Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Tectonophysics 621 (2014) 101122

Contents lists available at ScienceDirect

Tectonophysics
journal homepage: www.elsevier.com/locate/tecto

Rates of geodetic deformation across active faults in southern Italy


Luigi Ferranti a,, Mimmo Palano b, Flavio Cannav b, Maria Enrica Mazzella a,1, John S. Oldow c,
Erwan Gueguen d, Mario Mattia b, Carmelo Monaco e
a
DiSTAR Dipartimento di Scienze della Terra, delle Risorse e dell'Ambiente, Universit Federico II, Largo S. Marcellino 10, 80138 Napoli, Italy
b
Istituto Nazionale di Geosica e Vulcanologia, Osservatorio Etneo, Catania, Italy
c
Department of Geosciences, University of Texas at Dallas, Richardson, TX, USA
d
Istituto di Metodologie per l'Analisi Ambientale, C.N.R., Tito Scalo, PZ, Italy
e
Dipartimento di Scienze Biologiche, Geologiche e Ambientali, Universit di Catania, Italy

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

Article history: Active deformation in southern Italy is accommodated by a distributed number of faults with lowmoderate slip
Received 8 November 2013 rates. Outcropping extensional faults and mostly blind transcurrent faults are mapped within a western (or axial)
Received in revised form 10 February 2014 and an eastern domain, respectively. We use a combination of continuous (2001.002011.84) and episodic
Accepted 11 February 2014
(1995.682010.79) GNSS observations to rstly estimate the geodetic deformation rate on 32 faults. Geodetic re-
Available online 20 February 2014
sults were successively compared with geological displacement estimates. In agreement with seismological and
Keywords:
geological information, a net spatial segregation emerges between the extensional axial belt, and the eastern do-
GNSS velocity main where strikeslip faults are geodetically active. Although uncertainties are at times large, average displace-
Active fault ment rates show broadly consistent patterns within both domains. A longitudinal gradient in extension rate is
Geodetic slip rate observed for the axial fault array, with two sectors of higher magnitude (~0.81.7 mm/yr for individual faults).
Southern Italy This result is consistent with geological observations and supports the notion that extension occurs in discrete
patches. Faults of the eastern domain have lower (few 0.1 to ~1.2 mm/yr) strikeslip rates and an eastward-
decreasing extensional component, but signicant geodetic displacement is detected in areas lacking clear evi-
dence of activity. Few faults with 12 mm/yr extension rate are locally found in the eastern domain, but, based
on their limited length and on inconsistency with seismology and geology, they are considered as due to deep-
seated gravitational spreading. For crustal faults, although geodetic slip and moment rates are larger than geolog-
ical rates, the broad trend of long- to short-term rates is similar, indicating the feasibility of geodetic analysis to
contribute estimating fault slip rate and testing tectonic models in the region. Whereas the western domain
extension is thought to be controlled by potential energy related to the Tyrrhenian Moho uplift beneath the
Apennines, strikeslip in the east is related to shear on inherited faults within the Adriatic crust.
2014 Elsevier B.V. All rights reserved.

1. Introduction In southern Italy, active faults and focal mechanism solutions of


crustal earthquakes dene two laterally juxtaposed belts of contempo-
Southern Italy is characterized as one of the highest seismic hazard rary deformation, with strikeslip faulting beneath the Apulia plains
in the Mediterranean region, as documented by historical and instru- and plateau to the east, and extension along the axis of the Southern
mental catalogs where several large (6.5 M 7.5) and destructive Apennines chain in the west (Figs. 2, 3). The current deformation
events have occurred (Fig. 1; Rovida et al., 2011). Despite that the larg- regime in the Apennines started to form between Late PlioceneEarly
est earthquakes have been associated to a causative fault, large uncer- Pleistocene, when it replaced contractional tectonics responsible for
tainties still exist regarding denition of source geometric, kinematic Neogene accretion of the Southern Apennines fold-and-thrust belt
and energetic parameters (e.g. slip rates); this represents a severe draw- (Fig. 2, inset; Hyppolite et al., 1994; Patacca and Scandone, 2007).
back for the growing practice of seismic hazard estimation based on tec- Because of this relatively recent change in tectonic regime, the cur-
tonic information for individual seismogenic sources (e.g. Peruzza et al., rent deformation pattern in the Apennines is diffuse and is character-
2010). ized by low geodetic (~ 35 mm/yr) and geologic (12 mm/yr) rates
(Faure Walker et al., 2012; Ferranti et al., 2008; Giuliani et al., 2009;
Palano et al., 2011; Papanikolaou and Roberts, 2007; Serpelloni et al.,
Corresponding author at: Tel.: +39 812538180.
E-mail address: lferrant@unina.it (L. Ferranti).
2005), and the landscape still bears the dominant signature of contrac-
1
Present address: Dipartimento di Scienze della Terra, Universit degli Studi di Perugia, tional tectonics. Consequently, detection and characterization of active
Italy. and seismogenic faults present signicant challenges in the region,

http://dx.doi.org/10.1016/j.tecto.2014.02.007
0040-1951/ 2014 Elsevier B.V. All rights reserved.
102 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

Fig. 1. Instrumental seismicity (M 2.2) in southern Italy 19832012 (http://iside.rm.ingv.it). Blue squares are historical earthquakes from CPTI catalog (Rovida et al., 2011); Imax,
maximum macroseismic intensity.

and to this end the GNSS (Global Navigation Satellite System; formerly faults. Later, we compare the geodetic results with information obtained
known as GPS) observations can offer a vital contribution. from geological studies of active faults with the aim of reconciling con-
In the last decade, GNSS data analysis has contributed to the temporary and geologic displacements and provide insights into the
understanding of the pattern of active deformation in southern Italy pattern of moment accumulation and release over different time spans.
(D'Agostino et al., 2008; Devoti et al., 2011; Ferranti et al., 2008; Although our estimation of geodetic displacement is on occasion bi-
Hunstad et al., 2003; Oldow et al., 2002; Palano et al., 2011; Serpelloni ased by a considerable uncertainty, results of this study help to rene
et al., 2005). Although published geodetic studies have singled out the and implement the poorly constrained magnitude of slip rate on active
regional scale pattern of crustal deformation, the relation between sur- faults, which is a timely contribution for fault-based seismic hazard as-
face displacement and inter-seismic strain accumulation at the scale of sessment in this highly-populated region (e.g. Peruzza et al., 2010,
individual faults or fault arrays has been established only locally and 2011; Roberts et al., 2004). Combination of site velocities with local
with a signicant uncertainty (Ferranti et al., 2008; Giuliani et al., seismicity and structural studies as discussed in this paper indicates
2009). In addition, a comparison between geodetic and geologic rates that, although broadly driven by convergence between Europe and
of fault slip has not been attempted yet. The geodetic displacement re- Nubia (e.g. the African plate west of the East African Rift), the pattern
solved onto the fault may embed components of seismic and aseismic of crustal displacement in southern Italy is strongly inuenced by a
slip, and of strain accumulation because of block coupling on the fault combination of processes not immediately related to plate boundary
slip surface; whether this displacement pattern is however representa- forces, such as crustal delamination and mantle upwelling, and exploi-
tive of the long-term behavior of crustal faults requires comparison of tation of intra-plate anisotropy.
different datasets.
In this paper, we use a dense combination of: i) continuous GNSS 2. Regional setting
data collected between 2001.00 and 2011.84 on geodetic networks
installed in southern Italy, and ii) episodic GNSS measurements collect- Southern Italy is oored by the Southern Apennines fold-and-thrust
ed between 1995.68 and 2010.79, in order to provide the rst estima- belt in the western and central part, and by the Apulia foreland in the
tion of geodetic horizontal slip rate on a signicant number of active eastern part (Fig. 2, inset). The Southern Apennines accreted from
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 103

Fig. 2. Active faults in southern Italy. Compiled from the Database of Italian Seismogenic Sources (DISS; http://diss.rm.ingv.it/diss/), the GNDT (Gruppo Nazionale Difesa dai Terremoti)
database (Galadini et al., 2000), the Neotectonic map of Italy (Ambrosetti et al., 1987) and from recent literature as quoted in the text. The dashed white line in gure and in the inset
represents the boundary between the Adriatic and Tyrrhenian Moho (slightly modied from Di Stefano et al., 2011). Localities abbreviations are as follows: Ci, Cilento; CP, Campania
Plain; CV, Crati Valley; Da, Daunia Mountains; Ga, Gargano; Ir, Irpinia; Mu, Murge; MM, Matese Mountains; PhF, Phlegrean Fields; PM, Picentini Mountains; Pn, Pollino; Po, Potenza;
RV; Si, Sila; Sn, Sannio; SP, Sele Plain; Ta, Tavoliere; Tr, Tremiti Islands; VdA, Val d'Agri; VdD, Vallo di Diano; VV, Vesuvius Volcano. Inset shows the main tectonic provinces in southern
Italy, and the front of the fold and thrust belt.

Miocene to Early Pleistocene during the westward-directed subduction previously emplaced thin-skinned allochthon (Ferranti et al., 2009;
and delamination of the Adriatic slab, a small lithospheric fragment Monaco et al., 1998). These faults are at times seismically active in the
trapped within the articulated collision boundary between Africa and eastern part of southern Italy, in the Apulia foreland and beneath the
Europe (Faccenna et al., 2001; Malinverno and Ryan, 1986; Patacca frontal part of the belt (Figs. 2, 3).
et al., 1990). To the east, the Apulian sector of the Adriatic continental Since Late Pliocene, extension related to opening of the Tyrrhenian
margin served as the foreland to the eastward-migrating thrust belt. Sea and crustal delamination of the Adriatic slab affected the western
The Apulia foreland has an average ~ 30 km crustal thickness sector of the Apennines, forming coastal and intermontane basins con-
(Chiarabba and Amato, 1996) typical for continental crust, and is char- trolled by high-angle normal faults which crosscut the pre-existing
acterized by a ~6 km thick MesozoicCenozoic carbonate platform sed- thrust structures (Ferranti and Oldow, 2005; Hyppolite et al., 1994). Ex-
imentary cover (Mostardini and Merlini, 1986). To the west, contraction tension progressed from west to east, and starting from Early Pleisto-
was followed by extension, which since the Early Miocene resulted in cene an array of active normal faults was nested along the axial part of
hinterland stretching of continental crust and local emplacement of the belt and includes major recognized seismogenic structures (Fig. 2).
oceanic crust beneath the southern Tyrrhenian Sea (Fig. 1, inset; In western Campania region, extension was locally accompanied by vol-
Malinverno and Ryan, 1986). canism, whose emplacement was controlled by the intersecting fault
From Late Pliocene to Early Pleistocene, east-directed displacement networks (e.g. Milano et al., 2004).
of the contractional front in the Southern Apennines comes to an end The regional gravity anomaly map and Deep Seismic Soundings
(Fig. 2, inset; Patacca and Scandone, 2007). During this time, involve- studies outline the existence of a doubling of the Moho beneath the
ment of thick Apulian continental crust in the collision system caused Southern Apennines (Fig. 2, inset; Cassinis et al., 2003; Di Stefano
the (re-)activation of faults in the foreland plate with strikeslip and lo- et al., 2011; Morelli, 2000; Nicolich, 1989; Tiberti et al., 2005). This pat-
cally transpressional motion, that in the west broke through the tern is related to crustal delamination (Channell and Mareschal, 1989)
104 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

Fig. 3. Lower hemisphere, equal area projection of focal mechanisms compiled from recent literature and on-line catalogs for earthquakes with M 3.5; focal mechanisms are colored
according to rake: red, thrust faulting, blue, normal faulting, yellow, strikeslip faulting. Localities are as in Fig. 2. Dashed line is the same as in Fig. 2.

and impinging of the Tyrrhenian Moho, formed since the Miocene be- The swarm of active normal faults along the axis of the mountain
neath the stretched (thickness down to ~ 10 km) back-arc basin, over belt has been associated to most of the large historical earthquakes,
the sinking, old Moho of the Adriatic plate (Cassinis et al., 2003). which concentrate in the extensional domain (Figs. 1 to 3; DISS
Working Group, 2010). These faults are regarded as having an Early
3. Active deformation Pleistocene and in some cases an even younger age (Improta et al.,
2010; Maschio et al., 2005; Pantosti et al., 1993). Although the location
Active tectonics in southern Italy appears segregated in two discrete of the active normal faults is relatively well constrained, signicant
domains, which are characterized by different deformation mecha- uncertainties exist on the age initiation, geological displacements, as
nisms. The two domains are separated by a localized zone which runs well as on the rate of slip for individual structures.
along the axis of the mountain chain, and strikingly coincides with the Geodetic data so far available supply the rate of integrated extension
deep boundary between the Tyrrhenian and Adriatic Moho. In the west- in the Apennines but not the strain accumulation across individual
ern domain, active normal faults (Fig. 2) and seismicity (Figs. 1, 3) are normal faults. The geodetically-estimated rate of regional extension
concentrated within an ~50 km wide belt running along the mountain varies from ~23.5 (Ferranti et al., 2008; Serpelloni et al., 2005) to ~4
ridge (Papanikolaou and Roberts, 2007; Valensise and Pantosti, 2001). 6 mm/yr (Giuliani et al., 2009; Hunstad et al., 2003; Ward, 1994).
Within this belt, earthquakes have relatively shallow focal depths On the opposite, the eastern domain encompassing the foothills of
(~ 1015 km), and exhibit focal mechanisms with primarily normal the Apennines and the Apulia foreland locally show strikeslip and
faulting features and tensile axis roughly orthogonal to the Tyrrhenian subordinately thrust earthquakes (Fig. 3) with focal depths deeper
margin (Fig. 3). The belt of extensional earthquakes and active normal (~ 1535 km) than the extensional earthquakes in the west (Boncio
faults follows to the south the curvature of the orogenic arc. Seismolog- et al., 2007; Pondrelli et al., 2006). Within the Apulia foreland, whereas
ical observations are generally consistent with borehole breakouts the Gargano block and its offshore extension forms an active seismic
(Montone et al., 2012) and fault-slip data from Quaternary rocks belt with moderate-magnitude thrust and transpressional earthquakes,
(Faure Walker et al., 2012; Hyppolite et al., 1994; Maschio et al., 2005; the southern part including the Murge and Salento blocks lacks signi-
Papanikolaou and Roberts, 2007). cant seismicity (Figs. 1, 3). Notwithstanding, NESW convergence
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 105

Fig. 4. GNSS velocities and 95% condence ellipses referred to a xed: (a) Central Europe frame; and (b) Apulian frame (see text and Table 1 for additional details). Dashed line is the same
as in Fig. 2.
106 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

Table 1 network, surveyed for the rst time in 1995, was distributed in the en-
Relative GPSEuler vector and associated errors (95% of condence) for the xed Central tire southern peninsular Italy. Since then, the PTGA network has been
Europe (Eu) and the Apulia (Ap) reference frame. The Euler vector parameters (latitude
and longitude of pole, rotation rate) were estimated by minimizing, with a weighted least
surveyed in 1997, 2000, 2008 and 2010, and it has been expanded to
squares inversion (Palano et al., 2010), the adjustments to horizontal GNSS velocities aligned reach the current conguration of 42 benchmarks (Mazzella, 2011).
to a xed Central Europe reference frame (Nocquet and Calais, 2003). Sites used for the pole All the benchmarks were installed within hard bedrock not affected by
parameters computations are reported on Table S1 of the Supplementary section. fractures as far as possible. Particular care was devoted to locate obser-
Plate pair Latitude N Longitude E (/Myr) vation sites within coherent and rooted blocks with no evidence of
gravitational sliding. Overall, this analysis builds on 95 continuous and
EuAp 31.362 1.468 145.253 2.251 0.137 0.044
31 episodic measurement sites (Fig. 4).
GNSS data were processed by using the GAMIT/GLOBK software
(Herring et al., 2010). To tie the regional measurements to the
ITRF2005 reference frame (Altamimi et al., 2007), data coming from
between southern Apulia and the seismologically active front of the 12 continuously operating IGS (International GNSS Service) stations
contractional orogen in Albania and the Dinarids across the southern (AJAC, BRUS, CAGL, GRAS, GRAZ, JOZE, LAMP, MATE, MEDI, NOTO,
Adriatic Sea (Fig. 3) occurs at ~ 5 mm/yr (Hollenstein et al., 2003; NOT1 and ZIMM) were introduced in the processing. By using the
Jouanne et al., 2012; Serpelloni et al., 2005). South of Apulia, seismicity GLOBK software we combined, on a daily basis, our solutions (and
in the eastern domain resumes in north-eastern Calabria, where strike their covariance matrices) with global solutions (IGS1, IGS2, IGS3,
slip and subordinately thrust and extensional earthquakes are recorded IGS4, IGS5, IGS6 and EURA) provided by the Scripps Orbit and Perma-
along the continental margin, the shelf and into the Ionian Sea (Fig. 3). nent Array Center (SOPAC) to generate site time series. As a nal step,
Seismogenic deformation in the eastern domain is characterized by by using the GLORG module of GLOBK we combined these solutions
sequences, mainly segregated in the northern part of the region and their full covariance matrices to estimated a consistent set of posi-
(Pondrelli et al., 2003). These clusters, characterized by a sparse seis- tions and velocities in the ITRF2005 reference frame by minimizing
micity and by events with moderate magnitude, show an eastwest the horizontal velocity of the 12 continuously operating IGS stations
alignment (Fig. 3) which outlines the surface and subsurface location mentioned above. To account for correlated noise, we applied the real-
of broad systems of right-lateral faults which are thought to connect istic sigma algorithm of Herring (2003) to each of our time series, after
the deep part of the Apennines to the Apulia foreland (e.g. Boncio removing the best-tting annual and semi-annual signals and including
et al., 2007; Di Bucci et al., 2006). A northern array of right-lateral faults the estimated random walk component for every single component of
is buried at depth beneath the Daunia foothills, but resurfaces across the each station in our nal velocity solution.
Gargano upland (Fig. 2). To the south, two additional arrays of EW Our GNSS network shares some stations with those processed by
striking right-lateral faults are thought to join the Apennines with the D'Agostino et al. (2011) allowing for a rigorous integration of the two
Murge and Salento blocks of Apulia. Further south, NWSE striking estimated velocity elds. To this end, we aligned the published veloci-
transcurrent faults are mapped in northern Sila and in the Ionian Sea ties to the ITRF2005 reference frame by applying a seven-parameter
offshore the Pollino range (Fig. 2). Helmert transformation, obtained by nding the transformation that
Unlike the normal faults in the west, the location and geometry of minimizes the Root Mean Square (RMS) of differences between veloci-
active transcurrent faults in the eastern domain is not well understood, ties of common sites. The average discrepancies are small, and the RMS
because most of the faults are blind and supposedly segmented. These for the used 15 common stations is less than 0.4 mm/yr. The resulting
faults are thought to be Mesozoic or older extensional faults affecting velocity eld and additional details are reported in the Supplementary
the Apulia platform, and nowadays reactivated in strikeslip motion material section.
(Boncio et al., 2007; Di Bucci et al., 2006).
Boncio et al. (2007) evidenced that the 199091 sequence that
stroke the frontal belt of the Apennines near Potenza town in the east- 4.2. GNSS velocity elds
ern domain (Fig. 3), concentrated between a 1523 km depth range,
corresponding to the Apulia middle crust underthrust beneath the In order to analyze the crustal deformation pattern over the investi-
Apennines belt. The apparent anomalous earthquake depths (N15 km) gated area, the estimated GNSS velocities were aligned to a xed Central
and the connement within a relatively narrow depth range are Eurasia reference frame (Nocquet and Calais, 2003) and to an additional
explained by Boncio et al. (2007) according to the crustal rheology, local reference frame (that we label Apulia), whose Euler pole was
which consists of a strong mid crustal brittle layer sandwiched between estimated by using GNSS sites located in southern Apulia and eastern
two plastic horizons. Lucania regions (Tables 1; S1).
The subdue seismogenic role of faults within the eastern domain is In a Central Europe reference frame, the geodetic velocity eld
reected in the minor historical seismicity record when compared to depicts a general NW-directed motion at a rate of 13 mm/yr on
the western normal faults. In the eastern domain, strong historical the Tyrrhenian sector, and a N-to-NE-directed motion at a rate of
events are only sparsely documented and notably affect the northern 36 mm/yr on the Adriatic sector (Fig. 4a). The divergent velocity
part of the region only (Fig. 2). On the other hand, no major historical boundary is sharp and is positioned along the crest of the Apennines,
earthquakes are reported during the last 1000 yr in the central- where it coincides with the seismological and active faults boundary
southern Apulian foreland and eastern Lucania region (see 2011, for ad- (Figs. 2, 3). Further south, the geodetic boundary is loosely traced across
ditional details). central Calabria. Sites located in the Campania coastal plain show a
pattern that differs from the general trend and underlines local effects
4. GNSS velocity analysis deriving from volcanic centers on the western Campania margin
(Vesuvius and Phlegrean Fields, Fig. 2).
4.1. Network and data processing Velocities referred to the Apulia frame evidence, on the Tyrrhenian
side of Campania and Lucania regions, a pattern of WSW-oriented mo-
To improve the detail of the geodetic velocity eld in the investigat- tion at a rate of 1 to 4 mm/yr (Fig. 4b). The Tyrrhenian sector of Calabria
ed area, we updated the GNSS results reported in Palano et al. (2011) shows a similar pattern but with rates generally b1.5 mm/yr. On the
by extending the data processing up to the 2011.84 epoch and including Adriatic side of southern Italy, the velocity eld shows a broad ENE-
into the processing the periodic measurements carried out on the Peri- directed pattern with rates of about 14 mm/yr in the area between
Tyrrhenian GPS Array (PTGA; Ferranti et al., 2008). The 15-benchmark eastern Molise and northern Apulia. Sites in central-southern Apulia
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 107

Fig. 5. Geodetic strain-rate parameters: (a) rate of the areal change magnitude (background color code) and greatest extensional and contractional horizontal strain-rates (arrows);
(b) rate of maximum shear strain.
108
Table 2
Modeled active crustal faults of the western and eastern domain, and gravitational faults. For each structure, the geological source parameters, the kinematic and the Quaternary slip rate (bold for latest PleistoceneHolocene slip rate), and the source

L. Ferranti et al. / Tectonophysics 621 (2014) 101122


reference are listed. NMI = Neotectonic map of Italy (Ambrosetti et al., 1987); DISS = Database of Italian Seismogenic Sources, Version 3.1.1 (http://diss.rm.ingv.it/diss/); ITIS = Individual Seismogenic Source; ITCS = Composite Seismogenic
Source; GNDT = Gruppo Nazionale Difesa Terremoti; CNR MS = Consiglio Nazionale delle Ricerche, Modello Strutturale d'Italia.

ID Fault name Length Coordinate Down-dip Min depth Max depth Strike Dip Kinematic Dip Rake () Slip-(Throw, T)-rate Horizontal slip-rate Source
(km) width (km) (km) (km) () () geology direction (mm/yr) (mm/yr)
West tip East tip

Eastern domain
1 Biferno 30 414312 415848 18.3 2 16 55 50 n/a SE n/a n/a n/a NMI
144236 15112
2 Frosolone 36 414030 414030 14.9 11 25 269 70 RN N 230 0.11.0 0.11.0 DISS (ITIS095)
142031 144642
3 RipabottoniS. Severo 64 41417 414323 19.1 6 25 270 85 R N 180220 0.10.5 0.10.5 DISS (ITCS053)
144751 153347
4 S. Marco in LamisMattinata 45 414328 414252 25.1 0 25 270 85 RN N 200230 0.11 0.11 DISS (ITCS058),
153431 16539 0.8 0.8 (Tondi et al., 2005)
5 Castelluccio dei SauriTrani 93 411649 411628 12.2 11 23 270 80 R N 170190 0.10.5 0.10.5 DISS (ITCS004)
152458 163130
6 Mirabella EclanoMonteverde 60 41222 405821 16.6 1 16 280 65 RN N 230250 0.11.0 0.11.0 DISS (ITCS084)
145629 153843
7 Potenza 15 403951 40388 6 15 21 95 88 R S 175 0.10.5 0.10.5 DISS (ITIS084)
154723 155819
8 Palagianello 65 403931 403731 9 13 22 90 85 R S 170190 0.10.5 0.10.5 DISS (ITCS005)
161951 17535
9 AmendolaraSibari 68 394738 393343 26.9 3 22 286 45 TL N 235 0.71.0 0.71.0 Ferranti et al. (2009)
162910 171356
10 Rossano 40 393637 393352 28.3 2 22 280 45 LT N 220 1.05.0 (T) 1.02.0 GNDT (99),
162228 164950 (Van Dijk et al., 2000)
11 Cecita 20 392952 391928 25.4 0 22 162 60 NL SW 320 n/a n/a n/a
162952 16348
12 Lakes 30 391950 39657 25.4 0 22 142 60 NL SW 320 1.2 0.6 Galli and Bosi (2003)
163259 164601
13 LameziaCatanzaro 21 385534 385441 n/a n/a n/a 95 70 L S n/a n/a n/a GNDT (104)
161238 16273

Eastern domain, shallow-crustal creeping faults


14 Varano 35 415420 414546 5.8 0 5 315 60 N NE 270 n/a n/a Ithaca, (Bertotti
154153 16523 et al., 1999)
15 Castel del Monte 44 4174 405443 20.8 0 18 300 60 N NE 270 n/a n/a CNR MS, (Del Gaudio
161319 164034 et al., 2007)
16 AvenaLauropoli 28 394543 395645 4.6 0 4 222 60 N E 270 x 0.3 Ferranti et al. (2009)
162317 163616

Western domain
17 Acquae Iuliae 22 41322 41259 13.2 1 13 125 65 N SW 270 0.20.4 0.10.2 GNDT (52) (Galli and
14231 141530 0.451.9 0.20.8 Naso, 2009)
18 Matese Lake 36 413022 412133 14.3 0 13 120 65 N SW 270 0.51.0 0.20.4 GNDT (53), [this paper]
14847 143219
19 Boiano Basin 24 413122 41245 13.8 1 12.3 55 55 N NE 270 0.11.0 (T) 0.10.6 DISS (ITIS004), (Galli
14195 143356 0.50.9 0.30.5 and Galadini, 2003;
Pordo et al., 2002)
20 Volturno Plain N 15 411560 411132 13.9 1 13 120 65 N SW 270 0.51.5 (T) 0.20.5 (T) 0.20.6 GNDT (65)
14349 141318 0.10.2
21 Ariano Irpino 25 412540 411323 14.9 11 25 277 70 RN N 230 0.11.0 0.10.3 DISS (ITIS092)
14458 145216
22 Lauro 11 405455 405132 13.9 1 13 125 60 N SW 270 0.20.5 (T) 0.10.2 GNDT (76) (Ambrosetti
143225 143911 et al., 1987)
23 Volturara 30 40545 404631 13.9 1 13 310 60 N NE 270 0.3 (T) 0.2 Papanikolaou and
145054 15948 Roberts (2007)
24 Irpinia Fault (Colliano segment) 28 405112 404128 15 1 14 310 60 N NE 270 0.40.6 0.20.3 DISS (ITIS077)
1584 152317 (Papanikolaou and
Roberts, 2007)

L. Ferranti et al. / Tectonophysics 621 (2014) 101122


25 Irpinia Fault (S. Gregorio segment) 9 404015 403749 15 1 14 300 60 N NE 270 0.40.6 0.20.3 DISS (ITIS078)
152438 153011
26 MelandroPergola 18 403351 402646 11.3 1 10.8 317 60 N NE 270 0.11.0 0.10.5 DISS (ITIS010)
15304 153844
27 East Agri 20 402455 401929 13.1 5 15 120 50 N SW 270 0.71.3 0.6 GNDT (90), (Papanikolaou
154240 15556 0.83 (T) 0.5 and Roberts, 2007)
28 M. Maddalena 23 402353 401458 13.5 1 12.7 316 60 N NE 270 0.11.0 0.10.5 DISS (ITIS008),
153918 155037 0.2 0.2 (Maschio et al., 2005)
29 South Diano 15 401954 401235 14.4 1 12 140 50 N SW 270 0.50.7 (T) 0.30.4 GNDT (86), (Villani and
153856 154340 Pierdominici, 2010)
30 Mercure Basin 32 40552 395540 15.0 0 13 120 60 N SW 270 0.20.4 (T) 0.10.2 Papanikolaou and
155230 161042 Roberts (2007)
31 Pollino 32 395915 394923 17.3 0 15 125 60 N SW 270 0.20.5 (T) 0.10.2 GNDT (97) (Papanikolaou
15593 161732 0.20.4 (T) 0.10.2 and Roberts, 2007)
32 Crati Valley 51 393615 39846 10.4 1 10 175 60 N W 270 0.11.0 0.10.5 DISS (ITCS015),
161555 161843 0.8 (Galli et al., 2007)
Adopted value.

109
110 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

have an E-directed motion at a rate b 1 mm/yr, and the Ionian side of Adriatic and Ionian coastal region that shows locally moderate dilata-
northern Calabria displays a SE-directed velocity pattern at b 1.5 mm/yr. tion rates (Fig. 5a).

4.3. Geodetic strain-rate eld 5. Geodetic slip rate across active faults

To compute the 2D strain-rate tensor over the studied area, starting 5.1. Geometric and kinematic parameters
from the observed horizontal velocity eld and associated covariance
information, we derived a continuous velocity gradient tensor on a reg- In order to estimate the geodetic slip rate and later the moment rate
ular 0.2 0.2 grid using a spline in tension technique (Wessel and for active faults, in a rst step we derived the fault geometric and kine-
Bercovici, 1998). The tension is controlled by a factor T, where T = 0 matic information from published and unpublished studies and data-
leads to a minimum curvature (i.e. natural bicubic spline), while T = bases (Table 2). In particular, based on the location of GNSS stations,
1 leads to a maximum curvature, allowing for maxima and minima we have singled out 32 structures equally distributed within the west-
only at observation points. In the computations we set T = 0.4, slightly ern and eastern domain. For each structure we stipulated length (L),
different from the value (T = 0.3) suggested by Wessel and Bercovici strike (), dip (), minimum (Dmin) and maximum (Dmax) depth,
(1998) for topographic interpolation. As a nal step, we computed the down-dip width (H; dened as H = d / sin where d is the difference
average 2D strain-rate tensor as derivative of the velocities at the between Dmax and Dmin values), rake and slip rate. Parameters of
nodes of each grid cell. The estimated strain-rates are shown in Fig. 5a many structures (particularly for those of the eastern domain) were ex-
as greatest extensional ( H max ) and contractional ( h min ) horizontal tracted from the Database of Italian Seismogenic Sources (DISS) Version
strain-rates and rate of the areal change, and in Fig. 5b as rate of the 3.1.1 (http://diss.rm.ingv.it/diss/). Two types of seismogenic sources
maximum shear strain. were drawn from the DISS database, namely Composite and Individual
We present here a rened version of the results shown by Palano sources (Table 2). Unlike Individual sources, which are relatively well
et al. (2011). A continuous belt of positive area change with high parameterized faults which can produce characteristic earthquake
strain-rate values (up to ~ 110 nanostrains/yr) and NESW-oriented ruptures, Composite sources are viewed as seismogenic volumes
H max is recognized along the crest of the Apennines (Fig. 5a), in agree- which incorporate an unspecied number of Individual sources, and
ment with active crustal stretching documented by geological and seis- thus their characteristic rupture length is poorly dened or unknown.
mological data. However, stretching does not appear quantitatively Although this may imply a smaller accuracy of source description, they
uniform, with areas of relative maxima and minima alternating are nonetheless used to achieve completeness of the record of potential
along the belt of extensional strain and positive area change rate. earthquake sources, and thus contribute to the development of regional
In detail, from north to south, the border between central and south- probabilistic seismic hazard assessment (Basili et al., 2008). In the
ern Apennines (at the boundaries of Molise, Lazio and Campania present study, because of the limited number of stations used to
regions, Fig. 5a) is characterized by a rate of area change of up to constrain the geodetic slip rate, both Composite and Individual sources
~ 80 nanostrains/yr with a NESW trending H max and a locally positive were treated as single structures. Parameters of some faults were
h min indicating a bi-axial extensional pattern. Southward, the belt of extracted from the GNDT (Gruppo Nazionale Difesa dai Terremoti)
high-extension rate follows the curvature of the chain, with principal database (Galadini et al., 2000), or from individual papers and maps
axis having a constant NESW trend, and theh min swinging from nega- (Table 2). For these structures, source information is less complete
tive to positive. The rate of area change, which is slightly reduced in the relative to the DISS database, particularly as regards the fault depth.
border sector between Campania and Apulia, increases again at the For few faults, the parameters were rstly dened based on our detailed
boundary between Campania and Lucania and has two peaks up to or reconnaissance investigation.
~ 110 nanostrains/yr. Southward, the belt of rapid area dilation and The modeled deformation sources have length typically in the ~10
high extension rate sharply disappears at the CalabriaLucania border 35 km range, with only some of them, and namely the Composite
(Fig. 5a). sources or some other regional fault array, exceeding 40 km. Dmin ranges
A second belt of area dilatation, albeit of minor magnitude typically from null (outcropping fault) to 1 km for the major and minor
(~ 30 nanostrains/yr, with local peaks up to ~ 60 nanostrains/yr) and faults of the extensional domain, respectively. Faults of the foreland part
with E- to SE trendingH max, characterizes the eastern coast of northern of the transcurrent domain have the top at shallow depth (1 km) or are
Calabria, and pushes northward into central Apulia where the maxi- exposed, as occurring in the Gargano promontory. On the opposite,
mum principal axis progressively turns to a NE trend (Fig. 5a). Further faults of the eastern domain which underlay at depth the foothills of
north, this positive areal change belt is offset to the left and crosses the Apennines and are structurally covered by the thrust belt have
the Gargano promontory, with H max axis keeping to a NE trend. Dmin at 611 km, rarely less (Table 2). In northern Calabria, strikeslip
Areas of negative area change rate are limited. Apart from the coastal faults outcrop or have tips at shallow depths below the coastal plain
sector of the Campania Plain, where alternating patches of negative and or the continental shelf and slope.
positive areal changes may testify local volcanic processes, in the The Dmax values stipulated in the sourced datasets, which is related
Apennines they are restricted to an arcuate belt that follows the front to the crustal seismogenic thickness (hereinafter Hs), was calibrated
of the chain between Molise and Lucania, with prevailing NE- to E- through the analysis of the depth distribution of seismicity. To this
trending h min axis and up to 80 nanostrains/yr (Fig. 5a). A marginally aim, we dened four different sectors (Fig. 6a), one in the western do-
signicant contraction was already evidenced in this area by Giuliani main and three in the eastern domain. The three sectors within the east-
et al. (2009). ern domain are: the foothills belt in the Apennines, the foreland in the
The maximum shear strain rate map (Fig. 5b) shows highest values Apulia region and the eastern part of Calabria. By taking into account
along the crest of the Apennines, which spatially coincide with those all earthquakes with M 2.2 occurred in the investigated area since
areas that in the last 103 yr were struck by strong earthquakes and cur- 1983 (http://iside.rm.ingv.it), we computed a cumulative frequency
rently experience a diffuse low-to-moderate seismic release (Figs. 1, 3). plot of the depth-distribution of instrumental seismicity for each sector.
Particularly, the CampaniaMolise border in the north and the Campa- The value of 2.2, usually dened as magnitude of completeness, repre-
niaLucania border in the south are the sectors with highest shear strain sents the lowest magnitude for which the catalog can be considered
of up to 80 nanostrains/yr. High maximum shear strain-rate up to 60 complete (see Palano et al., 2011 for additional details). The plot
nanostrains/yr can be locally recognized at the volcanic districts of the shows that 90% of earthquakes (e.g., Marone and Scholz, 1988) occur
Tyrrhenian margin of Campania as well. On the opposite, the shear in the upper crust within a depth that increases from the western to
strain-rate in northern Calabria, Lucania and Apulia is low, even in the the eastern domain (Fig. 6b). The limiting depth for 90% of M 2.2
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 111

earthquakes is 13, 17 and 19 km in the western, frontal and foreland stations have an observational record that partly span the 2002 seismic
part of the Apennines; the limiting depth in the Calabria frontal belt is events. Thus, large co-seismic offset related to these earthquakes do not
22 km. These results are in good agreement with conclusions achieved signicantly affect our geodetic velocity eld.
in the study of local seismic sequences beneath the frontal part of the The geodetic displacement resolved onto the faults may embed
Apennines (Boncio et al., 2007; Di Luccio et al., 2005; Latorre et al., components of both seismic and aseismic slip, and of strain accumula-
2010). The Dmax listed in the DISS database is slightly deeper than our tion because of friction on the fault plane (e.g. Savage, 1983). Sites
estimate of the limiting depth for 90% of M 2.2 earthquakes, because used here to compute slip rates for individual faults are reasonably far
the former is based on the depth of the most energetic shocks within the from any of the faults (typically N1 km), and thus we assume that our
sequence associated to the seismogenic source. We mostly adopted the slip estimate reects the steady surface deformation. Based on this
maximum depth value reported in the DISS where available, or alterna- reasoning, we compared the observed interseismic displacement to
tively, used our estimate. the geologically recorded slip rate. We assume the long-term or geologic
When not listed, strike and dip of faults were estimated based on (over many earthquake cycles) fault displacement means that the slip has
published and unpublished geological information. For the purpose of occurred on frictionless fault planes.
computation, strikes were averaged between the mapped surface tips
of faults or fault systems (compare Figs. 2 and 6). Although this simpli- 6. Scalar geodetic and geologic moment rate
cation may introduce errors for low-angle dipping faults, the modeled
structures have all moderate- to high-angle dips and thus errors related To better understand the relation between long-term and contem-
to wrong strikes are negligible. Faults of the eastern domain have porary deformation on the faults for which geodetic slip rate were com-
assigned dips between 70 and 90, a range of value consistent for puted, we estimated in a second step the scalar geological and geodetic
strikeslip and oblique-reverse faults. Fault dips for the western domain moment rates on these faults (Table 3).
are in the range of 5565 typical for normal faults. The geological moment rate was computed using the Brune (1968)
The kinematics of modeled faults rely on structural analysis carried formulation:
out in recent deposits and on stress tensor data derived from available
focal mechanisms. We list in Table 2 the general kinematic role of M geol HLu 1
each structure as reported in the consulted databases and studies, and
the rake stipulated in published accounts or based on our knowledge. where is the shear modulus of the rocks involved in faulting with
Finally, because we are mostly interested at scalar rates, we draw assigned value of 30 GPa, L and H are length and down-dip width, re-
from published sources the geologic slip rate of faults, which are typical- spectively, of the dislocation along which faulting occurred and u is
ly averaged on a Pleistocene or on a Holocene time interval (the two the average displacement on the dislocation, which corresponds to the
different time spans are reported in Table 2). It must be noted that the stipulated geologic slip rate (Table 2). We estimated both the Pleisto-
geological estimates are in some cases loosely constrained particularly cene and Holocene moment rate depending on the available
for blind faults of the eastern domain. For sake of comparison with the information.
geodetic rates, the slip rate reported in consulted repositories was The geodetic moment rate was estimated using the Savage and
converted to the horizontal slip rate by using the dip angle. For the Simpson (1997) formulation:
strikeslip faults, the indicated value represents the horizontal compo-  
nent of slip. M geod 2H s A Max H max ; h min ; H max h min 2

5.2. Computation of geodetic slip rates where is the shear modulus of rocks within the focal volume, with
assigned value of 30 GPa, Hs is the seismogenic thickness (Table 2), A
By selecting the closest GPS stations straddled by each fault, we com- is the surface area of the investigated region and H max and h min are
puted the parallel (strike) and perpendicular (tensile or shortening) the principal horizontal strain-rates previously described. For each
horizontal components of the geodetic slip rate (Table 3) by using a fault listed in Table 2 we dened a rectangular-hull with surface area
geometrical approach that projects the vector decomposition of GNSS dened as A = (4)(L + 2) where 4 and L + 2 are the short and
stations velocities onto the fault trace. Only in ten cases we used two the long side of the rectangular-hull, L is the length of the fault and ,
stations for each pair of blocks across the fault in order to compute the with arbitrary assigned value of 8 km, represents the minimum distance
horizontal slip rate. For the remaining faults, selected stations were between the fault trace and the side of the rectangular-hull (Fig. 7). By
three or more (see Table 3 for details). taking into account the observed horizontal velocity eld and associated
For each fault, the kinematic estimation was carried out by using a covariance information, in a rst step we computed the expected veloc-
Monte Carlo simulation (Metropolis and Ulam, 1949) to a set of selected ity values at the nodes of each rectangular-hull and as a second step we
stations surrounding the fault. During the iterations of Monte Carlo sim- derived the average 2D strain-rate tensor at the center of each
ulation, values of station velocities were sampled at random from their rectangular-hull. Finally, by adopting the formulation described in (2)
probability distributions. The geometrical approach was used to calcu- and parameters reported in Table 2, we computed the geodetic moment
late the kinematic parameters at each iteration. After 10,000 iterations rate (Table 3).
the result was a probability distribution of possible parameters. The We did not incorporate in this study an estimation of the seismic
kinematic parameter estimations were the expected values for the moment rates because of the limited time-span (~ 30 yr) covered by
obtained probability distributions and the associated errors were the the instrumental seismicity catalog, which is too short to cover the
related standard deviations (Table 3). recurrence time of most of the large earthquakes and consequently is
Within the time period considered in this study, no signicant earth- not an adequate measure of the seismically-released strain.
quake has occurred in southern Italy nor in the immediate surround-
ings, apart for the 2002 Molise sequence in the Daunia sector of the 7. Results
eastern domain (Fig. 3), which had a cumulative co-seismic offset of
120 mm for the two main shocks (both with Mw = 5.7) as recorded 7.1. Strikeslip faults of the eastern domain
by GNSS data (Giuliani et al., 2007). Within the western domain, no es-
timates are available for the offset associated to the 1998 Mw = 5.6 In the eastern part of southern Italy, GNSS velocities depict substan-
Mercure earthquake. The permanent GNSS stations considered here tial displacement across a deformation belt spanning between the
were installed after both earthquakes, and only some episodic PTGA Apennines and northern Apulia (Fig. 8). Moving from west to east, the
112 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

b
Fig. 6. a) Sketch map of the four seismogenic domains identied in southern Italy; b) Cumulative frequency plot of the depth-distribution of instrumental seismicity with M 2.2 for the
four seismogenic domains. Earthquakes having depth xed at the default values of 5 and 10 km (usually adopted for mid-continental areas when the depth of the earthquake is poorly
constrained by available seismic data) were also included in the computation.

rst signicant velocity gradient is encountered across an ~30 km NE Pleistocene strikeslip rate estimate range, is consistent with the geo-
SW-trending boundary spatially coinciding with the Biferno River detic result for the whole fault system (Fig. 9a). The eastern extension
Valley (n. 1, Fig. 8) in the Daunia sector. Here, no known seismogenic of the deformation belt is tracked by high-resolution seismic reection
source is listed in available databases, but our analysis reveals 0.9 proles offshore Gargano, where it deforms the sea-bed with right-
0.5 mm/yr right-lateral slip with a minor contraction component lateral displacement which decreases to ~ 0.2 mm/yr (Di Bucci et al.,
(Fig. 9a) across a structure inferred to run beneath the valley. It is possi- 2009; Ridente and Trincardi, 2006).
ble that this structure, reported in generalized structural maps To the south, residual GNSS velocities across the EW striking fault
(Ambrosetti et al., 1987), is part of a broader deformation belt running system separating Murge and the lowland south of Gargano
offshore the Adriatic Sea north of the Gargano upland and characterized (Castelluccio dei SauriTrani, n. 5; Fig. 8) indicate 1.2 0.3 mm/yr
by thrust and strikeslip earthquakes (Fig. 3; Del Gaudio et al., 2007). right-lateral slip. The fault system is buried under the lowland and
Along the offshore extension of the Biferno structure, oblique thrust does not show seismogenic activity (Figs. 2, 3), but the geodetic dis-
faults accommodate uplift of the Tremiti Island high (Fig. 2) and fold placement is over twofold the estimated maximum geologic slip rate
the sea-oor (Ridente and Trincardi, 2006). Thus, geodetic result on- (Tables 2, 3; Fig. 9a).
land is consistent with marine geophysical data. An ~EW trending fault system further south is located beneath the
The transpressional structure underlying the Biferno Valley merges foothills of the Apennines (Mirabella EclanoMonteverde, n. 6, Fig. 8),
to the southwest with the ~ 150 km long EW striking array of right- and is characterized by 1.1 0.5 mm/yr right-lateral slip with a slight
lateral faults connecting the Apennines to the Apulia foreland (Fig. 2). extensional component. The geodetic signal is in broad agreement
The location of these structures has been proposed in the aftermath of with the inferred geological slip rate and kinematics (Fig. 9a), and
the 2002 seismic sequence, caused by ruptures along right-lateral with the proposition of normal-right oblique slip during the MS = 6.6,
blind fault segments (Fig. 3; Di Luccio et al., 2005). At the western part 1930 earthquake (Pino et al., 2008) on a fault spatially coinciding with
of the array, the geodetic signal is however inadequate to precisely esti- the location of source n. 6. GNSS network conguration and observed
mate the horizontal slip rate across the Frosolone (n. 2, Fig. 8) and velocities are not adequate to estimate a possible eastern extension of
RipabottoniS. Severo sources (n. 3) of the DISS database (Fig. 9a). this fault system in the Murge plateau, however.
The geodetic signal, on the other hand, is better in the eastern part of Further south, an ~80 km long system of deeply buried (top at 1315
the array, within the Gargano highland, where it suggests right-oblique depth) EW striking faults has been proposed (DISS Working Group,
transtensional displacement at ~ 0.5 mm/yr across the exposed 2010) to stretch from the Lucania Apennines to the Apulia foreland at
MattinataS. Marco in Lamis Fault (n. 4). The fault spatially coincides the border between the Murge and Salento blocks (sources n. 7 and 8,
with a seismically active belt characterized by a mixture of thrust and Fig. 8). Geodetic estimate for the western part of the system (Potenza
strikeslip earthquakes (Fig. 3). The structure consists of three seg- Fault), are of a small right-normal motion (Fig. 9a). This result is consis-
ments, all of them straddled by GNSS sites. Across the eastern segment, tent with the seismological characterization of the structure responsible
Holocene dextral slip at ~0.8 mm/yr is recorded by paleoseismological for several seismic sequences in the last two decades, with the larger
studies (Piccardi, 1998; Tondi et al., 2005), and, together with the shocks occurring on EW striking, right-lateral fault planes (Boncio
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 113

Table 3
Modeling results for the active faults (same as in Table 2), listing for each fault the GNSS sites assigned to individual blocks straddled by the fault, the geodetic slip rate and kinematics
(positive for right-lateral and extensional motion, and negative for left-lateral and contractional motion), and the geodetic and geological moment rate (bold for moment-rate estimations
derived from latest Pleistocene-Holocene geologic slip rate). Question mark on kinematics indicates a more uncertain estimate.

ID Fault name GPS sites Geodetic slip-rate (mm/yr) Kinematic geodesy Moment rate (Nm/yr)

Block 1 Block 2 Strikeslip Opening Geodetic Geologic

Eastern domain
1 Biferno TRIV, MAUR, FRES, PETC LARI, MELA 0.92 0.50 0.29 0.47 R 5.64E + 15 n/a
2 Frosolone BSSO, CABA TRIV, MAUR 0.07 0.96 0.43 0.35 RN? 2.36E + 16 1.61E + 151.61E + 16
3 RipabottoniS. Severo FOGG, MOCO, SVIT MELA, LARI, SPCI 0.13 0.98 1.12 0.92 RN? 2.81E + 16 3.66E + 151.83E + 16
4 S. Marco in LamisMattinata ACQ1, MSAG LASE, SGRT 0.45 0.57 0.53 0.29 RN 7.02E + 15 3.39E + 153.39E + 16
2.71E + 16
5 Castelluccio dei SauriTrani FOGG, MARG CADM, MRVN 1.16 0.28 0.23 0.39 R 2.49E + 16 3.40E + 151.70E + 16
6 Mirabella EclanoMonteverde ANDR, CAFE, GROT ACCA, SERB, SGTA 1.05 0.54 0.70 0.58 RN 1.49E + 16 2.98E + 152.98E + 16
7 Potenza TITO CAMA 0.28 0.28 0.26 0.24 RN 3.72E + 15 2.70E + 141.35E + 15
8 Palagianello SVTO, MATE, GINO NOCI, FASA 1.03 0.66 0.10 0.29 R 1.52E + 16 1.76E + 158.81E + 15
9 AmendolaraSibari TREB ROSS 1.01 1.12 0.76 0.60 LN? 1.31E + 16 3.84E + 165.48E + 16
10 Rossano ROSS PIPA 1.22 0.37 0.78 0.38 RT? 9.00E + 15 3.39E + 16
11 Cecita LUZX, CECI, CAMO PIPA 0.31 0.22 0.81 0.25 RN 5.12E + 15 n/a
12 Lakes CRLM, SERS CCRI, STSV 1.06 0.78 1.05 0.29 RN 1.55E + 16 2.74E + 16
1.37E + 16
13 LameziaCatanzaro SERS, AMAN LAME 0.14 0.32 0.02 0.18 RN? 2.79E + 15 n/a

Eastern domain, shallow-crustal creeping faults


14 Varano ISCH LASE, MSAG 0.16 0.46 2.12 0.32 N 3.05E + 15 n/a
15 Castel del Monte MRVN, CADM, AMUR VALZ 0.98 0.33 1.51 0.31 NR 1.26E + 16 n/a
16 AvenaLauropoli TIMP TREB 0.07 0.44 1.09 0.44 N 3.26E + 15 1.16E + 15

Western domain
17 Acquae Iuliae VAGA LONG 0.37 0.23 1.21 0.23 N 8.10E + 15 1.75E + 153.50E + 15
3.93E + 151.66 + 16
18 Matese Lake LONG, CAMP VAGA, PTRJ 0.65 0.37 0.76 0.52 NL 8.99E + 15 7.75E + 151.55E + 16
19 Boiano Basin LONG, CAMP BSSO, CABA 1.53 0.90 1.74 0.54 NR 6.87E + 15 1.21E + 151.21E + 16
4.97E + 158.94E + 15
20 Volturno Plain N VAGA MODR 0.80 0.21 0.27 0.21 N 4.57e + 015 3.29E + 159.86E + 15
1.31E + 154.60E + 15
21 Ariano Irpino PSB1, SACR SVIT, MOCO, CICL 0.15 0.54 2.06 0.81 N 1.31E + 16 1.34E + 151.34E + 16
22 Lauro AVEL PACA 0.59 0.37 1.61 0.30 N 2.20E + 15 1.06E + 152.64E + 15
23 Volturara BAGN, MCRV TOBE 1.86 0.38 1.30 0.87 NR 1.11E + 16 4.32E + 15
24 Irpinia Fault ANDR, SNAL, MRLC BAGN, MCRV 0.51 0.35 1.52 1.06 NR 1.08E + 16 5.04E + 157.57E + 15
(Colliano segment)
25 Irpinia Fault GIUL DONG, CDRU 0.08 0.32 1.0 0.27 N 6.72E + 15 1.62E + 152.43E + 15
(S. Gregorio segment)
26 MelandroPergola CDRU, DONG TITO 0.45 0.37 2.83 0.22 N 7.47E + 15 6.11E + 146.11E + 15
27 East Agri VOLX GRUM, MCEL 1.63 0.52 0.66 0.29 RT 5.67E + 15 5.48E + 151.02E + 16
8.49E + 15
28 M. Maddalena MTSN, SLCN MCEL, GRUM 1.23 0.64 1.69 1.43 NL 6.08E + 15 9.32E + 149.32E15
2.80E + 15
29 South Diano MTSN CARM 0.43 0.30 1.05 0.30 N 4.13E + 15 4.22E + 155.90E + 15
30 Mercure Basin TIMP CUCC 0.47 0.27 1.54 0.27 N 4.48E + 15 3.33E + 156.66E + 15
31 Pollino SERR, CASS ACRI 0.32 0.34 0.09 0.57 RN? 5.71E + 15 3.84E + 159.60E + 15
3.84E + 157.68E + 15
32 Crati Valley BISI, LUZZ, LUZX CETR, CARO 0.35 0.45 0.16 0.45 RN? 3.20E + 15 1.59E + 151.59E + 16
2.54E + 16

et al., 2007; Di Luccio et al., 2005). To the east, a parallel structure by geodetic data, although biased by uncertainty (Fig. 9a), is in agree-
(Palagianello source) has been proposed to connect the mountain belt ment with seismic prole analysis and focal mechanisms of middle-
and the foreland plateau (Fig. 8); geodetic data indicate possibly sub- upper crustal earthquakes, that along the Amendolara Ridge are charac-
stantial right-lateral slip at 1.0 0.7 mm/yr (Fig. 9a), in agreement terized by left-slip on NW-striking nodal planes (Ferranti et al., 2009,
with the inferred seismogenic source kinematics albeit at slightly larger under revision; Fig. 3).
rate. Further south, on the northern border of the Sila massif, GNSS veloc-
On the Ionian margin of southern Italy, signicant differential mo- ities document that the NWSE striking RossanoS. Nicola Fault array
tion exists between GNSS sites in eastern Lucania and northern Calabria (n. 10), has a right-lateral motion at 1.2 0.4 mm/yr, coupled to a
and those further south (Fig. 8). Sites ROSS and PIPA in northern Sila shortening component at 0.8 0.4 mm/yr (Fig. 9a), which may result
have left-directed motion with respect to sites in and around the Pollino from sites ROSS and PIPA being located at a restraining bend of the
Range. The differential motion occurs across a deformation belt which fault system (Figs. 2, 8). This nding is consistent with studies proposing
borders the southern side of Pollino and continues offshore along the a transpressive character of the fault system (Ferranti et al., 2009; Tansi
southern ank of the Amendolara Ridge (Fig. 8; n. 9). Morphotectonic et al., 2007; Van Dijk et al., 2000), and with focal solutions of the most
and marine geophysical studies (Ferranti et al., 2009, under revision) energetic earthquakes (Fig. 3), although the right lateral geodetic mo-
documents that this ~70-km long fault system (AmendolaraSibari) is tion is not consistent with geology and seismology. The magnitude of
active and involves SW-oriented, left-oblique transpressional blind seg- the geodetic estimation is at the lower boundary of the stipulated geo-
ments extending down to ~10 km depth. Thanks to the irregular coast- logic rate (15 mm/yr; Galadini et al., 2000). However, we believe
line and the geodetic network conguration, the GNSS data are able to that the geological slip rate, when compared to all other faults of the re-
track displacement on the offshore fault. The broad left slip estimated gion, is probably overestimated. We have arbitrarily reduced the upper
114 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

structure, possibly reactivating the old thrust surface, which accommo-


dates gravity-controlled, seaward motion of the eastern part of Gargano.
Accordingly, we assign to the fault a shallow rooting depth of 5 km
(Table 2).
To the south, on approaching the Adriatic coast of Murge, NE diver-
gence occurs between sites straddled by a NWSE striking discontinuity
which cuts through the plateau (here baptized Castel del Monte Fault,
n. 15; Fig. 8). This structure may coincide with the location of the
small-energy 1991 Castel del Monte seismic sequence (Mmax = 3.1)
which was characterized by prevailing extensional focal mechanisms
(Del Gaudio et al., 2005). Like for the Varano Fault, the signicant
horizontal extension (1.5 0.3 mm/yr) across the Castel del Monte
Fault is suggestive of a contribution from gravity-controlled surface
Fig. 7. Draft of the rectangular hull used for the computation of the geodetic moment-
rates. The upper surface area (A) of the rectangular-hull is dened as A = (4)(L + 2) motion toward the bathymetric low.
where 4 and L + 2 are the short and the long side of the rectangular-hull, L is the length On the Ionian side of northern Calabria, signicant divergence is ob-
of the fault and , with arbitrary assigned value of 8 km, represents the minimum distance served between sites TREB on the coast and TIMP in the Pollino Moun-
between the fault trace and the side of the rectangular-hull. H, dened as H = d / sin tain range, and indicates horizontal extension at 1.4 0.5 mm/yr across
(where d is the difference between Dmax and Dmin values; see main text and Table 2 for de-
tails) represents the down-dip width of the fault; Hs is the crustal seismogenic thickness.
a NESW striking fault system (AvenaLauropoli, n. 16; Fig. 8). The geo-
detic results are consistent with geological analysis, which documented
that the AvenaLauropoli Fault is composed of discontinuous and small-
bound to 2.0 mm/yr (Table 2) to maintain coherency with the regional length fault segments rather than a through-going lineament, and is
displacement eld. accompanied by locally dramatic stratal titling indicative of rooting at
In the interior of the Sila massif, the NNWSSE striking Sila Fault sys- shallow depths (Ferranti et al., 2009). The fault was interpreted by
tem (Fig. 8) has been proposed, based on trench investigations, to be re- Ferranti et al. (2009) as part of a supercial extensional detachment
sponsible for large (M 6.26.7) historical earthquakes (Galli and Bosi, related to deep uplift of the Pollino range above transpressional faults.
2003; Galli and Scionti, 2006). According to Galli and Scionti (2006),
the system is composed of two segments, the Cecita and Lakes Faults 7.3. Normal faults of the western domain
to the NW and SE, respectively, and of minor segments further SE, that
have normal-left oblique kinematics. In the last four centuries, this sys- Moving from the northwest along the axis of the Apennines, the rst
tem ruptured almost entirely from NW toward SE (163817441832), major extensional structures are encountered at the Matese massif, lo-
with the exception of the north-western segment, the silent Cecita cated at the CampaniaMolise boundary. On the southwest side of the
Fault. Geodetic observations, on the other hand, indicate right-oblique massif, sites LONG and VAGA support a geodetic extension of 1.2
transtensional motion for both faults (n. 11, 12, Figs. 8, 9a). The geodetic 0.2 mm/yr on the Acquae Iuliae Fault (n. 17, Figs. 8, 9b), which is viewed
strikeslip rate for the Lakes Fault is consistent with the reported as the source of large historical and Holocene earthquakes, and was al-
Holocene rate, but no geological information available for the Cecita ready identied as the location of a signicant geodetic velocity gradient
Fault. (Giuliani et al., 2009). Our geodetic extension estimate is by far lesser
In the central part of Calabria, a main EW morphological boundary than the value (2.53 mm/yr) shown by Giuliani et al. (2009), but it is
corresponds to the trace of the LameziaCatanzaro Fault (n. 13; Fig. 8), broadly consistent with the upper bound (0.8 mm/yr) Late Holocene
which is thought to limit northward the active subduction zone beneath range proposed on the base of trench data by Galli and Naso (2009).
the Calabrian Arc (Faccenna et al., 2011). Disagreement exists on the re- Within the Matese massif, geodetic data indicate divergence across a
cent activity of the fault, which is alternatively considered to have a left- NWSE striking and SW-dipping fault system (Matese Lake Fault, n. 18)
lateral (Tansi et al., 2007) or a normal (Faccenna et al., 2011) kinematics. which cuts along the central part of the range. The fault is thought to
GNSS sites show only a very small differential motion across the fault have Holocene activity (Bousquet, 1973; Galadini et al., 2000), but it is
(Fig. 9a), however. not well parameterized nor included in the DISS database. Geodetic ve-
locities of sites straddled by the fault, however, are consistent with a
7.2. Shallow-crustal creeping faults of the eastern domain normal-left oblique horizontal motion at ~0.7 mm/yr. This estimate is
moderately larger than the geologic extension computed from parame-
Within the eastern domain, few faults characterized by high-rate ters drawn from a published cross-section (Calabr et al., 2003), which
geodetic extension are locally encountered within the transcurrent yield component rate of ~ 0.2 to ~ 0.4 mm/yr, assuming an Early
fault arrays. These faults have also an inferred or proved geologic exten- (1.8 Myr) to Middle (0.8 Myr) Pleistocene inception of faulting, respec-
sion, but they are not typically associated to extensional earthquakes or tively (Fig. 9b).
to large seismic events. On the contrary, agreement exists on the seismogenic role of the
In northern Apulia, on the northeastern tip of Gargano, a well- NE-dipping Boiano normal fault, located at the northern border of
constrained diverging velocity gradient documents horizontal exten- the Matese massif (n. 19, Fig. 8), which is considered responsible
sion of up to 2.1 0.3 mm/yr across the NWSE striking Varano Fault for the Me = 6.6, 1805 earthquake (Fig. 1; DISS Working Group,
which dissects the Gargano promontory (n. 14; Fig. 8). The contradicto- 2010). Geodetic analysis indicates signicant extension (1.7
ry pattern of slip when compared to the extent of strikeslip faulting in 0.5 mm/yr) and a component of right-lateral slip (1.5 0.9) for
the region is reected in the sharp variation in area change pattern at the fault (Fig. 9b). Our extension slip estimation is moderately larger
Gargano from negative in the west to positive in the east (Fig. 5a), and than the inferred Pleistocene and Holocene (Galli and Galadini,
clearly calls for different processes at hand. Although geologic slip 2003) rate and broadly consistent with the estimate of Giuliani
rates are not available for structure n. 14, the high geodetic extension et al. (2009).
is not consistent with the limited length of the fault to be a crustal Between the Matese massif and the Tyrrhenian coastline of northern
seismogenic structure. According to Bertotti et al. (1999), the Gargano Campania region, the horizontal GNSS velocities indicate an additional
promontory was structured by Neogene thrusts and transpressional albeit minor right-lateral slip and extension (Figs. 8; 10). However,
faults rooted above a shallow-crustal NNE-dipping detachment. Based the stations distribution is not adequate to assess where the gradient
on the above arguments, we regard the Varano Fault as an ancillary is precisely occurring, and thus to constrain the location of the active
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 115

Fig. 8. Modeled active faults and seismogenic sources (numbers according to Tables 2, 3), and labeled GNSS sites and related velocity in the Apulian reference frame. Localities as in Fig. 2.

fault(s). These are commonly thought to lie at the inner border of the fault parameters proposed in the DISS database make this fault more
northern part of the Campania plain (Volturno Plain Fault, n. 20, akin to the eastern domain (Table 1), our results are more consistent
Galadini et al., 2000) where they form a NWSE trending, ~ 80 km with it being part of the western, extensional domain.
long system of right-hand en-echelon strands of limited length Within the Picentini Mountains, albeit constrained by few GNSS
(Fig. 8). If the Volturno Plain Fault is the active faults, then the strain ac- sites, geodetic velocities suggest motion across some faults which
cumulation on it is consistent with the Pleistocene and Holocene rates. however lack evidence of Holocene activity. Both the Lauro (n. 22)
In the Sannio area, geodetic extension at 2.1 0.8 mm/yr is ob- and the Volturara (n. 23) Faults (Fig. 8) have a geodetic extension
served across the Ariano Irpino Fault (n. 21; Figs. 8, 10). This source of ~1.6 mm/yr, the latter associated to a signicant (2.2 0.2 mm/yr)
lies at the boundary between the western and eastern domain, and is right-lateral slip. For both faults, the geodetic estimate is larger than the
thought to be a blind normal-oblique structure (Table 2) associated to geological slip rate (Fig. 9b).
destructive historical earthquakes (Fig. 1). Note that, because velocities East of the Picentini Mountains, the Irpinia Fault is a well character-
of sites SACR and PSB1 are similar and are both different from SVIT, ized structure because it slipped during the Mw = 6.9, 1980 earthquake
MOCO and CICL, we slightly turned counter-clockwise the reported (Pantosti et al., 1993). Co-seismic rupture observations showed that the
strike of the fault to let SACR and PSB1 in the western or footwall fault is segmented, and paleo-seismological investigations have sug-
block and the remaining sites in the hanging-wall block. Although the gested a 0.40.6 mm/yr extension rate (DISS Working Group, 2010).
116 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

a) b)

Fig. 9. Horizontal fault-parallel and fault-normal components of the geodetic and geologic slip rate for individual faults in the: a) eastern domain; and b) western domain of southern Italy.
Geologic and geodetic rates drawn from Tables 2 and 3, respectively. The geologic rate for source n. 28 (thicker dashed line) includes the estimate for the MelandroPergola (DISS Working
Group, 2010, Table 2) Fault, and an additional ~0.4 mm/yr including the summed contribution from the Vallo di Diano (VD) and Alburni (A) Faults (Papanikolaou and Roberts, 2007),
assuming a ~60 dip.

GNSS data indicate 1.5 1.1 and 1.0 0.3 mm/yr extension for the PTGA sites CARM and MTSN straddling the southern strand of the
Colliano (n. 24) and S. Gregorio (n. 25) segments of the fault, respective- Vallo di Diano Fault (n. 29, Fig. 8), which limit the eastern side of the
ly, which are larger than the trench data (Tables 2, 3, Fig. 10). Probably, Vallo di Diano basin, have a velocity residual indicative of 1.1
the estimate for the S. Gregorio segment include contributions from 0.3 mm/yr extension. This result appears consistent with structural
other faults southwest of the Irpinia Fault and north of sites DONG and paleoseismological investigations (Spina et al., 2008; Villani and
and CDRU. Pierdominici, 2010) although at a slightly higher geodetic value
The Cilento promontory, which faces the Tyrrhenian Sea in southern (Fig. 9b). Unfortunately, we have no sites straddling the central part of
Campania, appears to behave as a geodetically and geologically coher- the Vallo di Diano Fault to assess its geodetic activity.
ent block. The boundary sector between the Cilento promontory and Regarding the Val d'Agri basin, uncertainty exists on the geodetic
the Apennines is marked by Quaternary extensional basins (Vallo di horizontal slip rate for the NE-dipping Monti della Maddalena
Diano and Val d'Agri, Fig. 2). Different views exist about the location Fault (n. 28; Fig. 8), where eld studies and detailed seismological ob-
of the master fault controlling the destructive seismogenic activity in servations (Improta et al., 2010; Maschio et al., 2005; Valoroso et al.,
the area (Cello et al., 2003; DISS Working Group, 2010; Maschio et al., 2009) record active extension controlling the recent evolution of the
2005), specically whether it is NE- or SW-dipping structure. basin. Geodetic observations broadly document an extension (1.7
In the northern part of the boundary sector between Cilento and the 1.4 mm/yr) and a left-lateral slip component (1.2 0.6 mm/yr) across
Apennines, substantial divergent motion is recorded in a sector span- the M. Maddalena Fault, at the upper bound of the Pleistocene and
ning sites CDRU and DONG in the west and TITO in the east (Fig. 8). Ex- Holocene estimate (Fig. 9b).
tension is here modeled as largely accommodated by the NE-dipping To the eastern border of the Val d'Agri basin, the SW-dipping East
MelandroPergola Fault (n. 26), although motion could be also occur- Agri Fault (n. 27) has been proposed to be an active normal to left-
ring across the northern strand of the 45-km long Vallo di Diano Fault, oblique fault (Benedetti et al., 1998; Cello et al., 2003). Our analysis is
and on the Alburni Fault (Fig. 2). The geodetic extension (2.8 not consistent with this view, but rather indicates right-lateral slip at
0.2 mm/yr) is larger than the geological estimate (Fig. 9b), even when 1.6 0.5 mm/yr and a shortening component for the fault (Fig. 9b).
the additional contribution of the Vallo di Diano and Alburni Faults The southward continuation of the active extensional belt in south-
(Papanikolaou and Roberts, 2007) are included, and suggests current ern Lucania is not well constrained either by focal mechanisms or by ac-
slip on all the faults plus additional internal strain in the surrounding tive fault studies. Diverging GNSS velocities between sites of the axial
crust. zone (TIMP) and of the Tyrrhenian margin (CUCC) suggests current
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 117

extension (1.5 0.3 mm/yr) accommodated across inferred active nor- 8. Discussion
mal faults of the Mercure basin (n. 30; Fig. 8). The basin has been struck
by the 1998 Mw 5.6 Castelluccio Earthquake and in the last few years it 8.1. Lateral variability of geodetic deformation in southern Italy
has been experiencing an intense, low-energy seismic sequence (Arrigo
et al., 2006; Totaro et al., 2013). The geodetic estimate is substantially It is a common believe that the current tectonic regime in southern
larger than the Holocene estimates (Fig. 9b), but the signicant distance Italy is dominated by extension along the axis of the Apennines
between the GNSS sites (Fig. 8) could hinder the existence of additional (e.g. Montone et al., 2012), with localized strikeslip faulting in
sources and/or diffuse strain. northern Apulia. Our analysis, supported by a detailed inspection of
To the southeast of the Mercure Basin, along the south-western bor- geological and seismological data, is rather consistent with a sharp
der of the Pollino range, a number of PTGA sites constrain motion on the and continuous spatial segregation between the western and eastern
WNW-striking Pollino Fault (n. 31, Fig. 8), which is regarded as the main parts of the region, which are characterized by extension and strike
Neogene transcurrent shear zone of northern Calabria (Ghisetti and slip deformation, respectively (Figs. 2 to 4). In the following, we
Vezzani, 1982; Van Dijk et al., 2000). Investigators proposed a main discuss the broad pattern of geodetic horizontal slip rates for fault
left-lateral kinematics (Bousquet and Gueremy, 1969; Catalano et al., systems within both domains, and specically how do they vary
1993; Ghisetti and Vezzani, 1982), and some suggested it was across and along the orogen. For sake of simplicity, we refer to the
reactivated as normal since the Middle Pleistocene (Schiattarella, average geodetic estimate and neglect uncertainties.
1988). Based on paleoseismological investigations, it was argued that Within the western sector, we nd that the extension rate on indi-
the central part of the fault is active in extension together with the vidual normal faults is broadly within the ~ 0.81.7 mm/yr range
NNWSSE striking Castrovillari Fault to the south, but the whole fault (Fig. 9b). Three faults, located at the tip (n. 31, 32) or off (n. 20) the
system is currently characterized by an important seismic gap (Cinti main axial fault array, have lower extension rates of ~ 0.10.3 mm/yr.
et al., 1997, 2002; Michetti et al., 1997). Although uncertainty is signi- Only two sources (n. 21 and 26) have larger rates, but the geodetic esti-
cant, our geodetic observations are not supportive of the left-lateral slip, mate probably includes contributions from more than a single fault and/
rather they possibly indicate minor right-lateral slip and extension or signicant internal block strain. A mostly right-lateral strikeslip
(Fig. 9b), in agreement with the reactivation model. Geodetic data, on component is found for faults of the western domain, but it is limited
the other hand, are consistent with the geological slip rate estimate at ~0.40.6 mm/yr, and, given uncertainty, is of low statistical signi-
for the fault (Fig. 9b). cance. Only few faults have a signicant (up to ~1.5 mm/yr) component
South of the Pollino Fault, the Crati Valley is bordered on both sides of strikeslip motion (Fig. 9b).
by normal faults (Monaco and Tortorici, 2000; Tortorici et al., 1995), co- Although we acknowledge that extension in the western sector of
incident with strong historical and moderate instrumental earthquakes southern Italy is focused within the high mountains, we rene this
(Figs. 2, 3). GNSS station distribution allows to estimate extension on view and surmise that it is not concentrated on a single, through-
the eastern border fault (n. 32, Fig. 8), although the activity of the north- going fault. Rather, it appears that at least two parallel fault swarms
ern part of the western fault system, suggested by geologic studies along the axis of the Apennines (e.g. Matese, Irpinia, Val d'Agri sectors,
(Monaco and Tortorici, 2000), cannot be neglected. A small right- Fig. 8) are needed to accommodate most (~2/3 to ~9/10) of the regional
oblique extension displacement is detected (n. 30), in agreement with geodetic strain gradient. Other faults to the west, located at the bor-
fault-kinematic studies (Faccenna et al., 2011) and Pleistocene rate der of the coastal plains (e.g. n. 20, 22), accommodate the remaining
estimate. strain.

a) b)

Fig. 10. Geodetic and geologic moment rates computed for individual faults in the: a) western and, b) eastern domain of southern Italy. Geologic and geodetic rates drawn from Tables 2
and 3, respectively.
118 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

More signicantly, the magnitude of geodetic extension and strain eld study (e.g. Ferranti et al., 2009, for source n. 16), of their gravitational
appears to vary systematically moving along the axis of the chain paral- nature.
lel to the main faults arrays (Figs. 8, 10). Within the CampaniaMolise
sector in the north, the aggregated extension accommodated by faults 8.2. Relation between geodetic and geologic deformation
at the Matese massif (faults n. 17 to 19) is ~3.6 mm/yr, and an addition-
al 0.3 mm/yr extension may occur west of this location (n. 20). Moving Geodetic horizontal slip rates and moment rates computed in this
to the southeast, extension decreases to ~ 2 mm/yr in Sannio (n. 21), work are generally larger than the reported geological rates, and only
and then to ~ 2.3 mm/yr in the Picentini Mountains (n. 2223 at in some cases are at the upper bound of their uncertainty range, partic-
~1.5 mm/yr, with an additional ~0.8 mm/yr extension accommodated ularly for extensional faults of the western domain (Figs. 9 to 10). As al-
by oblique motion on fault n. 6 to the east) and ~11.5 mm/yr in Irpinia ready pointed out elsewhere (e.g. Friedrich et al., 2003), there may be
(n. 2425). Thus, a 1.52 mm/yr decrease in extension rate from various reasons for the observed discrepancy between geodetic and
~ 3.5 mm/yr to ~ 1.52 mm/yr characterizes the transition from geological observations. On the one hand, there could be missing faults,
northern (Matese) to central (SannioPicentiniIrpinia) Campania or signicant internal strain between the GNSS stations and the
(Fig. 9b). modeled fault trace. This non-recoverable deformation may result
A second sector of relatively greater extension (~ 2.8 mm/yr) is from slip on multiple small faults within the blocks that are too small
found in the Vallo di DianoVal d'Agri area at the CampaniaLucania and have too low slip rates (1 mm/yr) to be detected by GNSS obser-
border (aggregated contribution of n. 2829, and n. 26 with additional vations. Based on the amount of research carried so far, it is unlikely that
strain admittedly taken by faults west of it). South of here, in southern important active faults have gone unnoticed in the Apennines. For ex-
Lucania, the magnitude of extension decreases progressively, but ample, the Volturara Fault (n. 23) has a higher geodetic than geologic
rapidly to ~ 1.5 mm/yr (n. 30) and down to near null in northern rate with no additional large fault intervening between the GNSS sites,
Calabria (n. 3132). In southern Lucania and northern Calabria, it and this is possibly indicative of diffuse internal strain.
appears that only a single fault swarm is capable of accommodating Another reason for the discrepancy could be the known possibility of
the extension (Fig. 8). The variation in extension magnitude along time uctuations in the geodetic record, which in several cases turns to
the axis of the chain, with two patches of greatest values at the be unrepresentative of the longer-term deformation (e.g. Friedrich et al.,
MoliseCampania and CampaniaLucania borders, is reected in 2003). These uctuations may record mechanical transients during the
the pattern of area and shear strain change (Fig. 5a, b). seismic cycle (interseismic recoverable strain or post-seismic rebound
The observed trend in geodetic data is consistent with the long-term effects) which are smoothed out in the permanent strain recorded by
slip rate pattern (Fig. 9b), and appears a persistent feature of crustal the geological offsets. Perhaps, post-seismic effects associated to the M
faults in the Southern Apennines (Papanikolaou and Roberts, 2007). 6.9, 1980 Irpinia earthquake, which may derive by processes occurring
The latter authors studied the active faults between Irpinia and northern at the ductile layer depth (e.g. Della Via et al., 2003), conspire with
Calabria (roughly between our sources n. 23 and 31, encompassing the other uncertainty sources to augment the discrepancy between the geo-
southern sector of greater extension) and found individual and aggre- detic and geological record in the Irpinia sector of central Campania
gated slip rates which are greater at the center and diminish toward (faults n. 24, 25).
the tips of the individual faults and regional fault array. We found a sim- Finally, it could be argued that part of the discrepancy may be caused
ilar pattern in geodetic extension not only for the southern fault array, by sliding processes at some of the GNSS sites which led to an overesti-
but also for the northern one from Matese to Irpinia (Fig. 9b). Although mation of extension magnitude. However, because almost all measure-
when compared to geologic rates, the geodetic rates are one to twofold ment sites are installed in hard bedrock (see Section 4.1), we exclude
higher (Fig. 9b), our ndings lend support to the conclusion that region- that gravitational sliding is present in our estimates.
al extension in the Apennines occurs in discrete patches, and crustal Clearly, our geodetic estimates have a yet large range, because a
faults interact to maintain the established geometry and slip rate large number of GNSS velocity estimation are characterized by uncer-
pattern of the array (Papanikolaou and Roberts, 2007). The possible rea- tainties of ~ 0.3 mm/yr, which is in the same range of the geological
sons for the magnitude discrepancy between geologic and geodetic slip-rate of most of the faults studied here. Because uncertainties are
rates are discussed in the next section. inversely proportional to the observation time span, the current
When compared to normal faults of the western domain, faults of or planned increment of the observation time interval at these
the eastern sector have lower strikeslip rates distributed within two sites will lead to a reduction of uncertainty values at 0.1 mm/yr level
average clusters of ~0.10.3 mm/yr and ~11.2 mm/yr (Fig. 9a). Unlike (e.g. Herring, 2003).
the western domain, the lower slip rates and the higher uncertainty for Notwithstanding geodetic uncertainty, the geodetic and geological
faults of the eastern domain hinders the detection of possible longitudi- estimate still will not match. We are aware that a sound interpretation
nal gradients in magnitude. The associated fault-normal component of of the relation between the geodetic and geologic deformation relies
the horizontal slip rate appears larger in northern Calabria, and smaller on a knowledge of the seismic versus aseismic behavior of the fault,
in the Apulian sector, where it mostly characterizes faults closer to the which may be provided by a quantitative appraisal of the historical seis-
western domain. The fault-normal component typically changes from micity associated to the fault, far beyond the scope of the paper. We can
extensional close to the western boundary (n. 2, 3 and 6, but also n. only make a qualitative estimation of the seismic behavior of faults by
21 at the border with the western domain), to null or slightly contrac- inspection of their spatial relation with the historical seismicity release
tional in the Apulia foreland (n. 1, 2, 8). A similar pattern was already pattern (Fig. 1).
noted in seismological data (Boncio et al., 2007) for the Apulian faults, Within the axial part of the chain, where destructive earthquakes are
and supports the correlation between surface geodetic observations reasonably attributed to extensional faults (Figs. 1, 2), and seismic ef-
and deep seismogenic processes. ciency on the 600 yr time-window is up to 100% (Palano et al., 2011), a
Finally, extension at ~12 mm/yr is extracted for some faults within discrepancy between geodetic and geologic rates has different implica-
the eastern domain (n. 14 to 16, Fig. 8) which are: a) located near the tions depending on the timing of the last earthquake rupture.
coast; b) of limited length; c) surrounded by seismologically active Geodetically active faults at the CampaniaMolise boundary
strikeslip and reverse faults. These faults do not form a through- (Acquae Iuliae, Matese Lake, Boiano n. 17 to 19), have geodetic rates
going belt, but rather they are isolated segments. The observed only slightly larger than, or comparable to geologic rates. Because the
extension magnitude is clearly off-scale when compared to their Boiano Fault lastly ruptured in the Me = 6.6, 1805 earthquake (DISS
length according to well-established empirical relations (Wells and Working Group, 2010), this result would suggest transient effects or dif-
Coppersmith, 1994), and rmly sustain the interpretation, supported by fuse strain. On the opposite, the Acquae Iuliae Fault, which is viewed as
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 119

the source of the Me = 6.3, 1349 event (Galli and Naso, 2009), may be of The direction of geodetic extension (Fig. 5a) is ~ NESW along the
some concern given the substantial time elapsed since. The Matese Lake Southern Apennines, but turns to ~EW in Calabria following the oro-
Fault, for which large historical events are not reported, might be ef- graphic bend, the rotation of seismological axes (Fig. 3) and the curva-
ciently releasing by means of creep or low seismicity (note however, ture of the deep Moho boundary (Fig. 2). Although numerical models
that the location on a high mountain terrain renders the historical seis- are needed to quantitatively appraise the relation between processes
micity record for the fault more uncertain). While this paper was in the acting at different depths, these observations are taken to imply that ex-
review process, a seismic sequence (Mmax = 4.95) occurred in the area, tension in the high Apennines is a direct result of the impingement of
whose epicentral distribution (http://iside.rm.ingv.it/) coincides with the Tyrrhenian Moho and delamination of the Adriatic crust, and is
the Matese Lake Fault. This occurrence allows to argue that if the fault thus controlled by an excess potential energy. Geodetic observations
is (at least partially) locked, the signicant geodetic strain accumulation are consistent with seismological data in supporting the notion that sur-
on the fault may be of concern, in light of its recent seismic activation. face and deep crustal processes are intimately twinned along the axis of
Faults in central and southern Campania, on the contrary, have geo- the Apennines.
detic moments by far larger than geological moment rates (n. 21 to 25, What is remarkable is the clear spatial segregation, within the axial
Fig. 10b). For some faults, post-seismic effects (Irpinia Fault segments) part of the western domain, of two discrete fault arrays with the highest
or diffuse strain may be postulated. Because no historical earthquakes slip and strain rates localized at the CampaniaMolise and Campania
are associated to the Lauro (n. 22) and Volturara (n. 23) Faults, signi- Lucania boundary, respectively (Figs. 5, 10). As already noted, the geo-
cant seismic loading can be excluded, and the discrepancy may result detic pattern is persistent in the geologic record (Papanikolaou and
from transient creep. On the other hand, the high geodetic rates for Roberts, 2007). The two fault arrays and shear strain regions in the
the Lucania faults may be of concern considering the location in a de- Apennines are mirrored to the east by the physiographic difference
structive seismic area (Maschio et al., 2005; Villani and Pierdominici, between Gargano and Murge, with an intervening lowland. We
2010). speculate that a deep segmentation between the two foreland blocks
On the opposite, the geodetic and geologic rates are broadly compa- (e.g. Doglioni et al., 1994; Ferranti and Oldow, 2005) has caused a
rable in Calabria (Fig. 10b). This result may be insightful for the Pollino spatially shifted encroachment of the Tyrrhenian Moho beneath the
Fault (n. 31), which is thought to be coinciding with a seismic gap based Apennines, which in turn has controlled the position and evolution
on trench investigations (Cinti et al., 1997, 2002; Michetti et al., 1997). of the two sectors of localized extension in the upper crust of the
Although the geodetic and geologic rates are both low, their similarity Apennines.
might be of concern. On the other hand, different causes must be invoked for the geodetic
Conversely, in regions of low to null seismicity, if the geodetic and activity of strikeslip faults in the eastern domain. The concealed and
geologic moment rates are equal, creep or very long recurrence intervals diffuse pattern of low slip rate faults are consistent with a model of re-
for seismic events may be postulated. This is the case of the extensional activation of Mesozoic or older extensional faults affecting the continen-
array located at the border of the coastal plains (n. 20, 22), which show a tal interior of the Adriatic micro-plate (Boncio et al., 2007; Di Bucci et al.,
minor albeit appreciable geodetic signal. We do not know whether this 2006). Although seismological data mostly indicate active displacement
motion is accommodated aseismically (e.g. Ferranti et al., 2008), as sug- only in Gargano and northern Calabria (Figs. 1, 3), geodetic observations
gested by the lack of signicant seismic events (Figs. 1, 3), or faults of indicate a more diffuse pattern of strain accumulation and release.
the Tyrrhenian margin have recurrence times longer than the complete- Overall, the transcurrent and locally oblique motion currently accom-
ness interval of the historical catalogs. Certainly, these faults have ac- modated on these faults may be seen as a far-eld response to the colli-
commodated the most severe displacement of the Apennines upper sion between the southern part of Adria, and the Dinarids and Albanian
crust down to the Tyrrhenian Sea during the last 23 Myr, and thus a sector of Europe (Hollenstein et al., 2003; Jouanne et al., 2012; Oldow
long-recurrence activation cannot be neglected. and Ferranti, 2005; Oldow et al., 2002).
Similarly, the eastern domain, which is characterized by diffuse
strikeslip deformation and low to moderate seismicity, may be charac- 9. Concluding remarks
terized by faults with efcient seismic release or by long-recurrence
earthquakes. For instance, signicant geodetic motion, comparable Despite the limited time-span covered by the GNSS data (about 85%
with geologic rates, is found for sources in northern Apulia and in the of the permanent sites have an observation interval shorter than 5 yr,
Apennines foothills (n. 2 to 6, Fig. 10a). On the one hand, it is possible and episodic sites have observation interval of ~15 yr), we observe for
that signicant creep is occurring on these faults. However, because it several faults a good agreement between the pattern of geodetic and
was proposed that large past earthquakes may have activated some geological deformation, albeit the geodetic rates are largely or at the
fault segments (e.g. Fracassi and Valensise, 2007), our observation raises upper bound of geological rates.
the possibility of future ruptures. Notwithstanding the locally signicant uncertainty and the likely
existence of transient uctuations in geodetic slip rate estimation, our
results supply vital information for rening the understanding of active
8.3. Different engines for the active deformation domains faults in the region. For most of these faults, the geological slip rate is to
date poorly constrained (Table 2). The results presented here are partic-
Appraisal of independent geodetic, geological and seismological ularly encouraging for the study of the eastern domain strikeslip faults,
datasets shows a remarkable consistency within both the western and where the geological slip rate is difcult to assess due to their deep blind
eastern tectonic domains. The boundary between the two domains is nature and the low deformation rates.
sharp and runs along the high crest of the Apennines, where active nor- Of importance, we detect signicant right-lateral geodetic displace-
mal faults accommodate large part of the geodetic extension (Figs. 2, 3, ment in areas of the eastern sector where no active structures are re-
4). Noteworthy, this boundary spatially coincides with the deep contact ported in available datasets (e.g. n. 1), or where the inferred sources
between the young and uplifted Tyrrhenian Moho to the west, and the show a weak evidence of activity (e.g. n. 5). Although we are not able
subjacent, older Adriatic Moho to the east (Fig. 2, inset). The eastward at this stage to discern whether it could represent a seismic hazard
encroachment of the Tyrrhenian Moho is thought to be a result of the threat or not, this outcome is relevant for near future active tectonic
progressive roll-back and delamination of the Adriatic slab and accom- analysis.
panying migration of the locus of stretching and asthenospheric upwell- The geodetically-based renement of slip rate for active faults has
ing in the Tyrrhenian back-arc basin (Cassinis et al., 2003; Di Stefano important implications for both fault-based and time-dependent seis-
et al., 2011; Malinverno and Ryan, 1986). mic hazard assessment models. At present, the national seismic hazard
120 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

map (Meletti et al., 2008) does not include individual seismogenic D'Agostino, N., Avallone, A., Cheloni, D., D'Anastasio, E., Mantenuto, S., Selvaggi, G., 2008.
Active tectonics of the Adriatic region from GPS and earthquake slip vectors.
sources, mostly for the difculty of characterizing their seismicity be- J. Geophys. Res. 113, B12413. http://dx.doi.org/10.1029/2008JB005860.
cause of the large uncertainty on the deformation rate (see Table 2). D'Agostino, N., D'Anastasio, E., Gervasi, A., Guerra, I., Nedimovi, M.R., Seeber, L.,
Thus, incorporation of geodetic slip rates may effectively contribute to- Steckler, M., 2011. Forearc extension and slow rollback of the Calabrian Arc
from GPS measurements. Geophys. Res. Lett. 38, L17304. http://dx.doi.org/10.
ward a fault-based seismic hazard assessment (Peruzza et al., 2010, 1029/2011GL048270.
2011; Roberts et al., 2004), even in this region of low and diffuse Del Gaudio, V., Pierri, P., Calcagnile, G., Venisti, N., 2005. Characteristics of the low energy
deformation. seismicity of central Apulia (southern Italy) and hazard implications. J. Seismol. 9,
3959. http://dx.doi.org/10.1007/s10950-005-1593-9.
Supplementary data to this article can be found online at http://dx. Del Gaudio, V., Pierri, P., Frepoli, V., Calcagnile, G., Venisti, N., Cimini, G.B., 2007. A critical
doi.org/10.1016/j.tecto.2014.02.007. revision of the seismicity of the Northern Apulia (Adriatic microplate? Southern Italy)
and implications for the identication of seismogenic structures. Tectonophysics 436,
935. http://dx.doi.org/10.1016/j.tecto.2007.02.013.
Acknowledgments Della Via, G., De Natale, G., Troise, C., Pingue, F., Obrizzo, F., Riva, R., Sabadini, R., 2003. First
evidence of post-seismic deformation in the central Mediterranean: crustal viscoelas-
tic relaxation in the area of the 1980 Irpinia earthquake (southern Italy). Geophys.
This research was conducted under the auspices of a collaborative J. Int. 154, F9F14.
research initiative funded by various sources in Italy and in the United Devoti, R., Esposito, A., Pietrantonio, G., Pisani, A.R., Riguzzi, F., 2011. Evidence of large
scale deformation patterns from GPS data in the Italian subduction boundary. Earth
States. The GNSS campaigns could not have been carried out without
Planet. Sci. Lett. 311, 230241. http://dx.doi.org/10.1016/j.epsl.2011.09.034.
the diligent work of many undergraduate and graduate students from Di Bucci, D., Ravaglia, A., Seno, S., Toscani, G., Fracassi, U., Valensise, G., 2006.
Italy and the United States. The careful reviews of B. Pace and I. Seismotectonics of the southern Apennines and Adriatic foreland: Insights on active
Papanikolaou were instrumental in clarifying several aspects of the regional EW shear zones from analogue modeling. Tectonics 25, TC4015. http://
dx.doi.org/10.1029/2005TC001898.
work. Di Bucci, D., Ridente, D., Fracassi, U., Trincardi, F., Valensise, G., 2009. Marine
palaeoseismology from very high resolution seismic imaging: the Gondola Fault
Zone (Adriatic foreland). Terra Nova 21, 393400. http://dx.doi.org/10.1111/j.1365-
References 3121.2009.00895.x.
Di Luccio, F., Fukuyama, E., Pino, N.A., 2005. The 2002 Molise earthquake sequence: what
Altamimi, Z., Collilieux, X., Legrand, J., Garayt, B., Boucher, C., 2007. ITRF2005: a new re- can we learn about the tectonics of southern Italy. Tectonophysics 405 (14),
lease of the International Terrestrial Reference Frame based on time series of station 141154. http://dx.doi.org/10.1016/j.tecto.2005.05.024.
positions and earth orientation parameters. J. Geophys. Res. 112, B09401. http:// Di Stefano, R., Bianchi, I., Ciaccio, M.G., Carrara, G., Kissling, E., 2011. Three dimensional
dx.doi.org/10.1029/2007JB004949. Moho topography in Italy: new constraints from receiver functions and controlled
Ambrosetti, P., Bosi, C., Carraro, F., Ciaran, N., Panizza, M., Papani, G., Vezzani, L., source seismology. Geochem. Geophys. Geosyst. 12, Q09006. http://dx.doi.org/
Zanferrari, A., 1987. Neotectonic map of Italy, in: Quaderni della Ricerca Scientica, 10.1029/2011GC003649.
vol. 4, scale 1:500,000, CNR, Rome. DISS Working Group, 2010. Database of Individual Seismogenic Sources (DISS), version
Arrigo, G., Roumelioti, Z., Benetatos, C., Kiratzi, A., Bottari, A., Neri, G., Termini, D., Gorini, 3.1.1: a compilation of potential sources for earthquakes larger than M 5.5 in Italy
A., Marcucci, S., 2006. A source study of the 9 September 1998 (Mw 5.6) Castelluccio and surrounding areas. Ist. Naz. Geos. Vulcanol. http://dx.doi.org/10.6092/INGV.IT-
Earthquake in Southern Italy using teleseismic and strong motion data. Nat. Hazards DISS3.1.1 (http://diss.rm.ingv.it/diss/).
37 (3), 245262. http://dx.doi.org/10.1007/s11069-005-4644-1. Doglioni, C., Mongelli, F., Pieri, P., 1994. The Puglia uplift (SE Italy): an anomaly in the fore-
Basili, R., Valensise, G., Vannoli, P., Burrato, P., Fracassi, U., Mariano, S., Tiberti, M.M., land of the Apenninic subduction due to buckling of a thick continental lithosphere.
Boschi, E., 2008. The Database of Individual Seismogenic Sources (DISS), version 3: Tectonics 13, 13091321.
summarizing 20 years of research on Italys earthquake geology. Tectonophysics Faccenna, C., Becker, T.W., Lucente, F.P., Jolivet, L., Rossetti, F., 2001. History of subduction
453, 2043. http://dx.doi.org/10.1016/j.tecto.2007.04.014. and back-arc extension in the central Mediterranean. Geophys. J. Int. 145, 809820.
Benedetti, L., Tapponier, P., King, G.C.P., Piccardi, L., 1998. Surface rupture of the 1857 http://dx.doi.org/10.1046/j.0956-540X.2001.01435.x.
southern Italian earthquake. Terra Nova 10 (4), 206210. Faccenna, C., Molin, P., Orecchio, B., Olivetti, V., Bellier, O., Funiciello, F., Minelli, L.,
Bertotti, G., Casolari, E., Picotti, V., 1999. The Gargano Promontory, a contractional belt in Piromallo, C., Billi, A., 2011. Topography of the Calabria subduction zone (southern
the Adriatic plate. Terra Nova 11, 168173. http://dx.doi.org/10.1046/j.1365- Italy): clues for the origin of Mt. Etna. Tectonics 30, TC1003. http://dx.doi.org/
3121.1999.00243.x. 10.1029/2010TC002694.
Boncio, P., Mancini, T., Lavecchia, G., Selvaggi, G., 2007. Seismotectonics of strikeslip Faure Walker, J.P., Roberts, G.P., Cowie, P.A., Papanikolaou, I., Michetti, A.M., Sammonds,
earthquakes within the deep crust of southern Italy: geometry, kinematics, stress P., Wilkinson, M., McCaffrey, M.J.W., Phillips, R.J., 2012. Relationship between topog-
eld and crustal rheology of the Potenza 19901991 seismic sequence (Mmax 5.7). raphy, rates of extension and mantle dynamics in the actively-extending Italian
Tectonophysics 445, 281300. http://dx.doi.org/10.1016/j.tecto.2007.08.016. Apennines. Earth Planet. Sci. Lett. 325326, 7684. http://dx.doi.org/10.1016/j.epsl.
Bousquet, J.C., 1973. La tectonique rcente de l'Apennin calabro-lucanien dans son cadre 2012.01.028.
gologique et gophysique. Geol. Romana 12, 1103. Ferranti, L., Oldow, J.S., 2005. Latest Miocene to Quaternary horizontal and vertical
Bousquet, J.C., Gueremy, P., 1969. Quelques phnomnes de notectonique displacement rates during simultaneous contraction and extension in the Southern
dans l'Apennin calabro-lucanien et leurs consquences morphologiques. II Apennines orogen, Italy. Terra Nova 17, 209214. http://dx.doi.org/10.1111/j.1365-
L'escarpement mridional du Pollino et son piedmont. Rev. Gogr. Phys. Gol. 3121.2005.00593.x.
Dyn. 11, 223236. Ferranti, L., Oldow, J.S., D'Argenio, B., Catalano, R., Lewis, D., Marsella, E., Avellone, G.,
Brune, J.N., 1968. Seismic moment, seismicity and rate of slip along major fault zones. Maschio, L., Pappone, G., Pepe, F., Sulli, A., 2008. Active deformation in Southern
J. Geophys. Res. 73 (2), 777784. http://dx.doi.org/10.1029/JB073i002p00777. Italy, Sicily and southern Sardinia from GPS velocities of the Peri-Tyrrhenian Geodetic
Calabr, R.A., Corrado, S., Di Bucci, D., Robustini, P., Tornaghi, M., 2003. Thin-skinned vs. Array (PTGA). Boll. Soc. Geol. Ital. (Ital. J. Geosci.) 127 (2), 299316.
thick-skinned tectonics in the Matese Massif, Central-Southern Apennines (Italy). Ferranti, L., Santoro, E., Mazzella, M.E., Monaco, C., Morelli, D., 2009. Active transpression
Tectonophysics 377, 269297. http://dx.doi.org/10.1016/j.tecto.2003.09.010. in the northern Calabria Apennines, Southern Italy. Tectonophysics 476, 226251.
Cassinis, R., Scarascia, S., Lozej, A., 2003. The deep crustal structure of Italy and surround- http://dx.doi.org/10.1016/j.tecto.2008.11.010.
ing areas from seismic refraction data. A new synthesis. Boll. Soc. Geol. Ital. 122, Ferranti, L., Burrato, P., Pepe, F., Santoro, E., Mazzella, M.E., Morelli, D., Passaro, S.,
365376. Vannucci, G., 2014. Geometry and modelling of an active submarine transpressional
Catalano, S., Monaco, C., Tortorici, L., Tansi, C., 1993. Pleistocene strikeslip tectonics in the belt from integrated marine geophysical data: the Amendolara Ridge, southern
Lucanian Apennine (Southern Italy). Tectonics 12, 656665. http://dx.doi.org/ Italy. Tectonics (under revision).
10.1029/92TC02251. Fracassi, U., Valensise, G., 2007. Unveiling the sources of the catastrophic 1456 multiple
Cello, G., Tondi, E., Micarelli, L., Mattioni, L., 2003. Active tectonics and earthquake sources earthquake: Hints to an unexplored tectonic mechanism in southern Italy. Bull.
in the epicentral area of the 1857 Basilicata earthquake (southern Italy). J. Geodyn. Seismol. Soc. Am. 97 (3), 725748. http://dx.doi.org/10.1785/0120050250.
36, 3750. http://dx.doi.org/10.1016/S0264-3707(03)00037-1. Friedrich, A.M., Wernicke, B.P., Niemi, N.A., Bennett, R.A., Davis, J.L., 2003. Comparison of
Channell, J.E.T., Mareschal, J.C., 1989. Delamination and asymmetric lithospheric thicken- geodetic and geologic data from the Wasatch region, Utah, and implications for the
ing in the development of the Tyrrhenian Rift. In: Coward, M.P., Dietrich, D., Park, R.G. spectral character of Earth deformation at periods of 10 to 10 million years.
(Eds.), Alpine Tectonics. Geol. Soc. London Spec. Publ., 45, pp. 285302. http:// J. Geophys. Res. 108 (B4), 2199. http://dx.doi.org/10.1029/2001JB000682.
dx.doi.org/10.1144/GSL.SP.1989.045.01.16. Galadini, F., Meletti, C., Rebez, A. (Eds.), 2000. Le ricerche del GNDT nel campo della
Chiarabba, C., Amato, A., 1996. Crustal velocity structure of the Apennines (Italy) from pericolosit sismica (19961999). CNR-Gruppo Nazionale per la Difesa dai Terremoti,
P-wave travel time tomography. Ann. Geophys. 39 (6), 11331148. Roma.
Cinti, F.R., Cucci, L., Pantosti, D., D'Addezio, G., Meghraoui, M., 1997. A major seismogenic Galli, P., Bosi, V., 2003. Catastrophic 1638 earthquakes in Calabria (southern Italy): new
fault in a silent area: the Castrovillari Fault (southern Apennines, Italy). Geophys. insights from paleoseismological investigation. J. Geophys. Res. 108 (B1). http://
J. Int. 130, 595605. http://dx.doi.org/10.1111/j.1365-246X.1997.tb01855.x. dx.doi.org/10.1029/2001JB001713.
Cinti, F.R., Moro, M., Pantosti, D., Cucci, L., D'Addezio, G., 2002. New constraints on the Galli, P., Galadini, F., 2003. Disruptive earthquakes revealed by faulted archaeological
seismic history of the Castrovillari fault in the Pollino gap (Calabria, southern Italy). relics in Samnium (Molise, southern Italy). Geophys. Res. Lett. 30, 1266. http://
J. Seismol. 6, 199217. http://dx.doi.org/10.1023/A:1015693127008. dx.doi.org/10.1029/2002GL016456.
L. Ferranti et al. / Tectonophysics 621 (2014) 101122 121

Galli, P.A.C., Naso, J.A., 2009. Unmasking the 1349 earthquake source (southern Italy): Nocquet, J.M., Calais, E., 2003. Crustal velocity eld of Western Europe from permanent
palaeoseismological and archaeoseismological indications from the Acquae Iuliae GPS array solutions, 19962001. Geophys. J. Int. 154, 7288. http://dx.doi.org/
fault. J. Struct. Geol. 31, 128149. http://dx.doi.org/10.1016/j.jsg.2008.09.007. 10.1046/j.1365-246X.2003.01935.x.
Galli, P., Scionti, V., 2006. Two unknown M N 6 historical earthquakes revealed by Oldow, J.S., Ferranti, L., 2005. Fragmentation of Adria and active decollement tectonics
palaeoseismological and archival researches in eastern Calabria (southern Italy). within the southern Peri-Tyrrhenian orogen, Italy. In: Pinter, N., Grenerczy, G.,
Seismotectonic implications. Terra Nova 18, 4449. http://dx.doi.org/10.1111/ Medak, D., Stein, S., Weber, J.C. (Eds.), The Adria Microplate: GPS Geodesy, Tectonics
j.1365-3121.2005.00658.x. and Hazards. Kluwer Academic Publishers, Dordrecht, pp. 255272.
Galli, P., Scionti, V., Spina, V., 2007. New paleoseismic data from the Lakes and Serre faults: Oldow, J.S., Ferranti, L., Lewis, D.S., Campbell, J.K., D'Argenio, B., Catalano, R., Pappone, G.,
seismotectonic implications for Calabria (Southern Italy). Boll. Soc. Geol. It. (Ital. Carmignani, L., Conti, P., Aiken, C.L.V., 2002. Active fragmentation of Adria, the North
J. Geosci.) 126 (2), 347364. African promontory, central Mediterranean orogen. Geology 30, 779782. http://
Ghisetti, F., Vezzani, L., 1982. Strutture tensionali e compressive indotte da meccanismi dx.doi.org/10.1130/0091-7613(2002)030b0779:AFOATNN2.0.CO;2.
profondi lungo la linea del Pollino (Appennino meridionale). Boll. Soc. Geol. Ital. Palano, M., Rossi, M., Cannav, F., Bruno, V., Aloisi, M., Pellegrino, D., Pulvirenti, M.,
101, 385440. Siligato, G., Mattia, M., 2010. Etn@ref, a geodetic reference frame for Mt. Etna GPS net-
Giuliani, R., Anzidei, M., Bonci, L., Calcaterra, S., D'Agostino, N., Mattone, M., Pietrantonio, works. Ann. Geophys. 53 (4), 4879. http://dx.doi.org/10.4401/ag-4879.
G., Riguzzi, F., Selvaggi, G., 2007. Co-seismic displacements associated to the Molise Palano, M., Cannav, F., Ferranti, L., Mattia, M., Mazzella, M.E., 2011. Strain and stress
(Southern Italy) earthquake sequence of OctoberNovember 2002 inferred from elds in the Southern Apennines (Italy) constrained by geodetic, seismological and
GPS measurements. Tectonophysics 432, 2135. http://dx.doi.org/10.1016/ borehole data. Geophys. J. Int. 187, 12701282. http://dx.doi.org/10.1111/j.1365-
j.tecto.2006.11.005. 246X.2011.05234.x.
Giuliani, R., D'Agostino, N., D'Anastasio, E., Mattone, M., Bonci, V., Calcaterra, S., Gambino, Pantosti, D., Schwartz, D.P., Valensise, G., 1993. Paleoseismology along the 1980 surface
P., Merli, K., 2009. Active crustal extension and strain accumulation from GPS data in rupture of the Irpinia Fault: implications for earthquake recurrence in the southern
the Molise region (central-southern Apennines, Italy). Boll. Geos. Teor. Appl. 50 (2), Apennines. Italy. J. Geophys. Res. 98 (B4), 65616577. http://dx.doi.org/10.1029/
145156. 92JB02277.
Herring, T.A., 2003. MATLAB Tools for viewing GPS velocities and time series. GPS Solu- Papanikolaou, I.D., Roberts, G.P., 2007. Geometry, kinematics and deformation rates
tions 7 (3), 194199. http://dx.doi.org/10.1007/s10291-003-0068-0. along the active normal fault system in the southern Apennines: implication
Herring, T.A., King, R.W., McClusky, S.C., 2010. Introduction to GAMIT/GLOBK, Release for fault growth. J. Struct. Geol. 29, 166188. http://dx.doi.org/10.1016/
10.4. MIT, Cambridge, MA (48 pp.). j.jsg.2006.07.009.
Hollenstein, Ch., Kahle, H.-G., Geiger, A., Jenny, S., Goes, S., Giardini, D., 2003. New GPS Patacca, E., Scandone, P., 2007. Geological interpretation of the CROP-04 seismic line
constraints on the AfricaEurasia plate boundary zone in southern Italy. Geophys. (southern Apennines, Italy). Boll. Soc. Geol. Ital. (Ital. J. Geosci.) 7, 297315.
Res. Lett. 30 (18), 1935. http://dx.doi.org/10.1029/2003GL017554. Patacca, E., Sartori, R., Scandone, P., 1990. Tyrrhenian basin and Apenninic arcs. Kinematic
Hunstad, I., Selvaggi, G., D'Agostino, N., England, P., Clarke, P., Pierozzi, M., 2003. Geodetic relations since late Tortonian times. Mem. Soc. Geol. Ital. 45, 425451.
strain in peninsular Italy between 1875 and 2001. Geophys. Res. Lett. 30 (4), 1181. Peruzza, L., Pace, B., Cavallini, F., 2010. Error propagation in time-dependent probability of
http://dx.doi.org/10.1029/2002GL016447. occurrence for characteristic earthquakes in Italy. J. Seismol. 14, 119141.
Hyppolite, J.C., Angelier, J., Roure, F., 1994. A major geodynamic change revealed by Qua- Peruzza, L., Pace, B., Visini, F., 2011. Fault-based earthquake rupture forecast in central
ternary stress patterns in the Southern Apennines (Italy). Tectonophysics 230, Italy: remarks after the L'Aquila Mw 6.3 event. Bull. Seismol. Soc. Am. 101,
199210. http://dx.doi.org/10.1016/0040-1951(94)90135-X. 404412. http://dx.doi.org/10.1785/0120090276.
Improta, L., Ferranti, L., De Martini, P.M., Piscitelli, S., Bruno, P.P., Burrato, P., Civico, R., Piccardi, L., 1998. Cinematica attuale, comportamentosismico e sismologia storica della
Giocoli, A., Iorio, M., D'Addezio, G., Maschio, L., 2010. Detecting young, slow- faglia di Monte Sant'Angelo (Gargano, Italia): La possibile rottura superciale del
slipping active faults by geologic and multidisciplinary high-resolution geophysical leggendario terremoto del 493 d.C. Geogr. Fis. Din. Quat. 21, 155166.
investigations: a case study from the Apennine seismic belt, Italy. J. Geophys. Res. Pino, N.A., Palombo, B., Ventura, G., Perniola, B., Ferrari, G., 2008. Waveform modeling of
115, B11307. http://dx.doi.org/10.1029/2010JB000871. historical seismograms of the 1930 Irpinia earthquake provides insight on blind
Jouanne, F., Mugnier, J.L., Koci, R., Bushati, S., Matev, K., Kuka, N., Shinko, I., Kociu, S., Duni, faulting in Southern Apennines (Italy). J. Geophys. Res. 113, B05303. http://
L., 2012. GPS constraints on current tectonics of Albania. Tectonophysics 554557, dx.doi.org/10.1029/2007JB005211.
5062. Pondrelli, S., Di Luccio, F., Fukuyama, E., Mazza, S., Olivieri, M., Pino, N.A., 2003. Fast deter-
Latorre, D., Amato, A., Chiarabba, C., 2010. High-resolution seismic imaging of the Mw 5.7, mination of moment tensors for the recent Molise (southern Italy) seismic sequence.
2002 Molise, southern Italy, earthquake area: evidence of deep fault reactivation. Tec- ORFEUS Electron. Newslett. 5 (1), 12.
tonics 29, TC4014. http://dx.doi.org/10.1029/2009TC002595. Pondrelli, S., Salimbeni, S., Ekstrm, G., Morelli, A., Gasperini, P., Vannucci, G., 2006. The
Malinverno, A., Ryan, W.B.F., 1986. Extension in the Tyrrhenian Sea and shortening in the Italian CMT dataset from 1977 to the present. Phys. Earth Planet. Inter. 159,
Apennines as result of arc migration driven by sinking of the lithosphere. Tectonics 5 286303. http://dx.doi.org/10.1016/j.pepi.2006.07.008.
(2), 227245. http://dx.doi.org/10.1029/TC005i002p00227. Pordo, S., Esposito, E., Vittori, E., Tranfaglia, G., Michetti, A.M., Blumetti, M., Ferreli, L.,
Marone, C., Scholz, C.H., 1988. The depth of seismic faulting and the upper transition from guerriei, L., Serva, L., 2002. Areal Distribution of Ground Effects Induced by Strong
stable to unstable slip regimes. Geophysical Research Letters 15 (621624). http:// Earthquakes in the Southern Apennines (Italy). Surveys in Geophysics 23 (6),
dx.doi.org/10.1029/88GL02148. 529562. http://dx.doi.org/10.1023/A:1021278811749.
Maschio, L., Ferranti, L., Burrato, P., 2005. Active extension in Val d'Agri area, Southern Ap- Ridente, D., Trincardi, F., 2006. Active foreland deformation evidenced by shallow folds
ennines, Italy: implications for the geometry of the seismogenic belt. Geophys. J. Int. and faults affecting late Quaternary shelf-slope deposits (Adriatic Sea, Italy). Basin
162, 591609. http://dx.doi.org/10.1111/j.1365-246X.2005.02597.x. Res. 18, 171188. http://dx.doi.org/10.1111/j.1365-2117.2006.00289.x.
Mazzella, E., 2011. Tettonica attiva nel settore peri-tirrenico meridionale tramite Roberts, G.P., Cowie, P., Papanikolaou, I., Michetti, A.M., 2004. Fault scaling relationships,
integrazione di dati geodetici e geologici. PhD thesis Universit degli Studi di Napoli deformation rates and seismic hazards: an example from the LazioAbruzzo
Federico II, Napoli, Italy. Apennines, central Italy. J. Struct. Geol. 26, 377398.
Meletti, C., Galadini, F., Valensise, G., Stucchi, M., Basili, R., Barba, S., Vannucci, G., Boschi, Rovida, R., Camassi, P., Gasperini, P., Stucchi, M. (Eds.), 2011. CPTI11, The 2011 Version of
E., 2008. A seismic source zone model for the seismic hazard assessment of the Italian the Parametric Catalogue of Italian Earthquakes. http://dx.doi.org/10.6092/INGV.IT-
territory. Tectonophysics 450, 85108. CPTI11,Milano,Bologna (http://emidius.mi.ingv.it/CPTI).
Metropolis, N., Ulam, S., 1949. The Monte Carlo method. J. Am. Stat. Assoc. 44 (247), Savage, J.C., 1983. A dislocation model of strain accumulation and release at a subduction
335341. http://dx.doi.org/10.2307/2280232. zone. J. Geophys. Res. 88, 49844996.
Michetti, A.M., Ferreli, L., Serva, L., Vittori, E., 1997. Geological evidence for strong histor- Savage, J.C., Simpson, R.W., 1997. Surface strain accumulation and the seismic moment
ical earthquakes in an aseismic region: the Pollino case (southern Italy). J. Geodyn. tensor. Bull. Seism. Soc. Am. 87, 13451353.
24 (14), 6786. http://dx.doi.org/10.1016/S0264-3707(97)00018-5. Schiattarella, M., 1988. Quaternary tectonics of the Pollino Ridge, CalabriaLucania
Milano, G., Petrazzuoli, S., Ventura, G., 2004. Effects of the hydrothermal circulation on the boundary, southern Italy. In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.),
strain eld of the Campanian Plain (southern Italy). Terra Nova 16, 205209. http:// Continental Transpressional and Transtensional Tectonics. Geological Society,
dx.doi.org/10.1111/j.1365-3121.2004.00556.x. London, Special Publications, vol. 135, pp. 341354.
Monaco, C., Tortorici, L., 2000. Active faulting in the Calabrian arc and eastern Sicily. Serpelloni, E., Anzidei, M., Baldi, P., Casula, G., Galvani, A., 2005. Crustal velocity and
J. Geodyn. 29, 407424. http://dx.doi.org/10.1016/S0264-3707(99)00052-6. strain-rate elds in Italy and surrounding regions: new results from the analysis of
Monaco, C., Tortorici, L., Paltrinieri, W., 1998. Structural evolution of the Lucanian Apen- permanent and non permanent GPS networks. Geophys. J. Int. 161, 861880.
nines, Southern Italy. J. Struct. Geol. 20 (5), 617638. http://dx.doi.org/10.1016/ http://dx.doi.org/10.1111/j.1365-246X.2005.02618.x.
S0191-8141(97)00105-3. Spina, V., Tondi, E., Galli, P., Mazzoli, M., Cello, G., 2008. Quaternary fault segmentation
Montone, P., Mariucci, M.T., Pierdominici, S., 2012. The Italian present-day stress and interaction in the epicentral area of the 1561 earthquake (Mw = 6.4), Vallo di
map. Geophys. J. Int. 189, 705716. http://dx.doi.org/10.1111/j.1365-246X. Diano, southern Apennines, Italy. Tectonophysics 453, 233245. http://dx.doi.org/
2012.05391.x. 10.1016/j.tecto.2007.06.012.
Morelli, C., 2000. The themes of crustal research in Italy and the role of DSS-WA seismics. Tansi, C., Muto, F., Critelli, S., Iovine, G., 2007. NeogeneQuaternary strikeslip tectonics in
Boll. Soc. Geol. Ital. 119, 141148. the central Calabrian Arc (southern Italy). J. Geodyn. 43, 393414. http://dx.doi.org/
Mostardini, F., Merlini, S., 1986. Appennino centro-meridionale. Sezioni Geologiche e 10.1016/j.jog.2006.10.006.
Proposta di Modello Strutturale. Mem. Soc. Geol. Ital. 35, 177202. Tiberti, M.M., Orlando, L., Di Bucci, D., Bernarbini, M., Parotto, M., 2005. Regional gravity
Nicolich, R., 1989. Crustal structure from seismic studies in the frame of the European anomaly map and crustal model of the CentralSouthern Apennines (Italy).
Geotraverse (southern segment) and CROP Projects. In: Boriani, A., Bonafede, M., J. Struct. Geol. 40, 7391. http://dx.doi.org/10.1016/j.jog.2005.07.014.
Piccardo, G.B., Vai, G.B. (Eds.), The Lithosphere in Italy, Advances in Earth Science Re- Tondi, E., Piccardi, L., Cacon, S., Kontny, B., Cello, G., 2005. Structural and time constraints
search, Italian National Committee for the International Lithosphere Program, Atti for dextral shear along the seismogenic Mattinata Fault (Gargano, southern Italy).
Convegni Lincei, vol. 80, pp. 4161 (Roma). J. Geodyn. 40, 134152. http://dx.doi.org/10.1016/j.jog.2005.07.003.
122 L. Ferranti et al. / Tectonophysics 621 (2014) 101122

Tortorici, L., Monaco, C., Tansi, C., Cocina, O., 1995. Recent and active tectonics in the model for the northern sector of the Calabrian Arc (southern Italy). Tectonophysics
Calabrian Arc (Southern Italy). Tectonophysics 243, 3749. http://dx.doi.org/ 324, 267320. http://dx.doi.org/10.1016/S0040-1951(00)00139-6.
10.1016/0040-1951(94)00190-K. Villani, F., Pierdominici, S., 2010. Late Quaternary tectonics of the Vallo di Diano basin
Totaro, C., Presti, D., Billi, A., Gervasi, A., Orecchio, B., Guerra, I., Neri, G., 2013. The ongoing (southern Apennines, Italy). Quat. Sci. Rev. 29 (2324), 31673183. http://
seismic sequence at the Pollino Mountains, Italy. Seismol. Res. Lett. 84 (6), 955962. dx.doi.org/10.1016/j.quascirev.2010.07.003.
http://dx.doi.org/10.1785/0220120194. Ward, S.N., 1994. Constraints on the seismotectonics of the central Mediterranean from
Valensise, G., Pantosti, D., 2001. Database of potential sources for earthquakes larger than very long-base-line interferometry. Geophys. J. Int. 117, 441452. http://dx.doi.org/
M 5.5 in Italy. Ann. Geophys. 44, 1180. 10.1111/j.1365-246X.1994.tb03943.x.
Valoroso, L., Improta, L., Chiaraluce, L., Di Stefano, R., Ferranti, L., Govoni, A., Chiarabba, C., Wells, D.L., Coppersmith, K.J., 1994. New empirical relationship among magnitude,
2009. Active faults and induced seismicity in the Val d'Agri area (Southern rupture length, rupture width, rupture area, surface displacement. Bull. Seism. Soc.
Apennines, Italy). Geophys. J. Int. 178, 488502. http://dx.doi.org/10.1111/j.1365- Am. 84, 9741002.
246X.2009.04166.x. Wessel, P., Bercovici, D., 1998. Interpolation with splines in tension: a Green's
Van Dijk, J.P., Bello, M., Brancaleoni, G.P., Cantarella, G., Costa, V., Frixa, A., Golfetto, F., function approach. Math. Geol. 30, 7793. http://dx.doi.org/10.1023/A:
Merlini, S., Riva, M., Torricelli, S., Toscano, C., Zerilli, A., 2000. A regional structural 1021713421882.

You might also like