bedrock of the offshore
study area, possibly lowered by the action of faults and
covered by middle Miocene–Holocene sediments [Milia
et al., 2009; Mattei et al., 2002].
[16] Mass susceptibilities of rock samples were measured
at the INGV paleomagnetic laboratory of Rome. Subse-
quently, volume susceptibilities were calculated by consid-
ering average rock density values available in the literature
(Table 1). Volume susceptibilities were determined for a
total of 24 samples including 11 sedimentary samples of
sandstone, sand, carbonate, and clay successions belonging
to terrigenous marine deposits and marine terraces; 9 sam-
ples of the metamorphic and plutonic complex such as the
granites and phyllades of Le Serreand Mesima units; 3 vol-
canic samples set up by pumices with different granulo-
metries deriving from outcrops lying in the CV area, in the
Gioia Tauro basin, and in the Mesima graben (Figure 2b);
and 1 sample of metagabbro from an Oligocene shear zone
separating nappes of the Hercinian basement [Langone
et al., 2006].
[17] The sedimentary and metamorphic/plutonic speci-
mens show very low volume susceptibility values (10−6
up
Figure 3. (a) Total intensity magnetic anomalies of the Capo Vaticano (CV) and offshore areas. Dotted
lines mark the Calabria coast line. Inset in Figure 3a represents the aeromagnetic survey flight lines
(black) and the tie lines (red). (b) Reduced‐to‐the‐pole magnetic anomaly field assuming geomagnetic
inclination of 90° and declination of 0°; the dotted red lines refer to the AB and CD magnetic profiles
reported in Figures 5a and 5b. The main anomalies are highlighted with numbers from 1 to 5.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
7 of 18
to 10−4
SI) (Table 1). The volcanic samples also exhibit
low susceptibility (10−4
SI), probably because of their high
silica content and clastic fabric. Our measurements are
consistent with those summarized in Barberi et al. [1994],
who report susceptibilities of the Calabrian Arc rocks less
than 7 × 10−7
SI. An isolated exception is our metagabbro
sample, which has a susceptibility value as high as 10−2
SI.
However, this lithology forms small (decimetric to metric)
lenses corresponding to a low‐intensity, long‐wavelength
magnetic anomaly developing 15 km northwest of CV. We
have not measured the magnetic remanence of the Calabrian
Arc rocks because the rock susceptibility values are gener-
ally very low, implying a very poor content of magnetite and
thus negligible magnetic remanence. In addition, the rocks
are Meso‐Cenozoic in age and are arranged in a chaotic
tectonic setting with the result that their polarity will be
mixed, and the reconstruction of the rock volumes with
constant magnetic polarity it is not possible. The measure of
remanence of these rocks has no practical bearings. More-
over, in situ samples of the CV volcanics are also lacking
and the on‐land outcropping pumices, which have been
reworked in shallow marine environments, show SI values
in the order of 10−4
(see Table 1), suggesting low magnetic
attitude. In conclusion, the CV positive magnetic anomaly
system represents a clear contrast between an unknown
magnetic source and the virtually nonmagnetic rocks of the
Calabrian Arc.
4. Data Acquisition and Processing
[18] A high‐resolution aeromagnetic survey was recently
carried out by the Airborne Geophysics Science Team of
INGV for a detailed characterization of the offshore–onshore
CV anomaly field. Thirty‐four measurement profiles were
flown in the N‐S direction across the CV morphological
ridge with a line spacing of about 2 km apart (Figure 3a,
inset). Tie lines were also flown in the E‐W direction to
level the survey lines. The survey area extends from the
Mesima graben to the Panarea and Stromboli eastern marine
areas. In the offshore sectors, the survey was flown at con-
stant barometric height (500 m), whereas the profiles cross-
ing the CV promontory were draped, reaching a maximum
flight altitude of 1000 m. Total intensity anomaly data were
acquired with an optically pumped cesium magnetometer at
the sampling rate of 10 Hz. A GPS receiver in differential
Figure 3. (continued)
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
8 of 18
mode was used for accurate survey navigation, and a laser
altimeter recorded the flight altitude.
[19] We performed the reduction of the aeromagnetic data
by removing the effects of the time‐varying external fields
and computed the anomalies obtained by removing the
Geomagnetic Reference Field [IAGA, 2005]. To produce the
total intensity magnetic anomaly map, the data sets were
merged and gridded at the spacing of 1,000 m by using
a minimum curvature algorithm. Finally, the data set was
microleveled to remove a few residual errors from the
gridded data; the field obtained is shown in Figure 3a.
[20] To facilitate the forward modeling and its geological
interpretation (see section 5.2), the total field anomalies
have been reduced‐to‐pole (RTP) [Baranov and Naudy,
1964]. This analytical transformation sets the magnetic
anomalies above their relative sources and transforms each
dipolar anomaly into positive or negative. This has the
advantage of considerably simplifying the magnetic mod-
eling because the magnetic contacts in the model would
thus not be substantially far from the real position. As a
consequence, the relation between anomaly and source in
the map is more evident. Using the RTP algorithm, we have
assumed that the ambient magnetic field and the magneti-
zation have constant directions along the studied area. The
smaller the area, the more valid is this assumption. For the
70 × 55 km coverage of this survey, we are confident that
this assumption holds. The RTP magnetic anomaly map is
shown in Figure 3b. A WNW‐ESE elongated positive
anomaly characterizes the CV area (anomaly 1 in Figure 3).
However, a negative anomaly also appears south of the CV
anomaly 1 in the RTP map (Figure 3b). Because this neg-
ative feature overlaps the CV WNW‐trending fault (see
Figure 2), we investigated its characteristics and meaning
(see section 5.2).
5. Magnetic Modeling
[21] The final goal of potential field surveys is to get
quantitative information about the causative sources and to
interpret them in the frame of the geological knowledge of
the area. Here, we have chosen to carry out both inverse and
direct approaches. To build a reliable 3‐D geological model
of the subsurface CV structure, aeromagnetic data have the
fundamental role of providing a quantitative constraint on
the distribution of the susceptibility rock properties on a
wide offshore area that is otherwise not directly accessible.
Table 1. Summary of Rock Magnetic Properties Measured in Calabria in the Areas of Capo Vaticano, Mesima Graben, and Le Serre
Chain Along With the Sample Description and Geographic Position
Sample Latitude Longitude Lithology Description Location
Volume
Susceptibility
(SI)
Sedimentarian
Units
1 38°33′24″ 16°07′37″ Sandstone Pleistocenic seabord coarse
silicocalstic sandstone
Mesima graben,
Marepotamo valley
−9.51 × 10−6
2 38°39′16″ 15°50′46″ Sandstone Tortonian medium to coarse
silicocalstic sandstone
Mt. Poro‐Capo Vaticano 1.22 × 10−5
3 38°33′27″ 16°07′42″ Sand Fine sand Mesima graben,
Marepotamo valley
1.20 × 10−4
4 38°38′16″ 15°54′00″ Carbonates Messinian carbonates and
marly carbonates
Mt. Poro ‐ Spilinga 8.79 × 10−5
5 38°23′19″ 16°23′52″ Sandstone Medium to lower Miocene
banded sandstone, turbidites sequence
Westward Caulonia village 1.54 × 10−4
6 38°33′29″ 16°06′42″ Clays Marly and silty clay succession Mesima basin 4.37 × 10−5
7 38°23′12″ 16°23′59″ Sandstone Lower Miocene silicocalstic sandstone Westward Caulonia village 5.76 × 10−5
8 38°22′29″ 16°24′58″ Sand Lower Pliocene sand Le Serre chain, western side,
Caulonia village
7.71 × 10−5
10 38°22′27″ 16°25′48″ Clays “Varicolori” clay; red, green, and
gray clay succession
Le Serre chain, western side,
Caulonia village
1.39 × 10−4
23 38°39′27″ 15°56′32″ Sands Quaternari marine terraces sands Spilunga‐Zungari 1.54 × 10−5
19 38°22′27″ 16°25′48″ Clays Varicolori clays: numidic olistonites Caulonia 9.97 × 10−5
Metamorphic and
Plutonic Units
9 38°25′16″ 16°22′31″ Phyllade Metamorphic complex
(from metarenites to paragneiss)
F.va Allaro 1.33 × 10−4
11 38°37′06″ 15°48′44″ Dark granite Melanocratic facies granite Capo Vaticano 1.34 × 10−4
12 38°37′06″ 15°48′44″ Bright granite Leucocratic facies granite Capo Vaticano 1.9 × 10−4
13 38°31′15″ 16°08′28″ Granite Le Serre chain, western side 1.04 × 10−4
14 38°25′16″ 16°22′31″ Phyllade Metamorphic complex
(from metarenites to paragneiss)
F. va Allaro 2.38 × 10−4
15 38°23′37″ 16°23′10″ Phyllade Gray phyllade Le Serre eastern side,
Popolli‐Serra Schiavella
1.60 × 10−4
16 38°24′57″ 16°21′33″ Granite–aplite Aplitic sills inside Le Serre granites F. va Allaro 6.95 × 10−6
17 38°24′57″ 16°21′33″ Granite F. va Allaro 1.30 × 10−4
18 38°31′15″ 16°08′28″ Granite Le Serre chain, western side 4.50 × 10−4
24 38°49′18″ 16°17′51″ Metagabbro Curinga 1.10 × 10−2
Volcanic Units
20 38°37′28″ 16°06′29″ Pumice Granulometry: 1 mm 1.66 × 10−4
21 38°37′28″ 16°06′29″ Pumice Granulometry: 4 mm 2.34 × 10−4
22 38°37′28″ 16°06′29″ Pumice Granulometry: 8 mm 2.28 × 10−4
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
9 of 18
As a first‐order of approximation, lithology controls mag-
netic properties through its mineralogy, and abrupt variation
in rock properties may highlight lithological contacts. Since
the backbone of southern Calabria is mainly composed of
nonmagnetic rocks (see section 3), the presence of the CV
positive magnetic anomaly allows us to infer the presence of
a magnetized body not related to the surrounding exposed
units. To get a more reliable interpretation of the CV
anomaly system causative bodies, we carried out data inver-
sion, obtaining the representative magnetization value, shape,
depth, and horizontal extent of the body. Furthermore, a
2.75‐D forward model has been carried out with the aim of
adding further geostructural constraints to the inversion
results by using measured susceptibility values for the back-
ground rock properties (Table 1). The CV anomaly system
shows a simple pattern lightly affected by the regional field,
which influences the intensities of the short‐wavelength
components (anomalies 2, 3, and 4 in Figure 3). The regional
field is not completely sampled due to the shorter dimension
of the surveyed area, and it is characterized by the over-
lapping effects of sources lying in south Calabria (e.g., in the
Catena Costiera and Sila areas) (Figure 3). The influence of
such sources is not homogeneous and difficult to evaluate
using analytical methods.
5.1. Inversion Modeling
[22] We applied a 3‐D inversion approach to the magnetic
anomaly data of the CV area with the aim of modeling the
subsurface bodies generating the observed anomaly. As is
well known, the mathematical inversion approach is defined
as the quantitative estimation of some models parameters
starting from their effect (e.g., magnetic measures), knowing
the mathematical relationship that links the models para-
meters with their spatial effect (forward problem). From a
mathematical point of view, the inversion applied to surface
potential data field is an ill‐posed problem that suffers from
instability and algebraic ambiguity of solutions [Blakely,
1995], so that model parameters are not analytically invert-
ible from anomaly data. A common procedure to reduce those
ambiguities and obtain realistic geological sources is the
introduction of additional information in the form of a priori
geological data or mathematical regularization. A common
approach to solve these kinds of problems in potential fields
inversion is to divide the subsoil, where the real geological
source is supposed to be located (a priori information), into
a number of homogeneous magnetized rectangular blocks
(cells) for which the direction of magnetization is supposed
to be known and equal for each block. Such “discretization”
defines the exact geometry of each single subsurface cell so
that it is possible to generate the observed potential field
measurement by applying an appropriate magnetization
value for each cell [Bear et al., 1995]. In this form, the
inversion problem is linearized so that the unknown
parameter to be determined is the magnetization module of
each cell. In our analysis, we adopted the forward solution
following Sharma [1986]. Considering the vector of spatial
measures b and the magnetization vector x of N cells, the
linear forward model evaluating the effect of the whole cells
on ith station point is expressed by
bi Âź
X
N
jÂź1
Aij  xj;
where j is the cells index, Aij is the sensitivity matrix
coefficient that evaluates the contribution on the ith station
point of the jth cell, considering unitary magnetization. To
resolve this linear system in the x vector unknowns, we
adopted the iterative regularization minimization method of
Levenberg and Marquardt [Press et al., 1992]. This method
simultaneously minimizes the norm of the solution and the
misfit between the observed and the computed synthetic
magnetic anomalies generated by the magnetization solution
vector. We also applied a positivity control on the magne-
tization solution because of the CV positive dipolar anomaly
geometry.
[23] Taking into account the spatial extent of the CV mag-
netic anomaly (Figures 2b and 3), we built a 3‐D block mesh
made with cells of dimensions 2,000 m E × 2,000 m N ×
500 m below the surface. The maximum depth is at 3,500 m.
The inversion setup corresponds to a total number of
3,600 cells. We adopted a regular measure grid made by
2,000 m spaced nodes calculated by gridding the original
magnetic data. To stabilize the inversion model results, we
used the same dimensions for the block cells mesh and the
measured grid. In calculating the sensitivity matrix, it is
necessary to define the vector values of block magnetiza-
tion; we assume that the total magnetization vector of the
CV body is subparallel to the Total Magnetic Field direc-
tion and hence apply a magnetic inclination of 55° and a
declination of 0°. To better constrain the resulting model,
we defined a bathymetric surface (data from Tortorici et al.
[2003]) above which cells were not used in the inversion
calculation.
[24] Figure 4a shows the resulting model from inversion
procedure applied to the magnetic anomaly data, Figure 4b
shows the measured magnetic data used in inversion pro-
cess, and Figure 4c shows the synthetic anomaly map. The
differences between the synthetic and observed data are
reported in Figure 4d. The CV anomaly is well recovered by
the inversion model. However, regional longer wavelengths
extend well beyond the relatively limited survey area, as
shown on the regional anomaly map in Figure 2c [Caratori
Tontini et al., 2004], and the inversion approach cannot
resolve this problem. The occurrence of such long‐wavelength
anomalies could cause a misfit between the measured mag-
netic anomaly data and the synthetic anomaly values gener-
ated by the recovered inversion model. Taking into account
Figure 4. (a) Magnetic source model from inversion process obtained from the sea level down to 3,000 m below sea level;
distances are expressed in meters. (b) Observed magnetic anomaly data. (c) Synthetic anomaly map obtained for the mag-
netization distribution shown with I = 54° and D = 2° inducing field. (d) Differences between synthetic and observed data
shown in Figures 4b and 4c. (e) Field (in color) generated by the susceptibility model in Figure 4a with a I = 90° and D = 0°
inducing field (the synthetic magnetic effect that this source model would have at the magnetic North Pole). The RTP field
of real data of Figure 3b is reported as isoline contouring.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
10 of 18
Figure 4
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
11 of 18
this possible misfit, we focused on the fitting quality of the
CV anomaly. The differences between the synthetic and
observed data (Figure 4d) show a long‐wavelength misfit
trend oriented N‐NE, S‐SW with amplitude between 0 and
−10 nT (about 5% of the maximum amplitude of the CV
anomaly). In the surrounding area, the misfit shows a clear
long‐wavelength pattern with amplitude oscillating between
−20 and −10 nT. The inversion does not well describe the
east side of the inversion area because of the positive
magnetic anomalies generated by geological sources located
in Calabria. This misfit is described in Figure 4d by negative
anomalies with a minimum of −40 nT in a very limited area
in the eastern side of the map. The above summarized misfit
values are very low, thus suggesting the goodness and
Figure 4. (continued)
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
12 of 18
reliability of the inversion model. The recovered inversion
model reported in Figure 4a describes a positive magnetic
body approximately 20–25 km long WNW‐ESE striking
with maximum magnetization values of 1.5 A/m.
[25] We carried out a further model to investigate the
nature of the RTP negative anomaly appearing just to the
south of positive anomaly 1 (Figure 3b). This negative
anomaly represents an interesting issue to investigate, since
it precisely overlaps the offshore prolongation of the CV
fault. We computed the field generated by the susceptibility
model of Figure 4a with a I = 90° and D = 0°–inducing field
(the synthetic magnetic effect that this source model would
have at the magnetic North Pole). The results are shown as a
color map in Figure 4e, where the RTP field of real data
(Figure 3b) is also reported as isoline contouring. The close
matching of these two fields indicates that the negative
feature depicted in the RTP map of Figure 3b results from
the RTP computation [see also Silva, 1986] and belongs to
the anomaly 1 source recomputed at the North Magnetic
Pole.
5.2. Forward Modeling
[26] We carried out the forward modeling along two key
profiles crossing the area in the WNW‐ESE and NNE‐SSW
directions (Figures 3b and 5). The measured rock suscepti-
bility values (Table 1) were used to characterize the exposed
and deep stratigraphic sequences shown in the geological
profile (Figure 2b). Tectonic units depicted in the sections
are taken from the structural model of Italy [Finetti, 2005].
The modeled body (position, geometry, and extension) was
drawn on the basis of the inversion results. In both profiles
shown in Figure 5, the structures were further constrained
with the available surface geology information. In addition,
the uppermost crust beneath the CV morphological ridge
was constrained with the seismic profiles crossing the area
[Trincardi et al., 1987; Milia et al., 2009]. The extension of
the modeled bodies along the profiles are reported in Table 2.
[27] The A‐B profile (Figure 5) was chosen with the goal
of crossing symmetrically the CV anomaly along its major
axis and sampling its intensity peak. The C‐D profile was
oriented orthogonally to the previous one. This profile was
chosen to further constrain the modeling and to interpret the
low‐intensity, short‐wavelength anomalies lying southward
of the CV fault and not imaged by the inversion results
(Figure 5). In fact, our high‐resolution, near‐surface survey
has also detected the higher‐frequency components in the
anomaly field that were not visible in the previous magnetic
surveys [Agip, 1981; Chiappini et al., 2000; Caratori
Tontini et al., 2004]. These minor magnetic signatures are
here interpreted in the same framework of the CV anomaly
causative body and, for this reason, are called the CV
magnetic anomaly system. We achieved the modeling by
using an implementation of a 2.75‐D forward modeling
found on the expressions of Talwani and Heirtzler [1964],
Rasmussen and Pedersen [1979] and on the algorithm
proposed by Won and Bevis [1987]. The proposed models,
summarized in Figure 5, are discussed in section 6.
6. Discussion and Conclusion
[28] The results of the magnetic modeling highlight the
occurrence of a WNW‐trending magnetized body about
20 km long (Figures 4 and 5). This body, which is directly
magnetized, extends from the seafloor to about 3 km bsl,
and intrudes toward ESE, between the CV granites and the
underlying granulites of the Calabrian Arc. The WNW flank
of the body is covered by sediments that, on the basis of
available seismic profiles, consist of sands and silts younger
than 0.7 Myr [Milia et al., 2009]. According to the models
shown in Figures 4 and 5, the width is between 3.5 and
5 km. The reconstructed geometry is consistent with that
of a semitabular, flat, elongated sill. Less extended, 2 to
3 km wide, laccolith‐like highly magnetized bodies also
occur between 3 and 4 km of depth. These minor bodies are
intruded within the granulites of the Calabrian Arc. The
magnetization values needed to explain the observed anoma-
lies in the CV area are between 0.7 and 1.5 A/m (Figures 4
and 5). This magnetization range encompasses that mea-
sured in the evolved rocks (rhyolites and trachytes) of the
Aeolian Islands (0.6–1.6 A/m) [Blanco‐Montenegro et al.,
2007, and references therein]. The collected magnetic data
show that the CV anomaly reaches magnetic anomaly values
(150–220 nT, Figure 3b) that are significantly higher than
those observed in the rest of Calabria (30–50 nT) [Agip,
1981; Caratori Tontini et al., 2004]. In addition, the CV
magnetic anomaly values are of the same magnitude as
those of Panarea Island, Lametini Seamount, Alcione Sea-
mount, and Stromboli canyon, where volcanic rocks outcrop
and have been dragged [Caratori Tontini et al., 2004;
Finetti, 2005]. The anomaly of the Stromboli canyon, which
is located 20 km eastward from Panarea Island, is also
visible in the westernmost side of the CV area survey
(anomaly 5 in Figure 3). The magnetic signature of the CV
promontory turns out to be completely different from the
magnetic pattern of the Calabrian Arc region for both inten-
sity and shape.
[29] The inverse model (Figure 4a) depicts a crustal vol-
ume with an increase of the magnetization (values 0.9 A/m)
in correspondence of the top of the CV bathymetric ridge.
We propose that this 5–6 km long and about 2–3 km wide
rock volume with higher magnetization is representative of
the shallower portion of the plumbing system of the volcano
from which WNW‐ESE elongated apophyses with lower
magnetization depart, probably representing dikes. Besides,
the forward model (Figure 5), which is characterized by a
constant magnetization of the represented bodies, requires a
vertical extension in correspondence to the abovementioned
bathymetric ridge. Both models are able to fit the observed
data correctly, represent two possible settings of the mag-
netization, and be coherent with the occurrence of a volcanic
body. In the first case (Figure 4), the higher magnetization
represents the plumbing systems zone, as reported above;
in the second one (Figure 5), a vertical extension of the body
is required to correctly fit the data to define the plumbing
system geometrically. Therefore, both models highlight a
magnetized, shallow, sill‐like body, the topographic expres-
sion of which is represented by the CV morphological ridge
(Figures 4 and 5). In fact, the results reveal that the apex of
the sill overlaps the subcircular truncated top of the CV ridge
located at a 200 m bsl (Figure 2b). As result of the above
observations, the CV sill could be interpreted as the remnant
of a shallow magmatic body characterized by a central plug
roughly extending to a depth between 500 and 1500 m bsl,
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
13 of 18
Figure 5. Crustal 2.75‐D magnetic models for profiles crossing the CV anomaly in (a) WNW‐ESE direction, A‐B profile;
(b) N20°E direction, C‐D profile. The trace of profiles is shown in Figure 3b (dotted red lines). Numbers identify marked
anomalies in the reduced‐to‐the‐pole magnetic anomaly field (Figure 3). Faults are indicated by red lines. The size of the
bodies in the direction perpendicular to the profile is reported in Table 2.
DE
RITIS
ET
AL.:
A
BURIED
VOLCANO
BY
AEROMAGNETIC
DATA
B11101
B11101
14
of
18
Figure
5.
(continued)
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
15 of 18
and by WNW‐trending minor lens, possibly representing dikes
intruding the Calabrian Arc.
[30] We propose that the CV sill may represent the source
area of the 0.67 to 1.06 Myr old widespread tephra fall
deposits from plinian columns poorly resedimented in shal-
low marine environment outcropping in western Calabria
(section 2.2). The occurrence of (1) maximum thickness of
such deposits in the CV and Mesima areas, (2) progressive
thinning of such deposits toward the southwest (i.e., Reggio
Calabria basin), (3) impact sags related to the emplacement
of ballistic ejecta in the CV area, and (4) sedimentary facies
consistent with a proximal origin for the tephra layers
strongly suggests that the CV magnetic anomaly sector may
correspond to the source area of the middle Pleistocene
volcaniclastic deposits of Calabria [De Rosa et al., 2008].
As a result, the shallow, WNW‐ESE elongated sill depicted
in Figures 4 and 5 is easy to use to represent the shallow
plumbing system of a middle Pleistocene volcano. Such
interpretation is consistent with the occurrence, in the sector
of the CV magnetic anomaly, of a 0–5 km deep, high Vp/Vs
body [Barberi et al., 2004]. The magnetization values
obtained from inversion of the data (Figure 4a) suggest that
the CV sill is composed of rocks with magnetization values
comparable to those of rhyolites and dacites of the Aeolian
Islands [Blanco‐Montenegro et al., 2007]. This consider-
ation corroborates the above interpretation of the CV sill as a
plumbing system of a volcano responsible for the emplace-
ment of the Pleistocene widespread rhyolitic to dacitic
pumices outcropping in Calabria, which show a HK calc‐
alkaline affinity compatible with that of the rocks of the
eastern Panarea sector of the Aeolian Islands [De Rosa et al.,
2008]. On the basis of the stratigraphic, magnetic, and geo-
chronological data, the CV ridge formed before 0.7 Ma,
being partly covered by Pleistocene sediments younger than
0.7 Myr [Milia et al., 2009]. The normal polarity magnetic
signature of the CV body suggests an emplacement during
the Bruhnes (0.78 Ma) [Cande and Kent, 1995; Gradstein
et al., 2005] or Jaramillo (0.99–1.07 Ma) Chrons. On the
other hand, the widespread pumices and volcanic ash layers
outcropping in the studied area deposited in a time inter-
val between 0.67 and 1.07 Ma [Cornette et al., 1987].
Therefore, the CV sill and the Pleistocene volcanics were
emplaced either between 0.67 and 0.78 Ma or between 0.99
and 1.07 Ma. These age intervals roughly overlap the older
ages of Aeolian volcanism, which started between 1.3 Ma
in the western, Alicudi–Filicudi sector and 0.8–0.5 Ma in
the eastern, Stromboli–Panarea sector [De Astis et al.,
2003]. However, the composition of the older dragged
Aeolian rocks is basaltic, whereas the more evolved prod-
ucts, with composition similar to the CV tephra, are younger
than 100 kyr. This indicates that evolved magmas charac-
terized the early, submarine, and possibly explosive, phases
of the Aeolian volcanism. In addition, the rhyolitic and dacitic
composition of the CV Pleistocene magmas volcano, along
with the results of our magnetic modeling, testify to the
formation of sill‐like, shallow magma reservoirs. Accord-
ing to the geometry depicted in Figure 5 and the available
geological data (Figure 2b), the storage depth of such reser-
voirs is controlled by the major discontinuity between the
lower crust granulites of the underlying, Carboniferous–
Permian granites of the upper crust. The shape of the CV
magnetic anomaly clearly shows that the sill geometry has
also been controlled by a WNW‐trending discontinuity. This
is also the strike of the fault affecting the southern coastline
and offshore of the CV promontory (Figure 2). This fault is
possibly seismically active [De Astis et al., 2003] and was
responsible for the destructive 1905 Calabria earthquake,
M = 6.3–6.7, according to Cucci and Tertulliani [2006].
Therefore, we conclude the CV volcano was emplaced on
the footwall of this still active tectonic structure. This feature
is structurally compatible with numerical models of magma
emplacement along active faults [e.g., Ellis and King, 1991;
Pascal and Cloetingh, 2002]. Such models reveal that magma
uprising to the surface along a faulting zone accumulates at
the footwall of a normal fault because the local extensional
strain concentrates at the footwall, whereas a compressive
strain occurs in the hanging wall.
[31] The collected data and the above observations sug-
gest that the recognized CV volcanism is consistent, in
terms of age, magnetic signature, and geochemical affinity
of the erupted products, with that of the Aeolian volcanoes.
The central sector of the Aeolian Arc, which includes the
islands of Salina, Lipari, and Vulcano, develops along a
tectonic structure that it is in a transversal position with
respect to the Aeolian volcanic ring. Our data and inter-
pretation suggest that, indeed, NW‐SE faults (i.e., trans-
versal structures to the general trend of the Calabria block)
must be considered as an important part of the Pleistocene
(and possibly active) tectonics of Calabria. It is also likely
that the seismogenic potential of such structures is at present
underevaluated [Basili et al., 2008] and should be better
assessed.
[32] The WNW‐trending fault that controls the CV
fissural‐like volcano is also transversal to the eastern sector
of the Aeolian volcanoes, which are controlled by NE‐SW
structures. However, this fault does not appear to represent a
slab tear fault [e.g., Chiarabba et al., 2008]. In any case, it
may represents a structure along which magma uprose and
stopped, at least in the last 1–0.7 Ma. In this period, a sig-
nificant (up to 1 mm/yr) uplift affected the Calabria region
and the Aeolian Islands [Westaway, 1993]. The uplift, whose
geodynamic origin is a subject of debate [e.g., Gvirtzman and
Nur, 2001], is maximized in the CV promontory (4 mm/yr
in the last 124 kyr) [Tortorici et al., 2003]. Different uplift
rates of segments of the Calabrian Arc could be responsible
for fragmentation of the upper crust and formation of
transversal faults. As a result, magma emplacement may
Table 2. The Strike of the Structures for the Two Profiles A‐B
and C‐D Is 90°a
Body y1 (km) y2 (km)
Granulites section A‐B 10 10
Sands and sandstone section A‐B 10 10
Granites section A‐B 10 10
Volcanic body section A‐B
Anomaly 1 1.7 1.7
Sands and sandstone section C‐D 10 10
Volcanic body section C‐D
Anomaly 1 5 15
Anomaly 4 3 3
Anomaly 2 3 3
a
y1 is the length toward the west‐northwest in profile A‐B and toward the
N20°E in profile C‐D, whereas y2 is the length toward the east‐southeast in
profile A‐B and toward the N290°E in profile C‐D.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
16 of 18
occur along structures like the WNW‐trending fault of the
CV. According to Tortorici et al. [2003], such structures
also act as transfer faults that allow the regional extensional
strain to be transferred from offshore, where it is accom-
modated by the ENW‐WSW trending normal faults, to the
Calabrian Arc, where it is accommodated by NE–SW
trending faults. In this general framework, the composition
of the CV tephra suggest that between 0.67 and 0.97 Ma,
while mafic products were erupting in the Aeolian area,
dacites and rhyolites characterized the explosive activity of
the CV volcano. From a structural point of view, we note
that the CV fissural‐like volcano is similar to the vents of
Lipari and Vulcano islands, which also depict a fissural‐like
structure [De Astis et al., 2003]. As a result, the transversal
faults of the Aeolian Arc appear to control the development
of fissure‐like eruptive centers.
[33] Acknowledgments. This study has been supported with FIRB‐
MUIR and INGV funds. Thanks to numerous colleagues for discussions,
comments, and suggestions. We are grateful to two anonymous referees,
as well as to the Journal of Geophysical Research Associate Editor and
the Editor (Andre Revil) for providing useful comments to our manuscript.
References
Acquafredda, P., S. Lorenzoni, N. Minzoni, and E. Zanettin Lorenzoni
(1987), Paleozoic sequence in the Stilo‐Bivongi area (central Calabria),
Mem. Sci. Geol., 39, 117–127.
Acquafredda, P., S. Lorenzoni, N. Minzoni, and E. Zanettin Lorenzoni
(1989), Stratigraphic correlation form of the stilo area (Serre region, cen-
tral Calabria, Italy), Rend. Soc. Geol. Ital., 12, 103–105.
Acquafredda, P., S. Lorenzoni, and E. Zanettin Lorenzoni (1994), Paleo-
zoic sequence and evolution of the Calabrian‐Peloritan Arc (Southern
Italy), Terranova, 6, 582–594.
Amodio Morelli, L., et al. (1976), L’arco Calabro‐peloritano nell’ orogene
Appenninico‐Maghrebide, Mem. Sci. Geol., 17, 1–60.
Antonioli, F., L. Ferranti, K. Lambeck, S. Kershaw, V. Verrubbi, and
G. Dal Pra (2006), Late Pleistocene to Holocene record of changing uplift
rates in southern Calabria and northeastern Sicily (southern Italy, central
Mediterranean Sea), Tectonophysics, 422, 23–40.
Ayuso, R. A., A. Messina, B. De Vivo, S. Russo, L. G. Woodruff, S. F.
Sutter, and H. E. Belkin (1994), Geochemistry and argon thermochronol-
ogy of the Variscan Sila batholith, southern Italy: Source rocks and
magma evolution, Contrib. Miner. Petrol., 117, 87–109.
Azzaro, R., D. Bella, L. Ferreli, A. Michetti, F. Santagati, L. Serva, and
E. Vittori (2000), First study of fault trench stratigraphy at Mt. Etna
volcano, Southern Italy: Understanding Holocene surface faulting along
the Moscarello fault, J. Geodyn., 29, 187–210.
Barberi, F., A. Gandino, A. Gioncada, P. La Torre, A. Sbrana, and
C. Zenucchini (1994), The deep structure of the Eolian arc (Filicudi‐
Panarea‐Vulcano sector) in light of gravity, magnetic and volcanological
data, J. Volcanol. Geotherm. Res., 61, 189–206.
Barberi, G., M. T. Cosentino, A. Gervasi, I. Guerra, G. Neri, and B. Orecchio
(2004), Crustal seismic tomography in the Calabrian Arc region, south
Italy, Phys. Earth Planet. Int., 147, 297–314.
Baranov, V., and H. Naudy (1964), Numerical calculation of the formula of
reduction to the magnetic pole, Geophysics, 29, 67–79.
Basili, R., G. Valensise, P. Vannoli, P. Burrato, U. Fracassi, S. Mariano,
M. M. Tiberti, and E. Boschi (2008), The Database of Individual Seismo-
genic Sources (DISS), version 3: Summarizing 20 years of research on
Italy’s earthquake geology, Tectonophysics, 453, 20–43, doi:10.1016/j.
tecto.2007.04.014.
Bear, G. W., H. J. Al‐Shukri, and A. J. Rudman (1995), Linear inversion of
gravity data for 3‐D density distribution, Geophysics, 60, 1354–1364.
Blakely, R. J. (1995), Potential Theory in Gravity and Magnetic Applica-
tions, 441 pp., Cambridge Univ. Press, Cambridge.
Blanco‐Montenegro, I., R. De Ritis, and M. Chiappini (2007), Imaging
and modelling the subsurface structure of volcanic calderas with high‐
resolution aeromagnetic data at Vulcano (Aeolian Islands, Italy), Bull.
Volcanol., 69(6), 643–659, doi:10.1007/s00445-006-0100-7.
Bonardi, G., A. Messina, V. Perrone, S. Russo, and A. Zuppetta (1984),
L’unità di stilo nel settore meridionale dell’arco calabro peloritano, Boll.
Soc. Geol. Ital., 103, 279–309.
Bonardi, G., W. Cavazza, V. Perrone, and S. Rossi (2001), Calabria‐
Peloritani terrane and northern Ionian Sea, in Anatomy of an Orogen:
The Apennines and Adjacent Mediterranean Basins, edited by G. B.
Vai and I. P. Martini, pp. 287–306, Kluwer, Dordrecht.
Bouillin, J. P., S. Baudelot, and C. Majestè‐Menjoulas (1984), Mise en
evidence du Cambro.Ordovicien en calibre centrale (Italie). AffinitĂŠs
palĂŠogĂŠographiques et consĂŠquences structurales, C. R. Acad. Sci.
Paris, II, 298, 88–92.
Bouillin, J. P., M. Menjoulas, S. Baudelot, C. Cygan, and C. Fournier
Vinas (1987), Les formations paleozoîques de l’Arc Calabro‐Peloritain
dans leur cadre structural, Boll. Soc. Geol. Ital., 106, 683–698.
Caggianelli, A., and G. Prosser (2002), Modelling the thermal perturbation of
the continental crust after intraplating of thick granitoid sheets: A compar-
ison with the crustal sections in Calabria (Italy), Geol. Mag., 139, 699–706.
Cande, S. C., and D. V. Kent (1995), Revised calibration of the geomagnetic
polarity time scale for the Late Cretaceous and Cenozoic, J. Geophys.
Res., 100, 6093–6095.
Caratori Tontini, F., P. Stefanelli, I. Giori, O. Faggioni, and C. Carmisciano
(2004), The revised aeromagnetic map of Italy, Ann. Geophys., 47,
1547–1556.
Cavazza, W. (1989), Detritial modes and provenance of the Stilo‐capo
d’Orlando Formation (Miocene), southern Italy, Sedimentology, 36,
1077–1090.
Cavazza, W., and P. G. DeCelles (1993), Upper Messinian siliciclastic
rocks in southeastern Calabria (southern Italy): Palaeotectonic and
eustatic implications for the evolution of the central Mediterranean
region, Tectonophysics, 298, 223–241.
Cavazza, W., J. Blenkinsop, P. G. De Celles, R. Timothy Patterson, and
E. G. Reinhardt (1997), Stratigrafia e sedimentologia della sequenza
sedimentaria oligocenico‐quaternaria del bacino calabro‐ionico, Boll.
Soc. Geol. Ital., 116, 51–77.
Chiappini, M., A. Meloni, E. Boschi, O. Faggioni, N. Beverini, C. Carmisciano,
and I. Marson (2000), Shaded relief magnetic anomaly map of Italy and
surrounding marine areas, Ann. Geophys., 43, 983–989.
Chiappini, M., F. Ferraccioli, E. Bozzo, and D. Damaske (2002), Regional
compilation and analysis of aeromagnetic anomalies for the Transantarctic
Mountains‐Ross Sea sector of the Antarctic, Tectonophysics, 347, 121–137.
Chiarabba, C., P. De Gori, and F. Speranza (2008), The Southern Tyrrhe-
nian Subduction Zone: Deep geometry, magmatism and Plio‐Pleistocene
evolution, Earth Planet. Sci. Lett., 268, 408–423.
Cifelli, F., M. Mattei, and F. Rossetti (2007a), Tectonic evolution of arcuate
mountain belts on top of a retreating subduction slab: The example of
the Calabria Arc, J. Geophys. Res., 112, B09101, doi:10.1029/
2006JB004848.
Cifelli, F., M. Mattei, and F. Rossetti (2007b), The architecture of brittle
postorogenic extension: Results from an integrated structural and paleo-
magnetic study in north Calabria (southern Italy), Geol. Soc. Am. Bull.
119(1), 221–239.
Colonna, V., S. Lorenzoni, and E. Zanettin Lorenzoni (1973), Sull’esistenza
di due complessi metamorfici lungo il bordo sud‐orientale del massiccio
“granitico” delle Serre (Calabria), Boll. Soc. Geol. Ital., 92, 801–830.
Cornette, Y., P. Y. Gillot, P. Barrier, and F. Jehenne (1987), DonnĂŠes
radiométriques preliminaries (Potassium‐Argon) sur des cinérites plio‐
pléistocènes du Détroit de Messine, Doc. Trav. Inst. Geol. Albert‐de‐
Lapparent, 11, 97–99.
Cucci, L., and A. Tertulliani (2006), I terrazzi marini nell’area di Capo
Vaticano (arco calabro): Solo un record di sollevamento regionale o anche
di deformazione cosismica?, Quaternario, 19, 89–101.
De Astis, G., G. Ventura, and G. Vilardo (2003), Geodynamic significance
of the Aeolian volcanism (Southern Tyrrhenian Sea, Italy) in light of
structural, seismological and geochemical data, Tectonics, 22(4), 1040,
doi:10.1029/2003TC001506.
Del Ben, A., C. Barnaba, and A. Taboga (2008), Strike‐slip systems as the
main tectonic features in the Plio‐Quaternary kinematics of the Calabrian
Arc, Mar. Geophys. Res., 29, 1–12.
De Ritis, R., I. Blanco‐Montenegro, G. Ventura, and M. Chiappini (2005),
Aeromagnetic data provide new insights on the volcanism and tectonics
of Vulcano Island and offshore areas (southern Tyrrhenian Sea, Italy),
Geophys. Res. Lett., 32, L15305, doi:10.1029/2005GL023465.
De Ritis, R., G. Ventura, and M. Chiappini (2007), Aeromagnetic anoma-
lies reveal hidden tectonic and volcanic structures in the central sector of
the Aeolian Islands, southern Tyrrhenian Sea, Italy, J. Geophys. Res.,
112, B10105, doi:10.1029/2006JB004639.
De Rosa, R., R. Dominici, and M. Sonnino (2001), Evidenze di vulcanismo
sinsedimentario nella successione pleistocenica del graben del Mèsima
(Calabria centro‐occidentale), Quaternario, 14(2), 81–91.
De Rosa, R., R. Dominici, P. Donato, and D. Barca (2008), Widespread
syn‐eruptive volcaniclastic deposits in the Pleistocenic basins of
South‐Western Calabria, J. Volcanol. Geotherm. Res., 177, 155–169.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
17 of 18
Doglioni, C., F. Innocenti, and S. Mariotti (2001), Why Mt. Etna?, Terra
Nova, 13, 25–31.
Dumas, B., P. GuÊrÊmy, R. LhÊnaff, and J. Raffy (1982), Le soulèvement
Quaternaire de la Calabre meridionale, Rev. Gèol. Dyn. GÊogr. Phys., 23,
27–40.
Ellis, M., and G. King (1991), Structural control of flank volcanism in con-
tinental rift, Science, 254, 839–843.
Faccenna, C., T. W. Becker, F. P. Lucente, L. Jolivet, and F. Rossetti
(2001), History of subduction and back‐arc extension in the Central
Mediterranean, Geophys. J. Int., 145, 809–820.
Finetti, I. (2005), Deep Seismic Exploration of the Central Mediterranean
and Italy, 794 pp., Elsevier, Amsterdam.
Finn, C., T. W. Sisson, and M. Deszcz‐Pan (2001), Aerogeophysical mea-
surements of collapse‐prone hydrothermally altered zones at Mount
Rainer volcano, Nature, 409, 600–603.
Ghisetti, F. (1981), L’evoluzione strutturale del bacino plio‐pleistocenico di
Reggio Calabria nel quadro geodinamico dell’Arco Calabro, Boll. Soc.
Geol. Ital., 100, 433–466.
Ghisetti, F., and L. Vezzani (1980), Contribution of structural analysis to
understanding the geodynamic evolution of the Calabrian arc (Southern
Italy), J. Struct. Geol., 3, 371–381.
Gradstein, F. M., J. G. Ogg, and A. G. Smith (2005), A Geologic Time
Scale, 610 pp., Cambridge Univ. Press, Cambridge.
Graessner, T., and V. Schenk (2001), An exposed Hercynian deep conti-
nental crustal section in the Sila massif of northern Calabria: Mineral
chemistry, petrology and a P–T path of granulite facies metapelitic mig-
matites and metabasites, J. Petrol., 42, 931–961.
Guarnieri, P. (2006), Plio‐Quaternary segmentation of the south Tyrrhenian
forearc basin, Int. J. Earth Sci., 95, 107–118.
Gvirtzman, Z., and A. Nur (2001), Residual topography, lithospheric struc-
ture and sunken slabs in the Central Mediterranean, Earth Planet. Sci.
Lett., 187, 117–130.
IAGA (International Association of Geomagnetism and Aeronomy)
(2005), The 10th‐Generation International Geomagnetic Reference Field,
Geophys. J. Int., 161, 561–565.
Jolivet, L., and C. Faccenna (2000), Mediterranean extension and the
Africa‐Eurasia collision, Tectonics, 19, 1095–1106.
Knott, S. D., and E. Turco (1991), Late Cenozoic kinematics of the Calab-
rian Arc, southern Italy, Tectonics, 10, 1164–1172.
Langone, A., E. Gueguen, G. Prosser, A. Caggianelli, and A. Rottura
(2006), The Curinga‐Girifalco fault zone (northern Serre, Calabria) and
its significance within the Alpine tectonic evolution of the western Med-
iterranean, J. Geodyn., 42, 140–158
Lenat, J. F., B. Gibert‐Malengreau, and A. Galdeano (2001), A new struc-
tural model for the evolution of the volcanic island of Reunion (Indian
Ocean), J. Geophys. Res., 106(B5), 8645–8663.
Malinverno, A., and W. B. F. Ryan (1986), Extension in Tyrrhenian Sea
and shortening in the Apennines as result of arc migration driven by sink-
ing of the lithosphere, Tectonics, 5, 227–254.
Mattei, M., F. Speranza, A. Argentieri, F. Rossetti, L. Sagnotti, and
R. Funiciello (1999), Extensional tectonics in the Amantea basin (Calab-
ria, Italy): A comparison between structural and magnetic anisotropy
data, Tectonophysics, 307, 33–49.
Mattei, M., P. Cipollari, D. Cosentino, A. Argentieri, F. Rossetti, and
F. Speranza (2002), The Miocene tectonic evolution of the Southern Tyr-
rhenian Sea: Stratigraphy, structural and paleomagnetic data from the
on‐shore Amantea basin (Calabrian Arc, Italy), Basin Res., 14, 147–168.
Milia, A., E. Turco, P. P. Pierantoni, and A. Schettino (2009), Four‐
dimensional tectono‐stratigraphic evolution of the Southeastern peri‐
Tyrrhenian basins (Margin of Calabria, Italy), Tectonophysics, 476,
41–56, doi:10.1016/j.tecto.2009.02.030
Miyauchi, T., G. Dal Pra, and S. S. Labini (1994), Geochronology of Pleis-
tocene marine terraces and regional tectonics in the Tyrrhenian coast of
South Calabria, Italy, Quaternario, 7, 17–34.
Monaco, C., and L. Tortorici (2000), Active faulting in the Calabrian Arc
and eastern Sicily, J. Geodyn., 29, 407–424.
Mulargia, F., P. Baldi, V. Achilli, and F. Broccio (1984), Recent crustal
deformations and tectonics of the Messina Strait area, Geophys. J. R.
Astron. Soc., 76, 369–381.
Neri, G., B. Orecchio, C. Totaro, G. Falcone, and D. Presti (2009), Subduc-
tion beneath southern Italy close the ending: Results from seismic tomog-
raphy, Seism. Res. Lett., 80, 63–70.
Nicolosi, I., F. Speranza, and M. Chiappini (2006), Ultrafast oceanic
spreading of the Marsili Basin, southern Tyrrhenian Sea: Evidence from
magnetic anomaly analysis, Geology, 34, 717–720.
Pascal, C., and S. A. P. L. Cloetingh (2002), Rifting in heterogeneous litho-
sphere: Inferences from numerical modeling of the northern North Sea
and the Oslo Graben, Tectonics, 21(6), 1060, doi:10.1029/2001TC901044.
Patterson, R. T., J. Blenkinsop, and W. Cavazza (1995), Planktic foraminiferal
biostratigraphy and 87Sr/86Sr isotopic stratigraphy of the Oligocene‐to‐
Pleistocene sedimentary sequence in the south‐eastern calabria micro-
plate, southern Italy, J. Paleontol., 69, 7–20.
Press, W., S. Teulosky, W. Vetterling, and B. Flannery (1992), Numerical
Recipes in C: The Art of Scientific Computing, 965 pp., Cambridge Univ.
Press, Cambridge.
Rasmussen, R., and L. B. Pedersen (1979), End corrections in potential
field modelling. Geophys. Prospect., 27, 749–760.
Roda, C. (1965), Il calcare portlandiano a Dasycladacee di M. Mutolo,
Reggio Calabria, Geol. Rom., 4, 259–290.
Rollin, P. J., J. Cassidy, C. A. Locke, and H. Rymer (2000), Evolution of
the magmatic plumbing system at Mt Etna: New evidence from gravity
and magnetic data, Terra Nova, 12, 193–198.
Rosenbaum, G., M. Gasparon, F. P. Lucente, A. Peccerillo, and M. S.
Miller (2008), Kinematics of slab tear faults during subduction segmen-
tation and implications for Italian magmatism, Tectonics, 27, TC2008,
doi:10.1029/2007TC002143.
Sagnotti, L., F. Speranza, A. Winkler, M. Mattei, and R. Funicello (1998),
Magnetic clay of clay sediments from the external northern Apennines
(Italy), Phys. Earth Planet. Int., 105, 73–93.
Schenk, V. (1984), Petrology of felsic granulites, metapelites, metabasic,
ultramafic, and metacarbonates from southern Calabria (Italy): Prograde
metamorphism, uplift and cooling of former lower crust, J. Petrol., 25,
225–298.
Schenk, V. (1990), The exposed crustal cross section of southern Calabria,
Italy: Structure and evolution of a segment of Hercynian crust, in
Exposed Cross‐Section of Continental Crust, edited by M. H. Salisbury
and D. M. Fountain, pp. 21–42, Kluwer, Dordrecht.
Sharma, P. V. (1986), Geophysical Methods in Geology, 442 pp., Elsevier,
New York.
Silva, J. B. C. (1986), Reduction to the pole as an inverse problem and its
application to low‐latitude anomalies, Geophysics, 51, 369–382.
Spakman, W., S. van der Lee, and R. van der Hilst (1993), Travel‐time
tomography of the European‐Mediterranean mantle down to 1400 km,
Phys. Earth Planet. Int., 79, 3–74.
Talwani, M., and J. R. Heirtzler (1964), Computation of magnetic anoma-
lies caused by two‐dimensional bodies of arbitrary shape, in Computers
in the Mineral Industries (Part 1), vol. 9, edited by G. A. Parks, pp. 464–
480, Stanford Univ. Publ. Geological Sciences, Stanford, Cal.
Tansi, C., F. Muto, S. Critelli, and G. Iovine (2007), Neogene‐Quaternary
strike‐slip tectonics in the central Calabrian Arc (southern Italy), J. Geo-
dyn., 43, 393–414.
Tortorici, L., C. Monaco, C. Tansi, and O. Cocina (1995), Recent and
active tectonics in the Calabrian arc (southern Italy), Tectonophysics,
243, 37–55.
Tortorici, G., M. Bianca, G. de Guidi, C. Monaco, and L. Tortorici (2003),
Fault activity and marine terracing in the Capo Vaticano area (southern
Calabria) during the Middle‐Late Quaternary, Quat. Int., 101, 269–278.
Trincardi, F., M. Cipolli, P. Ferretti, J. La Morgia, M. Ligi, G. Marozzi,
V. Palumbo, M. Taviani, and N. Zitellini (1987), Slope basin evolution
on the Eastern Tyrrhenian margin: Preliminary report, G. Geol., 49, 1–9.
Van Dijk, J. P. (1994), Late Neogene kinematics of intra‐arc oblique shear
zones: The Petilia–Rizzuto Fault Zone (Calabrian Arc, central Mediterra-
nean), Tectonics 13, 1201–1230.
Van Dijk, J. P., and M. Okkes (1991), Neogene tectonostratigraphy and
kinematics of Calabrian basin: Implications for the geodynamics of the
central Mediterranean, Tectonophysics, 196, 23–60.
Westaway, R. (1993), Quaternary uplift of southern Italy, J. Geophys. Res.,
98, 21741–21772.
Won, I. J., and M. Bevis (1987), Computing the gravitational and mag-
netic anomalies due to a polygon: Algorithms and Fortran subroutines,
Geophysics 52, 232–238.
Zanella, E. (1995), Studio delle variazioni paleosecolari del campo magne-
tico terrestre registrate nelle vulcaniti Quaternarie dell’area tirrenica e del
canale di Sicilia, PhD Thesis, 198 pp., UniversitĂ  di Torino, Italy.
M. Chiappini, R. De Ritis, I. Nicolosi, F. Speranza, and G. Ventura,
Instituto Nazionale di Geofisica e Vulcanologia, via di Vigna Murata
605, 00143 Rome, Italy. (riccardo.deritis@ingv.it)
R. De Rosa, P. Donato, R. Dominici, and M. Sonnino, Department of
Earth Science, UniversitĂ  della Calabria, 87036 Arcavacata di Rende, Italy.
DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101
B11101
18 of 18

More Related Content

Similar to A Buried Volcano In The Calabrian Arc (Italy) Revealed By High-Resolution Aeromagnetic Data

articoloBSGI_SileoEtAl_2007
articoloBSGI_SileoEtAl_2007articoloBSGI_SileoEtAl_2007
articoloBSGI_SileoEtAl_2007Giancanio Sileo
 
Festa et al., 2013 - Draft
Festa et al., 2013 - DraftFesta et al., 2013 - Draft
Festa et al., 2013 - DraftGianvito Teofilo
 
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...iosrjce
 
Moin CR Italy_Gath LR
Moin CR Italy_Gath LRMoin CR Italy_Gath LR
Moin CR Italy_Gath LREldon Gath
 
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİAli Osman Öncel
 
Marmara Sea: Tarihsel Deprem SenaryolarÄą
Marmara Sea: Tarihsel Deprem SenaryolarÄąMarmara Sea: Tarihsel Deprem SenaryolarÄą
Marmara Sea: Tarihsel Deprem SenaryolarıAli Osman Öncel
 
A research on resistivity measurements on wadi el nakhel
A research on  resistivity measurements on wadi el nakhelA research on  resistivity measurements on wadi el nakhel
A research on resistivity measurements on wadi el nakhelHussein Abd Elhafeez
 
A Morphological and Spatial Analysis of Volcanoes on Venus
A Morphological and Spatial Analysis of Volcanoes on VenusA Morphological and Spatial Analysis of Volcanoes on Venus
A Morphological and Spatial Analysis of Volcanoes on VenusSĂŠrgio Sacani
 
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİAli Osman Öncel
 
Simulation of tsunami generation, propagation and coastal inundation in the E...
Simulation of tsunami generation, propagation and coastal inundation in the E...Simulation of tsunami generation, propagation and coastal inundation in the E...
Simulation of tsunami generation, propagation and coastal inundation in the E...Jose Avila De Tomas
 
Josh Matthew's Project In Science
Josh Matthew's Project In ScienceJosh Matthew's Project In Science
Josh Matthew's Project In Sciencevhidha
 
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...iosrjce
 
Subduction Zones and their Associated Features
Subduction Zones and their Associated FeaturesSubduction Zones and their Associated Features
Subduction Zones and their Associated FeaturesRuhr Universität Bochum
 
Topographic Analysis Linkages among Climate, Erosion and Tectonics
Topographic Analysis Linkages among Climate, Erosion and TectonicsTopographic Analysis Linkages among Climate, Erosion and Tectonics
Topographic Analysis Linkages among Climate, Erosion and TectonicsShahadat Hossain Shakil
 
jennifer_digiulio_56x36in_poster
jennifer_digiulio_56x36in_posterjennifer_digiulio_56x36in_poster
jennifer_digiulio_56x36in_posterJennifer DiGiulio
 
Fluvial Features of Mars
Fluvial Features of MarsFluvial Features of Mars
Fluvial Features of MarsNam Le
 
The Geology of North West Milos, Greece.
The Geology of North West Milos, Greece.The Geology of North West Milos, Greece.
The Geology of North West Milos, Greece.John Love
 
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...Dario Chisari
 

Similar to A Buried Volcano In The Calabrian Arc (Italy) Revealed By High-Resolution Aeromagnetic Data (20)

articoloBSGI_SileoEtAl_2007
articoloBSGI_SileoEtAl_2007articoloBSGI_SileoEtAl_2007
articoloBSGI_SileoEtAl_2007
 
Festa et al., 2013 - Draft
Festa et al., 2013 - DraftFesta et al., 2013 - Draft
Festa et al., 2013 - Draft
 
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...
An Integrated Study of Gravity and Magnetic Data to Determine Subsurface Stru...
 
Moin CR Italy_Gath LR
Moin CR Italy_Gath LRMoin CR Italy_Gath LR
Moin CR Italy_Gath LR
 
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
 
Marmara Sea: Tarihsel Deprem SenaryolarÄą
Marmara Sea: Tarihsel Deprem SenaryolarÄąMarmara Sea: Tarihsel Deprem SenaryolarÄą
Marmara Sea: Tarihsel Deprem SenaryolarÄą
 
Ai36204209
Ai36204209Ai36204209
Ai36204209
 
Wilson200722222222222
Wilson200722222222222Wilson200722222222222
Wilson200722222222222
 
A research on resistivity measurements on wadi el nakhel
A research on  resistivity measurements on wadi el nakhelA research on  resistivity measurements on wadi el nakhel
A research on resistivity measurements on wadi el nakhel
 
A Morphological and Spatial Analysis of Volcanoes on Venus
A Morphological and Spatial Analysis of Volcanoes on VenusA Morphological and Spatial Analysis of Volcanoes on Venus
A Morphological and Spatial Analysis of Volcanoes on Venus
 
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİÖNCEL AKADEMİ: İSTANBUL DEPREMİ
ÖNCEL AKADEMİ: İSTANBUL DEPREMİ
 
Simulation of tsunami generation, propagation and coastal inundation in the E...
Simulation of tsunami generation, propagation and coastal inundation in the E...Simulation of tsunami generation, propagation and coastal inundation in the E...
Simulation of tsunami generation, propagation and coastal inundation in the E...
 
Josh Matthew's Project In Science
Josh Matthew's Project In ScienceJosh Matthew's Project In Science
Josh Matthew's Project In Science
 
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...
Structural Analysis and Tectonic Evolution based on Seismic Interpretation in...
 
Subduction Zones and their Associated Features
Subduction Zones and their Associated FeaturesSubduction Zones and their Associated Features
Subduction Zones and their Associated Features
 
Topographic Analysis Linkages among Climate, Erosion and Tectonics
Topographic Analysis Linkages among Climate, Erosion and TectonicsTopographic Analysis Linkages among Climate, Erosion and Tectonics
Topographic Analysis Linkages among Climate, Erosion and Tectonics
 
jennifer_digiulio_56x36in_poster
jennifer_digiulio_56x36in_posterjennifer_digiulio_56x36in_poster
jennifer_digiulio_56x36in_poster
 
Fluvial Features of Mars
Fluvial Features of MarsFluvial Features of Mars
Fluvial Features of Mars
 
The Geology of North West Milos, Greece.
The Geology of North West Milos, Greece.The Geology of North West Milos, Greece.
The Geology of North West Milos, Greece.
 
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...
A Comparison of Structural Styles and Prospectivity along the Atlantic Margin...
 

More from Andrea Porter

Taking Notes With Note Cards
Taking Notes With Note CardsTaking Notes With Note Cards
Taking Notes With Note CardsAndrea Porter
 
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHow
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHowHow To Write A TOK Essay 15 Steps (With Pictures) - WikiHow
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHowAndrea Porter
 
How To Write A Good Topic Sentence In Academic Writing
How To Write A Good Topic Sentence In Academic WritingHow To Write A Good Topic Sentence In Academic Writing
How To Write A Good Topic Sentence In Academic WritingAndrea Porter
 
Narrative Essay Argumentative Essay Thesis
Narrative Essay Argumentative Essay ThesisNarrative Essay Argumentative Essay Thesis
Narrative Essay Argumentative Essay ThesisAndrea Porter
 
A Good Introduction For A Research Paper. Writing A G
A Good Introduction For A Research Paper. Writing A GA Good Introduction For A Research Paper. Writing A G
A Good Introduction For A Research Paper. Writing A GAndrea Porter
 
How To Find The Best Paper Writing Company
How To Find The Best Paper Writing CompanyHow To Find The Best Paper Writing Company
How To Find The Best Paper Writing CompanyAndrea Porter
 
Primary Handwriting Paper - Paging Supermom - Free Printable Writing
Primary Handwriting Paper - Paging Supermom - Free Printable WritingPrimary Handwriting Paper - Paging Supermom - Free Printable Writing
Primary Handwriting Paper - Paging Supermom - Free Printable WritingAndrea Porter
 
Write An Essay On College Experience In EnglishDesc
Write An Essay On College Experience In EnglishDescWrite An Essay On College Experience In EnglishDesc
Write An Essay On College Experience In EnglishDescAndrea Porter
 
Ghost Writing KennyLeeHolmes.Com
Ghost Writing KennyLeeHolmes.ComGhost Writing KennyLeeHolmes.Com
Ghost Writing KennyLeeHolmes.ComAndrea Porter
 
How To Write A Good Concluding Observation Bo
How To Write A Good Concluding Observation BoHow To Write A Good Concluding Observation Bo
How To Write A Good Concluding Observation BoAndrea Porter
 
How To Write An Argumentative Essay
How To Write An Argumentative EssayHow To Write An Argumentative Essay
How To Write An Argumentative EssayAndrea Porter
 
7 Introduction Essay About Yourself - Introduction Letter
7 Introduction Essay About Yourself - Introduction Letter7 Introduction Essay About Yourself - Introduction Letter
7 Introduction Essay About Yourself - Introduction LetterAndrea Porter
 
Persuasive Essay Essaypro Writers
Persuasive Essay Essaypro WritersPersuasive Essay Essaypro Writers
Persuasive Essay Essaypro WritersAndrea Porter
 
Scholarship Personal Essays Templates At Allbusinesst
Scholarship Personal Essays Templates At AllbusinesstScholarship Personal Essays Templates At Allbusinesst
Scholarship Personal Essays Templates At AllbusinesstAndrea Porter
 
How To Create An Outline For An Essay
How To Create An Outline For An EssayHow To Create An Outline For An Essay
How To Create An Outline For An EssayAndrea Porter
 
ESSAY - Qualities Of A Good Teach
ESSAY - Qualities Of A Good TeachESSAY - Qualities Of A Good Teach
ESSAY - Qualities Of A Good TeachAndrea Porter
 
Fountain Pen Writing On Paper - Castle Rock Financial Planning
Fountain Pen Writing On Paper - Castle Rock Financial PlanningFountain Pen Writing On Paper - Castle Rock Financial Planning
Fountain Pen Writing On Paper - Castle Rock Financial PlanningAndrea Porter
 
Formatting A Research Paper E
Formatting A Research Paper EFormatting A Research Paper E
Formatting A Research Paper EAndrea Porter
 
Business Paper Examples Of Graduate School Admissio
Business Paper Examples Of Graduate School AdmissioBusiness Paper Examples Of Graduate School Admissio
Business Paper Examples Of Graduate School AdmissioAndrea Porter
 
Impact Of Poverty On The Society - Free Essay E
Impact Of Poverty On The Society - Free Essay EImpact Of Poverty On The Society - Free Essay E
Impact Of Poverty On The Society - Free Essay EAndrea Porter
 

More from Andrea Porter (20)

Taking Notes With Note Cards
Taking Notes With Note CardsTaking Notes With Note Cards
Taking Notes With Note Cards
 
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHow
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHowHow To Write A TOK Essay 15 Steps (With Pictures) - WikiHow
How To Write A TOK Essay 15 Steps (With Pictures) - WikiHow
 
How To Write A Good Topic Sentence In Academic Writing
How To Write A Good Topic Sentence In Academic WritingHow To Write A Good Topic Sentence In Academic Writing
How To Write A Good Topic Sentence In Academic Writing
 
Narrative Essay Argumentative Essay Thesis
Narrative Essay Argumentative Essay ThesisNarrative Essay Argumentative Essay Thesis
Narrative Essay Argumentative Essay Thesis
 
A Good Introduction For A Research Paper. Writing A G
A Good Introduction For A Research Paper. Writing A GA Good Introduction For A Research Paper. Writing A G
A Good Introduction For A Research Paper. Writing A G
 
How To Find The Best Paper Writing Company
How To Find The Best Paper Writing CompanyHow To Find The Best Paper Writing Company
How To Find The Best Paper Writing Company
 
Primary Handwriting Paper - Paging Supermom - Free Printable Writing
Primary Handwriting Paper - Paging Supermom - Free Printable WritingPrimary Handwriting Paper - Paging Supermom - Free Printable Writing
Primary Handwriting Paper - Paging Supermom - Free Printable Writing
 
Write An Essay On College Experience In EnglishDesc
Write An Essay On College Experience In EnglishDescWrite An Essay On College Experience In EnglishDesc
Write An Essay On College Experience In EnglishDesc
 
Ghost Writing KennyLeeHolmes.Com
Ghost Writing KennyLeeHolmes.ComGhost Writing KennyLeeHolmes.Com
Ghost Writing KennyLeeHolmes.Com
 
How To Write A Good Concluding Observation Bo
How To Write A Good Concluding Observation BoHow To Write A Good Concluding Observation Bo
How To Write A Good Concluding Observation Bo
 
How To Write An Argumentative Essay
How To Write An Argumentative EssayHow To Write An Argumentative Essay
How To Write An Argumentative Essay
 
7 Introduction Essay About Yourself - Introduction Letter
7 Introduction Essay About Yourself - Introduction Letter7 Introduction Essay About Yourself - Introduction Letter
7 Introduction Essay About Yourself - Introduction Letter
 
Persuasive Essay Essaypro Writers
Persuasive Essay Essaypro WritersPersuasive Essay Essaypro Writers
Persuasive Essay Essaypro Writers
 
Scholarship Personal Essays Templates At Allbusinesst
Scholarship Personal Essays Templates At AllbusinesstScholarship Personal Essays Templates At Allbusinesst
Scholarship Personal Essays Templates At Allbusinesst
 
How To Create An Outline For An Essay
How To Create An Outline For An EssayHow To Create An Outline For An Essay
How To Create An Outline For An Essay
 
ESSAY - Qualities Of A Good Teach
ESSAY - Qualities Of A Good TeachESSAY - Qualities Of A Good Teach
ESSAY - Qualities Of A Good Teach
 
Fountain Pen Writing On Paper - Castle Rock Financial Planning
Fountain Pen Writing On Paper - Castle Rock Financial PlanningFountain Pen Writing On Paper - Castle Rock Financial Planning
Fountain Pen Writing On Paper - Castle Rock Financial Planning
 
Formatting A Research Paper E
Formatting A Research Paper EFormatting A Research Paper E
Formatting A Research Paper E
 
Business Paper Examples Of Graduate School Admissio
Business Paper Examples Of Graduate School AdmissioBusiness Paper Examples Of Graduate School Admissio
Business Paper Examples Of Graduate School Admissio
 
Impact Of Poverty On The Society - Free Essay E
Impact Of Poverty On The Society - Free Essay EImpact Of Poverty On The Society - Free Essay E
Impact Of Poverty On The Society - Free Essay E
 

Recently uploaded

What is Model Inheritance in Odoo 17 ERP
What is Model Inheritance in Odoo 17 ERPWhat is Model Inheritance in Odoo 17 ERP
What is Model Inheritance in Odoo 17 ERPCeline George
 
ENGLISH6-Q4-W3.pptxqurter our high choom
ENGLISH6-Q4-W3.pptxqurter our high choomENGLISH6-Q4-W3.pptxqurter our high choom
ENGLISH6-Q4-W3.pptxqurter our high choomnelietumpap1
 
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...Nguyen Thanh Tu Collection
 
ACC 2024 Chronicles. Cardiology. Exam.pdf
ACC 2024 Chronicles. Cardiology. Exam.pdfACC 2024 Chronicles. Cardiology. Exam.pdf
ACC 2024 Chronicles. Cardiology. Exam.pdfSpandanaRallapalli
 
Difference Between Search & Browse Methods in Odoo 17
Difference Between Search & Browse Methods in Odoo 17Difference Between Search & Browse Methods in Odoo 17
Difference Between Search & Browse Methods in Odoo 17Celine George
 
Full Stack Web Development Course for Beginners
Full Stack Web Development Course  for BeginnersFull Stack Web Development Course  for Beginners
Full Stack Web Development Course for BeginnersSabitha Banu
 
How to Configure Email Server in Odoo 17
How to Configure Email Server in Odoo 17How to Configure Email Server in Odoo 17
How to Configure Email Server in Odoo 17Celine George
 
Influencing policy (training slides from Fast Track Impact)
Influencing policy (training slides from Fast Track Impact)Influencing policy (training slides from Fast Track Impact)
Influencing policy (training slides from Fast Track Impact)Mark Reed
 
Roles & Responsibilities in Pharmacovigilance
Roles & Responsibilities in PharmacovigilanceRoles & Responsibilities in Pharmacovigilance
Roles & Responsibilities in PharmacovigilanceSamikshaHamane
 
Field Attribute Index Feature in Odoo 17
Field Attribute Index Feature in Odoo 17Field Attribute Index Feature in Odoo 17
Field Attribute Index Feature in Odoo 17Celine George
 
Proudly South Africa powerpoint Thorisha.pptx
Proudly South Africa powerpoint Thorisha.pptxProudly South Africa powerpoint Thorisha.pptx
Proudly South Africa powerpoint Thorisha.pptxthorishapillay1
 
Introduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxIntroduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxpboyjonauth
 
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptx
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptxMULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptx
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptxAnupkumar Sharma
 
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPT
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPTECONOMIC CONTEXT - LONG FORM TV DRAMA - PPT
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPTiammrhaywood
 
Alper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentAlper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentInMediaRes1
 
Computed Fields and api Depends in the Odoo 17
Computed Fields and api Depends in the Odoo 17Computed Fields and api Depends in the Odoo 17
Computed Fields and api Depends in the Odoo 17Celine George
 

Recently uploaded (20)

What is Model Inheritance in Odoo 17 ERP
What is Model Inheritance in Odoo 17 ERPWhat is Model Inheritance in Odoo 17 ERP
What is Model Inheritance in Odoo 17 ERP
 
ENGLISH6-Q4-W3.pptxqurter our high choom
ENGLISH6-Q4-W3.pptxqurter our high choomENGLISH6-Q4-W3.pptxqurter our high choom
ENGLISH6-Q4-W3.pptxqurter our high choom
 
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...
HỌC TỐT TIẾNG ANH 11 THEO CHƯƠNG TRÌNH GLOBAL SUCCESS ĐÁP ÁN CHI TIẾT - CẢ NĂ...
 
ACC 2024 Chronicles. Cardiology. Exam.pdf
ACC 2024 Chronicles. Cardiology. Exam.pdfACC 2024 Chronicles. Cardiology. Exam.pdf
ACC 2024 Chronicles. Cardiology. Exam.pdf
 
Difference Between Search & Browse Methods in Odoo 17
Difference Between Search & Browse Methods in Odoo 17Difference Between Search & Browse Methods in Odoo 17
Difference Between Search & Browse Methods in Odoo 17
 
Full Stack Web Development Course for Beginners
Full Stack Web Development Course  for BeginnersFull Stack Web Development Course  for Beginners
Full Stack Web Development Course for Beginners
 
How to Configure Email Server in Odoo 17
How to Configure Email Server in Odoo 17How to Configure Email Server in Odoo 17
How to Configure Email Server in Odoo 17
 
Influencing policy (training slides from Fast Track Impact)
Influencing policy (training slides from Fast Track Impact)Influencing policy (training slides from Fast Track Impact)
Influencing policy (training slides from Fast Track Impact)
 
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
Model Call Girl in Tilak Nagar Delhi reach out to us at 🔝9953056974🔝
 
Roles & Responsibilities in Pharmacovigilance
Roles & Responsibilities in PharmacovigilanceRoles & Responsibilities in Pharmacovigilance
Roles & Responsibilities in Pharmacovigilance
 
Field Attribute Index Feature in Odoo 17
Field Attribute Index Feature in Odoo 17Field Attribute Index Feature in Odoo 17
Field Attribute Index Feature in Odoo 17
 
Proudly South Africa powerpoint Thorisha.pptx
Proudly South Africa powerpoint Thorisha.pptxProudly South Africa powerpoint Thorisha.pptx
Proudly South Africa powerpoint Thorisha.pptx
 
Introduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptxIntroduction to AI in Higher Education_draft.pptx
Introduction to AI in Higher Education_draft.pptx
 
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptx
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptxMULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptx
MULTIDISCIPLINRY NATURE OF THE ENVIRONMENTAL STUDIES.pptx
 
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdfTataKelola dan KamSiber Kecerdasan Buatan v022.pdf
TataKelola dan KamSiber Kecerdasan Buatan v022.pdf
 
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPT
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPTECONOMIC CONTEXT - LONG FORM TV DRAMA - PPT
ECONOMIC CONTEXT - LONG FORM TV DRAMA - PPT
 
Alper Gobel In Media Res Media Component
Alper Gobel In Media Res Media ComponentAlper Gobel In Media Res Media Component
Alper Gobel In Media Res Media Component
 
9953330565 Low Rate Call Girls In Rohini Delhi NCR
9953330565 Low Rate Call Girls In Rohini  Delhi NCR9953330565 Low Rate Call Girls In Rohini  Delhi NCR
9953330565 Low Rate Call Girls In Rohini Delhi NCR
 
Raw materials used in Herbal Cosmetics.pptx
Raw materials used in Herbal Cosmetics.pptxRaw materials used in Herbal Cosmetics.pptx
Raw materials used in Herbal Cosmetics.pptx
 
Computed Fields and api Depends in the Odoo 17
Computed Fields and api Depends in the Odoo 17Computed Fields and api Depends in the Odoo 17
Computed Fields and api Depends in the Odoo 17
 

A Buried Volcano In The Calabrian Arc (Italy) Revealed By High-Resolution Aeromagnetic Data

  • 1. A buried volcano in the Calabrian Arc (Italy) revealed by high‐resolution aeromagnetic data R. De Ritis,1 R. Dominici,2 G. Ventura,1 I. Nicolosi,1 M. Chiappini,1 F. Speranza,1 R. De Rosa,2 P. Donato,2 and M. Sonnino2 Received 26 November 2009; revised 8 June 2010; accepted 18 June 2010; published 4 November 2010. [1] Aeromagnetic data collected between the Aeolian volcanoes (southern Tyrrhenian Sea) and the Calabrian Arc (Italy) highlight a WNW‐ESE elongated positive magnetic anomaly centered on the Capo Vaticano morphological ridge (Tyrrhenian coast of Calabria), characterized by an apical, subcircular, flat surface. Results of forward and inverse modeling of the magnetic data show a 20 km long and 3–5 km wide magnetized body that extends from sea floor to about 3 km below sea level. The magnetic properties of this body are consistent with those of the medium to highly evolved volcanic rocks of the Aeolian Arc (i.e., dacites and rhyolites). In the Calabria mainland, widespread dacitic to rhyolitic pumices with calc‐alkaline affinity of Pleistocene age (1–0.7 Ma) are exposed. The tephra falls are related to explosive activity and show a decreasing thickness from the Capo Vaticano area southeastward. The presence of lithics indicates a provenance from a source located not far from Capo Vaticano. The combined interpretation of the magnetic and available geological data reveal that (1) the Capo Vaticano WNW‐ESE elongated positive magnetic anomaly is due to the occurrence of a WNW‐ESE elongated sill; (2) such a sill represents the remnant of the plumbing system of a Pleistocene volcano that erupted explosively producing the pumice tephra exposed in Calabria; and (3) the volcanism is consistent with the Aeolian products, in terms of age, magnetic signature, and geochemical affinity of the erupted products,. The results indicate that such volcanism developed along seismically active faults transversal to the general trend of the Aeolian Arc and Calabria block, in an area where uplift is maximized (∟4 mm/yr). Such uplift could also be responsible for fragmentation of the upper crust and formation of transversal faults along which seismic activity and volcanism occur. Citation: De Ritis, R., R. Dominici, G. Ventura, I. Nicolosi, M. Chiappini, F. Speranza, R. De Rosa, P. Donato, and M. Sonnino (2010), A buried volcano in the Calabrian Arc (Italy) revealed by high‐resolution aeromagnetic data, J. Geophys. Res., 115, B11101, doi:10.1029/2009JB007171. 1. Introduction [2] Aeromagnetic surveying is a relatively new technique for studying volcanic or active tectonic areas through improve- ments in data acquisition, GPS measurements, and data‐ processing methods [e.g., Finn et al., 2001; Lenat et al., 2001; Chiappini et al., 2002; De Ritis et al., 2005]. In volcanic areas, aeromagnetic surveys allow researchers to decipher the inner structure of eruptive centers and, in marine environ- ments, to detect submerged, hidden volcanoes and vents [Rollin et al., 2000; Blanco‐Montenegro et al., 2007; De Ritis et al., 2007]. Due to the occurrence of widespread vol- canic deposits of uncertain origin in southern Calabria (Italy), particularly in the Capo Vaticano (CV) promontory [De Rosa et al., 2001, 2008], a recent high‐resolution aeromagnetic survey has been conducted by the Airborne Geophysics Sci- ence Team of Istituto Nazionale di Geofisica e Vulcano- logia (INGV), covering the offshore and onshore areas between CV and the calc‐alkaline to shoshonitic Stromboli and Panarea volcanoes of the Aeolian Islands (Figures 1 and 2). Results of the aeromagnetic survey reveal a 25 km WNW‐ESE elongated magnetic anomaly covering the CV on land and in offshore areas. This anomaly is noteworthy to be investigated because of (1) the vicinity of the Aeolian volcanic Arc, (2) the lack of high‐susceptibility geologic units outcropping in the CV area, and (3) the presence of volcanic products of unknown origin outcropping in western Calabria and mainly on CV. A geological model has been built on the basis of inverse and forward magnetic models together with the geotectonic and seismic knowledge of the area. Inversion of magnetic data set was carried out to define a 3‐D anomalous susceptibility contrast model. To further constrain the obtained model, a 2.75‐D forward modeling was also made that included the lithological and structural 1 Istituto Nazionale di Geofisica e Vulcanologia, Rome, Italy. 2 Department of Earth Science, UniversitĂ  della Calabria, Arcavacata di Rende, Italy. Copyright 2010 by the American Geophysical Union. 0148‐0227/10/2009JB007171 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B11101, doi:10.1029/2009JB007171, 2010 B11101 1 of 18
  • 2. information of the study area. Rock magnetic properties have been measured on samples representative of the main local lithological units. The results highlight the presence of a previously unknown volcanic body emplaced along the main fault located at the southern edge of the CV prom- ontory. This body could represent the source area of the widespread volcanic deposits exposed in CV and western Calabria. The results, discussed in light of the available geological and geophysical information, shed new light on (1) the tectonics of the Calabrian Arc and (2) the general control of preexisting structures and, in particular, of trans- versal structures in subduction setting on the emplacement of volcanoes. 2. Geodynamic and Geological Setting 2.1. Structure [3] The Calabrian Arc, Sicilian Maghrebides, and Apen- nines are mountain belts bounding the southern and eastern sector of the Tyrrhenian Sea back‐arc basin (Figure 1). The evolution of the Calabrian Arc and Tyrrhenian Sea back arc during Neogene and Quaternary periods was driven by the southeastward retreat of the Ionian slab [Malinverno and Ryan, 1986; Jolivet and Faccenna, 2000; Faccenna et al., 2001]. The Calabrian Arc is presently located on the top of a 200 km wide, subducting slab dipping 70° toward NE (Figure 1) [Spakman et al., 1993; Chiarabba et al., 2008; Neri et al., 2009]. Earthquake distribution and seismic tomog- raphy indicate that the northeastern boundary of the sub- ducting slab roughly corresponds on the surface to the boundary between southern Apennines and Calabrian Arc [Chiarabba et al., 2008, and references therein]. The western boundary of the slab occurs in correspondence of the NNW‐ SSE‐striking Tindari‐Letojanni fault system, which crosses the eastern margin of Sicily (Figure 1) [Barberi et al., 1994; De Astis et al., 2003]. This fault system constitutes the northern extent of the Malta Escarpment, which represents a transtensional (right‐lateral) transfer zone between the con- tinental crust of eastern Sicily (west) and the oceanic Ionian basin (east) [Doglioni et al., 2001, and references therein]. The Malta Escarpment transfer zone, along which the Mt. Etna volcano is emplaced, is characterized by vertical slip‐ rates of up to 2 mm/yr [Azzaro et al., 2000]; its movement is related to the larger retreat of the subduction hinge in the Ionian Sea with respect to the Sicily mainland. [4] The Aeolian Arc, which forms a ring‐like chain of volcanic islands and seamounts in the southern margin of the southern Tyrrhenian Sea oceanic basin (Marsili basin) [Nicolosi et al., 2006], formed during the last 1 Myr, and the volcanoes of Lipari, Vulcano, Panarea, and Stromboli Figure 1. Tectonic scheme of Italy (modified from Chiarabba et al. [2008]; Rosenbaum et al. [2008]). Dashed lines with numbers identify the depth of the slab below the Calabrian Arc as deduced from seismic data. TL, the Tindari‐Letojianni fault system; ME, the Malta Escarpment. Lines with triangles indicate reverse faults; lines with perpendicular marks indicate normal faults. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 2 of 18
  • 3. are still active today (Figure 2a). The evolution of the Aeolian volcanism, which includes calc‐alkaline to shosh- onitic and K‐alkaline products, has been related to both the subduction of the Ionian lithosphere [Chiarabba et al., 2008] and a postcollisional, rift‐type volcanism associated with the extensional tectonics that have affected the Calab- rian Arc in last 1–0.7 Myr [De Astis et al., 2003, and refer- ences therein]. Recently, Rosenbaum et al. [2008] also proposed that slab tear faults, which reflect the segmentation of the Apennines–Calabrian Arc–Tyrrhenian Sea subduction, control the emplacement of some Quaternary volcanoes and identify, in the Aeolian Islands–Calabrian Arc region, the Tindari‐Letojanni fault system as one of these slab tear faults (Figure 1). The internal structure and recent tectonics of the Calabria block is debated, and two main end‐member models have been proposed in the last years: (1) Calabria is frag- mented into several blocks undergoing differential displace- ments toward the trench and separated by NW‐SE striking faults [Knott and Turco, 1991; Van Dijk, 1994; Guarnieri, 2006; Tansi et al., 2007; Del Ben et al., 2008]; (2) alterna- tively, the Calabria block is solely characterized by exten- sional faulting, as syn‐sedimentary extensional features are documented in the whole middle to upper Miocene to Pleis- tocene sedimentary succession [Mattei et al., 1999, 2002; Monaco and Tortorici, 2000; Cifelli et al., 2007a, 2007b]. [5] The Calabrian Fore‐arc basin (CFB), located in the central‐southern sector of the Calabrian Arc, is composed by several NE‐SW elongated basins fed by sediments derived from Serre, Poro, and Aspromonte crystalline Mas- sifs (Figure 2b). The geometry of these basins is controlled by late Pliocene–Holocene NE‐SW and minor NW‐SE striking faults. CFB consists of different allochthonous struc- tural units (Figure 2b). A continuous crustal section from granulite grade through granitoids and amphibolite‐facies gneisses into unmetamorphosed Paleozoic and younger sediments crop out in the Serre Massif [Schenk, 1984, 1990; Amodio Morelli et al., 1976]. The granulite‐facies is over- lain by gneiss and paragneiss with granofels and siliceous marble (Polia‐Copanello unit). In the middle sector of the Serre and Poro Massifs, granitoids crop out over an area of about 1250 km2 . The granitoids overlie the Polia‐Copanello unit and intrude into the overlying ortho‐ and paragneisses (Aspromonte unit) and lower‐grade Paleozoic rocks (Stilo unit) [Schenk, 1984; Ayuso et al., 1994; Graessner and Schenk, 2001; Caggianelli and Prosser, 2002]. Along the south- eastern side of the Serre Massif, the granites come into contact with Mammola and Pazzano complexes [Colonna et al., 1973], which are made up by micaschists and para- gneisses with interbedded Ordovician tholeiitic metabasalts [Acquafredda et al., 1994]. The Pazzano complex consists of phyllites, with interbedded graywackes and marbles [Bouillin et al., 1984, 1987; Acquafredda et al., 1987, 1989]. A Mesozoic sequence made up of dolostone, calcirudite and calcarenite, breccia, and reef limestone [Roda, 1965; Bonardi et al., 1984] lies on the Stilo–Pazzano and Mammola com- plexes. Along the Ionian coast of Calabria, a succession of latest Chattian–Quaternary age rocks [Patterson et al., 1995] overlies the Stilo unit and Mesozoic carbonates. The Stilo–Capo d’Orlando formation is composed of red con- glomerate, sandstone, and pelites [Cavazza, 1989; Cavazza and DeCelles, 1993]. The unit is conformably overlain by a chaotic melange of pelitic matrix that encloses olistoliths of calciturbidites and Chattian–early Miocene quartzarenitic turbidites [Cavazza et al., 1997; Bonardi et al., 2001]. [6] The Serravallian–Tortonian sedimentary sequence con- sists of continental conglomerate, marine sandstone, marl, and shale, as well as minor shales, Messinian limestone, and gypsum. This unit is overlain by alluvial conglomerate and discontinuous shallow‐marine to continental sandstone and pelite. The upper portion of the basin filling is represented by marls of the Trubi formation overlain by shallow‐marine and transitional‐continental facies of late Pliocene–Quaternary age. Figure 2. (a) Location of the Capo Vaticano ridge. (b) Geological map of southern Calabria showing outcrops of the volcaniclastic deposits (stars) and WNW‐ESE geological section (profile A‐B). The bathymetry of the Capo Vaticano ridge is reported along with offshore active faults [data are from Cifelli et al., 2007a; De Rosa et al., 2008; Tortorici et al., 2003; and Milia et al., 2009). (c) Southern Calabria and northeastern Sicily magnetic anomaly field sketch map from Caratori Tontini et al. [2004]. The black dotted line is the geological profile shown in Figure 2b. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 3 of 18
  • 4. Figure 2. (continued) DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 4 of 18
  • 5. [7] Several orders of marine terraces distributed from about 50 up to 1,000 m above sea level (asl) indicate an uplift of the southern and western sectors of Calabria from lower to middle Pleistocene times [Ghisetti, 1981; Dumas et al., 1982; Mulargia et al., 1984; Westaway, 1993; Miyauchi et al., 1994; Tortorici et al., 1995; Antonioli et al., 2006]. The estimated uplift velocity is generally between 1.0 and 2.1 mm/yr in the last 124 kyr. In the CV area, however, the uplift velocity has reached 4.0 mm/yr [Westaway, 1993; Miyauchi et al., 1994]. Westaway [1993] estimated an uplift velocity of the Calabria area of 1 mm/yr for the last 0.9 Myr. Normal, active faults trending WNW or NE affect the onshore Figure 2. (continued) DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 5 of 18
  • 6. and offshore sectors of the CV promontory. A major WNW‐ ESE fault shows late Quaternary slip‐rates of up to 2.5 mm/yr and is responsible for the northeastward tilting of the CV promontory (Figure 2b) [Tortorici et al., 2003]. This fault also controls the morphology of the CV southern coastline and its offshore sector. The CV offshore is characterized by a WNW‐ESE elongated ridge with a flat, subcircular trun- cated top at 200 m below sea level (bsl) (Figure 2b). A seismic profile crossing the northern flank of the ridge evidences a chaotic basement covered by a thin succession of sediments younger than 0.7 Myr [Milia et al., 2009]. Tomographic studies [Barberi et al., 2004] reveal the occurrence of a subcircular body located between 0 and 5 km depth with Vp/Vs values (1.8) higher than those of the surrounding areas (Vp/Vs < 1.6–1.7). This body is centered on the CV morphological ridge. 2.2. Volcaniclastic Sedimentation in the Calabria Fore‐Arc Basins [8] The CFB and CV sedimentary evolution is charac- terized by an important middle Pleistocene volcanic and volcaniclastic input. Volcaniclastic deposits consist of iso- lated pumice swarms, ash layers, and thick pumiceous lapilli sequences occurring as distinct lithofacies within sedimen- tary successions of shallow marine environments [De Rosa et al., 2001]. According to De Rosa et al. [2008], these sequences have the characteristics of proximal tephra fall deposits from Plinian columns reworked in shore environ- ments. In the Mesima basin and Gioaia Tauro plain, the deposits have reached 6 and 21 m in thickness, respectively (Figure 2b). In the CV area, the occurrence of impact sags in the volcanic deposits attests the proximity of the vent area. The compositional homogeneity of fragments (calc‐alkaline dacites and rhyolites) and the lack of nonvolcanic detritus, soils, or organic matter indicate that the deposits were quickly reworked after their primary emplacements [De Rosa et al., 2001, 2008]. K‐Ar determinations of the glass from the pumices give ages in the range 0.67–1.07 Myr [Cornette et al., 1987]. The mineralogical assemblage and the trace element geochemistry of the pumices are consistent with a provenance from magmatic‐arc volcanism [De Rosa et al., 2008]. The decrease in the maximum size of pumice from Mesima basin‐CV area to the Gioia Tauro plain and Reggio Calabria basin (Figure 2b) and the presence of lithics of lavas suggest a provenance from a source located off- shore, not far from the present CV promontory [De Rosa et al., 2001, 2008]. De Rosa et al. [2008] propose two hypotheses for the source areas of the pumice deposits of CFB and CV: [9] 1. A provenance from Panarea Island (Aeolian Arc) during an early highly explosive phase, leading to the col- lapse of the summit of the first edifice. This hypothesis remains to be confirmed and does not agree with the lack of rhyodacitic, pumiceous deposits on other Aeolian terrains with age >100 kyr. In addition, the occurrence of impact sags in the CV deposits [De Rosa et al., 2001] is not con- sistent with a source located at a distance of about 70 km (i.e., the distance between Panarea and the CV promontory). [10] 2. The tephra falls are related to the explosive activity of a “buried” calc‐alkaline volcano located in the Tyr- rhenian Sea. According to De Rosa et al. [2001, 2008], the CV pumice deposits could represent the only evidence of a large explosive eruption in the southern Tyrrhenian Sea, the very proximal equivalent of which is not preserved. 3. Regional Magnetic Anomaly Field [11] To get a reliable interpretation of the CV magnetic anomaly system, it is necessary to arrange the study area in the frame of the regional knowledge described in section 2. Geologic and magnetic field maps [Agip, 1981; Chiappini et al., 2000; Caratori Tontini et al., 2004; Finetti, 2005] provide a general outline of the geostructural characteristics and long‐wavelength magnetic anomalies (Figure 2c) of southern Calabria and the southern Tyrrhenian Sea. Fifty kilometers west of the CV, the Aeolian Arc (Panarea Island and Lamentini seamount; Figure 2a) is set up by strongly magnetic Holocene–Pleistocene volcanics [Zanella, 1995] overlying the barely magnetic units of the Calabrian Arc [Barberi et al., 1994]. Consistently, this western region is characterized by high‐intensity positive magnetic anoma- lies corresponding to the volcanic islands and seamounts [Chiappini et al., 2000; De Ritis et al., 2005]. Eastward, the Calabride units [Schenk, 1990], characterized by low to null magnetic anomaly values, are exposed in the CV and Le Serre horsts and in the Mesima graben (Figure 2b). Westward, in the CV offshore sector, the Calabrian units are intensely faulted and lowered by normal faults, which have produced the nearby Paola and Gioia Tauro sedimentary basins. These depressions are filled with as much as 5,000 m of middle Miocene to Pleistocene sedimentary sequences that can be considered of low susceptibility, according to measurements carried out on similar sediments exposed elsewhere in Italy [Sagnotti et al., 1998]. In fact, in this sector, marine terrigenous series above the Calabrian metamorphic units have been interpreted through seismic profiles and dril- lings starting from the Serravallian age [Milia et al., 2009]. A wide region characterized by low to negative values of the magnetic anomaly field exists as a result of this regional setting (Figure 2c). It extends from the Stromboli canyon to the Ionian Sea (E‐W) and from the Messina Strait to the Catanzaro plain (N‐S). [12] Toward North and Northeast, two long‐wavelength, low‐intensity linear positive anomalies develop parallel to the main regional tectonic features of the Catena Costiera and Sila (Figure 2c) [Van Dijk and Okkes, 1991]. The anomalies express the presence of contrasts characterized by very low magnetization values of the Calabrian Arc units. [13] Southward, the Etna long‐wavelength magnetic anom- aly negative lobe contributes to lowering of the regional magnetic anomalies field (Figure 2c). Negative lobes of other high‐intensity magnetic anomalies lying in the Stromboli canyon affect the CV magnetic anomaly system. By means of synthetic models, we estimated the total entity of this magnetic anomaly lowering to be in the range of 15–20 nT. [14] The CV high‐intensity positive magnetic anomaly (Figures 2c, 3a, and 3b) lies in correspondence of the CV morphological ridge (Figure 2b), highlighting the presence of a noticeable magnetization contrast in an area strongly affected by tectonic structures. [15] Because of the lack of susceptibility measurements available in the literature for the Calabrian Arc, we carried DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 6 of 18
  • 7. out sampling of the main lithological units of the area. Then, we measured rock sample susceptibility in the paleomag- netic laboratory of INGV (Rome) to get reliable modeling and effective interpretation of magnetic anomalies. In addi- tion, because crustal magnetic anomalies are originated by 3‐D variations of rock properties at depth, we reconstructed the stratigraphic sequence reported in the geological pro- file of Figure 2b and sampled the most representative and widespread lithologies of the region. Southern Calabrian structures, from the Ionian to the Tyrrhenian sides (NW‐SE direction), are composed by conglomerates, sandstones, and clays of the Calabride units (Tortonian to lower Pliocene), granites, granodiorites, and tonalities of the Aspromonte Peloritani units (Permian–Carboniferous), and terrigenous marine deposits and calcarenites lower Messinian to lower Pleistocene in age. The CV promontory, in particular, is made up by intrusive igneous rocks and terrigenous marine deposits (Figure 2b). Therefore, it is reasonable to infer that these units are the representative bedrock of the offshore study area, possibly lowered by the action of faults and covered by middle Miocene–Holocene sediments [Milia et al., 2009; Mattei et al., 2002]. [16] Mass susceptibilities of rock samples were measured at the INGV paleomagnetic laboratory of Rome. Subse- quently, volume susceptibilities were calculated by consid- ering average rock density values available in the literature (Table 1). Volume susceptibilities were determined for a total of 24 samples including 11 sedimentary samples of sandstone, sand, carbonate, and clay successions belonging to terrigenous marine deposits and marine terraces; 9 sam- ples of the metamorphic and plutonic complex such as the granites and phyllades of Le Serreand Mesima units; 3 vol- canic samples set up by pumices with different granulo- metries deriving from outcrops lying in the CV area, in the Gioia Tauro basin, and in the Mesima graben (Figure 2b); and 1 sample of metagabbro from an Oligocene shear zone separating nappes of the Hercinian basement [Langone et al., 2006]. [17] The sedimentary and metamorphic/plutonic speci- mens show very low volume susceptibility values (10−6 up Figure 3. (a) Total intensity magnetic anomalies of the Capo Vaticano (CV) and offshore areas. Dotted lines mark the Calabria coast line. Inset in Figure 3a represents the aeromagnetic survey flight lines (black) and the tie lines (red). (b) Reduced‐to‐the‐pole magnetic anomaly field assuming geomagnetic inclination of 90° and declination of 0°; the dotted red lines refer to the AB and CD magnetic profiles reported in Figures 5a and 5b. The main anomalies are highlighted with numbers from 1 to 5. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 7 of 18
  • 8. to 10−4 SI) (Table 1). The volcanic samples also exhibit low susceptibility (10−4 SI), probably because of their high silica content and clastic fabric. Our measurements are consistent with those summarized in Barberi et al. [1994], who report susceptibilities of the Calabrian Arc rocks less than 7 × 10−7 SI. An isolated exception is our metagabbro sample, which has a susceptibility value as high as 10−2 SI. However, this lithology forms small (decimetric to metric) lenses corresponding to a low‐intensity, long‐wavelength magnetic anomaly developing 15 km northwest of CV. We have not measured the magnetic remanence of the Calabrian Arc rocks because the rock susceptibility values are gener- ally very low, implying a very poor content of magnetite and thus negligible magnetic remanence. In addition, the rocks are Meso‐Cenozoic in age and are arranged in a chaotic tectonic setting with the result that their polarity will be mixed, and the reconstruction of the rock volumes with constant magnetic polarity it is not possible. The measure of remanence of these rocks has no practical bearings. More- over, in situ samples of the CV volcanics are also lacking and the on‐land outcropping pumices, which have been reworked in shallow marine environments, show SI values in the order of 10−4 (see Table 1), suggesting low magnetic attitude. In conclusion, the CV positive magnetic anomaly system represents a clear contrast between an unknown magnetic source and the virtually nonmagnetic rocks of the Calabrian Arc. 4. Data Acquisition and Processing [18] A high‐resolution aeromagnetic survey was recently carried out by the Airborne Geophysics Science Team of INGV for a detailed characterization of the offshore–onshore CV anomaly field. Thirty‐four measurement profiles were flown in the N‐S direction across the CV morphological ridge with a line spacing of about 2 km apart (Figure 3a, inset). Tie lines were also flown in the E‐W direction to level the survey lines. The survey area extends from the Mesima graben to the Panarea and Stromboli eastern marine areas. In the offshore sectors, the survey was flown at con- stant barometric height (500 m), whereas the profiles cross- ing the CV promontory were draped, reaching a maximum flight altitude of 1000 m. Total intensity anomaly data were acquired with an optically pumped cesium magnetometer at the sampling rate of 10 Hz. A GPS receiver in differential Figure 3. (continued) DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 8 of 18
  • 9. mode was used for accurate survey navigation, and a laser altimeter recorded the flight altitude. [19] We performed the reduction of the aeromagnetic data by removing the effects of the time‐varying external fields and computed the anomalies obtained by removing the Geomagnetic Reference Field [IAGA, 2005]. To produce the total intensity magnetic anomaly map, the data sets were merged and gridded at the spacing of 1,000 m by using a minimum curvature algorithm. Finally, the data set was microleveled to remove a few residual errors from the gridded data; the field obtained is shown in Figure 3a. [20] To facilitate the forward modeling and its geological interpretation (see section 5.2), the total field anomalies have been reduced‐to‐pole (RTP) [Baranov and Naudy, 1964]. This analytical transformation sets the magnetic anomalies above their relative sources and transforms each dipolar anomaly into positive or negative. This has the advantage of considerably simplifying the magnetic mod- eling because the magnetic contacts in the model would thus not be substantially far from the real position. As a consequence, the relation between anomaly and source in the map is more evident. Using the RTP algorithm, we have assumed that the ambient magnetic field and the magneti- zation have constant directions along the studied area. The smaller the area, the more valid is this assumption. For the 70 × 55 km coverage of this survey, we are confident that this assumption holds. The RTP magnetic anomaly map is shown in Figure 3b. A WNW‐ESE elongated positive anomaly characterizes the CV area (anomaly 1 in Figure 3). However, a negative anomaly also appears south of the CV anomaly 1 in the RTP map (Figure 3b). Because this neg- ative feature overlaps the CV WNW‐trending fault (see Figure 2), we investigated its characteristics and meaning (see section 5.2). 5. Magnetic Modeling [21] The final goal of potential field surveys is to get quantitative information about the causative sources and to interpret them in the frame of the geological knowledge of the area. Here, we have chosen to carry out both inverse and direct approaches. To build a reliable 3‐D geological model of the subsurface CV structure, aeromagnetic data have the fundamental role of providing a quantitative constraint on the distribution of the susceptibility rock properties on a wide offshore area that is otherwise not directly accessible. Table 1. Summary of Rock Magnetic Properties Measured in Calabria in the Areas of Capo Vaticano, Mesima Graben, and Le Serre Chain Along With the Sample Description and Geographic Position Sample Latitude Longitude Lithology Description Location Volume Susceptibility (SI) Sedimentarian Units 1 38°33′24″ 16°07′37″ Sandstone Pleistocenic seabord coarse silicocalstic sandstone Mesima graben, Marepotamo valley −9.51 × 10−6 2 38°39′16″ 15°50′46″ Sandstone Tortonian medium to coarse silicocalstic sandstone Mt. Poro‐Capo Vaticano 1.22 × 10−5 3 38°33′27″ 16°07′42″ Sand Fine sand Mesima graben, Marepotamo valley 1.20 × 10−4 4 38°38′16″ 15°54′00″ Carbonates Messinian carbonates and marly carbonates Mt. Poro ‐ Spilinga 8.79 × 10−5 5 38°23′19″ 16°23′52″ Sandstone Medium to lower Miocene banded sandstone, turbidites sequence Westward Caulonia village 1.54 × 10−4 6 38°33′29″ 16°06′42″ Clays Marly and silty clay succession Mesima basin 4.37 × 10−5 7 38°23′12″ 16°23′59″ Sandstone Lower Miocene silicocalstic sandstone Westward Caulonia village 5.76 × 10−5 8 38°22′29″ 16°24′58″ Sand Lower Pliocene sand Le Serre chain, western side, Caulonia village 7.71 × 10−5 10 38°22′27″ 16°25′48″ Clays “Varicolori” clay; red, green, and gray clay succession Le Serre chain, western side, Caulonia village 1.39 × 10−4 23 38°39′27″ 15°56′32″ Sands Quaternari marine terraces sands Spilunga‐Zungari 1.54 × 10−5 19 38°22′27″ 16°25′48″ Clays Varicolori clays: numidic olistonites Caulonia 9.97 × 10−5 Metamorphic and Plutonic Units 9 38°25′16″ 16°22′31″ Phyllade Metamorphic complex (from metarenites to paragneiss) F.va Allaro 1.33 × 10−4 11 38°37′06″ 15°48′44″ Dark granite Melanocratic facies granite Capo Vaticano 1.34 × 10−4 12 38°37′06″ 15°48′44″ Bright granite Leucocratic facies granite Capo Vaticano 1.9 × 10−4 13 38°31′15″ 16°08′28″ Granite Le Serre chain, western side 1.04 × 10−4 14 38°25′16″ 16°22′31″ Phyllade Metamorphic complex (from metarenites to paragneiss) F. va Allaro 2.38 × 10−4 15 38°23′37″ 16°23′10″ Phyllade Gray phyllade Le Serre eastern side, Popolli‐Serra Schiavella 1.60 × 10−4 16 38°24′57″ 16°21′33″ Granite–aplite Aplitic sills inside Le Serre granites F. va Allaro 6.95 × 10−6 17 38°24′57″ 16°21′33″ Granite F. va Allaro 1.30 × 10−4 18 38°31′15″ 16°08′28″ Granite Le Serre chain, western side 4.50 × 10−4 24 38°49′18″ 16°17′51″ Metagabbro Curinga 1.10 × 10−2 Volcanic Units 20 38°37′28″ 16°06′29″ Pumice Granulometry: 1 mm 1.66 × 10−4 21 38°37′28″ 16°06′29″ Pumice Granulometry: 4 mm 2.34 × 10−4 22 38°37′28″ 16°06′29″ Pumice Granulometry: 8 mm 2.28 × 10−4 DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 9 of 18
  • 10. As a first‐order of approximation, lithology controls mag- netic properties through its mineralogy, and abrupt variation in rock properties may highlight lithological contacts. Since the backbone of southern Calabria is mainly composed of nonmagnetic rocks (see section 3), the presence of the CV positive magnetic anomaly allows us to infer the presence of a magnetized body not related to the surrounding exposed units. To get a more reliable interpretation of the CV anomaly system causative bodies, we carried out data inver- sion, obtaining the representative magnetization value, shape, depth, and horizontal extent of the body. Furthermore, a 2.75‐D forward model has been carried out with the aim of adding further geostructural constraints to the inversion results by using measured susceptibility values for the back- ground rock properties (Table 1). The CV anomaly system shows a simple pattern lightly affected by the regional field, which influences the intensities of the short‐wavelength components (anomalies 2, 3, and 4 in Figure 3). The regional field is not completely sampled due to the shorter dimension of the surveyed area, and it is characterized by the over- lapping effects of sources lying in south Calabria (e.g., in the Catena Costiera and Sila areas) (Figure 3). The influence of such sources is not homogeneous and difficult to evaluate using analytical methods. 5.1. Inversion Modeling [22] We applied a 3‐D inversion approach to the magnetic anomaly data of the CV area with the aim of modeling the subsurface bodies generating the observed anomaly. As is well known, the mathematical inversion approach is defined as the quantitative estimation of some models parameters starting from their effect (e.g., magnetic measures), knowing the mathematical relationship that links the models para- meters with their spatial effect (forward problem). From a mathematical point of view, the inversion applied to surface potential data field is an ill‐posed problem that suffers from instability and algebraic ambiguity of solutions [Blakely, 1995], so that model parameters are not analytically invert- ible from anomaly data. A common procedure to reduce those ambiguities and obtain realistic geological sources is the introduction of additional information in the form of a priori geological data or mathematical regularization. A common approach to solve these kinds of problems in potential fields inversion is to divide the subsoil, where the real geological source is supposed to be located (a priori information), into a number of homogeneous magnetized rectangular blocks (cells) for which the direction of magnetization is supposed to be known and equal for each block. Such “discretization” defines the exact geometry of each single subsurface cell so that it is possible to generate the observed potential field measurement by applying an appropriate magnetization value for each cell [Bear et al., 1995]. In this form, the inversion problem is linearized so that the unknown parameter to be determined is the magnetization module of each cell. In our analysis, we adopted the forward solution following Sharma [1986]. Considering the vector of spatial measures b and the magnetization vector x of N cells, the linear forward model evaluating the effect of the whole cells on ith station point is expressed by bi Âź X N jÂź1 Aij xj; where j is the cells index, Aij is the sensitivity matrix coefficient that evaluates the contribution on the ith station point of the jth cell, considering unitary magnetization. To resolve this linear system in the x vector unknowns, we adopted the iterative regularization minimization method of Levenberg and Marquardt [Press et al., 1992]. This method simultaneously minimizes the norm of the solution and the misfit between the observed and the computed synthetic magnetic anomalies generated by the magnetization solution vector. We also applied a positivity control on the magne- tization solution because of the CV positive dipolar anomaly geometry. [23] Taking into account the spatial extent of the CV mag- netic anomaly (Figures 2b and 3), we built a 3‐D block mesh made with cells of dimensions 2,000 m E × 2,000 m N × 500 m below the surface. The maximum depth is at 3,500 m. The inversion setup corresponds to a total number of 3,600 cells. We adopted a regular measure grid made by 2,000 m spaced nodes calculated by gridding the original magnetic data. To stabilize the inversion model results, we used the same dimensions for the block cells mesh and the measured grid. In calculating the sensitivity matrix, it is necessary to define the vector values of block magnetiza- tion; we assume that the total magnetization vector of the CV body is subparallel to the Total Magnetic Field direc- tion and hence apply a magnetic inclination of 55° and a declination of 0°. To better constrain the resulting model, we defined a bathymetric surface (data from Tortorici et al. [2003]) above which cells were not used in the inversion calculation. [24] Figure 4a shows the resulting model from inversion procedure applied to the magnetic anomaly data, Figure 4b shows the measured magnetic data used in inversion pro- cess, and Figure 4c shows the synthetic anomaly map. The differences between the synthetic and observed data are reported in Figure 4d. The CV anomaly is well recovered by the inversion model. However, regional longer wavelengths extend well beyond the relatively limited survey area, as shown on the regional anomaly map in Figure 2c [Caratori Tontini et al., 2004], and the inversion approach cannot resolve this problem. The occurrence of such long‐wavelength anomalies could cause a misfit between the measured mag- netic anomaly data and the synthetic anomaly values gener- ated by the recovered inversion model. Taking into account Figure 4. (a) Magnetic source model from inversion process obtained from the sea level down to 3,000 m below sea level; distances are expressed in meters. (b) Observed magnetic anomaly data. (c) Synthetic anomaly map obtained for the mag- netization distribution shown with I = 54° and D = 2° inducing field. (d) Differences between synthetic and observed data shown in Figures 4b and 4c. (e) Field (in color) generated by the susceptibility model in Figure 4a with a I = 90° and D = 0° inducing field (the synthetic magnetic effect that this source model would have at the magnetic North Pole). The RTP field of real data of Figure 3b is reported as isoline contouring. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 10 of 18
  • 11. Figure 4 DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 11 of 18
  • 12. this possible misfit, we focused on the fitting quality of the CV anomaly. The differences between the synthetic and observed data (Figure 4d) show a long‐wavelength misfit trend oriented N‐NE, S‐SW with amplitude between 0 and −10 nT (about 5% of the maximum amplitude of the CV anomaly). In the surrounding area, the misfit shows a clear long‐wavelength pattern with amplitude oscillating between −20 and −10 nT. The inversion does not well describe the east side of the inversion area because of the positive magnetic anomalies generated by geological sources located in Calabria. This misfit is described in Figure 4d by negative anomalies with a minimum of −40 nT in a very limited area in the eastern side of the map. The above summarized misfit values are very low, thus suggesting the goodness and Figure 4. (continued) DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 12 of 18
  • 13. reliability of the inversion model. The recovered inversion model reported in Figure 4a describes a positive magnetic body approximately 20–25 km long WNW‐ESE striking with maximum magnetization values of 1.5 A/m. [25] We carried out a further model to investigate the nature of the RTP negative anomaly appearing just to the south of positive anomaly 1 (Figure 3b). This negative anomaly represents an interesting issue to investigate, since it precisely overlaps the offshore prolongation of the CV fault. We computed the field generated by the susceptibility model of Figure 4a with a I = 90° and D = 0°–inducing field (the synthetic magnetic effect that this source model would have at the magnetic North Pole). The results are shown as a color map in Figure 4e, where the RTP field of real data (Figure 3b) is also reported as isoline contouring. The close matching of these two fields indicates that the negative feature depicted in the RTP map of Figure 3b results from the RTP computation [see also Silva, 1986] and belongs to the anomaly 1 source recomputed at the North Magnetic Pole. 5.2. Forward Modeling [26] We carried out the forward modeling along two key profiles crossing the area in the WNW‐ESE and NNE‐SSW directions (Figures 3b and 5). The measured rock suscepti- bility values (Table 1) were used to characterize the exposed and deep stratigraphic sequences shown in the geological profile (Figure 2b). Tectonic units depicted in the sections are taken from the structural model of Italy [Finetti, 2005]. The modeled body (position, geometry, and extension) was drawn on the basis of the inversion results. In both profiles shown in Figure 5, the structures were further constrained with the available surface geology information. In addition, the uppermost crust beneath the CV morphological ridge was constrained with the seismic profiles crossing the area [Trincardi et al., 1987; Milia et al., 2009]. The extension of the modeled bodies along the profiles are reported in Table 2. [27] The A‐B profile (Figure 5) was chosen with the goal of crossing symmetrically the CV anomaly along its major axis and sampling its intensity peak. The C‐D profile was oriented orthogonally to the previous one. This profile was chosen to further constrain the modeling and to interpret the low‐intensity, short‐wavelength anomalies lying southward of the CV fault and not imaged by the inversion results (Figure 5). In fact, our high‐resolution, near‐surface survey has also detected the higher‐frequency components in the anomaly field that were not visible in the previous magnetic surveys [Agip, 1981; Chiappini et al., 2000; Caratori Tontini et al., 2004]. These minor magnetic signatures are here interpreted in the same framework of the CV anomaly causative body and, for this reason, are called the CV magnetic anomaly system. We achieved the modeling by using an implementation of a 2.75‐D forward modeling found on the expressions of Talwani and Heirtzler [1964], Rasmussen and Pedersen [1979] and on the algorithm proposed by Won and Bevis [1987]. The proposed models, summarized in Figure 5, are discussed in section 6. 6. Discussion and Conclusion [28] The results of the magnetic modeling highlight the occurrence of a WNW‐trending magnetized body about 20 km long (Figures 4 and 5). This body, which is directly magnetized, extends from the seafloor to about 3 km bsl, and intrudes toward ESE, between the CV granites and the underlying granulites of the Calabrian Arc. The WNW flank of the body is covered by sediments that, on the basis of available seismic profiles, consist of sands and silts younger than 0.7 Myr [Milia et al., 2009]. According to the models shown in Figures 4 and 5, the width is between 3.5 and 5 km. The reconstructed geometry is consistent with that of a semitabular, flat, elongated sill. Less extended, 2 to 3 km wide, laccolith‐like highly magnetized bodies also occur between 3 and 4 km of depth. These minor bodies are intruded within the granulites of the Calabrian Arc. The magnetization values needed to explain the observed anoma- lies in the CV area are between 0.7 and 1.5 A/m (Figures 4 and 5). This magnetization range encompasses that mea- sured in the evolved rocks (rhyolites and trachytes) of the Aeolian Islands (0.6–1.6 A/m) [Blanco‐Montenegro et al., 2007, and references therein]. The collected magnetic data show that the CV anomaly reaches magnetic anomaly values (150–220 nT, Figure 3b) that are significantly higher than those observed in the rest of Calabria (30–50 nT) [Agip, 1981; Caratori Tontini et al., 2004]. In addition, the CV magnetic anomaly values are of the same magnitude as those of Panarea Island, Lametini Seamount, Alcione Sea- mount, and Stromboli canyon, where volcanic rocks outcrop and have been dragged [Caratori Tontini et al., 2004; Finetti, 2005]. The anomaly of the Stromboli canyon, which is located 20 km eastward from Panarea Island, is also visible in the westernmost side of the CV area survey (anomaly 5 in Figure 3). The magnetic signature of the CV promontory turns out to be completely different from the magnetic pattern of the Calabrian Arc region for both inten- sity and shape. [29] The inverse model (Figure 4a) depicts a crustal vol- ume with an increase of the magnetization (values 0.9 A/m) in correspondence of the top of the CV bathymetric ridge. We propose that this 5–6 km long and about 2–3 km wide rock volume with higher magnetization is representative of the shallower portion of the plumbing system of the volcano from which WNW‐ESE elongated apophyses with lower magnetization depart, probably representing dikes. Besides, the forward model (Figure 5), which is characterized by a constant magnetization of the represented bodies, requires a vertical extension in correspondence to the abovementioned bathymetric ridge. Both models are able to fit the observed data correctly, represent two possible settings of the mag- netization, and be coherent with the occurrence of a volcanic body. In the first case (Figure 4), the higher magnetization represents the plumbing systems zone, as reported above; in the second one (Figure 5), a vertical extension of the body is required to correctly fit the data to define the plumbing system geometrically. Therefore, both models highlight a magnetized, shallow, sill‐like body, the topographic expres- sion of which is represented by the CV morphological ridge (Figures 4 and 5). In fact, the results reveal that the apex of the sill overlaps the subcircular truncated top of the CV ridge located at a 200 m bsl (Figure 2b). As result of the above observations, the CV sill could be interpreted as the remnant of a shallow magmatic body characterized by a central plug roughly extending to a depth between 500 and 1500 m bsl, DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 13 of 18
  • 14. Figure 5. Crustal 2.75‐D magnetic models for profiles crossing the CV anomaly in (a) WNW‐ESE direction, A‐B profile; (b) N20°E direction, C‐D profile. The trace of profiles is shown in Figure 3b (dotted red lines). Numbers identify marked anomalies in the reduced‐to‐the‐pole magnetic anomaly field (Figure 3). Faults are indicated by red lines. The size of the bodies in the direction perpendicular to the profile is reported in Table 2. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 14 of 18
  • 15. Figure 5. (continued) DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 15 of 18
  • 16. and by WNW‐trending minor lens, possibly representing dikes intruding the Calabrian Arc. [30] We propose that the CV sill may represent the source area of the 0.67 to 1.06 Myr old widespread tephra fall deposits from plinian columns poorly resedimented in shal- low marine environment outcropping in western Calabria (section 2.2). The occurrence of (1) maximum thickness of such deposits in the CV and Mesima areas, (2) progressive thinning of such deposits toward the southwest (i.e., Reggio Calabria basin), (3) impact sags related to the emplacement of ballistic ejecta in the CV area, and (4) sedimentary facies consistent with a proximal origin for the tephra layers strongly suggests that the CV magnetic anomaly sector may correspond to the source area of the middle Pleistocene volcaniclastic deposits of Calabria [De Rosa et al., 2008]. As a result, the shallow, WNW‐ESE elongated sill depicted in Figures 4 and 5 is easy to use to represent the shallow plumbing system of a middle Pleistocene volcano. Such interpretation is consistent with the occurrence, in the sector of the CV magnetic anomaly, of a 0–5 km deep, high Vp/Vs body [Barberi et al., 2004]. The magnetization values obtained from inversion of the data (Figure 4a) suggest that the CV sill is composed of rocks with magnetization values comparable to those of rhyolites and dacites of the Aeolian Islands [Blanco‐Montenegro et al., 2007]. This consider- ation corroborates the above interpretation of the CV sill as a plumbing system of a volcano responsible for the emplace- ment of the Pleistocene widespread rhyolitic to dacitic pumices outcropping in Calabria, which show a HK calc‐ alkaline affinity compatible with that of the rocks of the eastern Panarea sector of the Aeolian Islands [De Rosa et al., 2008]. On the basis of the stratigraphic, magnetic, and geo- chronological data, the CV ridge formed before 0.7 Ma, being partly covered by Pleistocene sediments younger than 0.7 Myr [Milia et al., 2009]. The normal polarity magnetic signature of the CV body suggests an emplacement during the Bruhnes (0.78 Ma) [Cande and Kent, 1995; Gradstein et al., 2005] or Jaramillo (0.99–1.07 Ma) Chrons. On the other hand, the widespread pumices and volcanic ash layers outcropping in the studied area deposited in a time inter- val between 0.67 and 1.07 Ma [Cornette et al., 1987]. Therefore, the CV sill and the Pleistocene volcanics were emplaced either between 0.67 and 0.78 Ma or between 0.99 and 1.07 Ma. These age intervals roughly overlap the older ages of Aeolian volcanism, which started between 1.3 Ma in the western, Alicudi–Filicudi sector and 0.8–0.5 Ma in the eastern, Stromboli–Panarea sector [De Astis et al., 2003]. However, the composition of the older dragged Aeolian rocks is basaltic, whereas the more evolved prod- ucts, with composition similar to the CV tephra, are younger than 100 kyr. This indicates that evolved magmas charac- terized the early, submarine, and possibly explosive, phases of the Aeolian volcanism. In addition, the rhyolitic and dacitic composition of the CV Pleistocene magmas volcano, along with the results of our magnetic modeling, testify to the formation of sill‐like, shallow magma reservoirs. Accord- ing to the geometry depicted in Figure 5 and the available geological data (Figure 2b), the storage depth of such reser- voirs is controlled by the major discontinuity between the lower crust granulites of the underlying, Carboniferous– Permian granites of the upper crust. The shape of the CV magnetic anomaly clearly shows that the sill geometry has also been controlled by a WNW‐trending discontinuity. This is also the strike of the fault affecting the southern coastline and offshore of the CV promontory (Figure 2). This fault is possibly seismically active [De Astis et al., 2003] and was responsible for the destructive 1905 Calabria earthquake, M = 6.3–6.7, according to Cucci and Tertulliani [2006]. Therefore, we conclude the CV volcano was emplaced on the footwall of this still active tectonic structure. This feature is structurally compatible with numerical models of magma emplacement along active faults [e.g., Ellis and King, 1991; Pascal and Cloetingh, 2002]. Such models reveal that magma uprising to the surface along a faulting zone accumulates at the footwall of a normal fault because the local extensional strain concentrates at the footwall, whereas a compressive strain occurs in the hanging wall. [31] The collected data and the above observations sug- gest that the recognized CV volcanism is consistent, in terms of age, magnetic signature, and geochemical affinity of the erupted products, with that of the Aeolian volcanoes. The central sector of the Aeolian Arc, which includes the islands of Salina, Lipari, and Vulcano, develops along a tectonic structure that it is in a transversal position with respect to the Aeolian volcanic ring. Our data and inter- pretation suggest that, indeed, NW‐SE faults (i.e., trans- versal structures to the general trend of the Calabria block) must be considered as an important part of the Pleistocene (and possibly active) tectonics of Calabria. It is also likely that the seismogenic potential of such structures is at present underevaluated [Basili et al., 2008] and should be better assessed. [32] The WNW‐trending fault that controls the CV fissural‐like volcano is also transversal to the eastern sector of the Aeolian volcanoes, which are controlled by NE‐SW structures. However, this fault does not appear to represent a slab tear fault [e.g., Chiarabba et al., 2008]. In any case, it may represents a structure along which magma uprose and stopped, at least in the last 1–0.7 Ma. In this period, a sig- nificant (up to 1 mm/yr) uplift affected the Calabria region and the Aeolian Islands [Westaway, 1993]. The uplift, whose geodynamic origin is a subject of debate [e.g., Gvirtzman and Nur, 2001], is maximized in the CV promontory (4 mm/yr in the last 124 kyr) [Tortorici et al., 2003]. Different uplift rates of segments of the Calabrian Arc could be responsible for fragmentation of the upper crust and formation of transversal faults. As a result, magma emplacement may Table 2. The Strike of the Structures for the Two Profiles A‐B and C‐D Is 90°a Body y1 (km) y2 (km) Granulites section A‐B 10 10 Sands and sandstone section A‐B 10 10 Granites section A‐B 10 10 Volcanic body section A‐B Anomaly 1 1.7 1.7 Sands and sandstone section C‐D 10 10 Volcanic body section C‐D Anomaly 1 5 15 Anomaly 4 3 3 Anomaly 2 3 3 a y1 is the length toward the west‐northwest in profile A‐B and toward the N20°E in profile C‐D, whereas y2 is the length toward the east‐southeast in profile A‐B and toward the N290°E in profile C‐D. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 16 of 18
  • 17. occur along structures like the WNW‐trending fault of the CV. According to Tortorici et al. [2003], such structures also act as transfer faults that allow the regional extensional strain to be transferred from offshore, where it is accom- modated by the ENW‐WSW trending normal faults, to the Calabrian Arc, where it is accommodated by NE–SW trending faults. In this general framework, the composition of the CV tephra suggest that between 0.67 and 0.97 Ma, while mafic products were erupting in the Aeolian area, dacites and rhyolites characterized the explosive activity of the CV volcano. From a structural point of view, we note that the CV fissural‐like volcano is similar to the vents of Lipari and Vulcano islands, which also depict a fissural‐like structure [De Astis et al., 2003]. As a result, the transversal faults of the Aeolian Arc appear to control the development of fissure‐like eruptive centers. [33] Acknowledgments. This study has been supported with FIRB‐ MUIR and INGV funds. Thanks to numerous colleagues for discussions, comments, and suggestions. We are grateful to two anonymous referees, as well as to the Journal of Geophysical Research Associate Editor and the Editor (Andre Revil) for providing useful comments to our manuscript. References Acquafredda, P., S. Lorenzoni, N. Minzoni, and E. Zanettin Lorenzoni (1987), Paleozoic sequence in the Stilo‐Bivongi area (central Calabria), Mem. Sci. Geol., 39, 117–127. Acquafredda, P., S. Lorenzoni, N. Minzoni, and E. Zanettin Lorenzoni (1989), Stratigraphic correlation form of the stilo area (Serre region, cen- tral Calabria, Italy), Rend. Soc. Geol. Ital., 12, 103–105. Acquafredda, P., S. Lorenzoni, and E. Zanettin Lorenzoni (1994), Paleo- zoic sequence and evolution of the Calabrian‐Peloritan Arc (Southern Italy), Terranova, 6, 582–594. Amodio Morelli, L., et al. (1976), L’arco Calabro‐peloritano nell’ orogene Appenninico‐Maghrebide, Mem. Sci. Geol., 17, 1–60. Antonioli, F., L. Ferranti, K. Lambeck, S. Kershaw, V. Verrubbi, and G. Dal Pra (2006), Late Pleistocene to Holocene record of changing uplift rates in southern Calabria and northeastern Sicily (southern Italy, central Mediterranean Sea), Tectonophysics, 422, 23–40. Ayuso, R. A., A. Messina, B. De Vivo, S. Russo, L. G. Woodruff, S. F. Sutter, and H. E. Belkin (1994), Geochemistry and argon thermochronol- ogy of the Variscan Sila batholith, southern Italy: Source rocks and magma evolution, Contrib. Miner. Petrol., 117, 87–109. Azzaro, R., D. Bella, L. Ferreli, A. Michetti, F. Santagati, L. Serva, and E. Vittori (2000), First study of fault trench stratigraphy at Mt. Etna volcano, Southern Italy: Understanding Holocene surface faulting along the Moscarello fault, J. Geodyn., 29, 187–210. Barberi, F., A. Gandino, A. Gioncada, P. La Torre, A. Sbrana, and C. Zenucchini (1994), The deep structure of the Eolian arc (Filicudi‐ Panarea‐Vulcano sector) in light of gravity, magnetic and volcanological data, J. Volcanol. Geotherm. Res., 61, 189–206. Barberi, G., M. T. Cosentino, A. Gervasi, I. Guerra, G. Neri, and B. Orecchio (2004), Crustal seismic tomography in the Calabrian Arc region, south Italy, Phys. Earth Planet. Int., 147, 297–314. Baranov, V., and H. Naudy (1964), Numerical calculation of the formula of reduction to the magnetic pole, Geophysics, 29, 67–79. Basili, R., G. Valensise, P. Vannoli, P. Burrato, U. Fracassi, S. Mariano, M. M. Tiberti, and E. Boschi (2008), The Database of Individual Seismo- genic Sources (DISS), version 3: Summarizing 20 years of research on Italy’s earthquake geology, Tectonophysics, 453, 20–43, doi:10.1016/j. tecto.2007.04.014. Bear, G. W., H. J. Al‐Shukri, and A. J. Rudman (1995), Linear inversion of gravity data for 3‐D density distribution, Geophysics, 60, 1354–1364. Blakely, R. J. (1995), Potential Theory in Gravity and Magnetic Applica- tions, 441 pp., Cambridge Univ. Press, Cambridge. Blanco‐Montenegro, I., R. De Ritis, and M. Chiappini (2007), Imaging and modelling the subsurface structure of volcanic calderas with high‐ resolution aeromagnetic data at Vulcano (Aeolian Islands, Italy), Bull. Volcanol., 69(6), 643–659, doi:10.1007/s00445-006-0100-7. Bonardi, G., A. Messina, V. Perrone, S. Russo, and A. Zuppetta (1984), L’unitĂ  di stilo nel settore meridionale dell’arco calabro peloritano, Boll. Soc. Geol. Ital., 103, 279–309. Bonardi, G., W. Cavazza, V. Perrone, and S. Rossi (2001), Calabria‐ Peloritani terrane and northern Ionian Sea, in Anatomy of an Orogen: The Apennines and Adjacent Mediterranean Basins, edited by G. B. Vai and I. P. Martini, pp. 287–306, Kluwer, Dordrecht. Bouillin, J. P., S. Baudelot, and C. Majestè‐Menjoulas (1984), Mise en evidence du Cambro.Ordovicien en calibre centrale (Italie). AffinitĂŠs palĂŠogĂŠographiques et consĂŠquences structurales, C. R. Acad. Sci. Paris, II, 298, 88–92. Bouillin, J. P., M. Menjoulas, S. Baudelot, C. Cygan, and C. Fournier Vinas (1987), Les formations paleozoĂŽques de l’Arc Calabro‐Peloritain dans leur cadre structural, Boll. Soc. Geol. Ital., 106, 683–698. Caggianelli, A., and G. Prosser (2002), Modelling the thermal perturbation of the continental crust after intraplating of thick granitoid sheets: A compar- ison with the crustal sections in Calabria (Italy), Geol. Mag., 139, 699–706. Cande, S. C., and D. V. Kent (1995), Revised calibration of the geomagnetic polarity time scale for the Late Cretaceous and Cenozoic, J. Geophys. Res., 100, 6093–6095. Caratori Tontini, F., P. Stefanelli, I. Giori, O. Faggioni, and C. Carmisciano (2004), The revised aeromagnetic map of Italy, Ann. Geophys., 47, 1547–1556. Cavazza, W. (1989), Detritial modes and provenance of the Stilo‐capo d’Orlando Formation (Miocene), southern Italy, Sedimentology, 36, 1077–1090. Cavazza, W., and P. G. DeCelles (1993), Upper Messinian siliciclastic rocks in southeastern Calabria (southern Italy): Palaeotectonic and eustatic implications for the evolution of the central Mediterranean region, Tectonophysics, 298, 223–241. Cavazza, W., J. Blenkinsop, P. G. De Celles, R. Timothy Patterson, and E. G. Reinhardt (1997), Stratigrafia e sedimentologia della sequenza sedimentaria oligocenico‐quaternaria del bacino calabro‐ionico, Boll. Soc. Geol. Ital., 116, 51–77. Chiappini, M., A. Meloni, E. Boschi, O. Faggioni, N. Beverini, C. Carmisciano, and I. Marson (2000), Shaded relief magnetic anomaly map of Italy and surrounding marine areas, Ann. Geophys., 43, 983–989. Chiappini, M., F. Ferraccioli, E. Bozzo, and D. Damaske (2002), Regional compilation and analysis of aeromagnetic anomalies for the Transantarctic Mountains‐Ross Sea sector of the Antarctic, Tectonophysics, 347, 121–137. Chiarabba, C., P. De Gori, and F. Speranza (2008), The Southern Tyrrhe- nian Subduction Zone: Deep geometry, magmatism and Plio‐Pleistocene evolution, Earth Planet. Sci. Lett., 268, 408–423. Cifelli, F., M. Mattei, and F. Rossetti (2007a), Tectonic evolution of arcuate mountain belts on top of a retreating subduction slab: The example of the Calabria Arc, J. Geophys. Res., 112, B09101, doi:10.1029/ 2006JB004848. Cifelli, F., M. Mattei, and F. Rossetti (2007b), The architecture of brittle postorogenic extension: Results from an integrated structural and paleo- magnetic study in north Calabria (southern Italy), Geol. Soc. Am. Bull. 119(1), 221–239. Colonna, V., S. Lorenzoni, and E. Zanettin Lorenzoni (1973), Sull’esistenza di due complessi metamorfici lungo il bordo sud‐orientale del massiccio “granitico” delle Serre (Calabria), Boll. Soc. Geol. Ital., 92, 801–830. Cornette, Y., P. Y. Gillot, P. Barrier, and F. Jehenne (1987), DonnĂŠes radiomĂŠtriques preliminaries (Potassium‐Argon) sur des cinĂŠrites plio‐ plĂŠistocènes du DĂŠtroit de Messine, Doc. Trav. Inst. Geol. Albert‐de‐ Lapparent, 11, 97–99. Cucci, L., and A. Tertulliani (2006), I terrazzi marini nell’area di Capo Vaticano (arco calabro): Solo un record di sollevamento regionale o anche di deformazione cosismica?, Quaternario, 19, 89–101. De Astis, G., G. Ventura, and G. Vilardo (2003), Geodynamic significance of the Aeolian volcanism (Southern Tyrrhenian Sea, Italy) in light of structural, seismological and geochemical data, Tectonics, 22(4), 1040, doi:10.1029/2003TC001506. Del Ben, A., C. Barnaba, and A. Taboga (2008), Strike‐slip systems as the main tectonic features in the Plio‐Quaternary kinematics of the Calabrian Arc, Mar. Geophys. Res., 29, 1–12. De Ritis, R., I. Blanco‐Montenegro, G. Ventura, and M. Chiappini (2005), Aeromagnetic data provide new insights on the volcanism and tectonics of Vulcano Island and offshore areas (southern Tyrrhenian Sea, Italy), Geophys. Res. Lett., 32, L15305, doi:10.1029/2005GL023465. De Ritis, R., G. Ventura, and M. Chiappini (2007), Aeromagnetic anoma- lies reveal hidden tectonic and volcanic structures in the central sector of the Aeolian Islands, southern Tyrrhenian Sea, Italy, J. Geophys. Res., 112, B10105, doi:10.1029/2006JB004639. De Rosa, R., R. Dominici, and M. Sonnino (2001), Evidenze di vulcanismo sinsedimentario nella successione pleistocenica del graben del Mèsima (Calabria centro‐occidentale), Quaternario, 14(2), 81–91. De Rosa, R., R. Dominici, P. Donato, and D. Barca (2008), Widespread syn‐eruptive volcaniclastic deposits in the Pleistocenic basins of South‐Western Calabria, J. Volcanol. Geotherm. Res., 177, 155–169. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 17 of 18
  • 18. Doglioni, C., F. Innocenti, and S. Mariotti (2001), Why Mt. Etna?, Terra Nova, 13, 25–31. Dumas, B., P. GuĂŠrĂŠmy, R. LhĂŠnaff, and J. Raffy (1982), Le soulèvement Quaternaire de la Calabre meridionale, Rev. Gèol. Dyn. GĂŠogr. Phys., 23, 27–40. Ellis, M., and G. King (1991), Structural control of flank volcanism in con- tinental rift, Science, 254, 839–843. Faccenna, C., T. W. Becker, F. P. Lucente, L. Jolivet, and F. Rossetti (2001), History of subduction and back‐arc extension in the Central Mediterranean, Geophys. J. Int., 145, 809–820. Finetti, I. (2005), Deep Seismic Exploration of the Central Mediterranean and Italy, 794 pp., Elsevier, Amsterdam. Finn, C., T. W. Sisson, and M. Deszcz‐Pan (2001), Aerogeophysical mea- surements of collapse‐prone hydrothermally altered zones at Mount Rainer volcano, Nature, 409, 600–603. Ghisetti, F. (1981), L’evoluzione strutturale del bacino plio‐pleistocenico di Reggio Calabria nel quadro geodinamico dell’Arco Calabro, Boll. Soc. Geol. Ital., 100, 433–466. Ghisetti, F., and L. Vezzani (1980), Contribution of structural analysis to understanding the geodynamic evolution of the Calabrian arc (Southern Italy), J. Struct. Geol., 3, 371–381. Gradstein, F. M., J. G. Ogg, and A. G. Smith (2005), A Geologic Time Scale, 610 pp., Cambridge Univ. Press, Cambridge. Graessner, T., and V. Schenk (2001), An exposed Hercynian deep conti- nental crustal section in the Sila massif of northern Calabria: Mineral chemistry, petrology and a P–T path of granulite facies metapelitic mig- matites and metabasites, J. Petrol., 42, 931–961. Guarnieri, P. (2006), Plio‐Quaternary segmentation of the south Tyrrhenian forearc basin, Int. J. Earth Sci., 95, 107–118. Gvirtzman, Z., and A. Nur (2001), Residual topography, lithospheric struc- ture and sunken slabs in the Central Mediterranean, Earth Planet. Sci. Lett., 187, 117–130. IAGA (International Association of Geomagnetism and Aeronomy) (2005), The 10th‐Generation International Geomagnetic Reference Field, Geophys. J. Int., 161, 561–565. Jolivet, L., and C. Faccenna (2000), Mediterranean extension and the Africa‐Eurasia collision, Tectonics, 19, 1095–1106. Knott, S. D., and E. Turco (1991), Late Cenozoic kinematics of the Calab- rian Arc, southern Italy, Tectonics, 10, 1164–1172. Langone, A., E. Gueguen, G. Prosser, A. Caggianelli, and A. Rottura (2006), The Curinga‐Girifalco fault zone (northern Serre, Calabria) and its significance within the Alpine tectonic evolution of the western Med- iterranean, J. Geodyn., 42, 140–158 Lenat, J. F., B. Gibert‐Malengreau, and A. Galdeano (2001), A new struc- tural model for the evolution of the volcanic island of Reunion (Indian Ocean), J. Geophys. Res., 106(B5), 8645–8663. Malinverno, A., and W. B. F. Ryan (1986), Extension in Tyrrhenian Sea and shortening in the Apennines as result of arc migration driven by sink- ing of the lithosphere, Tectonics, 5, 227–254. Mattei, M., F. Speranza, A. Argentieri, F. Rossetti, L. Sagnotti, and R. Funiciello (1999), Extensional tectonics in the Amantea basin (Calab- ria, Italy): A comparison between structural and magnetic anisotropy data, Tectonophysics, 307, 33–49. Mattei, M., P. Cipollari, D. Cosentino, A. Argentieri, F. Rossetti, and F. Speranza (2002), The Miocene tectonic evolution of the Southern Tyr- rhenian Sea: Stratigraphy, structural and paleomagnetic data from the on‐shore Amantea basin (Calabrian Arc, Italy), Basin Res., 14, 147–168. Milia, A., E. Turco, P. P. Pierantoni, and A. Schettino (2009), Four‐ dimensional tectono‐stratigraphic evolution of the Southeastern peri‐ Tyrrhenian basins (Margin of Calabria, Italy), Tectonophysics, 476, 41–56, doi:10.1016/j.tecto.2009.02.030 Miyauchi, T., G. Dal Pra, and S. S. Labini (1994), Geochronology of Pleis- tocene marine terraces and regional tectonics in the Tyrrhenian coast of South Calabria, Italy, Quaternario, 7, 17–34. Monaco, C., and L. Tortorici (2000), Active faulting in the Calabrian Arc and eastern Sicily, J. Geodyn., 29, 407–424. Mulargia, F., P. Baldi, V. Achilli, and F. Broccio (1984), Recent crustal deformations and tectonics of the Messina Strait area, Geophys. J. R. Astron. Soc., 76, 369–381. Neri, G., B. Orecchio, C. Totaro, G. Falcone, and D. Presti (2009), Subduc- tion beneath southern Italy close the ending: Results from seismic tomog- raphy, Seism. Res. Lett., 80, 63–70. Nicolosi, I., F. Speranza, and M. Chiappini (2006), Ultrafast oceanic spreading of the Marsili Basin, southern Tyrrhenian Sea: Evidence from magnetic anomaly analysis, Geology, 34, 717–720. Pascal, C., and S. A. P. L. Cloetingh (2002), Rifting in heterogeneous litho- sphere: Inferences from numerical modeling of the northern North Sea and the Oslo Graben, Tectonics, 21(6), 1060, doi:10.1029/2001TC901044. Patterson, R. T., J. Blenkinsop, and W. Cavazza (1995), Planktic foraminiferal biostratigraphy and 87Sr/86Sr isotopic stratigraphy of the Oligocene‐to‐ Pleistocene sedimentary sequence in the south‐eastern calabria micro- plate, southern Italy, J. Paleontol., 69, 7–20. Press, W., S. Teulosky, W. Vetterling, and B. Flannery (1992), Numerical Recipes in C: The Art of Scientific Computing, 965 pp., Cambridge Univ. Press, Cambridge. Rasmussen, R., and L. B. Pedersen (1979), End corrections in potential field modelling. Geophys. Prospect., 27, 749–760. Roda, C. (1965), Il calcare portlandiano a Dasycladacee di M. Mutolo, Reggio Calabria, Geol. Rom., 4, 259–290. Rollin, P. J., J. Cassidy, C. A. Locke, and H. Rymer (2000), Evolution of the magmatic plumbing system at Mt Etna: New evidence from gravity and magnetic data, Terra Nova, 12, 193–198. Rosenbaum, G., M. Gasparon, F. P. Lucente, A. Peccerillo, and M. S. Miller (2008), Kinematics of slab tear faults during subduction segmen- tation and implications for Italian magmatism, Tectonics, 27, TC2008, doi:10.1029/2007TC002143. Sagnotti, L., F. Speranza, A. Winkler, M. Mattei, and R. Funicello (1998), Magnetic clay of clay sediments from the external northern Apennines (Italy), Phys. Earth Planet. Int., 105, 73–93. Schenk, V. (1984), Petrology of felsic granulites, metapelites, metabasic, ultramafic, and metacarbonates from southern Calabria (Italy): Prograde metamorphism, uplift and cooling of former lower crust, J. Petrol., 25, 225–298. Schenk, V. (1990), The exposed crustal cross section of southern Calabria, Italy: Structure and evolution of a segment of Hercynian crust, in Exposed Cross‐Section of Continental Crust, edited by M. H. Salisbury and D. M. Fountain, pp. 21–42, Kluwer, Dordrecht. Sharma, P. V. (1986), Geophysical Methods in Geology, 442 pp., Elsevier, New York. Silva, J. B. C. (1986), Reduction to the pole as an inverse problem and its application to low‐latitude anomalies, Geophysics, 51, 369–382. Spakman, W., S. van der Lee, and R. van der Hilst (1993), Travel‐time tomography of the European‐Mediterranean mantle down to 1400 km, Phys. Earth Planet. Int., 79, 3–74. Talwani, M., and J. R. Heirtzler (1964), Computation of magnetic anoma- lies caused by two‐dimensional bodies of arbitrary shape, in Computers in the Mineral Industries (Part 1), vol. 9, edited by G. A. Parks, pp. 464– 480, Stanford Univ. Publ. Geological Sciences, Stanford, Cal. Tansi, C., F. Muto, S. Critelli, and G. Iovine (2007), Neogene‐Quaternary strike‐slip tectonics in the central Calabrian Arc (southern Italy), J. Geo- dyn., 43, 393–414. Tortorici, L., C. Monaco, C. Tansi, and O. Cocina (1995), Recent and active tectonics in the Calabrian arc (southern Italy), Tectonophysics, 243, 37–55. Tortorici, G., M. Bianca, G. de Guidi, C. Monaco, and L. Tortorici (2003), Fault activity and marine terracing in the Capo Vaticano area (southern Calabria) during the Middle‐Late Quaternary, Quat. Int., 101, 269–278. Trincardi, F., M. Cipolli, P. Ferretti, J. La Morgia, M. Ligi, G. Marozzi, V. Palumbo, M. Taviani, and N. Zitellini (1987), Slope basin evolution on the Eastern Tyrrhenian margin: Preliminary report, G. Geol., 49, 1–9. Van Dijk, J. P. (1994), Late Neogene kinematics of intra‐arc oblique shear zones: The Petilia–Rizzuto Fault Zone (Calabrian Arc, central Mediterra- nean), Tectonics 13, 1201–1230. Van Dijk, J. P., and M. Okkes (1991), Neogene tectonostratigraphy and kinematics of Calabrian basin: Implications for the geodynamics of the central Mediterranean, Tectonophysics, 196, 23–60. Westaway, R. (1993), Quaternary uplift of southern Italy, J. Geophys. Res., 98, 21741–21772. Won, I. J., and M. Bevis (1987), Computing the gravitational and mag- netic anomalies due to a polygon: Algorithms and Fortran subroutines, Geophysics 52, 232–238. Zanella, E. (1995), Studio delle variazioni paleosecolari del campo magne- tico terrestre registrate nelle vulcaniti Quaternarie dell’area tirrenica e del canale di Sicilia, PhD Thesis, 198 pp., UniversitĂ  di Torino, Italy. M. Chiappini, R. De Ritis, I. Nicolosi, F. Speranza, and G. Ventura, Instituto Nazionale di Geofisica e Vulcanologia, via di Vigna Murata 605, 00143 Rome, Italy. (riccardo.deritis@ingv.it) R. De Rosa, P. Donato, R. Dominici, and M. Sonnino, Department of Earth Science, UniversitĂ  della Calabria, 87036 Arcavacata di Rende, Italy. DE RITIS ET AL.: A BURIED VOLCANO BY AEROMAGNETIC DATA B11101 B11101 18 of 18