Mapping Moho depth variations in central Italy from PsMoho-P delay times: evidence of an E-W transition in the Adriatic Moho at 42 N latitude

Similar documents
New insights on some structural geometries in Southern Apennines by multiscale analysis of potential fields

DR

VHR seismic imaging of displacement along an active off-shore fault system of the Adriatic foreland

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Mantle wedge dynamics vs crustal seismicity in the Apennines (Italy)

The Buried Fold-and-Thrust Belt in Sicily: Perspectives for Future Exploration*

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth.

A free online copy of this book is available at Hard copies can be order from

APPLICATION OF RECEIVER FUNCTION TECHNIQUE TO WESTERN TURKEY

Crustal and uppermost mantle structure in Italy from the inversion of P-wave arrival times: geodynamic implications

Receiver Function (RF) Estimation Using Short Period Seismological Data

SEISMOTECTONIC ANALYSIS OF A COMPLEX FAULT SYSTEM IN ITALY: THE

Extended-time multi-taper frequency domain cross-correlation receiver function estimation

Seismicity and crustal structure in the Italian region: a new review using a synthesis of DSS results and updated catalogues of earthquakes

Tomographic imaging of P wave velocity structure beneath the region around Beijing

USU 1360 TECTONICS / PROCESSES

Dynamic Subsidence and Uplift of the Colorado Plateau. Supplementary Material

Monitoring long-term ground movements and Deep Seated Gravitational

NATURAL ENVIRONMENT. Geophysics

Northern Sicily, September 6, 2002 earthquake: investigation on peculiar macroseismic effects

Earth Science, (Tarbuck/Lutgens) Chapter 10: Mountain Building

Dynamic Crust Practice

Laboratory #7: Plate Tectonics

NAME HOMEWORK ASSIGNMENT #4 MATERIAL COVERS CHAPTERS 19, 20, 21, & 2

Ionian Sea and Margins: Recent Prospections and Interpretations

Ambient Noise Tomography in the Western US using Data from the EarthScope/USArray Transportable Array

GEOPHYSICAL RESEARCH LETTERS, VOL. 37, L02304, doi: /2009gl041835, 2010

University of Leeds 3GP Geophysics Field Trip Lake Balaton, Hungary

Aegeis deep structure & slab rollback in the Mediterranean

GEO-DEEP9300 Lithosphere and Asthenosphere: Composition and Evolution

Seismicity and P-wave velocity image of the Southern Tyrrhenian subduction zone

Copyright McGraw-Hill Education. All rights reserved. No reproduction or distribution without the prior written consent of McGraw-Hill Education

overlie the seismogenic zone offshore Costa Rica, making the margin particularly well suited for combined land and ocean geophysical studies (Figure

Some aspects of seismic tomography

Broadband converted phases from midmantle discontinuities

A) B) C) D) 4. Which diagram below best represents the pattern of magnetic orientation in the seafloor on the west (left) side of the ocean ridge?

SUPPLEMENTARY INFORMATION

Overview of the Seismic Source Characterization for the Palo Verde Nuclear Generating Station

09/06/2017. The orogenic cycle in the Andes: Arc magmatism and delamination as control of the fold and thrust belts

Palaeozoic oceanic crust preserved beneath the eastern Mediterranean

Hints of active deformation in the southern Adriatic foreland: Holocene tectonics along the Apulia offshore (Italy)

APPLICATION OF A PASSIVE TOMOGRAPHY METHOD AND CORRELATION WITH ACTIVE SEISMIC OBSERVATIONS IN THE KYPARISSIAKOS GULF, SOUTHWESTERN HELLENIC ARC

THE SEISMICITY OF THE CAMPANIAN PLAIN: PRELIMINARY RESULTS

Received May 11, 1989; Accepted July 10, * To whom correspondence should be addresseḍ

Data Repository Item

Data Repository Item For: Kinematics and geometry of active detachment faulting beneath the TAG hydrothermal field on the Mid-Atlantic Ridge

Non-ideal Subduction

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Rocky Mountain evolution: Tying Continental Dynamics of the Rocky Mountains and Deep Probe seismic experiments with receiver functions

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Earthquakes in Barcelonnette!

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

Karl Mueller QUATERNARY RIFT FLANK UPLIFT OF THE PENINSULAR RANGES IN BAJA AND SOUTHERN CALIFORNIA BY REMOVAL OF MANTLE LITHOSPHERE

Plate Tectonics. entirely rock both and rock

Topic 12: Dynamic Earth Pracatice

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Depositional Environments and Hydrocarbon Potential of Northern Ionian Sea

12. The diagram below shows the collision of an oceanic plate and a continental plate.

Whole Earth Structure and Plate Tectonics

Lateral variation of the D 00 discontinuity beneath the Cocos Plate

Dipartimento di Scienze Biologiche, Geologiche e Ambientali, Sezione di Scienze della Terra, Università di Catania, Italy 2

Wide-angle observations of ALP 2002 shots on the TRANSALP profile: Linking the two DSS projects

29. IMPLICATIONS OF DEEP SEA DRILLING, SITES 186 AND 187 ON ISLAND ARC STRUCTURE

The Earth s Structure from Travel Times

High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

Our Dynamic Earth Unit Unit 5

CONTINENTAL PLATE BOUNDARY ZONES

Topic 5: The Dynamic Crust (workbook p ) Evidence that Earth s crust has shifted and changed in both the past and the present is shown by:

Kinematic inversion of pre-existing faults by wastewater injection-related induced seismicity: the Val d Agri oil field case study (Italy)

Practice Questions: Plate Tectonics

Evolution of Continents Chapter 20

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

Africa-Arabia-Eurasia plate interactions and implications for the dynamics of Mediterranean subduction and Red Sea rifting

The Dynamic Crust 2) 4) Which diagram represents the most probable result of these forces? 1)

General Oceanography Geology 105 Expedition 10 - Whole Lotta Shakin' Goin' On

The continental lithosphere

GEOLOGY Vol. II - Mantle Dynamics and Plate Kinematics - Carlo Doglioni, Roberto Sabadini

SUPPLEMENTARY INFORMATION

Continental Margin Geology of Korea : Review and constraints on the opening of the East Sea (Japan Sea)

We A10 12 Common Reflection Angle Migration Revealing the Complex Deformation Structure beneath Forearc Basin in the Nankai Trough

Mountain Building. Mountain Building

High resolution seismic imaging of the M w 5.7, 2002 Molise, southern Italy, earthquake area: Evidence of deep fault reactivation

6. In the diagram below, letters A and B represent locations near the edge of a continent.

Editors. Volume 1C. D.G. Roberts* A.W. Bally*

Moho (Mohorovicic discontinuity) - boundary between crust and mantle

Unit Topics. Topic 1: Earth s Interior Topic 2: Continental Drift Topic 3: Crustal Activity Topic 4: Crustal Boundaries Topic 5: Earthquakes

Inversion of travel times to estimate Moho depth in Shillong Plateau and Kinematic implications through stress analysis of Northeastern India

Strike-Slip Faults. ! Fault motion is parallel to the strike of the fault.

Figure 1. Examples of vector displacement diagrams for two and three-plate systems.

SCIENCE CHINA Earth Sciences

Earth Movement and Resultant Landforms

The Lithosphere and the Tectonic System. The Structure of the Earth. Temperature 3000º ºC. Mantle

Global surface-wave tomography

The thermal and density structure of passive margins

Imaging Moho topography beneath the Alps by multdisciplinary seismic tomography

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

ALPS VS APENNINES. CARMINATI E., DOGLIONI C. Dipartimento di Scienze della Terra, Università La Sapienza, Roma Italy

Normal faults and thrusts re-activated by deep fluids: the 6 April 2009 Mw 6.3 L Aquila earthquake, central Italy.

Northern Tanzanian Earthquakes: Fault orientations, and depth distribution

Transcription:

1 1 2 Mapping Moho depth variations in central Italy from PsMoho-P delay times: evidence of an E-W transition in the Adriatic Moho at 42 N latitude 3 4 Giuliana Mele 1,*, Emiliano Di Luzio 2, Cristina Di Salvo 3 5 6 7 8 9 10 11 12 13 1 Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna Murata, 605, 00143 Roma, Italy; tel: +39 06 51860416; email: giuliana.mele@ingv.it 2 Consiglio Nazionale delle Ricerche, Area della Ricerca Roma 1 - Via Salaria km. 29,300-00016 Monterotondo stazione (Roma), Italy; tel. +39 06 90672722; email: emiliano.diluzio@itabc.cnr.it 3 Istituto di Geologia Ambientale e Geoingegneria, CNR, Area della Ricerca Roma 1 - Via Salaria km. 29,300-00016 Monterotondo stazione (Roma), Italy; tel. +39 06 90672740; email: cristina.disalvo@igag.cnr.it 14 15 16 17 18 19 20 21 22 23 24 25 * Corresponding author: G. Mele, Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna Murata, 605, 00143 Roma, Italy; tel: +39 06 51860416 (email: giuliana.mele@ingv.it) 26

2 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 Abstract Along the Italian peninsula adjoin two crustal domains, peri-tyrrhenian and Adriatic, whose boundary is not univocal in central Italy. In this area, we attempt to map the extent of the Moho in the two terrains from variations of the travel time difference between the direct P wave and the P-to-S wave converted at the crust-mantle boundary. We use teleseismic receiver functions computed at 43 broad-band stations in this and previous studies, and assign each of the recording sites to the Adriatic or peri-tyrrhenian terrains based on station location, geologic and geophysical data and interpretation, and consistency of delays with the regional Moho trend. The results of the present study show that the Ps Moho arrival time varies from 2.3 s to 4.1 s in the peri-tyrrhenian domain and from 3.7 to 5.5 s in the Adriatic domain. As expected, the lowest time difference is observed along the Tyrrhenian coastline and the largest values are observed in the axial zone of the Apennine chain. A key new result of this study is a sharp E-W boundary in the Adriatic domain that separates a deeper Moho north of about 42 N latitude from a shallower Moho to the south. This feature is constrained for a length of about 40 km by the observations available in this study. The E-W boundary requires a revision of prior mapping of the Moho in central Italy and supports previous hypotheses of lithosphere segmentation. 43 44 45 46 47 48 49 50 1. Introduction Peninsular Italy extends in the Mediterranean Sea from 38 to 46 latitude North and 8 to 18 longitude East. Its geologic setting is dominated by the Apennine chain that extends along the whole peninsula. This chain built up mostly during the Neogene and early Pleistocene following the deformation of the African continental margin of the Tethyan ocean [e.g., Malinverno and Ryan, 1986; Albarello et al., 1995; Vezzani et al., 2010]. In peninsular Italy, the topography of the Moho discontinuity, that is the object of this

3 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 study, has been investigated through active seismic profiles collected during the DSS experiments in the 1960's-1990's [Cassinis et al., 2003, and references therein] and the CROP Project in the 1980's-1990's [Scrocca et al., 2003, and references therein], and passive seismology methods such as tomography and teleseismic receiver functions [e.g., Piana Agostinetti et al., 2002; Mele and Sandvol, 2003; Mele et al., 2006; Di Luzio et al., 2009; Di Stefano et al., 2009; Piana Agostinetti and Amato, 2009]. Active and passive seismic data have been combined in Di Stefano et al. [2011]. The Moho map proposed by Cassinis et al. [2003] had the merit, unlike the majority of the maps derived from other studies, of distinguishing the crustal domains that characterize Italy and surrounding areas: continental crust in the European and African/Adriatic domains; oceanic/suboceanic crust in the Ligurian and Tyrrhenian Seas; transitional crust in the peri- Tyrrhenian side of peninsular Italy and northern Sicily. The boundary between the Adriatic and peri-tyrrhenian crusts runs along peninsular Italy and northern Sicily (Figure 1). Recently, Di Stefano et al. [2009, 2011] have proposed two boundaries that differ from each other and from that of Cassinis et al. [2003] in central Italy, as shown in Figure 1. In this area, where the three boundaries deviate one from the other and one of them is partially unconstrained, we attempt to reconstruct the extent of the Adriatic and peri-tyrrhenian crust. To map the Adriatic and peri-tyrrhenian Moho, we use the teleseismic receiver functions method that is based on the identification of the P wave converted to S at the Moho discontinuity (called Ps Moho in the following). The delay time of the Ps Moho with respect to the direct P arrival is affected primarily by Moho depth: the larger/smaller the delay, the deeper/shallower the Moho beneath the recording site; therefore, we interpret variations in the Ps Moho time in terms of variations of Moho depth. We integrate the new data with previous receiver functions computed by Mele et al. [2006] and Di Luzio et al. [2009]. The 43 recording stations used in central Italy are assigned to one or the other crustal

4 76 77 78 domain based on location, geologic and geophysical data, and consistency with the regional trend of the Moho. The Ps Moho -P times are interpolated with the Ordinary Kriging statistical method to map the extent and the lateral variations of the Adriatic and peri-tyrrhenian Moho. 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 2. Geologic setting In peninsular Italy, the peri-tyrrhenian area is characterized by a stretched transitional crust with positive Bouguer anomalies [e.g., Morelli, 1981], high heat flow [e.g., Della Vedova et al., 2001] and relatively low uppermost mantle velocities [e.g., Mele et al., 1998]. On the contrary, the Adriatic domain is a more stable area with low heat flow, low-tomoderate positive Bouguer anomalies and normal-to-high uppermost mantle velocities. Since Mele and Sandvol [2003], the Adriatic Moho was inferred to deepen to about 50 km beneath the Apennine chain. Central Italy is characterized by Meso-Cenozoic platform and basin units of the Apennine chain verging NE-ward above the Bradanic foredeep and the Adriatic/Apulian foreland (Figure 2). To the west, Plio-Quaternary marine-to-continental deposits and Pleistocene volcanics cover large sectors of the internal Apennines that were downthrown by extensional faults since the late Miocene [e.g., Patacca et al., 1990]. In the study area, the foreland sequence outcrops in the Gargano promontory and Tremiti Islands (Figure 2), mainly characterized by the carbonate units of the Apulian Platform (AP). Part of the Apulian Platform was involved in the Apennine deformation during the Pliocene- Early Pleistocene; it is exposed in the Maiella Massif and surroundings (Apennine external units of Figure 2) [Bally et al., 1986; Mostardini and Merlini, 1986; Cipollari and Cosentino, 1995; Patacca et al., 2008; Cosentino et al., 2010]. The Apulian Platform Top (APT) is a regional key-horizon distinctive of the Adriatic crust; it was used to follow the westward dipping of the foreland monocline beneath the

5 101 102 103 104 105 106 107 foredeep and the Apennines [Mariotti and Doglioni, 2000]. This horizon, made of Miocene limestones and/or evaporites, is reached at depths ranging from about 1 km in the peri- Adriatic region to about 3 km in the axial zone of the Apennines by the exploration wells plotted in Figure 2. In the CROP11 profile, a high-amplitude pair of reflectors interpreted as the APT horizon is followed from the Adriatic coast to the Fucino basin [Scrocca et al., 2003; Patacca et al., 2008]; west of the Fucino basin the CROP11 profile is not interpreted. In this work, the APT will be used to constrain the extent of the Adriatic crust. 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 3. Method of analysis, seismologic data, and observations Since the first observations in the 1950's [Cook et al., 1962, and references therein], teleseismic P waves converted to S at major velocity discontinuities of the Earth were used to infer the gross seismic structure under a recording station. The Ps Moho is often the highestenergy signal in the coda of the direct P arrival due to the large velocity contrast between the crust and the mantle, and is used to build regional Moho maps [e.g., Priestley et al., 1988; Kind et al., 1995; Jones and Phinney, 1998; Al-Damegh et al., 2005; Lloyd et al., 2010]. Data usable for these studies are three-component, possibly broad-band recordings of teleseismic events with epicentral distance of 30 to 90. The Ps Moho phase arrives few seconds after the direct P and most of the times it is hard to observe in the seismogram. The method used to identify the Ps Moho consists in deconvolving the vertical component of the ground motion from the horizontal component rotated into the radial direction (source-to-receiver path) where Ps conversions have the largest amplitude [Langston, 1979]. Deconvolution filters out most of the common features such as source, travel path effects, and instrumental response, producing a simpler time series called receiver function. This last is composed by the first positive P pulse followed by Ps conversions and reverberations. Deconvolution also enables to compare receiver functions from various

6 126 127 128 129 seismic sources that are stacked together to enhance the coherent signals. The time delay between Ps Moho and P (t Ps hereinafter) can be used to estimate the depth of the Moho (H) for given bulk crustal velocities Vp and Vs and P-wave incidence angle (expressed through the ray parameter p): 130 H = t Ps (1/Vs 2 p 2 ) (1/Vp 2 p 2 ) (1) 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 In this work, we have collected teleseisms with minimum magnitude Mw 5.5 recorded in the 2004-2009 period by 29 permanent stations of the Italian Seismic Network. The epicentral distance is computed from the center of the study area. Given the abundance of seismic sources in the distance range 80 ±10, we selected these events because steeper incidence angles yield larger energy of the incoming P wave. After a selection of the recordings in terms of the signal-to-noise ratio, we cut a window of 30 s from the seismograms of 148 events (Figure 3a), starting 5 s before the P onset. To compute receiver functions, we applied the time-domain deconvolution technique of Ligorria and Ammon [1999]; a Gaussian low-pass filter with width parameter α=2.0 was used to remove the high-frequency noise. In the receiver functions of 24 stations, a positive peak arriving 2.3 to 5.2 s after P was interpreted as the Ps wave converted at the Moho discontinuity; 5 stations were discarded due to noisy or inconsistent observations. We also used the t Ps computed by Mele et al. [2006] at the permanent station AQU and 12 temporary stations installed for few months in 1995 (0-4C, 6-9C, 11C, 12C, 14C), and by Di Luzio et al. [2009] at the permanent station FRES (Figure 3b). For most of these stations, only events from the north-east and 80 ±10 distance were available [see Figure 5 of Mele et al., 2006].

7 149 150 151 152 153 154 155 156 157 In the present study, most of the observations are naturally clustered between 330 and 100 backazimuth (Figure 3a) and this prevented to analyze the crustal response as a continuous function of azimuth. For this reason, and for consistency with previous works, we stacked the receiver functions of events occurred in the NE quadrant. This ensures also to sample the same Moho structure beneath each station. Depending on the working state and quality of the recording site, the number of receiver functions varies from 4 (4C, GUAR, CIGN) to 56 (INTR). In Figure 3b are shown the stacks of 5 stations arranged along a SW-NE profile that crosses the boundaries between the peri- Tyrrhenian and Adriatic crusts. 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 4. Mapping the peri-tyrrhenian and Adriatic Mohos In order to estimate Moho depth from the Ps Moho delay, a bulk crustal velocity must be provided for all stations. However, previous works propose conflicting models in the study area, especially at mid-crustal depth. As an example, we show in Figure 4 two seismic tomography sections where high-velocity anomalies are imaged on both sides of the Fucino basin [Chiarabba et al., 2010] and two crustal sections interpreted from active seismic data where low velocity is inferred in the same area [Cassinis et al., 2003; Patacca et al., 2008]. Because of the uncertainty in the regional velocity structure, in the present study we use the Ps Moho delays as indicative of Moho depth variations. The delay of the Moho conversion is read from the stack trace of each station and mapped in Figure 5a. Ps Moho delays span from 2.3 to 5.5 s, and the conversion points at the Moho occur NE of the stations, at an average distance of 10 km. In this map, we attributed each station (i.e. observation points of t Ps ) to the Adriatic or peri-tyrrhenian terrain based on location with respect to the proposed boundaries; where the boundaries deviate from each other, the attribution is based on surface and shallow geology (well logs) or on the consistency of t Ps with the Moho trend defined by the

8 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 Tyrrhenian stations 0-4C and the Adriatic stations 6-14C and FRES [Mele et al., 2006; Di Luzio et al., 2009]. Stations located west of the three boundaries are assigned to the peri-tyrrhenian crust (from north to south: MAON, LATE, CESX, MNS, TOLF, MTCE, ROM9, RDP, CERT, GUAR, GIUL), while stations located east of the boundaries are assigned to the Adriatic crust (TERO, CAMP, CAFR, LPEL, CIGN, SGRT, MSAG). Other stations can be attributed to the Adriatic crust based on the following aspects: i) MIDA and CERA, located close to two explorations wells that reached the APT horizon at about 3 km of depth, and to outcrops of the deformed Apulian domain (see area framed in Figure 2); ii) INTR, located along the segment of the CROP11 profile where the reflection package interpreted as the Apulian Platform Top is recognized beneath the Apennine units [Patacca et al., 2008]; iii) CAMP and FAGN, where the relatively large t Ps (5.0 and 5.2 s) is consistent with the westward deepening Adriatic Moho. The attribution of stations matches the boundaries of Cassinis et al. [2003] and Di Stefano et al. [2009], while it is inconsistent with the boundary proposed by Di Stefano et al. [2011] (Figure 5a). Stations FIAM, VVLD, and POFI are uncertain because their location is not constrained by geologic evidence and the t Ps matches the trend of the Moho in both crustal domains. Figure 5b displays a contouring of t Ps obtained with ArcGIS Geostatistical Wizard [ESRI, 2009]. We used the Ordinary Kriging prediction method [Matheron, 1970] to model the spatial trend of a single variable; to avoid a-priori bias, data were interpolated without using barrier polylines between the Mohos. The basic assumption, when using statistics to handle heterogeneity in Earth systems, is that properties are not random, but have some spatial continuity or are correlated over some distance. The Geostatistical Analyst Extension module of ArcGIS examines the distribution of the data to create a semivariogram model that allows

9 199 200 201 202 203 204 205 206 207 208 to compute the parameter value in unsampled locations. The Kriging model generates the predicted surface after selecting the best suitable model based on regression statistics. Observed vs simulated t Ps resulting from the cross-validation procedure are plotted in the inset of Figure 5b. In the map of Figure 5b, smaller differential times occur in the western sector of the peninsula, characterized by brown colors (t Ps between 3.3 and 3.8 s), matching the attribution of most of the 16 peri-tyrrhenian stations. As to the Adriatic stations, the contouring highlights two regions with different t Ps that define a sharp transition of the Moho surface along the 42 N latitude: t Ps changes from 4.6-4.7 s to the north to 3.7-3.8 s to the south. The receiver function stacks of the 5 stations straddling the Moho transition are shown in Figure 5b. 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 5. Discussion In central Italy, we have distinguished stations located in the peri-tyrrhenian and in the Adriatic terrains to reconstruct the variations of the Moho in these crustal domains. A key finding of this study is a sharp variation of t Ps in the Adriatic domain, at about 42 N latitude: from north to south, t Ps changes from 4.6-4.7 s at stations 9C, 11C and 12C to 3.7-3.8 s at stations INTR and LPEL, within a distance of 15 km (Figures 5a,b). At stations 9C, 11C, and 12C, Di Luzio et al. [2009] have estimated a Moho depth of 38±1 km using a local bulk crustal Vp of 6.3 km/s derived from the interpretation of the CROP 11 profile. This is a good crustal average commonly used in literature. Adopting such Vp value in equation (1), we estimate a Moho depth of 30 and 31 km beneath stations LPEL and INTR, respectively, i.e. the Adriatic Moho is ~ 8 km shallower south of the 42 N parallel. The E-W Moho transition can be constrained for about 40 km with the observations available for this study (Figure 5b). It is worth to underline that the Ps Moho delays of the Adriatic stations are consistent on either side of the Moho transition: 4.6 to 5.5 s are observed at all stations

10 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 located north of 9C-12C while 3.8 to 4.2 s are observed at all stations located south of INTR and LPEL (Figures 5a,b). The E-W step of the Adriatic Moho supports previous ideas of lithosphere segmentation in central Italy [Royden et al., 1987; Doglioni et al., 1994]. Royden et al. [1987] based their model on the morphology of the Apennine foredeep basin (correlated with Bouguer gravity anomalies) and of the outermost thrust of the chain; both show differential offsets from north to south reflecting a different amount of lithosphere retreat (Figure 5c). Doglioni et al. [1994] hypothesized that a differential lithosphere rollback occurs between the central Adriatic and the Puglia region, caused by the difference in the lithospheric thickness inherited from the Mesozoic rifting: the downgoing of the 40-km thicker Puglia lithosphere slowed down since the middle Pleistocene favouring the uplift of the foreland and the Moho in the Gargano promontory (Figure 5d). The present study results confirm that the Moho is shallower over the whole sector below 42 N latitude, not only beneath the Gargano promontory, and is rather flat: stations MSAG and SGRT show the same t Ps of the nearby stations, including the one located in the Tremiti islands where Mele et al. [2006] estimated a Moho depth of 33 km. The step of the Moho in central Italy is not displayed in the Moho map of Piana- Agostinetti and Amato [2009], obtained with the receiver functions stacking technique of Zhu and Kanamori [2000]. The reason could be that this map is a smoothed image of Moho depth variations with less than 1/3 high-quality stations (class 1-2 defined by the authors). Additionally, the temporary stations 0-14C and the permanent station LPEL, i.e., 4 of the 5 stations that constrained the E-W Moho step, are not used by these authors; this produces a low-resolution image of the Adriatic Moho around the 42 N parallel. It is worth noting that INTR, that is the only station shared by the two studies around the Moho step, has the same average Moho depth (Table 1).

11 248 249 250 251 252 253 The Moho depths estimated by Piana-Agostinetti and Amato [2009] are used by Di Stefano et al. [2011] to integrate active seismic data and reconstruct the Moho topography in Italy. In central Italy, the Tyrrhenian/Adriatic boundary of Di Stefano et al. [2011] is in contrast with the Ps Moho delays: several stations located west of this boundary have t Ps of 5.0 s and more (Figure 5a), corresponding to Moho depths larger than 40 km, that cannot be associated with the peri-tyrrhenian Moho. 254 255 256 257 258 259 260 261 262 263 264 6. Conclusions We have presented a revised mapping of the peri-tyrrhenian and Adriatic Moho in central Italy supplementing previous receiver function studies (14 stations) with results obtained from 24 additional stations. We have compared the cumulative receiver function results with constraints from well data and active source imaging to assign each station to either crustal domain. The new result of the present study is evidence for a sharp E-W transition in the Adriatic Moho that rises of ~ 8 km south of ~42 N parallel. This feature can be constrained for a length of ~ 40 km with the data available in this study. The E-W transition requires a major revision to prior mapping of crustal domains and supports previously hypothesized lithosphere segmentation. 265

12 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 References Albarello, D., E. Mantovani, D. Babbucci, and C. Tamburelli (1995), Africa Eurasia kinematics: main constraints and uncertainties, Tectonophysics, 243, 25 36. Al-Damegh, K., E. Sandvol, and M. Barazangi (2005), Crustal structure of the Arabian plate: New constraints from the analysis of teleseismic receiver functions, Earth Planet. Sci. Lett., 231, 177 196. Bally, A. W., L. Burbi, C. Cooper, and R. Ghelardoni (1986), Balanced sections and seismic reflection profiles across the central Apennines, Mem. Soc. Geol. Ital., 35, 257 310. Cassinis, R., S. Scarascia, and A. Lozej (2003), The deep crustal structure of Italy and surroundings areas from seismic refraction data. A new synthesis, Boll. Soc. Geol. It., 122, 365 376. Chiarabba, C., S. Bagh, I. Bianchi, P. De Gori, and M. Barchi (2010), Deep structural heterogeneities and the tectonic evolution of the Abruzzi region (Central Apennines, Italy) revealed by microseismicity, seismic tomography, and teleseismic receiver functions, Earth Planet. Sci. Lett., 295, 462 476. Cipollari, P., and D. Cosentino (1995), Miocene unconformities in the central Apennines: Geodynamic significance and sedimentary basin evolution, Tectonophysics, 252, 375 389. Cook, K. L., S. T. Algermissen, and J. L. Costain (1962), The status of Ps converted waves in crustal studies, J. Geophys. Res., 67, 4769 4778. Cosentino, D., P. Cipollari, P. Marsili, and D. Scrocca (2010), Geology of the central Apennines: a regional review. In The Geology of Italy: tectonics and life along plate margins, edited by M. Beltrando, A. Peccerillo, M. Mattei, S. Conticelli, and C. Doglioni, Journal of the Virtual Explorer, Electronic Edition, ISSN 1441-8142, vol. 6, paper 12, doi:10.3809/jvirtex.2010.00223. Della Vedova, B., S. Bellani, G. Pellis, and P. Squarci (2001), Deep temperatures and surface heat flow distribution. In: G.B. Vai (eds), Anatomy of an Orogen: The Apennines and adjacent Mediterranean basins, Kluwer academic publishers, 65-76. Di Luzio, E., G. Mele, M. M. Tiberti, G. P. Cavinato, and M. Parotto (2009), Moho deepening and shallow upper crustal delamination beneath the central Apennines, Earth Planet. Sci. Lett., 280(3 4), 1 12. Di Stefano, R., E. Kissling, C. Chiarabba, A. Amato, and D. Giardini (2009), Shallow subduction beneath Italy: Three-dimensional images of the Adriatic European- Tyrrhenian lithosphere system based on high- quality P wave arrival times, J. Geophys. Res., 114, B05305, doi:10.1029/2008jb005641.

13 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 Di Stefano, R., I. Bianchi, M. G. Ciaccio, G. Carrara, and E. Kissling (2011), Three dimensional Moho topography in Italy: New constraints from receiver functions and controlled source seismology, Geochem. Geophys. Geosyst., 12, Q09006, doi:10.1029/2011gc003649. Doglioni, C., F. Mongelli, and P. Pieri (1994), The Puglia uplift (SE Italy): an anomaly in the foreland of the Appenine subduction due to buckling of a thick continental lithosphere, Tectonics, 13, 1309 1321. ESRI (Environmental Systems Research Institute) (2009), ArcMap 9.3.1 Redlands, CA, USA. Jones, C. H., and R. A. Phinney (1998), Seismic structure of the lithosphere from teleseismic converted arrivals observed at small arrays in the southern Sierra Nevada and vicinity, California, J. Geophys. Res., 103, 10065 10090. Kind, R., G. L. Kosarev, and N. V. Petersen (1995), Receiver functions at the stations of the German Regional Seismic Network (GRSN), Geophys. J. Int., 121, 191 202, doi:10.1111/j.1365-246x.1995.tb03520.x. Langston, C. A. (1979), Structure under Mount Rainier, Washington, inferred from teleseismic body waves, J. Geophys. Res., 84 (B9), 4749 4762. Ligorria, J. P., and C. J. Ammon (1999), Iterative deconvolution and receiver function estimation, Bull. Seismol. Soc. Am., 89, 1395 1400. Lloyd, S., S. van der Lee, G. S. França, M. Assumpção, and M. Feng (2010), Moho map of South America from receiver functions and surface waves, J. Geophys. Res., 115, B11315, doi:10.1029/2009jb006829. Malinverno, A., and W. B. F. Ryan (1986), Extension in the Tyrrhenian Sea and shortening in the Apennines as a result of arc migration driven by sinking of the lithosphere, Tectonics, 5, 227 245. Matheron, G (1970), The Theory of Regionalized Variables and its Applications, Les Cahiers du Centre de Morphologie mathématique, Fascicule V, Ecole des Mines de Paris, 211 pp. Mariotti, G., and C. Doglioni (2000), The dip of the foreland monocline in the Alps and Apennines, Earth Planet. Sci. Lett., 181(1-2), 191-202, doi: http://dx.doi.org/10.1016/s0012-821x(00)00192-8. Mele G., A. Rovelli, D. Seber, T.M. Hearn, and M. Barazangi (1998), Compressional velocity structure and anisotropy in the uppermost mantle beneath Italy and surrounding regions, J. Geophys. Res., 103(B6), 12,529-12,543. Mele, G., and E. Sandvol (2003), Deep crustal roots beneath the northern Apennines inferred from teleseismic receiver functions, Earth Planet. Sci. Lett., 211, 69 78.

14 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 Mele, G., E. Sandvol, and G. P. Cavinato (2006), Evidence of crustal thickening beneath the central Apennines (Italy) from teleseismic receiver functions, Earth Planet. Sci. Lett., 249(3-4), 425 435. Morelli, C. (1981), Gravity anomalies and crustal structures connected with the Mediterranean margins. In: Sedimentary Basins of Mediterranean Margins (F.C. Wezel, ed.), pp. 33 53. C.N.R. Italian Project of Oceanography, Tecnoprint, Bologna. Mostardini, F. and S. Merlini (1986), Appennino centro meridionale. Sezioni geologiche e proposta di modello strutturale, Mem. Soc. Geol. It., 35, 177 202. Patacca, E., R. Sartori, and P. Scandone (1990), Tyrrhenian basin and Apenninic arc: Kinematic relations since Late Tortonian times, Mem. Soc. Geol. Ital., 45, 425 451. Patacca, E., P. Scandone, E. Di Luzio, G. P. Cavinato, and M. Parotto (2008), Structural architecture of the central Apennines: interpretation of the CROP 11 seismic profile from the Adriatic coast to the orographic divide, Tectonics, 27, TC3006, doi:10.1029/2005tc001917. Piana Agostinetti, N., F. P. Lucente, G. Selvaggi, and M. Di Bona (2002), Crustal structure and Moho geometry beneath the Northern Apennines (Italy), Geophys. Res. Lett., 29(20), doi:10.1029/2002gl015109. Piana Agostinetti, N., and A. Amato (2009), Moho depth and Vp/Vs ratio in peninsular Italy from teleseismic receiver functions, J. Geophys. Res., 114, B06303, doi:10.1029/2008jb005899. Priestley, K., G. Zandt, and G. Randall (1988), Crustal structure in eastern Kazakh, U.S.S.R. from teleseismic receiver functions, Geophys. Res. Lett., 15, 613 616. Royden L., E. Patacca, and P. Scandone (1987), Segmentation and configuration of subducted lithosphere in Italy: an important control on thrust-belt and foredeep-basin evolution, Geology, 15, 714-717. Scrocca, D., C. Doglioni, F. Innocenti, P. Manetti, A. Mazzotti, L. Bertelli, L. Burbi, and S. D Offizi (Eds.) (2003), CROP Atlas: seismic reflection profiles of the Italian crust, Mem. Descr. Carta Geol. Ital., 62, pp. 194. Vezzani, L., A. Festa, and F. C. Ghisetti (2010), Geology and tectonic evolution of the Central-Southern Apennines, Italy, Geol. Soc. of America Special Paper, 469, 1 58. ViDEPI Project, Visibility of petroleum exploration data in Italy, http://unmig.sviluppoeconomico.gov.it/videpi/en/pozzi/pozzi.asp. Zhu, L., and H. Kanamori (2000), Moho depth variation in southern California from teleseismic receiver functions, J. Geophys. Res., 105 (B2), 2969 2980.

15 367 368 369 370 371 Figure Captions Figure 1. Moho isobaths and crustal domains of Italy and adjacent areas after Cassinis et al. [2003]. The boundaries between Tyrrhenian and Adriatic plates at Moho depth proposed by Di Stefano et al. [2009] and [2011] are superimposed for comparison. White segments along the boundary of Di Stefano et al. [2009] are poorly constrained. 372 373 374 375 376 377 Figure 2. Geologic sketch of central Italy (Fb=Fucino basin; Mm: Maiella massif). The exploration wells that drilled the Apulian Platform Top (APT) and the trace of the CROP11 deep reflection profile are shown. The red square indicates the most internal part of the Apennine chain where the APT, distinctive of the Adriatic crust, is drilled [ViDEPI Project: http://unmig.sviluppoeconomico.gov.it/videpi/]. 378 379 380 381 382 383 384 385 386 387 Figure 3. a) Azimuthal projection of the events used in the present study, centered in the study area. b) Topography map of central Italy showing the seismic stations used in this (29) and previous (14) studies; 5 stations were discarded because no clear identification of the Moho conversion could be made. The boundaries between Adriatic and peri-tyrrhenian Moho proposed by Cassinis et al. [2003] (CA03) and by Di Stefano et al. [2009, 2011] (DS09, DS11) are also shown. Receiver function stacks of 5 stations projected along the profile A-A' and the position of the Adriatic/peri-Tyrrhenian boundaries are shown in the upper panel. In the receiver functions, arrows mark the P onset (time=0) and the Ps Moho phase; n indicates the number of events used in the stack. 388 389 390 391 Figura 4. Upper panel: traces of active and passive seismic profiles in the study area. Lower panel: (left) Vp models obtained by Chiarabba et al. [2010] combining local earthquakes tomography and teleseismic receiver functions and (right) interpreted crustal sections along

16 392 393 394 the "Latina-Pescara" DSS profile [after Cassinis et al., 2003] and the CROP 11 profile [simplified after Patacca et al., 2008]. Profiles 2-2 and CROP11 are parallel, such as the profiles 6-6 and DSS. 395 396 397 398 399 400 401 402 403 404 405 406 407 Figure 5. a) Delay times of Ps waves converted from the Moho discontinuity beneath 43 stations. The recording sites are tentatively assigned to the Adriatic or peri-tyrrhenian crust. Ps Moho delays range from 2.3 s along the Tyrrhenian coastline to 5.5 s in the Apennine region. b) Contouring of Ps Moho delays interpolated with the Ordinary Kriging prediction method; the range of delays is divided into contour intervals assigned to different colors. From the seismic sources used in this study the Moho conversion occurs at 10±5 km from the station, depending on crustal thickness. The red segment indicates the offset of the Moho and the minimum extent that can be constrained with the data presented in this study; the receiver function stacks of the 5 closest stations are also shown. In the inset are plotted the observed vs simulated t Ps resulting from the cross-validation of the predictive model (root mean square is 0.325 s, mean error is 0.024 s, average standard error is 0.409 s). c) Sketch of lithosphere segmentation after Royden et al. [1987] and d) Doglioni et al. [1994]. 408 409 410 411 412 413 414 415 Table 1. Seismic stations used in this study listed in alphabetical order with Ps Moho time delays (t Ps ), assigned crustal domain (AD: Adriatic; TR: peri-tyrrhenian), and Moho depths computed in this study ( + ), Mele et al. [2006] ( x ), and Di Luzio et al. [2009] ( xx ). In the last column are listed for comparison the Moho depths of Piana-Agostinetti and Amato [2009] (PA-A 2009); in parenthesis is the quality class of each station defined by these authors, decreasing from 1 to 5. The Adriatic stations located within 50 km from INTR are highlighted in boldface. 416